Water resource assessment for the Southern Gulf catchments Australia’s National Science Agency A report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid Editors: Ian Watson, Caroline Bruce, Seonaid Philip, Cuan Petheram and Chris Chilcott ISBN 978-1-4863-2081-3 (print) ISBN 978-1-4863-2082-0 (online) Citation Watson I, Bruce C, Philip S, Petheram C and Chilcott C (eds) (2024) Water resource assessment for the Southern Gulf catchments. A report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Chapters should be cited in the format of the following example: Philip S, Watson I, Petheram C and Bruce C (2024) Chapter 1: Preamble. In: Watson I, Bruce C, Philip S, Petheram C, and Chilcott C (eds) (2024) Water resource assessment for the Southern Gulf catchments. A report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Copyright © Commonwealth Scientific and Industrial Research Organisation 2024. 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 Southern Gulf 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 (NT) and Queensland governments. The Assessment was guided by two committees: i. The Governance Committee: CRC for Northern Australia/James Cook University; CSIRO; National Water Grid (Department of Climate Change, Energy, the Environment and Water); Northern Land Council; NT Department of Environment, Parks and Water Security; NT Department of Industry, Tourism and Trade; Office of Northern Australia; Queensland Department of Agriculture and Fisheries; Queensland Department of Regional Development, Manufacturing and Water ii. The Southern Gulf catchments Steering Committee: Amateur Fishermen’s Association of the NT; Austral Fisheries; Burketown Shire; Carpentaria Land Council Aboriginal Corporation; Health and Wellbeing Queensland; National Water Grid (Department of Climate Change, Energy, the Environment and Water); Northern Prawn Fisheries; Queensland Department of Agriculture and Fisheries; NT Department of Environment, Parks and Water Security; NT Department of Industry, Tourism and Trade; Office of Northern Australia; Queensland Department of Regional Development, Manufacturing and Water; Southern Gulf NRM 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 their release. This report was reviewed by Mr Mike Grundy (Independent consultant). Individual chapters were reviewed by Dr Peter Wilson, CSIRO (Chapter 2); Dr Andrew Hoskins, CSIRO (Chapter 3); Dr Brendan Malone, CSIRO (Chapter 4); Dr James Bennett, CSIRO (Chapter 5); Dr Nikki Dumbrell, CSIRO (Chapter 6); Mr Darran King, CSIRO (Chapter 7). The material in this report draws largely from the companion technical reports, which were themselves internally and externally reviewed. For further acknowledgements, see page xxviii. 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 Saltwater Arm, a tributary of the Albert River. This view typifies the tidal rivers and estuaries along the southern coast of the Gulf of Carpentaria. Source: Shutterstock Director’s foreword Sustainable development and regional economic prosperity are priorities for the Australian, Queensland and Northern Territory (NT) governments. However, more comprehensive information on land and water resources across northern Australia is required to complement local information held by Indigenous Peoples and other landholders. Knowledge of the scale, nature, location and distribution of likely environmental, social, cultural and economic opportunities and the risks of any proposed developments is critical to sustainable development. Especially where resource use is contested, this knowledge informs the consultation and planning that underpin the resource security required to unlock investment, while at the same time protecting the environment and cultural values. In 2021, the Australian Government commissioned CSIRO to complete the Southern Gulf Water Resource Assessment. In response, CSIRO accessed expertise and collaborations from across Australia to generate data and provide insight to support consideration of the use of land and water resources in the Southern Gulf catchments. The Assessment focuses mainly on the potential for agricultural development, and the opportunities and constraints that development could experience. It also considers climate change impacts and a range of future development pathways without being prescriptive of what they might be. The detailed information provided on land and water resources, their potential uses and the consequences of those uses are carefully designed to be relevant to a wide range of regional-scale planning considerations by Indigenous Peoples, landholders, citizens, investors, local government, and the Australian, Queensland and NT governments. By fostering shared understanding of the opportunities and the risks among this wide array of stakeholders and decision makers, better informed conversations about future options will be possible. Importantly, the Assessment does not recommend one development over another, nor assume any particular development pathway, nor even assume that water resource development will occur. It provides a range of possibilities and the information required to interpret them (including risks that may attend any opportunities), consistent with regional values and aspirations. All data and reports produced by the Assessment will be publicly available. Chris Chilcott C:\Users\bru119\AppData\Local\Microsoft\Windows\Temporary Internet Files\Content.Word\C_Chilcott_high.jpg Project Director Key findings for the Southern Gulf catchments The Southern Gulf catchments have an area of 108,200 km2 across the Northern Territory (NT) and Queensland. They comprise the catchments of Settlement Creek (17,600 km2), Nicholson River (52,200 km2), Leichhardt River (33,400 km2) and, Morning Inlet (3,700 km2) and the Wellesley island groups (1,200 km2). The rivers of these catchments flow into the lower Gulf of Carpentaria, an important part of northern Australia’s marine environment with high ecological, cultural and economic values. Within the study area, 12% of the land is Aboriginal freehold tenure of which 33% is national park, including the UNESCO World Heritage-listed Australian Fossil Mammal Sites (Riversleigh, Queensland). The dominant land use by area is extensive grazing of beef cattle on native rangelands, constituting 77% of the study area. Mining occupies less than 0.05% of the land area and irrigated agriculture, occupies about 0.01% (~1,400 ha). The population of the catchments in 2021 was approximately 22,500 people, of whom about 27% were Indigenous Australians. In comparison, Indigenous Australians make up 25% of the population of the NT, 4% of the population of Queensland and 3% of Australia as a whole. There is one significant urban area, Mount Isa (population 18,000) and Doomadgee (population 1,387) is the only other settlement with a population greater than 1,000. Distinct demographic and socio- economic differences exist within the Southern Gulf catchments based on the differences between Mount Isa and the remainder of the study area. The population density outside Mount Isa is one of the lowest in Australia, and some communities in the study area are ranked as being among the most disadvantaged in Australia. Indigenous Peoples have continuously occupied and managed the Southern Gulf catchments for tens of thousands of years. They retain significant and growing rights and interests in land and water resources, including crucial roles in water and development planning and as co-investors in future development. Indigenous Peoples remain closely tied to their territory associations. The key language groups are Garawa, Gangalidda, Waanyi, Kukatj, Lardil, Yangkaal, Kaiadilt, Wakabunga, Nguburinji, Kalkadoon, Mitakoodi, Mayi-Kutuna, Mayi‑Thakurti, Mayi‑Yapi and Mayi-Yali. A number of related groups and subgroups occur within these regional ownership descriptors. The creeks and rivers of the Southern Gulf catchments contribute about 6% of the mean annual discharge into the Gulf of Carpentaria, supporting many important ecological assets, some of national significance, and existing industries such as the Northern Prawn Fishery. This fishery had a prawn catch across whole of northern Australia valued at $85 million in 2019–20. While the Leichhardt River is deeply incised over much of its length, high-flow distributary channels have formed where the Nicholson and Gregory rivers flow across the extensive Armraynald Plain. The Nicholson and Leichhardt rivers are the two largest rivers in the Southern Gulf catchments and have median annual discharges of 1,873 and 1,211 GL, respectively. The Settlement Creek catchment encompasses numerous relatively short rivers, including Running, Settlement, Lagoon, Eight Mile, Cliffdale and Moonlight creeks. The median annual discharge from the creeks within the Settlement Creek catchment is 1,304 GL. The Gregory River, which flows into the Nicholson ii | Water resource assessment for the Southern Gulf catchments River approximately 85 km from the coast, is the largest perennial river in semi‑arid Queensland. The Gregory River and its major tributary, Lawn Hill Creek, receive groundwater discharge throughout the dry season from the Cambrian Limestone Aquifer (CLA), which is comprised mostly of the Camooweal Dolostone and Thorntonia Limestone in the south of the study area. The concentration of mining and industrial activity around Mount Isa has resulted in sufficiently high water demand from high-value industries for the construction of five large, purpose-built reservoirs of at least 10 GL capacity in the Leichhardt catchment, including the Leichhardt Dam (Lake Moondarra) and Julius Dam (Lake Julius). Consequently, as a proportion of its median annual streamflow, the water in the Leichhardt River is more heavily regulated than most other rivers in northern Australia west of the Great Dividing Range. As a result of existing entitlements (105.7 GL/year) and licence conditions in the Leichhardt catchment, only limited quantities of water could be reliably available for irrigated agriculture from Julius Dam without affecting the reliability of water supply to existing water entitlement holders. The study area has numerous small dams (<10 GL) providing water to mines as well as farm-scale dams, several for irrigation, but the only other instream water structure of note is a low weir (<1 GL capacity) on the Nicholson River downstream of Doomadgee, which is used to supply water for the town. With irrigation, the Southern Gulf catchments have a climate that is suitable for a wide range of annual and perennial horticulture, and broadacre crops and forages. However, the opportunities and risks of irrigation development adjacent to each of the major rivers are starkly different. Adjacent to the east bank of the Gregory River downstream and slightly upstream of the township of Gregory, the grey cracking clay soils of the treeless Armraynald Plain are highly favourable for large-scale irrigation developments (based on instream dams) or small-scale irrigation developments (based on water harvesting). The large contiguous areas of soil suitable for broadacre irrigation and ringtanks (1.03 million ha) gently slope away from the Gregory River, enabling water to be efficiently distributed using gravity and natural distributary channels running away from the river. The most cost-effective potential site for a dam in the study area is located approximately 70 km upstream of the township of Gregory, but that potential dam’s reservoir would be located within a part of the Gregory River that is a nationally significant wetland, the Thorntonia Aggregation. Being one of the few perennial rivers in the Southern Gulf and neighbouring Flinders catchment region, the Gregory River is particularly highly valued by local residents. On the Nicholson River downstream of Doomadgee are contiguous areas of red sandy and loamy soils (totalling 23,000 ha) suitable for irrigated vegetables. However, in the absence of a dam upstream, on-farm storage of water for irrigated annual horticulture would be challenging due to the difficulty of constructing ringtanks on these sandy soils. These soils are unlikely to be suitable for perennial horticulture due to the risk of flooding. The Leichhardt catchment has opportunities for irrigated horticulture on the friable soils (103,000 ha) on the narrow levees of the Leichhardt River downstream of Kajabbi. Adjacent to these soils, and up to 1 km from the river, are friable clay soils suitable for broadacre irrigation. While these heavier soils are suitable for constructing ringtanks, conveying water pumped from the Leichhardt River across the sandy levee soils to ringtanks may be costly and inefficient in some locations. The proximity of the eastern side of the mid-reaches of the Leichhardt River to the Burke Developmental Road and Mount Isa would be an advantage to new irrigation developments compared with more remote parts of northern Australia. For example, the more remote Nicholson and Settlement Creek catchments are often inaccessible for long periods due to flooding during the wet season. It is physically possible that water harvesting could extract 150 GL of water in 75% of years, spread across the Nicholson and Leichhardt catchments, without affecting the reliability of supply of existing water entitlement holders. This extraction would reduce the median and mean annual discharge to the Gulf of Carpentaria from these catchments by 5% and 3%, respectively. It would be sufficient water to irrigate about 12,000 ha of mixed broadacre crops. A hypothetical total of 10,000 ha of irrigated mixed broadacre (85%) and horticulture (15%) agriculture in the Southern Gulf catchments, would result in total annual increased economic activity (direct and indirect) of potentially as much as $190 million and support up to 360 full-time- equivalent jobs. In reality, the nature and scale of future surface water development would depend heavily upon community and government values, acceptance of potential impacts to water-dependent ecosystems, and the likely beneficiaries. Economic data from the NT and north-western Queensland indicate that, like other industries such as mining, grazing and fishing, benefits arising from agriculture developments elsewhere in the NT and Queensland have been heavily skewed to non-Indigenous households relative to Indigenous households. The most promising groundwater resources in the Southern Gulf catchments are the regional- scale CLA and Gilbert River Aquifer (GRA). The CLA coincides with large, contiguous areas of cracking clay soils suitable for irrigated broadacre cropping (308,000 ha) along the south-western margin of the study area. However, groundwater discharge from the CLA provides the dry-season flow in Lawn Hill Creek and the Gregory and O’Shannassy rivers and some of their tributaries. The major rivers of the Southern Gulf catchments are also characterised by relatively extensive alluvial aquifers, up to 5 km either side of the river. These systems are poorly characterised but likely to be highly heterogeneous and connected to the rivers, meaning the alluvial aquifers could be recharged by and discharge into ‘connected’ rivers at different times of the year. In addition to the 3.5 GL of existing groundwater entitlements in the study area, it is physically possible that groundwater systems across the Southern Gulf catchments could supply up to approximately 30 GL of water per year for irrigated agriculture, sufficient to irrigate about 4,000 ha of mixed broadacre cropping. Up to 10 GL of additional groundwater per year could potentially be extracted for other uses from areas of hard rock (i.e. poor soil) and poor water quality. The actual quantity of water extracted, however, would depend on community and government acceptance of impacts to groundwater-dependent ecosystems (GDEs) and existing groundwater users. Irrigated agriculture and aquaculture in the Southern Gulf catchments is only likely to be financially viable where there is an alignment of good prices for high‑value crops and market advantages. This makes achieving scale challenging. Other factors include availability of suitable markets for the products, investment in fundamental infrastructure such as all-weather roads and bridges and land tenure arrangements. New agricultural developments in the study area, if they were to eventuate, are likely to start irrigating broadacre crops on heavier clay soils before progressing to higher‑value and higher-input enterprises, such as horticulture on sandy and loamy soils, as farmers build confidence in their skills and expertise in this largely greenfield region. Landholders could potentially transition from broadacre cropping to annual horticulture on the heavy clay soils and nearby sandy levees adjacent to the Leichhardt River. Along the lower coastal reaches, about 194,000 ha of land is suitable for prawn and barramundi aquaculture, using earthen ponds. For all of the area estimates noted above, the land is considered suitable but with minor to considerable limitations and would require careful soil management. Growing irrigated forages or hay to feed young cattle to improve market access or to enable sale at younger ages is unlikely to be financially viable. Irrigation increases beef production but requires high capital outlay, and gross margins would be reasonably similar to, or less than, baseline cattle operations. Consistent rainfed cropping in the catchment is likely to be opportunistic (only possible in suitable years) and depend upon farmers’ appetite for risk and future local demand. How land cultivated for rainfed agriculture is managed in the years when rainfall is insufficient for rainfed cropping will be crucial for sustainable farming operations and the industry’s social licence to operate. Changes to groundwater baseflow and streamflow under projected drier future climates are likely to be considerably greater than changes that would result from plausible groundwater and surface water developments. Of the global climate models examined, 44% projected a drier future climate over the Southern Gulf catchments and 40% projected ‘little change’. Adopting a conservatively dry global climate model projection of a 13% reduction in long‑term mean annual rainfall and a 9% increase in potential evaporation, the modelled reduction in median annual discharge to the Gulf of Carpentaria from the Nicholson and Leichhardt catchments was 30%. The Southern Gulf catchments, although not pristine, have many unique characteristics and valuable ecological assets, which support existing industries such as grazing, tourism and commercial fishing. Whether based on groundwater or offstream storage, irrigated agricultural development has a wide range of potential benefits and risks that differentially intersect diverse stakeholder views on ecology, economy and culture. The detailed reports upon which this summary is based provide information that can be used to help consider the trade-offs from potential developments. Leichhardt River flowing through the Isa Highlands Photo: CSIRO – Nathan Dyer Overview of the Southern Gulf catchments The Southern Gulf catchments comprise four river basins defined by the Australian Water Resource Council, referred to here as catchments, that discharge into the Gulf of Carpentaria: Settlement Creek, Nicholson, Leichhardt and Morning Inlet. The study area also encompasses the Wellesley islands groups. The Southern Gulf catchments have a highly variable climate Northern Australia’s tropical climate is notable for the extremely high variability of rainfall between seasons and especially between years. This has major implications for evaluating and managing risks to development, infrastructure and industry. The climate of the Southern Gulf catchments is hot and semi-arid to dry subhumid. Generally, these catchments are water-limited environments, so efficient and effective methods for capturing, storing and using water are critical. •The mean and median annual rainfall – averaged across the Southern Gulf catchments – are 602 mm and 554 mm, respectively. A strong rainfall gradient runs from Mornington Island in the north (1,150 mm annual median) to the south of Mount Isa (400 mm annual median). •Averaged across the catchments, 6% of the rainfall occurs in the dry season (May to October). Median annual dry‑season rainfall ranges from 15 mm in the south-east to 60 mm along the western boundary. •Annual rainfall totals in the Southern Gulf catchments are highly variable. Annual totals are approximately 1.3 times more variable than in comparable parts of the world. The seasonality of rainfall presents opportunities and challenges for both wet- and dry-season cropping. •Information about water availability (i.e. soil water and water in storages) helps minimise risk when it is known ahead of important agricultural decisions – before planting time for most dry-season planted crops. Such information allows farmers to manage risk by choosing crops that optimise use of the available water or by deciding to forego cropping for a season. Rainfall is difficult to store. •Mean annual potential evaporation is higher than rainfall, averaging about 1,850 mm over mostof the study area, with the highest values near the mouth of the Leichhardt River and MorningInlet. •Large farm-scale ringtanks lose about 30% to 50% of their water to evaporation and seepagebetween April and October. Deeper farm-scale gully dams lose about 20% to 40% of their waterover the same period. Using stored water early in the season is the most effective way to reducethese losses. The more promising agricultural land in the Southern Gulf catchments is in the middle to lower reaches of the rivers, which are the parts of the study area most susceptible to destructive cyclonic winds. •Of the 53 consecutive cyclone seasons prior to 2021–22, the Southern Gulf catchments had notropical cyclones in 60% of those seasons, one cyclone in 36% of seasons and two cyclones in 4% of seasons. More than twice as many global climate models project a drier future climate than a wetter future climate for the Southern Gulf catchments, thus it is prudent to plan for water scarcity. •For the Southern Gulf catchments, 44% of climate models project a drier future, 16% project awetter future and 40% project a future within ±5% of the historical mean, indicating ‘littlechange’. Other studies indicate that tropical cyclones will be fewer but more intense in thefuture, although uncertainties remain. •Palaeoclimate records indicate past climates have been both wetter and drier over the pastseveral thousand years. •Climate and hydrology data that support short- to medium-term water resource planning shouldcapture the full range of likely or plausible conditions and variability at different timescales, andparticularly for periods when water is scarce. These are the periods that most affect industryand the environment. When industries ‘compete’ with the environment for water when it ismost scarce, it can have profound long-term consequences to both if the scarcity is poorlyunderstood and not planned for. •Future changes in temperature, vapour pressure deficit, solar radiation, wind speed and carbondioxide concentrations will separately act to increase or decrease crop water demand and cropyield under irrigation in northern Australia. For example, during the cooler months of the dryseason, cotton in the more southerly parts of the Southern Gulf catchments (e.g. BarklyTableland) may be limited by minimum temperatures at key stages of cotton boll development. Higher temperatures may increase the irrigation water requirement, depending upon whetherother climate variables such as wind speed increase or decrease in the future. However, forsome crops, higher crop water demand, and hence smaller planted area, could be offset byhigher yields and lower risk of crop failure. •Changes under future climates to the amount of irrigation water required and crop yield arelikely to be modest compared to improvements arising from new crop varieties and technologyover the next 40 years. Historically, these types of improvements have been difficult to predict, but they are potentially large. The Southern Gulf catchments are comprised of many different types of rivers and creeks The rivers and creeks of the Southern Gulf catchments constitute about 6% of the mean annual flow into the Gulf of Carpentaria, an important part of northern Australia’s marine environment with high ecological and economic values. •The mean and median annual discharges from the Southern Gulf catchments into the Gulf ofCarpentaria are 6,759 and 4,961 GL, respectively. •The Nicholson and Leichhardt rivers are the two largest rivers in the Southern Gulf catchments. The Nicholson River has the 13th largest median annual discharge (1,873 GL) of all rivers thatflow into the Gulf of Carpentaria, and the Leichhardt River has the 16th largest (1,211 GL). •The Settlement Creek catchment encompasses numerous relatively short creeks and its medianannual discharge to the Gulf of Carpentaria is 1,304 GL. The median annual discharges from theMorning Inlet catchments and the Wellesley Island groups are 195 and 173 GL, respectively. •Annual variability in streamflow in the Southern Gulf catchments is comparable with other riversin northern Australia with similar mean annual runoff, but the annual variability in runoff is twoto three times greater than rivers in the rest of the world in similar climates. Broad-scale flooding occurs along the lower reaches of the rivers of the Southern Gulf catchments, particularly adjacent to the distributary channels from the Nicholson and Gregory rivers to the Albert River. •Between 1980 and 2023 (43 years), 51 streamflow events broke the banks of the Gregory Riverat Riversleigh Road crossing. All events occurred between November and April (inclusive), andabout 84% of events occurred between January and March (inclusive). Under a potential dry future climate (13% reduction in rainfall), median annual discharges from the Nicholson and Leichhardt rivers are projected to decrease by approximately 30%. The Leichhardt catchment is one of the most regulated catchments in northern Australia, west of the Great Dividing Range, based on extractions as a proportion of its median annual flow. •Current licensed surface water extractions in the Leichhardt catchment are about 105.7 GL/year(~7% of median annual discharge). However, actual mean annual water use is estimated to beless than half this amount. •Discounting tailings dams, the Leichhardt catchment contains five dams listed in the AustraliaNational Committee on Large Dams register: Julius Dam (107.5 GL capacity), Leichhardt RiverDam / Lake Moondarra (103 GL capacity), Greenstone Creek / Waggaboonya Lake (Greenstone) (13.6 GL capacity), East Leichhardt Dam / Lake Mary Kathleen (12.1 GL capacity) and Rifle CreekDam (9.5 GL capacity). With the exception of East Leichhardt Dam, which is now only used forrecreation, the dams supply water for mining, industry and town uses. •Water quality is variable across the Southern Gulf catchments. Studies in the late 2000s foundthe water quality of the upper Leichhardt River had been influenced by urban and miningactivity and exceeded a number of heavy metal water quality guideline trigger values, howeverthe drinking water source for Mount Isa (Lake Moondarra) was found to meet drinking waterguidelines. The Gregory River, which flows into the Nicholson River approximately 85 km from the coast, is the largest perennial river in semi‑arid Queensland. •Lawn Hill Creek and the Gregory and O’Shannassy rivers receive groundwater dischargethroughout the dry season from the CLA, which is composed mostly of the CamoowealDolostone and Thorntonia Limestone. •On average, approximately 87% of the streamflow in the Southern Gulf catchments occursbetween January and March. In the perennial Gregory River, however, approximately 78% ofstreamflow occurs over the same time period. The Southern Gulf catchments contain a significant diversity of species and habitats, including freshwater, terrestrial and marine habitats of great cultural, conservation and commercial importance The Southern Gulf catchments are largely intact, but they are not pristine. • A range of economic enterprises, infrastructure and human-induced impacts occur in the Southern Gulf catchments, and the nature and extent to which these have modified the habitats and affected species of the Southern Gulf catchments vary. • The near-coastal and estuary habitats of the Southern Gulf catchments support fishing industries, including a commercial barramundi fishery, prawns and mud crabs (mainly Scylla serrata). • In the Leichhardt catchment, Lake Julius and Lake Moondarra are listed in the Directory of Important Wetlands in Australia (DIWA). While the water in these two storages predominantly supplies urban, mining and industrial demand around Mount Isa and Cloncurry, the permanent water provides important dry-season refuge for waterbirds and supports a variety of freshwater fish species. • In the Southern Gulf catchments, feral horses, wild pigs and cane toads are among the introduced animals. • Weed species of interest in and around the Southern Gulf catchments include prickly acacia, buffel grass, salvinia rubber vine and water hyacinth. Floods have economic significance because they underpin the health of the recreational and commercial fisheries in the Gulf of Carpentaria, which include a barramundi fishery and the Northern Prawn Fishery, which had a catch of prawns worth $85 million in 2019–20. • Wet-season flooding inundates significant parts of the lower reaches of the rivers in the catchments, connecting wetlands to river channels, inundating floodplains, and allowing the exchange of fauna, flora and nutrients, which drives an ecological productivity boom. • The 13 DIWA-listed wetlands in the Southern Gulf catchments include a variety of wetland types, ranging from estuarine wetlands with salt flats and saltmarshes to artificial lakes and spring-fed creeks and rivers. • Protected areas located in the Southern Gulf catchments include the UNESCO World Heritage- listed Australian Fossil Mammal Sites (Riversleigh, Queensland), which is in Boodjamulla National Park (Queensland), and Finucane Island National Park (Queensland). There are also three Indigenous Protected Areas: Ganalanga‑Mindibirrina, NT; Nijinda Durlga, Queensland; and Thuwathu/Bujimulla, Queensland. • The marine, near-shore and estuarine environments of the Southern Gulf catchments have extensive intertidal flats and estuarine communities, including mangroves, salt flats, coral reefs and seagrass habitats. These habitats are highly productive, have high cultural value and include many areas of national significance. • Seagrass beds in the near-coastal Gulf of Carpentaria have high diversity and grow in vigorous stands. They provide important food and habitat for dugongs (Dugong dugon), green sea turtles (Chelonia mydas) and prawns (Penaeus spp.). • The dominant vegetation types in the catchments are open eucalypt woodlands, Melaleuca forests and woodlands, and, on the cracking clay plains, tussock grasslands characterised by the Mitchell grasses (Astrebla spp.). • The marine and near-shore marine areas of the catchments contain significant habitat for sea turtles, shorebirds and seabirds. • Protected species include freshwater or largetooth sawfish (Pristis pristis; Vulnerable, EPBC Act) and the Gulf snapping turtle (Elseya lavarackorum). • The Southern Gulf catchments provide important stopover habitat for migratory shorebird species listed under the EPBC Act, including the eastern curlew (Numenius madagascariensis; Critically Endangered) and the Australian painted snipe (Rostratula australis; Endangered). Indigenous values, rights and development goals Indigenous Peoples are a significant and increasing population of the Southern Gulf catchments. • Traditional Owners have recognised native title and cultural heritage rights, have Aboriginal freehold land ownership and control, or are the custodians of, significant natural and cultural resources, including land, water, coastline and sea. • Water-dependent fishing and hunting are key health, customary and economic roles for Indigenous Peoples in the Southern Gulf catchments. The rivers and underground water support food security and good nutrition, and are crucial to geographical and cultural relationships. • The history of pre-colonial and colonial patterns of land and natural resource use and preservation in the Southern Gulf catchments is important to understanding present circumstances. This history has shaped residential patterns, including across the Queensland–NT border, and it also informs responses by the Indigenous Peoples to future development possibilities. • The Indigenous population of the Southern Gulf catchments outside the Mount Isa urban centre is 60% of the total population. From an Indigenous perspective, ancestral powers are still present in the landscape, including underground water, and they intimately connect Peoples, Country and culture. • Those powers must be considered in any action that takes place on Country. • Patterns of ownership and language affiliation follow features of the landscape and waterways and are reflected in the place names and songs of significant Dreamings and totemic figures. • Water is central to the cultural landscape. Riverine and aquatic areas are known to be strongly correlated with cultural heritage sites. • Some current cultural heritage considerations restrict Indigenous capacity to respond to development proposals because some knowledge is culturally sensitive and cannot be shared with those who do not have the cultural right and authority to know. Catchment-wide deliberative processes will be vital to ensuring that Indigenous water rights and interests are actively engaged and included in future water-dependent development and planning. • Indigenous Peoples, especially those in the downstream parts of the catchments, see environmental impact assessments as crucial tools to assist them to make decisions about water-dependent development. • Indigenous Peoples have business and water development objectives designed to create opportunities for existing residential populations, including the supply of safe and secure community water to residents, and to aid the return of Peoples living elsewhere. • Indigenous Peoples want to be owners, partners, investors and stakeholders in any future development. This reflects their status as the longest-term residents with deep inter- generational ties to the catchment. Indigenous custodians in the catchments are engaged in diverse collaboration and partnership models with the private, non-government, government and research sectors. Opportunities for agriculture and aquaculture There is about 1,400 ha of irrigated agriculture and 4,000 ha of rainfed agriculture in the Southern Gulf catchments. Although an abundance of soil is suitable for irrigated agriculture in the Southern Gulf catchments, the area that can be irrigated is limited by water. • Up to about 5.1 million ha of soils in the Southern Gulf catchments are classified as moderately suitable with considerable limitations (Class 3) or better (Class 1 or Class 2) for irrigated agriculture, depending on the crop and irrigation method chosen. • Class 3 soils have considerable limitations that lower production potential or require more careful management than more suitable soils, such as Class 2 soils. • About 4.7 million ha of soils in the Southern Gulf catchments are rated as Class 3 or better for irrigated pulses (i.e. food legumes) using spray irrigation in the dry season. Most of this area is rated as Class 2. In comparison, about 1.7 million ha of soils are rated Class 3 or better using furrow irrigation in the dry season for the same crops. • About 4.7 million ha of soils in the Southern Gulf catchments are rated as Class 3 or better for cotton or grains crops using spray irrigation in the dry season. Under furrow irrigation in the dry season, there are about 1.8 million ha of soil suitable for cotton or grain crops, all of which are rated as Class 3. Bushfoods are an emerging niche industry across northern Australia. Kakadu plum, though not recorded as occurring in the Southern Gulf catchments, is one of the best known and has one of the most well-developed supply chains. However, most bushfoods continue to be wild‑harvested with very little grown commercially. Limited information on commercial bushfood operations is publicly available. The soils in different parts of the study area are starkly different. • The largest arable areas in the Southern Gulf catchments are the self-mulching cracking clay soils of the Armraynald Plain (1.03 million ha) and Barkly Tableland (308,000 ha). The cracking clay soils are likely to present severe trafficability constraints throughout much of the wet season. Irrigation of the Armraynald Plain and Barkly Tableland would be best suited to use of surface water and groundwater respectively. •Downstream of Doomadgee are large contiguous areas of red sandy and loamy soils suitable forirrigated vegetables (23,000 ha) along the Nicholson River. However, in the absence of a damupstream, on-farm storage of water for broad-scale irrigation would be challenging due to thedifficulty of constructing ringtanks on these sandy soils. These soils are unlikely to be suitable forperennial horticulture due to the risk of flooding. It may be possible to slightly raise and enhancethe existing weir on the Nicholson River near Doomadgee to provide some additional water toirrigate small-scale market gardens either on these sandy and loamy soils or on similar soilscloser to Doomadgee. •The Leichhardt catchment has opportunities for irrigated horticulture on the friable soils(103,000 ha) on the narrow levees of the Leichhardt River downstream of Kajabbi. Adjacent tothese soils, and up to 1 km from the river, are friable clay soils suitable for broadacre irrigation. While these heavier soils are suitable for constructing ringtanks, conveying water pumped fromthe Leichhardt River across the sandy levee soils to ringtanks may be costly and inefficient insome locations. Irrigation enables higher yields and more flexible and reliable production than rainfed crops •Many annual crops can be grown at most times of the year with irrigation in the Southern Gulfcatchments. Irrigation provides increased yields and flexibility in sowing date. •Sowing dates must be selected to balance the need for the best growing environment(optimising solar radiation and temperature) with water availability, pest avoidance, trafficability, crop sequences, supply chain requirements, infrastructure requirements, marketdemand, seasonal commodity prices and, in the case of genetically modified cotton, plantingwindows specified within the cotton industry. •Irrigated crops likely to be commercially viable with a dry-season planting (late March to August) include annual horticulture, cotton and mungbean. Irrigated crops likely to be commerciallyviable with a wet-season planting (December to early March) include cotton, forages andpeanuts. •Seasonal irrigation water applied to crops can vary enormously with crop type (e.g. due tovariations in duration of growth, rooting depth), season of growth, soil type and rainfallreceived. For example, cotton planted in the wet season and dry season requires about 5 ML/haand 6 ML/ha, respectively, of irrigation water in at least 50% of years. A high-yielding perennialforage such as Rhodes grass requires up to about 20 ML/ha each year, averaged across a fullproduction cycle. •Rainfed cropping is theoretically possible but most likely to be opportunistic in the Southern Gulfcatchments based on rainfall received and stored soil water.How cultivated land is managed inthe years when rainfall is insufficient for rainfed cropping will be crucial for sustainable farmingoperations and the industry’s social licence to operate. Excess rainfall can also constrain crop production on some soils. •High rainfall and possible inundation mean that wet-season cropping on the alluvial clay soils ofthe Armraynald Plain carries considerable risk due to potential difficulties with access topaddocks, trafficability, waterlogging of immature crops and deteriorating water quality where agricultural runoff is poorly managed.• Accumulation of soil salinity due to irrigation over time in these soils is currently unknown and would need to be monitored, especially in the imperfectly drained soils on the northern part of the Armraynald Plain. Establishing irrigated cropping in a new region (i.e. greenfield development) is challenging. It has high input costs, high capital requirements and requires an experienced skills set. • For broadacre crops, gross margins of the order of $4,000 per hectare per year are required to provide a sufficient return on investment where on-farm development capital costs are about $20,000/ha. Crops likely to achieve such a return include cotton and to a lesser extent Rhodes grass hay, noting that the gross margins of hay are highly sensitive to local demand, price and the transport costs. • Horticultural gross margins would have to be higher (of the order of $7,000 to $11,000 per hectare per year) to provide an adequate return on the higher capital costs of developing this more intensive type of farming (relative to broadacre). Profitability of horticulture is extremely sensitive to prices received, so the locational advantage of supplying out-of-season (winter) produce to southern markets is critical to viability. Annual horticulture row crops would be the most likely to achieve these returns in the Southern Gulf catchments, but would be marginal. Growing more than one crop per year may enhance the viability of greenfield irrigation development. • There are proven benefits to sequentially cropping more than one crop per year in the same field in northern Australia, particularly where additional net revenue can be generated from the same initial investment in farm development. • Numerous options for crop sequences could be considered, but these would need to be tested and adapted to the particular opportunities and constraints of the Southern Gulf catchments’ soils and climates. While somewhat opportunistic, the most likely sequential farming systems could be those combining short-duration crops such as annual horticulture (e.g. melons), mungbean, chickpea and grass forages. Following a rainfed wet-season crop with an irrigated dry-season crop might also be possible. • Trafficability constraints on the alluvial clay soils will limit the options for sequential cropping systems. The well‑drained loamy soils on elevated plains near Doomadgee and the friable soils adjacent to the Leichhardt River pose fewer constraints for scheduling sowing times and the farm operations required for sequencing two crops in the same field each year. Even so, sequential cropping systems would be very tight operationally. • Tight scheduling requirements mean that even viable crop sequences may be opportunistic. The challenges in developing locally appropriate sequential cropping systems, and the management packages and skills to support them, should not be under-estimated. Irrigated cropping has the potential to produce off-site environmental impacts, although these can be mitigated by good management and new technology. • The pesticide and fertiliser application rates required to sustain crop growth vary widely among crop types. Selecting crops and production systems that minimise the requirement for pesticides and fertilisers can simultaneously reduce costs and negative environmental impacts. •Refining application rates of fertiliser to better match crop requirements, using controlled- release fertilisers and improving irrigation management, are effective ways to minimise nutrientadditions to waterways and, hence, the risk of harmful microalgae blooms. •Adherence to well-established best management practices can significantly reduce erosionwhere intense rainfall and slope would otherwise promote risk. This would also decrease the riskof herbicides, pesticides and excess nitrogen entering the natural environment. •More than 99% of the cotton grown in Australia is genetically modified. The geneticmodifications have allowed the cotton industry to substantially reduce insecticide (by greaterthan 85%) and herbicide application to much lower levels than previously used. In addition toreducing the likelihood and severity of off-site impacts, genetically modified crops offer healthbenefits to farm workers who handle fewer chemicals. This technology has considerablerelevance to northern Australia. Irrigated forages can increase the number of cattle sold and the income of cattle enterprises. However, the increased income is usually offset by the high initial capital costs and ongoing costs of irrigating a forage crop. •The dominant beef production system in the Southern Gulf catchments is a cow–calf breedingenterprise, with several markets possible. There is limited fattening of cattle within the studyarea, and many cattle are moved to different parts of Queensland before their final destination. •While native pastures are generally well adapted to harsh environments, they imposeconstraints on beef production through their low productivity and digestibility and theirdeclining quality through the dry season. Growing irrigated forages and hay would allow higher- quality feed to be fed to specific classes of livestock to achieve higher production and/ordifferent markets. These species could include perennial grasses, forage crops and legumes. •Grazing of irrigated forages by young cattle, or feeding them hay, decreases the time they taketo reach sale weight and, in particular, increases their daily weight gain through the dry season. •While ostensibly simple, there are many unknowns regarding the best way to implement asystem whereby irrigated forages and hay are grown on-farm to augment an existing cattleproduction system. •Growing forages or hay to feed young cattle for the export market was not financially viable inthe modelled scenarios tested when capital costs were considered.While beef production andtotal income increased, gross margins were reasonably similar to, or less than, baseline cattleoperations. Pond-based black tiger prawns or barramundi (in saltwater) or red claw crayfish (in fresh water) offer potentially high returns Along the coastal reaches of the Southern Gulf catchments, approximately 194,000 ha of land is suitable for prawn and barramundi aquaculture, using earthen ponds. •A large (i.e. 100 ha) pond-based marine culture aquaculture farm typically uses less than 4 GL/year of freshwater. •Prawn and barramundi aquaculture elsewhere have proven land-based production practices andwell‑established markets for harvested products. These are not fully established for otheraquaculture species being trialled in northern Australia. •Prawns could potentially be farmed in either extensive (low-density, low-input) or intensive(higher-density, higher-input) pond-based systems. Land-based farming of barramundi wouldlikely be intensive. •The most suitable areas of land for pond-based marine aquaculture systems are restricted to theareas of the catchment under tidal influence and the river margins where cracking clay andseasonally or permanently wet soils dominate. •Annual operating costs for intensive aquaculture are so high that they can exceed the initial costof developing the enterprise. Operational efficiency is therefore the most importantconsideration for new enterprises, particularly the production efficiency in converting feed tosaleable product. Surface water storage potential Indigenous customary, residential and economic sites are usually concentrated along major watercourses and drainage lines. Consequently, potential instream dams are more likely to have an impact on areas of high cultural significance than are most other infrastructure developments of comparable size. •Complex changes in habitat resulting from inundation could create new habitat to benefit somespecies, while other species could experience a negative impact through loss of habitat. Many of the upland regions of the Southern Gulf catchments are topographically and geologically favourable for large instream dams. However, the semi-arid climate and small catchment areas would result in most potential dam sites being relatively low yielding. •The most cost-effective potential dam sites in the Southern Gulf catchments are on the Gregory River and Gunpowder Creek. In the 1960s and 1970s, the Queensland Water Supply Commission investigated potential dam sites on both drainage lines. •Potential dam sites on the Gregory River at natural constrictions upstream of the township of Gregory are the highest yielding and most cost-effective due to the large streamflow and proximity to large contiguous areas of black and grey cracking clay soils of the Armraynald Plain. The topography, sloping away from the Gregory River, enables cheap gravity-based methods of distributing water. •A hypothetical dam on the Gregory River upstream of Gregory with a reservoir full supply level such that it does not extend into the Boodjamulla National Park could yield 133 GL in 85% of years and cost $683 million (−20% to +50%) to construct, assuming favourable geological conditions. This would be sufficient water to support approximately 10,000 ha of dry-season cropping under surface irrigation, with the reticulation infrastructure costing approximately $38 million, or $3,800 per hectare of irrigated land. •A previous study found that residents of Mount Isa place high value on particular rivers. The Gregory River, being one of the few perennial rivers in north‑west Queensland, has a high value to Indigenous and non‑Indigenous residents. •A hypothetical dam on Gunpowder Creek could potentially yield 119 GL in 85% of years and cost$773 million. If used for agriculture, the nearest suitable soils are approximately 35 kmdownstream near the junction of Gunpowder Creek and the Leichhardt River. This would besufficient to irrigate approximately 11,000 ha of dry-season broadacre and horticultural cropsunder spray irrigation. However, the grade of the land and the sandy texture of the surface soilsmean that water would need to be piped from the river, greatly increasing the cost ofreticulation infrastructure (estimated to be about $320 million or $29,000 per hectare ofirrigated land). •The higher cost of the irrigation schemes associated with a potential dam at Gunpowder Creekcompared to the potential site on the Gregory River mean that high‑value annual horticulturewould need to be grown on the friable soils downstream of the former site to offset the highercosts, assuming there was a market for the produce. •Although there are some potential dam sites on the Nicholson River, this part of the study areais particularly remote, greatly increasing capital and ongoing costs. On-farm water storages may have more prospects of being commercially viable than large instream dams in the Southern Gulf catchments. •Suitably sited large farm-scale gully dams are a relatively cost-effective method of supplyingwater. Although most sites are located in the uplands, where the soil is rocky and shallow andless suitable for their construction and irrigated agriculture, some locations – such as thenorth‑western part of Settlement Creek, Mornington Island, the lower Leichhardt River andupper parts of Morning Inlet – have a coincidence of topography and soils that is favourable forgully dams and irrigated agriculture. The alluvial clay soils on the Armraynald Plain offer different opportunities and risks to the loamy soils near Doomadgee. •Approximately 21% of the Southern Gulf catchments has soils that are likely to be suitable forthe construction of ringtanks. The most promising areas are the cracking clay soils of theArmraynald Plain east of the Gregory River and downstream of Gregory, where the natural slopeaway from the river enables the cost-effective conveyance of water using gravity. •Except for in the lower reaches of the Nicholson River, the soils of the Nicholson catchment aregenerally too sandy for ringtanks. The Leichhardt River downstream of Kajabbi has soils suitablefor ringtanks; however, sandy levees adjacent to the river can be up to 1 km wide, increasingconveyance costs and/or losses. Further downstream, although the soils are suitable forringtanks adjacent to the Leichhardt River, to the west the land slopes towards the river, whichwould result in additional pumping costs. •With water harvesting it is physically possible to extract 150 GL of water in 75% of years, assuming an even split between the Nicholson and Leichhardt catchments, without havingimpacts on the reliability of water supply to existing licensed users. This quantity of water issufficient to irrigate about 12,000 ha of broadacre crops during the dry season by pumping ordiverting water and storing it in offstream storages such as ringtanks. This would result inreductions in the mean and median annual discharges to the Gulf of Carpentaria from theNicholson and Leichhardt catchments of about 3% and 5%, respectively. Energy infrastructure constraints are a key consideration for energy‑intensive enterprises in the Southern Gulf catchments •Mount Isa and immediate surrounds are provided power by the North West Power System. •As a grid isolated from the rest of the Queensland electricity grid, the availability of generationon the North West Power System is largely matched to the existing demand of the major miningcustomers. Dispatch protocols may place limits on availability of power to new large retailcustomers, such as a large irrigation development. •Capacity on the existing Ergon 220kV transmission line that runs from Mount Isa to CenturyMine is limited, and new developments would need to negotiate with both Ergon and majorcustomers on the line to obtain the required energy and capacity. •New energy-intensive enterprises would require development of dedicated electricaldistribution infrastructure such as new substations and distribution lines. •CopperString 2032 will connect the North West Power System to the National Electricity Market, alleviating any constraints in generation. However, constraints in the transmission anddistribution system would remain. Development of a large in-river storage solely for hydro-electric power purposes would not be commercially feasible in the Southern Gulf catchments. Other existing and potential industries that depend on water Mining The Southern Gulf catchments fall within the North West Minerals Province, which is considered to be one of the world’s most significant areas for producing base and precious metals. Mining is by far the largest industry in the study area and has an annual value more than 100 times that of grazing. •Mining provides about 28% of all jobs in the Southern Gulf catchments. •Mount Isa is in a period of transition, with major lead, copper and zinc mines in the regionrecently closing or scheduled to cease operations imminently. However, technological advances(e.g. batteries, super magnets, electronics, medical imaging) have resulted in increased demandfor rare earth elements. These are referred to as critical minerals and strategic materials, andthe NT and Queensland governments have programs to attract investment in exploration andinfrastructure. •Occurrences of critical minerals and strategic materials have been recorded in the Southern Gulfcatchments, and mineral and petroleum exploration leases cover 67% of the study area. Futuredemand for minerals is highly speculative. Across Australia, mining uses about one-tenth of the water used by irrigated agriculture, though water for mining is assigned a higher reliability than agriculture, which makes water use comparisons difficult. •The amount of water used for mining is highly variable depending upon a wide range of factors, including mining methods, ore types, ore grades and processing treatments. With the exceptionof coal mines, which are not found in the Southern Gulf catchments, and large gold mines, individual mines typically use less than 0.5 GL/year. •Many uses of water by mining enterprises (e.g. dust suppression, cooling, slurry and processwater) do not require the water to be of potable quality. •Because water is typically a very small fraction of the total input cost of a mine, and miningproduces high‑value products, mining enterprises have the capacity to develop their own watersupplies. In the Southern Gulf catchments, the historical concentration of mining, industrial andurban activity around Mount Isa resulted in sufficiently high water demand for the constructionof large, purpose-built reservoirs. •Recent water use from existing reservoirs in the Mount Isa region is estimated to be modestrelative to total supplemented entitlements (<35%), indicating that the existing dams havescope to supply water for a modest expansion of mining, depending upon the location ofpotential mines. •There is limited scope for existing or new dams in the Southern Gulf catchments to be used forboth mining and irrigated agriculture because high-value minerals are located in areas with hardrock, which tend to be distant from large contiguous areas of soil suitable for irrigatedagriculture, such as alluvial plains. Petroleum extraction and de-watering of mine pits produce water that needs to be treated and appropriately disposed. Tourism The Southern Gulf catchments have a highly seasonal and relatively low volume of visitation, largely due to their climate, remoteness, sparse population and limited tourism development. •Annual visitation to the study area is estimated to result in approximately $150 millionexpenditure. Most visitors go to Mount Isa, and a large proportion (~60%) of visitors travelling toMount Isa are on business (which cannot be separated in the visitation statistics). •Although tourism in northern Australia consumes very small quantities of water relative to otherindustries, the state of northern Australia’s tourism economy is closely tied to the state andperceived attractiveness of its ecosystems, including its rivers and water‑dependent ecosystems. •Self-drive is the predominant tourism market, representing 87% of visitors to outbackQueensland. Many self-drive visitors are motivated by experiencing an ancient, vast and ‘empty’landscape with opportunities for exploration and solitude. •With a large proportion of the Southern Gulf catchments in a relatively ‘undisturbed’ state, there is potential for growth in nature-based tourism, particularly as Mount Isa’s airportprovides a commercial gateway to the region and critical infrastructure that many other parts ofnorthern Australia lack. However, the remoteness of the region and lack of supportinginfrastructure beyond Mount Isa considerably constrain tourism’s potential, as does the highseasonality of visitation, which limits enterprise profitability and year-round employmentopportunities. •Much of the appeal of the Southern Gulf catchments to self-drive tourists is likely to be theabsence of human and commercial infrastructure, so development that alters the region’scurrent characteristics could be either appealing or alienating to current tourist markets. While water resource development for agriculture has the potential to negatively affect tourism opportunities in the Southern Gulf catchments, some developments and associated supporting infrastructure (e.g. roads, accommodation) may, conversely, present opportunities to foster tourism growth. •Lake Argyle in the east Kimberley region (WA) developed as an irrigation dam to supply the OrdRiver Irrigation Area. It is now advertised as one of WA’s most spectacular attractions, offering awide range of tourism activities. Studies have identified that Lake Argyle is perceived by visitorsin a similar fashion to some ‘natural’ local attractions such as billabongs. The cultivated land ofthe Ord River Irrigation Area, however, is reportedly perceived by visitors differently, as being‘domesticated’. •Due to the presence of Lake Moondarra and Lake Julius, the recreational value of additionaldams near Mount Isa is unlikely to make a discernible impact on the local economy, thoughdams in more remote regions of the Southern Gulf catchments could encourage self-drivetourists to stay in those regions longer. •Tourism has the potential to enable economic development within Indigenous communitiesbecause Indigenous tourism enterprises, which are usually micro businesses, often havecompetitive advantages. •While tourism offers economic and employment opportunities, it can also cause impacts such asclearing of vegetation and environmental damage due to foot, bike or vehicle traffic. This canreduce amenity for local residents and increase risks such as the potential for the spread ofweeds, pests, or root rot fungus. The Southern Gulf catchments have productive groundwater systems Groundwater systems in the Southern Gulf catchments are poorly studied. Nonetheless, it is estimated that they could potentially supply up to approximately 40 GL of water per year in addition to the 3.5 GL/year of existing licensed entitlements, depending on community and government acceptance of impacts to groundwater-dependent ecosystems (GDEs) and existing groundwater users. •Not all of this water could be used for irrigated agriculture, due to poor water quality, distanceto suitable soil, and/or low bore yields.• The most promising groundwater resource in theSouthern Gulf catchments is the regional-scale Cambrian Limestone Aquifer (CLA), which ismostly comprised of the Camooweal Dolostone and Thorntonia Limestone. With appropriatelysited groundwater borefields, it is physically possible that multiple small- to intermediate‑scale(1 to 3 GL/year) groundwater-based enterprises could extract approximately 10 to 20 GL/year ofwater from the CLA, depending on community and government acceptance of impacts to GDEsand existing groundwater users. •The mean annual recharge across the CLA within the Southern Gulf catchments is estimated tobe approximately 160 GL. The conservative mean annual recharge is estimated to be between40 and 80 GL. •The CLA outcrops along the south-western part of the Southern Gulf catchments. Groundwaterdischarge from the aquifer occurs from ecologically and culturally important springs and pointsof lateral outflow where the streams are incised into the CLA along Lawn Hill Creek and theGregory and O’Shannassy rivers. •Groundwater from the CLA varies from very fresh (<500 mg/L total dissolved solids (TDS)) tobrackish (<3,000 mg/L TDS), which is towards the upper limit of salinity for most crops. The regional-scale Gilbert River Aquifer (GRA) within the geological Carpentaria Basin underlies the central and northern parts of the study area. •The GRA does not outcrop within the Southern Gulf catchments, so there is no direct rechargeto this aquifer within the study area. Recharge predominantly occurs as lateral inflow ofgroundwater from the east of the Leichhardt catchment boundary. •With appropriately sited groundwater borefields, it is possible that multiple small- tointermediate-scale (1 to 3 GL/year) developments could extract up to approximately 5 GL/yearfrom the GRA. •The GRA is shallowest (<250 m below ground level (mBGL)) at its south-western basin margin inthe mid‑reaches of the study area. It hosts some fresh water (<1,000 mg/L TDS) in places alongits south-western margin between Settlement Creek and the Leichhardt River. The GRAincreases in depth in a north-east direction towards the coast (>400 mBGL), and the waterbecomes increasingly brackish (>2,000 mg/L TDS), presenting economic challenges forgroundwater infrastructure. •The GRA becomes artesian approximately 20 to 50 km from its southern basin margin, andartesian conditions extend into the Gulf of Carpentaria. •Potential opportunities for future groundwater resource development are likely to be limitednear existing licensed groundwater users, such as the communities of Burketown and Gununa, where the GRA is at depths of greater than 500 mBGL. Neither licence is currently in use, due tothe naturally poor water quality of the aquifer. •It is unknown whether the GRA contributes groundwater to springs and seeps within the studyarea, though it is unlikely, given the depth of the aquifer. Collectively, other groundwater systems in the Southern Gulf catchments may yield approximately 15 GL/year. •The Cenozoic alluvial aquifers in the mid- to lower reaches of the Nicholson, Gregory and Leichhardt rivers, and Settlement Creek and its tributaries, host local-scale groundwater systems. However, data are very sparse, and these water resources remain poorly understood. Imagery of the study area appears to indicate that alluvial material is more laterally extensive than found elsewhere in northern Australia. •Alluvial aquifers have potential for multiple small-scale (<0.5 GL/year) localised developments or as a conjunctive water resource where surface water is available. Opportunities are likely to be limited where the saturated thickness of the aquifer is thin (<10 m), as well as in areas within 1 km of the prescribed watercourses of the Nicholson, Gregory and Leichhardt rivers. Impacts on local GDEs would need to be evaluated, because streamflow and persistent waterholes in some rivers are potentially supported by groundwater discharge from the alluvium. •The sparse groundwater data available for the alluvial aquifers suggest that the groundwaterquality in these systems is good (<600 mg/L TDS) in places. •Fractured and weathered rock aquifers are hosted in a variety of Proterozoic rocks across thesouthern, central and northern parts of the study area. These aquifers are highly variable incomposition and water quality, and typically groundwater bores yield little water (<2 L/s). Where these aquifers occasionally exhibit higher yields, due to extensive fracturing and jointing, they are often storage limited. There is potential to undertake managed aquifer recharge in aquifers of the Cenozoic alluvium and in the CLA. •The most promising aquifers for infiltration-based managed aquifer recharge in the Southern Gulf catchments are aquifers within the Cenozoic alluvium, which is associated with many of the major rivers in the study area. However, there are few bore logs and limited groundwater‑level information available for the Cenozoic alluvium, and it is likely the opportunity for MAR may vary between locations. If an alluvial aquifer is found to be suitable for MAR, it will be more cost-effective to extract groundwater in the first instance. Changes in volumes and timing of river flows have ecological impacts •Although irrigated agriculture occupies only a small percentage of the landscape, relatively small areas of irrigation can use large quantities of water, and the resulting changes in the flow regime can have profound effects on flow-dependent flora and fauna and their habitats. •Changes in river flow may extend considerable distances downstream and onto the floodplain, including into the marine environment, and their impacts can be exacerbated by other changes, including changes to connectivity, water quality and invasive species. The magnitude and spatial extent of ecological impacts arising from water resource development are highly dependent on the type of development, location, timing, extraction volume and mitigation measures implemented. •Ecological impacts, inferred here by calculating change in ecological flow dependency for a range of fresh water–dependent ecological assets for the Nicholson and Leichhardt catchments, increases non-linearly with increasing scale of surface water development (i.e. large instream dams and water harvesting). A change in ecological flow dependency does not necessarily correlate to change in ecological condition. Change in ecological condition varies between assets depending on the importance of river flow to each asset’s condition and other drivers (e.g. rainfall, local runoff, groundwater). •At equivalent levels of water resource development (i.e. in terms of volume of water extracted), and without mitigation measures the mean impact on surface‑flow‑dependent ecology of an instream dam on a major tributary of the Leichhardt River was found to be less than water harvesting averaged across the Leichhardt catchment. The mean impact of an instream dam on the Gregory River averaged across the Gregory –Nicholson catchment was found to be larger than water harvesting. •Mud crabs, mullet, threadfin and prawn species are among the ecological assets most affected by flow changes for water harvesting. However, different types of development affected assets differently. •Water harvesting developments extracting a total of 50 to 300 GL/year of water from the Nicholson and Leichhardt catchments without mitigation strategies resulted in minor and negligible changes to ecology flow dependencies of freshwater assets when averaged across each catchment respectively. Local impacts below points of extraction, however, were generally moderate for freshwater assets at the higher extraction volumes and generally minor for near-shore marine assets at higher extraction volumes. Mitigation strategies that protect low flows and first flows of a wet season are successful in reducing impacts on ecological assets. These can be particularly effective if implemented for water harvesting developments. •At equivalent volumes of water extraction, imposing an end-of-system (EOS) annual flowrequirement, where water harvesting can only commence after a specified volume of water hasflowed past the EOS and into the Gulf of Carpentaria, is an effective mitigation measure forwater harvesting. However, the efficacy of this mitigation measure is reduced with a pump-startthreshold of 600 ML/day, as is required to ensure there are no impacts on existing consumptivewater users. •The additional impact caused by water harvesting developments extracting 150 GL/year andwith a pump-start threshold of 600 ML/day is less across most ecological assets than the impactof existing levels of development and use compared with pre‑European development. •A dry future climate has the potential to have a larger mean impact on ecological flowdependencies across the Nicholson and Leichhardt catchments than the largest physicallyplausible water resource development scenarios. However, the perturbations to flow arisingfrom a combined drier future climate and water resource development result in greater impactson ecology flow dependency than either factor on their own. For instream dams, location matters, and there is potential for high risks of local impacts. Improved outcomes are associated with maintaining attributes of the natural flow regime. •Potential dams may result in an extreme change in the ecological flow dependency of someassets immediately downstream of the dam. However, impacts reduce downstream with theaccumulation of additional tributary flows, so when averaged over the entire catchment ormeasured at the EOS, the change in ecological flow dependency is minor. •Providing translucent flows (flows allowed to ‘pass through’ the dam for ecological purposes) improves flow regimes for ecology, particularly for perennial rivers like the Gregory River. Itdoes, however, reduce the yield from the dam’s reservoir. But it is not just flow, other impacts and considerations are also important. •At catchment scales, the direct impacts of irrigated agriculture on the terrestrial environmentare relatively small. However, indirect impacts such as weeds, pests and landscapefragmentation may be considerable, particularly in riparian zones. •Loss of connectivity associated with new instream structures and changes in low flows may limitmovement patterns of many species within the catchment. Inefficient farm practices, poorly managed irrigation outflows and uncontrolled runoff from irrigation areas close to drainage lines may have a larger impact on ecological condition than likely changes in river flow patterns and volume in the Southern Gulf catchments. •Nitrogen, phosphorus and potassium are the three primary nutrients used in agriculturalfertilisers. Irrigation outflows and tailwater runoff from irrigation events can be high in thesenutrients as well as in pesticides, herbicides and total dissolved solids. •If best practice is not followed, the concentrations of these contaminants can be elevated inreceiving surface water and groundwater bodies. However, the extent to which irrigatedagriculture affects the quality of receiving waters is highly variable. It depends on a wide rangeof factors, including crop type, farm management and mitigation measures, type and scale ofdevelopment, water application method, proximity to drainage lines and environmental factorssuch as climate, soil type, topography, hydrogeochemistry and susceptibility of irrigated land toflooding. •Studies in parts of Queensland with a seasonal hydrology have found that first-flow eventsfollowing irrigation or rainfall play a critical role in determining water quality. Studies haveshown that pesticide concentrations in furrow irrigation runoff are highest following initialirrigation events and decrease in subsequent events. •When pesticide application rates are managed well and irrigation schedules are aligned withcrop growth stages, the concentration of pesticides in receiving waters is typically low, studies inthe Ord River Irrigation Area have found. •Vegetated areas can intercept agricultural runoff, reducing pesticide concentrations in surfacewaters approximately three times more than areas of bare soil. This highlights the importance ofmaintaining a wide riparian buffer zone. •Water quality issues will be most significant closest to the source of the contaminant, because ofdilution and naturally occurring processes by which aquatic systems can partially processcontaminants and regulate water quality. Such processes include denitrification of nitrogen andmicrobial degradation and ultraviolet photolysis of pesticides. There are no equivalent naturalprocesses for reducing phosphorus. Commercial viability and other considerations The total annual economic value of irrigated agriculture in the Southern Gulf catchments has the potential to increase substantially, from less than $1 million to over $190 million. •The estimated total annual gross value of agricultural production in the Southern Gulfcatchments in 2020–21 was $243.6 million. Livestock commodities accounted for most of this, and cropping accounted for only $0.9 million. •Agriculture is a relatively small employer in the Southern Gulf Catchments; it is not listed in thetop five industries of employment. Large dams could be marginally viable if public investors accepted a 3% discount rate or partial contributions to water infrastructure costs, similar to established irrigation schemes in other parts of Australia. •On-farm water storages provide better prospects and, where sufficiently cheap waterdevelopment opportunities can be found, could likely support viable broadacre farms andhorticulture with low development costs. •Proponents of large infrastructure projects have a systematic tendency to substantially underestimate development costs and risks and to over estimate the scale and rate at which benefitswill be achieved. This Assessment provides information on realistic unit costs and demandtrajectories to allow potential irrigation developments to be benchmarked and assessed on alike-for-like basis. •The viability of irrigated developments would be determined by: (i) markets and supply chainsthat can provide a sufficient price, scale and reliability of demand, (ii) farmers’ skills in managingthe operational and financial complexity of adapting crop mixes and production systems suitedto the environments of the Southern Gulf catchments, (iii) the nature of water resources interms of the volume and reliability of supply relative to optimal planting windows, (iv) thenature of the soil resources and their proximity to supply chains, and (v) the costs needed todevelop those resources and grow crops compared with alternative locations. It is prudent to stage developments to manage financial risk and to allow small‑scale trialling on new farms. •Farm productivity is subject to a range of risks, and setbacks that occur early have the greatesteffect on a development’s viability. A period of initial underperformance must be anticipated forestablishing greenfield farming in a new location, and this must be planned for. •There is a strong incentive to start any new irrigation development with well-established andunderstood crops, farming systems and technologies, and to incorporate lessons fromagricultural development elsewhere in northern Australia. Local experience and experience fromthe neighbouring Flinders catchment, complemented by learnings from long-establishedirrigation schemes such as the Ord River Irrigation Area and the Burdekin Haughton WaterSupply Scheme, would provide some guidance. •New agricultural developments in the study area are likely to start irrigating broadacre cropsand pastures on heavier clay soils before progressing to higher-value and more costlyenterprises such as horticulture on sandy and loamy soils. •Staging allows ‘learning by doing’ at a small scale, where risks can be contained while testinginitial assumptions of costs and benefits and while farming systems adapt to unforeseenchallenges in local conditions. Irrigated agriculture has a greater potential than rainfed production to generate economic and community activity. •Studies in the southern Murray–Darling Basin have shown that irrigation generates a level ofeconomic and community activity that is three to five times higher than would be generated byrainfed production. •Irrigated developments can unlock the economies of scale for supply chains and support servicesthat allow rainfed farming to establish more easily around the irrigated core. •A large proportion of increased economic activity during the construction phase of potentialirrigation developments in the Southern Gulf catchments would be expected to leak outside thestudy area. Assuming $250 million in capital costs, which could potentially enable 10,000 ha ofirrigated agriculture (~20 new farm-scale developments with on-farm water sources) the totalregional economic activity within the Southern Gulf catchments associated with the constructionphase would be approximately $200 million (assuming 50% leakage out of the study area). Additional benefits would flow to other regions, including Townsville, Brisbane and potentiallysome areas outside Queensland and the NT. •The total annual increased economic activity (direct and indirect) from 10,000 ha of irrigatedmixed broadacre (85%) and horticulture (15%) agriculture in the Southern Gulf catchments couldpotentially amount to $190 million, supporting up to 360 full‑time‑equivalent jobs. ••Based on economic data for North West Queensland, the additional income that flowed toIndigenous households from beef cattle developments was about 5% of that which flowed tonon-Indigenous households. For aquaculture, forestry and fishing together the figure was about2.5% and for mining about 6%. The additional income that flowed to Indigenous householdsfrom other agricultural developments (excluding beef) was about 1.5% of that which flowed tonon-Indigenous households. This indicates that, if agricultural developments in the SouthernGulf catchments are to equally benefit Indigenous and non‑Indigenous households, concertedaction will need to be taken by all stakeholders, including government, industry groups andproponents. Sustainable irrigated development requires resolution of diverse stakeholder values and interests. •Establishing and maintaining a social licence to operate is a precondition for substantialirrigation development. •The geographic, institutional, social and economic diversity of stakeholders increases theresources required to develop a social licence and reduces the size of the ‘sweet spot’ in which asocial licence can be established. •Key interests and values that stakeholders seek to address include the purpose and beneficiariesof development, the environmental conditions and environmental services that developmentmay alter, and the degree to which stakeholders are engaged. The Southern Gulf Water Resource Assessment Team Project Director Chris Chilcott Project Leaders Cuan Petheram, Ian Watson Project Support Caroline Bruce, Seonaid Philip Communications Emily Brown, Chanel Koeleman, Jo Ashley, Nathan Dyer Activities Agriculture and socio- economics Tony Webster, Caroline Bruce, Kaylene Camuti1, Matt Curnock, Jenny Hayward, Simon Irvin, Shokhrukh Jalilov, Diane Jarvis1, Adam Liedloff, Stephen McFallan, Yvette Oliver, Di Prestwidge2, Tiemen Rhebergen, Robert Speed3, Chris Stokes, Thomas Vanderbyl3, John Virtue4 Climate David McJannet, Lynn Seo Ecology Danial Stratford, Rik Buckworth, Pascal Castellazzi, Bayley Costin, Roy Aijun Deng, Ruan Gannon, Steve Gao, Sophie Gilbey, Rob Kenyon, Shelly Lachish, Simon Linke, Heather McGinness, Linda Merrin, Katie Motson5, Rocio Ponce Reyes, Jodie Pritchard, Nathan Waltham5 Groundwater hydrology Andrew R. Taylor, Karen Barry, Russell Crosbie, Margaux Dupuy, Geoff Hodgson, Anthony Knapton6, Stacey Priestley, Matthias Raiber Indigenous water values, rights, interests and development goals Pethie Lyons, Marcus Barber, Peta Braedon, Petina Pert Land suitability Ian Watson, Jenet Austin, Bart Edmeades7, Linda Gregory, Ben Harms10, Jason Hill7, Jeremy Manders10, Gordon McLachlan, Seonaid Philip, Ross Searle, Uta Stockmann, Evan Thomas10, Mark Thomas, Francis Wait7, Peter Zund Surface water hydrology Justin Hughes, Matt Gibbs, Fazlul Karim, Julien Lerat, Steve Marvanek, Cherry Mateo, Catherine Ticehurst, Biao Wang Surface water storage Cuan Petheram, Giulio Altamura8, Fred Baynes9, Jamie Campbell11, Lachlan Cherry11, Kev Devlin4, Nick Hombsch8, Peter Hyde8, Lee Rogers, Ang Yang Note: Assessment team as at September, 2024. All contributors are affiliated with CSIRO unless indicated otherwise. Activity Leaders are underlined. 1James Cook University; 2DBP Consulting; 3Badu Advisory Pty Ltd; 4Independent contractor; 5 Centre for Tropical Water and Aquatic Ecosystem Research. James Cook University; 6CloudGMS; 7NT Department of Environment, Parks and Water Security; 8Rider Levett Bucknall; 9Baynes Geologic; 10QG Department of Environment, Science and Innovation; 11Entura Acknowledgements A large number of people provided a great deal of help, support and encouragement to the Southern Gulf Water Resource Assessment (the Assessment) team over the past three years. Their contribution was generous and enthusiastic and we could not have completed the work without them. Each of the accompanying technical reports (see Appendix A) contains its own set of acknowledgements. Here we acknowledge those people who went ‘above and beyond’ and/or who contributed across the Assessment activities. The people and organisations listed below are in no particular order. The Assessment team received tremendous support from people in the NT Government, Queensland Government and associated agencies. They are too numerous to all be mentioned here but they not only provided access to files and reports, spatial and other data, information on legislation and regulations, groundwater bores and answered innumerable questions but they also provided the team with their professional expertise and encouragement. For the NT – Sally Heaton, Lauren Cooper, Brad Sauer, Nathaneal Mills, Brett Herbert, Rob Williams and Nerida Horner. For Queensland – Greg Mason, Michael Reid, Geoffrey Cahill, Graham Herbert, Ceri Pearce, Enrico Perotti, Michelle Smith, Wayne Vogler, Hayden Ferguson and Neale Searle. The Assessment acknowledges the staff of the Carpentaria Land Council Aboriginal Corporation who guided consultations and provided valuable assistance to plan and execute the on-ground meetings with the board members of the Prescribed Body Corporate entities it represents and services. It also acknowledges advice provided by staff of other Prescribed Body Corporate entities in the upper sections of the Southern Gulf catchments that influenced the focus of the Assessment.” A number of people in private industry, universities, local government and other organisations also helped us. They include Ernie Camp, Mark Thyer, Rob Worlein, Belinda Worlein, William Weaver and Craig Cahill. Land managers and landholders at several properties provided hospitality, support and information. They include Lorraine, Wernadinga, Cliffdale and Floraville. Our documentation, and its consistency across multiple reports, were much improved by a set of copy-editors and Word-wranglers who provided great service, fast turnaround times and patient application (often multiple times) of the Assessment’s style and convention standards. They include Joely Taylor, Margie Beilharz, Jeanette Birtles and Sally Woollett. Greg Rinder provided graphics assistance. Colleagues in CSIRO, both past and present, provided freely of their time and expertise to help with the Assessment. This was often at short notice and of sufficient scale that managing their commitment to other projects became challenging. The list is long, but we’d particularly like to thank (in no particular order) Tony Zhen, Ali Wood, Nikki Dumbrell, Daniel Grainger, Veronica Hannan, Mahdi Montazeri, Ash Shokri, Zaved Khan, Kellie Muffatti, Carmel Pollino, Sonja Serbov, Jai Vaze, Francis Chiew, Heather Stewart, Dilini Wijeweera, Jordan Marano, Juliet Morris, Tenneal Maskell, Rachel Harm, Jodie Hayward, Heather Stewart, June Chin, Christian Lawrence, Gillian Foley, Anne Freer, Sharon Hall, Sonja Heyenga, Amy Edwards, Sally Tetreault Campbell, Larissa Sherman, Phil Davies, Anna Rorke and Amy Nicholson. This project was funded through the National Water Grid’s Science Program, which sits within the Department of Climate Change, Energy, the Environment and Water. Staff in the Science Program supported the smooth administration of the Assessment despite the many challenges that arose during the project years. A long list of expert reviewers provided advice that improved the quality of our methods report, the various technical reports, the catchment report and the case study report. The Governance Committee and Steering Committee (listed on the verso pages) provided important input and feedback into the Assessment as it progressed. Finally, the complexity and scale of this Assessment meant that we spent more time away from our families than we might otherwise have chosen. The whole team recognises this can only happen with the love and support of our families, so thank you. Some of the CSIRO staff involved in field work in northern Australia water resource assessments who applied their experience to the Southern Gulf analyses Photo: CSIRO – Nathan Dyer Contents Director’s foreword .......................................................................................................................... i Key findings for the Southern Gulf catchments .............................................................................. ii Overview of the Southern Gulf catchments ....................................................................... vi Indigenous values, rights and development goals .............................................................. x Opportunities for agriculture and aquaculture .................................................................. xi Other existing and potential industries that depend on water....................................... xvii The Southern Gulf catchments have productive groundwater systems .......................... xix Changes in volumes and timing of river flows have ecological impacts .......................... xxi Commercial viability and other considerations .............................................................. xxiii The Southern Gulf Water Resource Assessment Team .............................................................. xxvi Acknowledgements .................................................................................................................... xxvii Part I Introduction and overview 1 1 Preamble ............................................................................................................................. 2 1.1 Context .................................................................................................................. 2 1.2 The Southern Gulf Water Resource Assessment .................................................. 4 1.3 Report objectives and structure ............................................................................ 9 1.4 Key background ................................................................................................... 12 1.5 References ........................................................................................................... 16 Part II Resource information for assessing potential development opportunities 18 2 Physical environment of the Southern Gulf catchments ................................................. 19 2.1 Summary .............................................................................................................. 20 2.2 Geology and physical geography of the Southern Gulf catchments ................... 22 2.3 Soils of the Southern Gulf catchments ................................................................ 31 2.4 Climate of the Southern Gulf catchments ........................................................... 46 2.5 Hydrology of the Southern Gulf catchments....................................................... 60 2.6 References ........................................................................................................... 94 3 Living and built environment of the Southern Gulf catchments .................................... 101 3.1 Summary ............................................................................................................ 102 3.2 Southern Gulf catchments and their environmental values ............................. 106 3.3 Demographic and economic profile .................................................................. 133 3.4 Indigenous values, rights, interests and development goals ............................ 169 3.5 Legal and policy environment ........................................................................... 182 3.6 References ......................................................................................................... 186 Part III Opportunities for water resource development 205 4 Opportunities for agriculture in the Southern Gulf catchments .................................... 206 4.1 Summary ............................................................................................................ 207 4.2 Land suitability assessment ............................................................................... 212 4.3 Crop and forage opportunities in the Southern Gulf catchments .................... 217 4.4 Crop synopses .................................................................................................... 245 4.5 Aquaculture ....................................................................................................... 279 4.6 References ......................................................................................................... 292 5 Opportunities for water resource development in the Southern Gulf catchments ...... 296 5.1 Summary ............................................................................................................ 297 5.2 Introduction ....................................................................................................... 301 5.3 Groundwater and subsurface water storage opportunities ............................. 302 5.4 Surface water storage opportunities ................................................................ 323 5.5 Water distribution systems – conveyance of water from storage to crop ....... 366 5.6 References ......................................................................................................... 373 Part IV Economics of development and accompanying risks 378 6 Overview of economic opportunities and constraints in the Southern Gulf catchments ........................................................................................................................... 379 6.1 Summary ............................................................................................................ 380 6.2 Introduction ....................................................................................................... 381 6.3 Balancing scheme-scale costs and benefits ...................................................... 383 6.4 Cost–benefit considerations for water infrastructure viability ......................... 399 6.5 Regional-scale economic impact of irrigated development ............................. 408 6.6 References ......................................................................................................... 414 7 Ecological, biosecurity, off-site and irrigation-induced salinity risks ............................. 417 7.1 Summary ............................................................................................................ 418 7.2 Introduction ....................................................................................................... 422 7.3 Ecological implications of altered flow regimes ................................................ 424 7.4 Biosecurity considerations ................................................................................ 451 7.5 Off-site and downstream impacts ..................................................................... 467 7.6 Irrigation-induced salinity.................................................................................. 474 7.7 References ......................................................................................................... 478 Appendices 495 ........................................................................................................................... 496 Assessment products ...................................................................................................... 496 ........................................................................................................................... 499 Shortened forms ............................................................................................................. 499 Units ........................................................................................................................... 502 ........................................................................................................................... 503 List of figures ................................................................................................................... 503 List of tables .................................................................................................................... 512 Part I Introduction and overview Chapter 1 provides background and context for the Southern Gulf Water Resource Assessment (referred to as the Assessment). This chapter provides the context for and critical foundational information about the Assessment, with key concepts introduced and explained. Groundwater fed O’Shannassy River near the Australian Fossil Mammal Sites at Riversleigh. Photo: CSIRO – Nathan Dyer 1 Preamble Authors: Seonaid Philip, Ian Watson, Cuan Petheram, Caroline Bruce 1.1 Context Sustainable development and regional economic prosperity are priorities for the Australian, Northern Territory (NT) and Queensland governments and a number of strategies have been prepared to progress this. For example, the Queensland Water Strategy looks to enable regional economic prosperity through a vision that ‘Sustainable and secure water resources are central to Queensland’s economic transformation and the legacy we pass on to future generations.’ (Queensland Government, 2023). Acknowledging the need for continued research, the NT Government (2023) announced a Territory Water Plan priority action to accelerate the existing water science program ‘to support best practice water resource management and sustainable development.’ For remote areas like the catchments of the Southern Gulf rivers, that is Settlement Creek, Gregory—Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley Island groupsFigure 1-1) the land, water and other environmental resources or assets will be key in determining how sustainable regional development might occur. Primary questions in any consideration of sustainable regional development relate to the nature and the scale of opportunities (e.g. how water might be sourced to grow crops and how much water could be extracted) and their risks. 1 (hereafter ‘Southern Gulf catchments’) ( 1 Only those islands greater than 1000 ha are mapped The Assessment recognises that sustainable development is not a finite concept; it depends on the different interests and perceptions brought by individuals and communities. How people perceive risks is critical, especially in the context of areas such as the Southern Gulf catchments, where about 27% of the population is Indigenous, and more than 60% for those parts of the catchments away from the population centre of Mount Isa (compared with 3.2% for Australia as a whole), and where many Indigenous Peoples still live on the same lands they have inhabited for tens of thousands of years. Irrespective of their perspective on development, most people would agree that having access to reliable information about land and water resources enables informed discussion and good decision making. Such information includes the amount and type of a resource or asset, where it occurs in relation to complementary resources; what commercial uses it might have, how the resource changes within a year and across years; the underlying socio-economic context; and the potential impacts of development on people, land and water. Most of northern Australia’s land and water resources have not been mapped sufficiently to reliably inform resource allocation, mitigate investment or environmental risks, or build policy settings that can support good decision making. The Southern Gulf Water Resource Assessment findings aim to partly address this gap, to enable better decision making on private investment and government expenditure, taking into account intersections between existing and potential resource users, and enabling net development benefits to be maximised. The Assessment differs somewhat from many resource assessments in that it considers a wide range of resources or assets, rather than being a single mapping exercise of, say, soils. It also provides a lot of contextual information about the socio-economic profile of the catchments and the economic possibilities and environmental impacts of development. Further, it considers many of the different resource and asset types in an integrated way, rather than separately. Figure 1-1 Map of Australia showing Assessment area (Southern Gulf catchments) and other recent or ongoing CSIRO Assessments The Murray–Darling Basin and major irrigation areas and major dams (>500 GL capacity) in Australia are shown for context. The Assessment does not take an advocacy position on development, or on particular opportunities or risks. Rather, the Assessment provides resource information in a way that can inform future decision making and policy development. The outcome of no change in land use or water resource development is also valid. CSIRO has been leading similar assessments since 2012 (Figure 1-1). At that time, the Australian Government commissioned CSIRO to undertake the Flinders and Gilbert Agricultural Resource Assessment in northern Queensland as part of the North Queensland Irrigated Agriculture Strategy, a joint Australian Government and Queensland Government initiative. This assessment had a strong agricultural focus and developed fundamental soil and water datasets, providing a comprehensive and integrated evaluation of the feasibility, economic viability and sustainability of agricultural development in two catchments in northern Queensland (Petheram et al., 2013a, Australia and WRAs overview map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\1_All\1_GIS\1_Map_docs\Re-A-503_Map_Australia_and_river_basins_SG_Vic_Preamble_V1_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au 2013b). Through this work and in response to two Australian Government white papers from 2015 (the White Paper on Developing Northern Australia (PMC, 2015) and the Agricultural Competitiveness White Paper (Commonwealth of Australia, 2015)) the Australian Government commissioned CSIRO in 2016 to undertake additional, more water-focused assessments, in the Fitzroy catchment in WA (Petheram et al., 2018a) four catchments around Darwin in the NT (Petheram et al., 2018b) and the Mitchell catchment in Queensland (Petheram et al., 2018c). Collectively these three assessments are known as the Northern Australia Water Resource Assessment (NAWRA). More recently, an assessment was released for the catchment of the Roper River in the NT (Watson et al., 2023) and simultaneous assessments have been undertaken for the catchment of the Victoria River in the NT (Petheram et al., 2024) and for the Southern Gulf catchments of the NT and Queensland, which is summarised in this catchment report. These last three assessments have also been commissioned by the Australian Government through the National Water Grid’s Science Program, which sits within the Department of Climate Change, Energy, the Environment and Water. While land, water and other environmental resources and/or assets can be put to a variety of uses (including the option of ‘no change in use’), this assessment was primarily concerned with how the land and water might be used for irrigated agriculture, since that is the most likely pathway to intensified use of these resources in the coming years. 1.2 The Southern Gulf Water Resource Assessment The Southern Gulf Water Resource Assessment has synthesised and summarised information from existing water, soil and other environmental resources in order to support regional and Country planning, resource management and sustainable regional development. The Southern Gulf catchments were identified by the Australian Government as being a suitable candidate for a large-scale assessment of the water and soil resources. This was due to both interest in, and concerns about, the development of irrigated agriculture in the catchment, and interest in diversifying the economy of the region. With Mount Isa, a major mining, industrial and service centre, situated in the headwaters of the Leichhardt River, the area is seen as having the potential for overcoming some of the challenges that typify northern Australia. For example, the Queensland Government released an economic diversification strategy for North West Queensland (Department of State Development, Manufacturing, Infrastructure and Planning, 2019), which includes mining and mineral processing; beef cattle production, cropping and commercial fishing; tourism with an outback focus; and small business, supply chains and emerging industry sectors. In its 2024–25 Budget the Australian Government announced large investment in renewable hydrogen, low-carbon liquid fuels, critical minerals processing and clean energy processing (Anon, 2024). This includes investing in regions such as the North West Minerals Province, situated mostly within the Southern Gulf catchments. The Assessment aimed to: • improve baseline datasets of water, soil and other environmental resources and/or assets • understand the nature and scale of potential water resource development options • assess the potential environmental, social and economic impacts and risks of water resource and irrigation development. Indigenous engagement, and understanding Indigenous views about and interests in development, were also high priorities for the Assessment. It is important to note that, although these three aims are listed sequentially above, activities in one part of the Assessment often informed (and hence influenced) activities in an another part. For example, understanding ecosystem water requirements (described in Part IV of this report) was particularly important in establishing rules around water extraction and diversion (i.e. how much water can be taken and when it should be taken; described in Part III of this report). Thus, the procedure of assessing a study area inevitably involved iterative steps, rather than a simple linear process. The techniques and approaches used in the Assessment were specifically tailored to the study area. In covering the aims listed above, the Assessment was designed to: •explicitly address the needs of and aspirations for local development by providing objectiveassessment of resource availability, with consideration of environmental and cultural issues •meet the information needs of governments as they assess sustainable and equitablemanagement of public resources, with due consideration of environmental and cultural issues •address the due diligence requirements of private investors, by exploring questions ofprofitability and income reliability of agricultural and other developments. The Assessment built on previous studies, synthesised some new data and employed an experienced science team, with quality assured through peer-review processes. The Assessment consulted with Traditional Owner representative agencies and reviewed literature relating to the Southern Gulf catchments and from previous Assessments within catchments of the Gulf of Carpentaria. Under the guidance of the Carpentaria Land Council Aboriginal Corporation, the team remodelled its method of consultation to engage each Prescribed Body Corporate with information about the Assessment and to address questions about their preferred approach for consultation and potential future collaboration on local issues. The Southern Gulf Water Resource Assessment, which incurred delays in 2021 due to the COVID-19 pandemic, took just over 3 years to complete, between 1 July 2021 and 30 September 2024. 1.2.1 Scope of work The Assessment comprised activities that together were designed to explore the scale of the opportunity for water resource development in the Southern Gulf catchments. A set of technical reports was produced as part of the Assessment (listed in Appendix A) from which the material in this catchment report was largely drawn. Functionally, the Assessment adopted an activities-based approach to the work (which is reflected in the content and structure of the outputs and products, as per Section 1.2.3) with the following activity groups: land suitability; surface water hydrology and climate; groundwater hydrology; agriculture and socio-economics; surface water storage; Indigenous water values, rights, interests and development goals; and ecology. In stating what the Assessment did, it is equally instructive to state what it did not do. The Assessment did not seek to advocate irrigation development or assess or enable any particular development; rather, it identified the resources that could be deployed in support of potential irrigation enterprises, evaluated the feasibility of development (at a catchment scale) and considered the scale of the opportunities that might exist. In doing so, the Assessment examined the monetary and non-monetary values associated with existing use of those resources, to enable a wide range of stakeholders to assess for themselves the costs and benefits of given courses of action. The Assessment is fundamentally a resource evaluation, the results of which can be used to inform planning decisions by citizens, investors, a range of organisations and the various tiers of government: local council, and the NT and/or Queensland governments and the Australian Government. The Assessment does not replace, or seek to replace, any planning processes; it does not recommend changes to existing plans or planning processes. The Assessment sought to lower the barriers to investment in the Assessment area by addressing many of the questions that potential investors would have about production systems and methods, crop yield expectations and benchmarks, and potential profitability and reliability. This information base was established for the Assessment area as a whole, not for individual paddocks, projects or businesses. The Assessment identified those areas that are most suited for new agricultural or aquaculture developments and industries, and, by inference, those that are not well suited. It did not assume that particular sections of the study area were in or out of scope. For example, the Assessment was blind to issues such as land-clearing regulations that may exclude land from development now, but might change in the future. The Assessment identified the types and scales of water storage and access arrangements that might be possible, and the likely consequences (both costs and benefits) of pursuing these possibilities. It did not assume that particular types or scales of water storage or water access were preferable to others, nor did it recommend preferred development possibilities. The Assessment examined resource use unconstrained by legislation or regulations, to allow the results to be applied to the widest range of uses, for the longest time frame possible. In doing so, it did not assume a particular future regulatory environment, but did consider a range of existing legislation, regulation and policy, and the impact of these on development. It was not the intention – and nor was it possible – for the Assessment to address all aspects of water, irrigation and agriculture development in northern Australia. Important aspects not addressed by the Assessment include the impacts of irrigation development on terrestrial ecology. 1.2.2 Plausibility of development pathways To understand how the hydrology, ecology and economic factors in the Southern Gulf catchments interact with and respond to various types and scales of development, a wide range of potential development scenarios were examined. These ranged from small incremental increases in surface and groundwater extraction to water volumes defined only by the physical limits of the catchment. These scenarios disregarded regulatory considerations (related to, for example, water, land tenure or land clearing) that may exclude land from development now but might change over time to permit new prospects in the future. The likelihood of various scenarios will be strongly influenced by the regulatory framework at any point in time and by community acceptance of irrigated agriculture, and its benefits and risks. One way of understanding the nature and likely scale and rate of change in irrigated agricultural development, and to have meaningful discussions about future prospects in the Southern Gulf catchments, is to examine the scale and historical rate of change in irrigated agriculture across northern Australia. Preliminary data from a recent analysis by the Assessment team show that in 2023 there were about 62,000 ha of irrigated agriculture across the 310 million ha of northern Australia, as defined below. This is equivalent to about 0.02% of the land area. There are many definitions of northern Australia. The one used for these area estimations is defined as that part of northern Australia west of the Great Dividing Range and north of the Tropic of Capricorn (Figure 1-1) but including all of the NT, and all of the Gascoyne catchment, which includes the Carnarvon Irrigation District. The definition includes the intensively developed Ord River Irrigation Area (ORIA) in WA. However, the intensively developed catchments east of the Great Dividing Range that flow into the Great Barrier Reef lagoon (such as the Burdekin catchment) were not included because their biophysical and socio-economic settings are very different (Petheram and Bristow, 2008). For example, this eastern area contains cities such as Townsville and Cairns, and large irrigation areas such as the Burdekin Delta and Burdekin Haughton Water Supply Scheme. By comparison with the 62,000 ha of irrigated agriculture noted above, there are more than 350,000 ha of land developed for irrigation in these eastward-flowing catchments and about 2.4 million ha of land that has been developed for irrigated agriculture in the Murray–Darling Basin. There was a net increase of approximately 1300 ha per year of irrigated land across northern Australia (as defined above) during the 24 years between 1999 and 2023. About 26% of this increase was in the ORIA (WA), and about 18% in the Daly River catchment (NT), with the remainder of the increase across 18 other catchments. There are few reasons to suggest that the average rate of increase in irrigated land over the next few years will be very different to that seen between 1999 and 2023, notwithstanding that the NT Land Corporation announced a preferred developer in early 2022 of 67,500 ha of land in the NT (considered as Ord Stage 3), which is likely to be a mix of irrigated and mostly rainfed cropping land, but is dependent on existing water capture and storage as part of the ORIA. There are also signs that the northern jurisdictions are taking a more conservative approach to release of water than they have in the past. For example, the NT Government’s (2024) policy for taking surface water in the wet season allows for a default maximum take of 5% ‘of the 25th percentile of total flows for the three highest flow months of the year based on the previous 50 years flow or modelled rainfall data of the river basin.’ This is a reduction from its previous policy of 20% of river flows at any time in any part of a river. Similarly, the Western Australian Government has taken a conservative approach to water planning in the Fitzroy catchment in the Kimberley, and the Queensland Water Strategy (Queensland Government, 2023) now has a priority to ‘Increase First Nations’ access to and ownership of water, and greater inclusion of cultural values and traditional knowledge in water decisions.’ Figure 1-2 shows the number of large dams (defined here as having a storage capacity of 10 GL or greater and are listed in the Australian National Committee on Large Dams database) constructed across Australia and northern Australia (west and east of the Great Dividing Range) over time. Over the past 40 years only nine large dams have been constructed across all of northern Australia (including the east coast), and only three of these nine dams were constructed for supplying water for irrigation, rather than for supplying water for mining or urban use. One of the three dams was also listed as having the purposes of flood mitigation, recreation and water supply for urban use. All three of the dams constructed to supply water for irrigation are east of the Great Dividing Range. No large dam has been constructed anywhere in northern Australia for the supply of water for irrigation for more than 25 years. Figure 1-2 Number of large dams constructed in Australia and northern Australia over time Large dams are defined as dams with a storage capacity of 10 GL or greater and are listed in the Australian National Committee on Large Dams database. Irrespective of the physical resources that may support water and irrigated agricultural development in the Southern Gulf catchments, if the future trajectory of irrigation development is similar to historical trends, the scale of future irrigation development in the Southern Gulf catchments is likely to be modest and unlikely to encompass large dam development. 1.2.3 Assessment products The Assessment produced written and internet-based products. These are summarised below, and the written products are listed in full in Appendix A. Downloadable reports and other outputs can be found at: https://www.csiro.au/southerngulf Written products The Assessment produced the following documents: • technical reports, which present scientific work in sufficient detail for technical and scientific experts to independently verify the work • a catchment report, which combines key material from the technical reports, providing well- informed but non-scientific readers with the information required for informed judgments about For more information on this figure please contact CSIRO on enquiries@csiro.au 06012018024018301850187018901910193019501970199020102030Number of damsYear completedNorthern AustraliaAustralia the general opportunities and risks for, and costs and benefits associated with, water resource development, including irrigated agriculture or aquaculture •a summary report, which is provided for a general public audience •a factsheet, which provides a summary of the key findings for the Southern Gulf catchments fora general public audience. Audiovisual product The following audiovisual product was produced by the Assessment: •a video, providing an overview of the work. Internet-based platforms The following internet-based platforms were used to deliver information generated by the Assessment: •CSIRO Data Access Portal – a portal that enables the user to download key research datasetsgenerated by the Assessment •NAWRA Explorer – a web-based tool that enables the user to visualise and interrogate keyspatial datasets generated by the Assessment •internet-based applications that enable the user to run selected models generated by theAssessment. 1.3 Report objectives and structure This is the catchment report for the Southern Gulf catchments. It summarises information from the technical reports for each activity and provides tools and information to enable stakeholders to see the opportunities for development and the risks associated with them. Using the establishment of a ‘greenfield’ (not having had any previous development for irrigation) irrigation development as an example, Figure 1-3 illustrates many of the complex considerations required for such development; key report sections that inform these considerations are also indicated. The structure of the Southern Gulf catchments report is as follows: •Part I (Chapter 1) provides background, context and a general overview of the Assessment. •Part II (Chapter 2 and Chapter 3) looks at current resources and conditions within thecatchments. •Part III (Chapter 4 and Chapter 5) considers the opportunities for water, agricultural andaquaculture development based on the available resources. •Part IV (Chapter 6 and Chapter 7) provides information on the economics of development and arange of risks of development, as well as on those risks that might accompany development. Figure 1-3 Schematic of key components and concepts in the establishment of a greenfield irrigation development Numbers shown in blue refer to sections of this report. 1.3.1 Part I – Introduction This part of the report provides a general overview of the Assessment. Chapter 1 (this chapter) covers the background and context of the Assessment. Key findings can be found in the front materials of this report. 1.3.2 Part II – Resource information for assessing potential development opportunities Chapter 2 is concerned with the physical environment, seeking to describe the soil and water resources present in the Southern Gulf catchments, including: •geology and physical geography: focusing on those aspects that are important for understandingthe distribution of soils, groundwater flow systems, suitable water storage locations and geologyof economic significance •soils: covering the soil types within the catchment, the distribution of key soil attributes andtheir general suitability for irrigated agriculture •climate: outlining the general climatic circulatory systems affecting the catchment and providinginformation on key climate parameters of relevance to irrigation under current and futureclimates •hydrology: describing and quantifying the surface water and groundwater hydrology of thecatchment. Chapter 3 is concerned with the living and built environment, providing information about the people and the ecology of the Southern Gulf catchments and the institutional context of the catchments, describing: •ecology: ecological systems and assets of the Southern Gulf catchments, including the keyhabitats, key biota and their important interactions and connections •socio-economic profile: current demographics, and existing industries and infrastructure ofrelevance to water resource development in the Southern Gulf catchments •Indigenous values, rights, interests and development objectives. 1.3.3 Part III – Opportunities for water resource development Chapter 4 presents information about the opportunities for irrigated agriculture and aquaculture in the Southern Gulf catchments, describing: •land suitability for a range of crop × season × irrigation type combinations, and for aquaculture, including key soil-related management considerations •cropping and other agricultural opportunities, including crop yields and water use •gross margins at the farm scale •prospects for integration of forages and crops into existing beef enterprises •aquaculture opportunities. Chapter 5 presents information about opportunities for extracting and/or storing water for use in the Southern Gulf catchments, describing: •groundwater and subsurface storage opportunities •surface water storage opportunities in the Southern Gulf catchments, including major dams, large farm-scale dams and natural water bodies •estimates of the quantity of water that could be pumped or diverted from the Nicholson, Gregory and Leichhardt rivers and their major tributaries •water distribution systems (i.e. for conveyance of water from a dam and application to a crop) •costs of potential broad-scale irrigation development. 1.3.4 Part IV – Economics of development and accompanying risks Chapter 6 covers economic opportunities and constraints for water resource development, describing: •balance of scheme-scale costs and benefits •cost–benefit considerations for water infrastructure viability •regional-scale economic impacts of irrigated development. Chapter 7 discusses a range of risks of development, including those that might accompany development, describing: •ecological impacts of altered flow regimes on aquatic, riparian and near-shore marine ecology •biosecurity risks to agricultural or aquaculture enterprises •potential off-site impacts (due to sediment, nutrients and agri-pollutants) to receiving waters inthe catchments •irrigation-induced salinity due to rising watertables. 1.3.5 Appendices This report contains three appendices: Appendix A – list of information products Appendix B – shortened forms and units Appendix C – list of figures and list of tables. 1.4 Key background 1.4.1 The Southern Gulf catchments The Southern Gulf catchments (Figure 1-4) have an area of 108,200 km2 and encompass the Settlement Creek (17,600 km2), Gregory–Nicholson (52,200 km2), Leichhardt (33,400 km2) and Morning Inlet (3700 km2) catchments and some of the islands of the Wellesley Island groups2. The majority of the published information from these catchments is focused on the Queensland portion of the area (79% of the Assessment area). Less well known, and more remote, is the 21% of the study area in the NT. The Assessment area has a population of approximately 22,500, the majority residing in the major regional service centre Mount Isa (18,000) and smaller centres at Doomadgee (1400), Gununa (1025) and Burketown (200) (ABS, 2021). The nearest major city and population centre is Townsville (192,768 in the 2021 Census), approximately 900 km by road from Mount Isa. The climate of the Southern Gulf catchments is hot and semi-arid to dry subhumid. The majority of streams and rivers are ephemeral, with a notable exception being the Gregory River and its tributaries Lawn Hill Creek and O’Shannassy River, into which groundwater discharges from karstic carbonate rocks in their headwaters. As a proportion of its median annual streamflow, the water in the Leichhardt River is more heavily regulated than most other rivers in northern Australia west of the Great Dividing Range. Several large reservoirs in the catchment of the Leichhardt River, namely Lake Julius (108 GL capacity) and Lake Moondarra (107 GL capacity) are used conjunctively to supply water for urban, mining and industrial use around Mount Isa and Cloncurry. Lake Moondarra is also used by Mount Isa residents and others for recreation. A low weir (of less than 1 GL capacity) upstream of the road crossing impounds water on the Nicholson River downstream of Doomadgee. 2 Only those islands greater than 1000 ha are mapped. Figure 1-4 The Southern Gulf catchments ALRA = Aboriginal Land Rights (Northern Territory) Act 1976. IPA = Indigenous Protected Area; NP = national park; NR = nature refuge. Southern Gulf catchments overview map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\CR-S-Ch1_500_SouGulf_overview_v01.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au The main land uses in the Assessment area are grazing native vegetation (77% of the Assessment area), and conservation and natural environments (16%). Cropping (both rainfed and irrigated) are sparsely practised (<0.04%). Mining, while a significant economic activity in the Assessment area, uses less than 1% of the land area. Protected areas located in the Southern Gulf catchments include the United Nations Educational, Scientific and Cultural Organization World Heritage–listed Australian Fossil Mammal Sites (Riversleigh); three Indigenous Protected Areas, namely Ganalanga–Mindibirrina, Nijinda Durlga and Thuwathu/Bujimulla; and Boodjamulla (Lawn Hill) and Finucane Island national parks and other conservation parks. In addition to these protected areas, the Southern Gulf catchments contain 13 nationally significant wetlands listed in the Directory of Important Wetlands in Australia. 1.4.2 Wet–dry seasonal cycle: the water year Northern Australia has a highly seasonal climate, with most rain falling during the 4-month period from December to March. Unless specified otherwise, this Assessment defines the wet season as being the 6-month period from 1 November to 30 April, and the dry season as the 6-month period from 1 May to 31 October. However, it should be noted that the transition from the dry to the wet season typically occurs in October or November, and the definition of the northern wet season commonly used by meteorologists is 1 October to 30 April. All results in the Assessment are reported over the water year, defined as the period 1 September to 31 August, unless specified otherwise. This allows each individual wet season to be counted in a single 12-month period, rather than being split over two calendar years (i.e. counted as two separate seasons). This is more realistic for reporting climate statistics from hydrological and agricultural assessment viewpoints. 1.4.3 Scenario definitions The Assessment considered four scenarios, reflecting combinations of different levels of development, and historical and future climates, much like those used in the Northern Australia Sustainable Yields project (CSIRO, 2009a, 2009b, 2009c), the Flinders and Gilbert Agricultural Resource Assessment (Petheram et al., 2013a, 2013b), the Northern Australia Water Resource Assessments (Petheram et al., 2018a, 2018b, 2018c) and the Roper River Water Resource Assessment (Watson et al., 2023): • Scenario A – historical climate and current development • Scenario B – historical climate and future development • Scenario C – future climate and current development • Scenario D – future climate and future development. Scenario A Scenario A and its subsets, Scenario AN and Scenario AE, assume a historical climate. The historical climate series is defined as the observed climate (rainfall, temperature and potential evaporation for water years from 1 September 1890 to 31 August 2022). All results presented in this report were calculated over this period, unless otherwise specified. Historical tidal data were used to specify downstream boundary conditions for the flood modelling. Scenario A assumes current levels of surface water development and that existing water licence holders use their full entitlement. For more information see the companion technical report on river model simulation in the Southern Gulf catchments (Gibbs et al., 2024). Strategic reserves are not included under Scenario A. Scenario AN assumes no water resource development. Scenario AE assumes current levels of surface water development, estimates actual use of the existing entitlement holders and applies this over the entire historical period. Strategic reserves are not included. The results under scenarios AN and AE only differ from the results under Scenario A in the Leichhardt catchment, because other catchments in the study area have no or very small (≤5 GL/year) entitlement volumes. Scenario B Scenario B is historical climate and future hypothetical development assessed at approximately 2060. Scenario B uses the same historical climate series as Scenario A. River inflow was modified to reflect potential future development. Potential development options are entirely hypothetical and were devised to assess the response of hydrological, ecological and economic systems to future development ranging from small incremental increases in surface water through to extraction volumes representative of the likely physical limits of the Southern Gulf catchments (i.e. considering the co-location of suitable soil and water). Price and cost information was indexed to December 2023 unless otherwise specified. The impacts of future hypothetical developments on existing licence holders in the Leichhardt catchment are reported. The impacts of changes in flow due to this future development were assessed, including impacts on: •instream, riparian and near-shore ecosystems •Indigenous water values •economic costs and benefits •opportunity costs of expanding irrigation •institutional, economic and social considerations that may impede or enable adoption ofirrigated agriculture. Scenario C Scenario C is future climate and current levels of surface water assessed at approximately 2060. Future climate impacts on water resources were explored within a sensitivity analysis framework by applying percentage changes in rainfall and potential evaporation to modify the 132-year historical climate series (as in Scenario A). The percentage change values adopted were informed by projected changes in rainfall and potential evaporation under Shared Socio-economic Pathways (SSP) 2-4.5 and 5-8.5 as defined in the IPCC Sixth Assessment Report on Climate change (IPCC, 2022). SSP 2-4.5 is broadly considered representative of a likely projection given current global commitments to reducing emissions, and SSP 5-8.5 is representative of an (unlikely) upper bound. Scenario D Scenario D is future climate and future hypothetical development. It uses the same future climate series as Scenario C. River inflow was modified to reflect future hypothetical developments, as in Scenario B. Therefore, in this report, the climate data for scenarios A and B are the same (historical observations from 1 September 1890 to 31 August 2022), and the climate data for scenarios C and D are the same (the above historical data scaled to reflect a plausible range of future climates). 1.5 References Australian Bureau of Statistics (ABS) (2021) Census of population and housing time series profile. Catalogue number 2003.0 for various SA regions falling partly within the Southern Gulf catchments. Australian Bureau of Statistics, Canberra. Viewed 29 October 2023, https://www.abs.gov.au/census/find-census-data/search-by-area. Anon (2024) Budget 2024-25. A future made in Australia. Viewed 11 September 2024, https://budget.gov.au/content/factsheets/download/factsheet-fmia.pdf. Commonwealth of Australia (2015) Agricultural Competitiveness White Paper, Canberra. Viewed 24 September 2024, https://www.agriculture.gov.au/sites/default/files/documents/ag- competitiveness-white-paper_0.pdf CSIRO (2009a) Water in the Gulf of Carpentaria Drainage Division. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO Water for a Healthy Country Flagship, Australia. https://doi.org/10.4225/08/5859749d4c71e. CSIRO (2009b) Water in the Timor Sea Drainage Division. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO Water for a Healthy Country Flagship, Australia. https://doi.org/10.4225/08/585ac5bf09d7c. CSIRO (2009c) Water in the Northern North-East Coast Drainage Division. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO Water for a Healthy Country Flagship, Australia. https://doi.org/10.4225/08/585972c545457. Department of State Development, Manufacturing, Infrastructure and Planning (2019) North west Queensland economic diversification strategy 2019. Viewed 12 September 2024, https://www.statedevelopment.qld.gov.au/regions/regional-priorities/a-strong-and- prosperous-north-west-queensland/north-west-queensland-economic-diversification- strategy. Gibbs M, Hughes J, Yang A, Wang B, Marvanek S and Petheram C (2024) River model scenario analysis for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Intergovernmental Panel on Climate Change (IPCC) (2022) Climate Change 2022: Impacts, adaptation, and vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York, NY. NT Government (2023) Territory Water Plan. A plan to deliver water security for all Territorians, now and into the future. Viewed 6 September 2024, https://watersecurity.nt.gov.au/__data/assets/pdf_file/0003/1247520/territory-water- plan.pdf. NT Government (2024) Surface water take – wet season flow policy. Viewed 25 July 2024, https://nt.gov.au/__data/assets/pdf_file/0008/1348190/surface-water-take-wet-season- flow-policy.pdf. Petheram C and Bristow K (2008) Towards an understanding of the hydrological factors, constraints and opportunities for irrigation in northern Australia: a review. Science Report No. 13/08. CRC for Irrigation Futures Technical Report No. 06/08. CSIRO Land and Water, Australia. Petheram C, Watson I and Stone P (eds) (2013a) Agricultural resource assessment for the Flinders catchment. A 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 Healthy Country and Sustainable Agriculture flagships, Australia. Petheram C, Watson I and Stone P (eds) (2013b) Agricultural resource assessment for the Gilbert catchment. A 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 Healthy Country and Sustainable Agriculture flagships, Australia. Petheram C, Bruce C, Chilcott C and Watson I (eds) (2018a) Water resource assessment for the Fitzroy catchment. A 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. Petheram C, Chilcott C, Watson I, Bruce CI (eds) (2018b) Water resource assessment for the Darwin catchments. A 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. Petheram C, Philip S, Watson I, Bruce C and Chilcott C (eds) (2024) Water resource assessment for the Victoria catchment. A report from the CSIRO Victoria River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Petheram C, Watson I, Bruce C and Chilcott C (eds) (2018c) Water resource assessment for the Mitchell catchment. A 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. Prime Minister and Cabinet (PMC) (2015) Our North, Our Future: White Paper on Developing Northern Australia, Prime Minister and Cabinet, Commonwealth Government of Australia, 2015. Viewed 24 September, https://www.infrastructure.gov.au/sites/default/files/documents/nawp-fullreport.pdf Queensland Government (2023) Queensland Water Strategy. Water. Our life resource. Viewed 11 September 2024, https://www.rdmw.qld.gov.au/qld-water-strategy/strategic-direction. 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. Part II Resource information for assessing potential development opportunities Chapters 2 and 3 provide baseline information that readers can use to understand what soils and water resources are present in the Southern Gulf catchments and the current living and built environment of the Southern Gulf catchments. This information covers: •the physical environment (Chapter 2) •the people, ecology and institutional context (Chapter 3). The Leichhardt River downstream of its junction with Gunpowder Creek. Loamy and clayey-surfaced friable soils are adjacent to the river. Treeless alluvial clay plains can be seen in the distance. Photo: CSIRO – Nathan Dyer 2 Physical environment of the Southern Gulf catchments Authors: Matthias Raiber, Matt Gibbs, Peter Zund, Andrew R Taylor, Seonaid Philip, Steve Marvanek, David McJannet, Fazlul Karim, Bill Wang, Cuan Petheram, Russell Crosbie, Justin Hughes Chapter 2 examines the physical environment of the catchments of the Southern Gulf rivers – that is, Settlement Creek, Gregory–Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groupsFigure 2-1. 1 – and seeks to identify the available soil and water resources. It provides fundamental information about the geology, soil, climate, and the river and groundwater systems of the catchment. These resources underpin the natural environment and existing industries, providing physical bounds to the potential scale of irrigation development. Key components and concepts are shown in 1 Only those islands greater than 1000 ha are mapped. Figure 2-1 Schematic diagram of key natural components and concepts in the establishment of a greenfield irrigation development "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\10_Reporting\1_All\9_Graphics_artist\3_Vic and SoG\C Bruce Vic CR Chp2_8_2024.jpg" Numbers in blue refer to sections in this report. 2.1 Summary This chapter provides a resource assessment of the geology, soil, climate, groundwater and surface water resources of the Southern Gulf catchments. No attempt is made in this chapter to calculate physically plausible areas of land or volumes of water that could potentially be used for agriculture or aquaculture developments. Those analyses are reported in chapters 4 and 5. 2.1.1 Key findings Soils The soils with potential for agriculture in the Southern Gulf catchments are dominated by cracking clay soils (23% of the catchment), which are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. They typically have a moderate to high agricultural potential, are deep with a high soil available water capacity (AWC) and are suited to a wide variety of irrigated dry-season crops. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain, and crop tolerance to poor soil drainage conditions restricts wet-season cropping in these areas. High salt levels within the soil profile in areas of the Armraynald Plain need management consideration. Friable, non-cracking clay soils (3% of the catchments) and loamy soils (~3% of the catchments) make up substantial areas. Both of these soils are well drained, and the friable, non-cracking clay soils have a generally high agricultural potential while the loamy soils have moderate to high agricultural potential. Sandy soils (10% of the catchment) have moderate agricultural potential with spray or trickle irrigation, although low to moderate water-holding capacity and hardsetting surface soils are common constraints. Shallow and/or rocky soils make up just over 56% of the catchment and have been assessed as having limited to no potential for agricultural development. These areas do, however, have substantial habitat value and broader biodiversity and cultural values. Climate The Southern Gulf catchments have a hot and arid climate that is highly seasonal with an extended dry season. They receive a mean rainfall of 602 mm/year, 94% of which falls during the wet season. Mean daily temperatures and potential evaporation are high relative to other parts of Australia. On average, potential evaporation is approximately 1900 mm/year. Overall, the climate of the Southern Gulf catchments generally suits a wide range of crops, though in most years rainfall would need to be supplemented with irrigation. The variation in rainfall from one year to the next is moderately high compared to elsewhere in northern Australia and high compared to other parts of the world with similar mean annual rainfall. The Southern Gulf catchments do not experience unusually long sequences of dry years, and the magnitude of dry spells is similar to many areas in the Murray–Darling Basin and the east coast of Australia. Since the 1969–70 water year (1 September to 31 August), the Southern Gulf catchments have experienced one tropical cyclone in 36% of cyclone seasons and two tropical cyclones in 4% of seasons. Approximately 10% of the global climate models (GCMs) project an increase in mean annual rainfall by more than 5%, 22% project a decrease in mean annual rainfall by more than 5% and about 73% indicate ‘little change’. Surface water and groundwater The timing and event-driven nature of rainfall events and high potential evaporation rates across the Southern Gulf catchments have important consequences for the catchments’ hydrology. Approximately 97% of runoff occurs during the wet season (November to April, inclusive), and 87% of all runoff occurs during the 3-month period from January to March, which is very different to southern Australia where rainfall and runoff are less seasonal. This means that, in the absence of groundwater, water storages are essential for dry-season irrigation. The major aquifers in the Southern Gulf catchments occur within dissolution features in the dolostones and limestones of the Cambrian (541 to 485 million years ago (Ma)) Georgina Basin in the west of the Southern Gulf catchments, the Late Jurassic to Early Cretaceous Gilbert River Formation and, to a lesser extent, the Normanton Formation within the Carpentaria Sub-basin of the Great Artesian Basin (GAB). The carbonate rocks of the Georgina Basin, in particular the Camooweal Dolostone and Thorntonia Limestone, form part of a complex, interconnected and highly productive regional-scale groundwater system (about 460,000 km2) that extends for hundreds of thousands of square kilometres west, south and north of the south-western boundaries of the Southern Gulf catchments. Mean annual volumetric recharge over the aquifers of the Georgina Basin within the Southern Gulf catchments is estimated to be 122 GL/year. Bore yields are variable due to the complex nature of the karstic aquifer, but often range from less than 1 to 20 L/second. The carbonate rocks of the Georgina Basin are a complex regional-scale groundwater system due to the variability and interconnectivity between fractures, fissures and karsts. Currently about 1.8 GL/year of groundwater is licensed to be extracted from the Camooweal Dolostone and Thorntonia Limestone within the Southern Gulf catchments (see Section 3.3.4). The Gilbert River Formation and Normanton Formation host the most regionally extensive aquifer systems within the Southern Gulf catchments, where they extend over the eastern part of the catchment in Queensland and are entirely covered by the Cenozoic (66 Ma to present) Karumba Basin. Bore yields of the Gilbert River Formation are variable, ranging from less than 1 to 46 L/second. The aquifers of the geological Carpentaria Basin form part of the 1.7 million km2 large GAB, which reaches far to the south and east of the Southern Gulf catchments and includes the Eromanga and Surat basins and parts of the Clarence–Moreton Basin. Other local- to intermediate-scale aquifer systems are likely to exist in other geological formations within the Southern Gulf catchments. For example, the Cenozoic alluvium may offer potential in terms of water quality (median total dissolved solids (TDS) of approximately 600 mg/L). However, only limited data are available to assess lithological variability, hydraulic properties and the spatial extent of potential productive aquifers. The Southern Gulf catchments consist of various rivers and streams that discharge into the southern Gulf of Carpentaria. The most substantial of these are the Leichhardt, Gregory and Nicholson rivers. The median and mean annual discharges from the Southern Gulf catchments into the Gulf of Carpentaria are 4961 and 6759 GL/year, respectively. The majority of this volume is relatively evenly split between the Leichhardt and Gregory-Nicholson River catchments. Most rivers cease to flow over the dry season (27% to 79% of the time), however the Gregory River is a notable exception with perennial flow resulting from discharge from the Thorntonia Limestone hydrogeological unit. Current surface water licences across the study area total about 114 GL (i.e. 2.3% of the median annual discharge from the study area), with the majority of these licences in the Leichhardt River catchment, including those supplied from Lake Moondarra and Lake Julius (Section 3.34). Many rivers in the catchments, particularly those in the southern parts of the Southern Gulf catchments, are ephemeral and are reduced to a few scarce and vulnerable waterholes during the dry season. Some waterholes and river reaches, particularly those in the upper reaches of the Gregory River, are permanent and are replenished by groundwater (see Section 2.5.4). 2.1.2 Introduction This chapter seeks to address the question: What soil and water resources are available for irrigated agriculture in the Southern Gulf catchments? The chapter is structured as follows: • Section 2.2 examines the geology of the Southern Gulf catchments, which is important in understanding the distribution of groundwater, soil and areas of low and high relief, which in turn influence flooding and the deposition of soil. • Section 2.3 examines the distribution and attributes of soils in the Southern Gulf catchments and discusses management considerations. • Section 2.4 examines the climate of the Southern Gulf catchments, including historical data and future projections of patterns in rainfall. • Section 2.5 examines the groundwater and surface water hydrology of the Southern Gulf catchments, including groundwater recharge, streamflow and flooding. 2.2 Geology and physical geography of the Southern Gulf catchments 2.2.1 Geological history The geological history of an area describes the major periods of deposition and tectonics (i.e. major structural changes) as well as weathering and erosion. These processes are closely linked to the physical environment that influences the evolution and formation of resources such as valuable minerals, coal, groundwater and soil. Geology also determines topography, which in turn is a key factor in the location of potential dam sites, flooding and deposition of soil. These resources are all important considerations when identifying suitable locations for large water storages and when understanding past and present ecological systems and patterns of human settlement. The oldest rocks in the Southern Gulf catchments are late Proterozoic (1780 to 1400 million years old). They consist of repeated thick sequences of sedimentary and metamorphic rocks and volcanics that include numerous prominent beds of sandstone (Figure 2-2). These sequences were deposited in a series of basins (e.g. the McArthur and South Nicholson basins and the Isa Superbasin) extending across the area and then folded, faulted and intruded (i.e. broken through) by igneous rocks to form mountain chains. Towards the end of the Proterozoic, the mountain chains had been eroded down to a level not far above that of the current topography. Surface geology map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS510-S_Catchment_1M_Geology_overview_v02_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-2 Surface geology of the Southern Gulf catchments Source: adapted from Raymond (2012) During the Neoproterozoic to late Palaeozoic (850 to 350 Ma), the limestones, dolomites and minor sandstones of the Georgina Basin were deposited on a tectonically inactive platform. The Cambrian strata provide an important regional groundwater source in the west and south-west of the Southern Gulf catchments and are largely concealed by overlying Cretaceous sediments. The Jurassic to Cretaceous (125 to 100 Ma) geological Carpentaria Basin, a sub-basin of the GAB, hosts sequences of interbedded sandstones, mudstones and siltstones. These sequences underlie most of the eastern part of the Southern Gulf catchments and extend and thicken offshore. They include deposits of the Gilbert River Formation (GRF), which is composed of fluvial quartzose sandstones and forms the major GAB aquifer and is an important groundwater resource within the Southern Gulf catchments. Following deposition of the GRF, widespread transgression and then major regression in the late Middle Cretaceous led to deposition of the thick mudstone successions of the Wallumbilla Formation and erosion and deposition of a thin succession of Cretaceous shallow marine sandstones, conglomerates and mudstones. Extensive Cenozoic alluvial plains deposits unconformably overlie the geological Carpentaria Basin and mostly correspond to the Cenozoic Karumba Basin, which is persistent across much of northern Queensland and extends into eastern parts of the NT. Overlying the Karumba Basin are the youngest sediments in the catchments: the alluvial sands, silts and gravels associated with the beds, channels and floodplains of the catchments’ rivers and creeks and their tributaries. The present landscape has been produced by warping and dissection of a series of erosion surfaces formed during several cycles of erosion that started in the Late Cretaceous about 70 Ma and ended in the mid-Cenozoic era about 25 Ma. During this time, stable crustal conditions and subaerial exposure led to patchy erosion of the Cretaceous rocks and prolonged subaerial weathering of the remaining Cretaceous and Proterozoic rocks, resulting in the formation of deep weathering profiles and associated iron-cemented capping. Between the mid-Cenozoic and the present day, there has been gentle uplift and warping of the various surfaces and their weathered cappings. Continued erosion has led to the emergence of the present-day landscape, and extensive floodplains and coastal deposits were built up on the margins of modern drainage systems and the coastline, respectively, in the Southern Gulf catchments. 2.2.2 Physiography of the Southern Gulf catchments Ten physiographic units have been identified based on the geological controls outlined in Section 2.2.1. They are described in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024) and shown in Figure 2-3. Physiographic map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-501_location_v2_v11_Arc10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-3 Physiographic units of the Southern Gulf catchments Physiographic units as described in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024) and based on Grimes (1974). Significant settlements and roads are overlaid on hillshaded terrain relief. The mainland Assessment areas can be split into the uplands and the Carpentaria Plains (Grimes, 1974). The upland area in the south and west reaches 620 metres above sea level (mASL) and is the headwaters for Assessment area catchments. The uplands can be divided into four physiographic units: Isa Highland, Barkly Tableland, Dissected Barkly Tableland and Gulf Fall (Figure 2-3). The oldest and most elevated and rugged unit is the Isa Highland (Twidale, 1956). It consists of Precambrian (>545 Ma) volcanic and sedimentary rocks that have been metamorphosed, weathered and eroded. Soil parent materials within the Isa Highland from west to east include rhyolite, basalt, dolomitic sediments, siltstone, meta basalt, granite, quartzite and metasediments. Land surface relief is moderate (200 to 230 m) and generally in a south–north alignment. The next most elevated upland physiographic unit is a small part of Barkly Tableland to the west of the Isa Highland. The tableland started out as a sedimentary basin in Precambrian times and then was uplifted, folded and eroded. During the Cambrian period, seas transgressed the area and deposited carbonate sediments in the depressions. Later the Cambrian-period sediments were exposed and eroded. During the Mesozoic, isolated lakes and swamps developed (Randal, 1966), and subsequently, during the Cenozoic period, the upland areas experienced deep weathering and laterisation. However, areas covered by lakes and swamps did not undergo strong leaching, and as the landscape dried, the current cracking clay soils formed on relatively fresh sediments (Christian et al., 1954). The clay soils overlie dolomitic rocks. Relief is very low (9 to 30 m), and Mitchell grasslands dominate. Since the Cambrian period, the drainage network that flows towards the Gulf of Carpentaria has dissected the tableland, leaving remnant land features defined by deep narrow gorges. This area is mapped as the Dissected Barkly Tableland physiographic unit in Figure 2-3. Dissection has been amplified because the underlying rocks formed from dolomitic sediments are relatively soluble compared to surrounding rocks. These gorges have intersected the groundwater systems of the tableland, resulting in spring-fed permanent creeks and rivers such as the O’Shannassy, Gregory and Lawn Hill Creek subcatchments. The remaining parts of the uplands, comprising mainly Mesozoic sedimentary formations (sandstones), have been eroded into a complex pattern of easterly flowing streams and valleys separated by ranges and outcrops of sedimentary formations (Mullera Formation, Constance Sandstone and Fickling Beds; Smith and Roberts (1972)). This physiographic unit is known as the Gulf Fall, and the Nicholson and South Nicholson rivers are the primary systems draining this area. Musselbrook, Lagoon, Settlement, Gold and Running creeks also drain this area. To the east of the uplands are the Carpentaria Plains comprising a series of plains, pediments and remanent plateaux that can be divided into six physiographic units: Cloncurry Plain, Doomadgee Plain, Armraynald Plain, Karumba Plain, Donors Plateau and Mornington Plateau (Figure 2-3). The most elevated sedimentary plain (30 to 150 mASL) is the Cloncurry Plain physiographic unit immediately east of the uplands. It consists of gently sloping colluvial and fluvial sedimentary plains and pediments with isolated low hills of Precambrian rock. Streams are few and incised into the pediments with narrow alluvial plains (Grimes, 1974). The Cloncurry Plain unit extends from the middle reach of the Leichhardt River to Lawn Hill Creek. In the northern Assessment area, the Doomadgee Plain physiographic unit lies below and adjacent to the Cloncurry Plain. It is predominantly a sandy, gently undulating plain overlying a deeply weathered Cenozoic land surface. Low eucalypt and paperbark scrub cover the lands. Widely spaced creeks drain the plains, currently in a radial north-westerly direction towards the coast. This suggests that the underlying old land surface could have been a large sedimentary fan. Prior streams of sandier soils, shallow swampy and water-filled depressions (particularly between Lilly and Moonlight creeks) and small pits caused by ferricrete subsidence occur throughout the plains (Grimes, 1974). In the southern half of the Assessment area, the Armraynald Plain physiographic unit lies below and adjacent to the Cloncurry Plain unit. It consists of argillaceous Cenozoic (Quaternary period) sediments (Armraynald Beds) that form black soils covered in grasslands. Stream channels are few, widely spaced and deeply incised due to sea-level changes. The plains extend up the Lawn Hill Creek, and the Gregory and Leichhardt valleys. Lawn Hill Creek and Gregory River are spring-fed permanent streams. The Gregory River splits into a giant braid (20 km at its widest) of permanent streams consisting of the Gregory River, Beames Brook, Barkly River and Running Creek downstream of the Gregory Crossing. Monsoonal rainforest grows immediately adjacent to these permanent streams that cross the otherwise grassland plains (Grimes, 1974). Downslope of both the Doomadgee Plain and Armraynald Plain lies the coastal Karumba Plain physiographic unit. This coastal unit extends 10 to 35 km inland from the Gulf of Carpentaria coast and is widest near the Albert River mouth (upstream called the Nicholson and Gregory rivers). This plain consists of Holocene beach ridges and tidal and extratidal flats and plains. Some of the inland plains only flood when the rivers are in spate or when the north-westerly winds cause exceptionally high tides during the monsoon. Because the plain is generally flat and wide and the tidal range is moderate (about 3.5 m), tidal waters can rapidly inundate the land. Mangroves and tidal flats dominate the coastline; beaches are few and consist of white shelly sand. Small crescent dunes have formed in places from wind action (Grimes, 1974). Strong north-easterly winds across the bare plains, especially in November, may cause a fog-like effect in Burketown from suspended particles. Due to the flatness of the plain, streams meander in complex patterns. To the east of the Armraynald Plain lies the Donors Plateau physiographic unit. This slightly elevated unit (10 to 80 mASL) forms a watershed between the catchments of the Leichhardt and Flinders rivers and forms the eastern boundary of the Assessment area. The plateau consists of siliceous sediments laid down in the Early Cretaceous epoch from upland sediment sources of the Normanton Formation. The plain, which was once more extensive, was deeply weathered and lateralised in the highest elevation parts in the Tertiary period and has subsequently been stripped away in parts, leaving today’s Donors Plateau and exposed older Cretaceous sediments. In parts, the older sediments have been re-covered by sediments laid down during the Pleistocene, forming the Armraynald Plain (Ingram, 1972). Much of the Wellesley Islands in the Gulf of Carpentaria represent remnants of a mainland laterised Cretaceous period plain called the Mornington Plateau physiographic unit. This unit is 5 to 20 mASL. Dissection of the unit is not as extensive as in the Donors Plateau unit. The Mornington Plateau unit is generally fringed with marine plains consisting of coastal sediments or dune fields lower in the landscape that support small mangroves; higher up are sea cliffs and wavecut platforms. The cliff faces exhibit well-developed laterite profiles. Some of the coastal features in the unit are 5 m above current high-tide levels, which is explained by changes in sea level and upwarping of the islands (Grimes, 1974). Potentially feasible dam sites occur where resistant ridges of rock that have been incised by the river systems outcrop on both sides of river valleys. The rocks are generally weathered to varying degrees, and the depth of weathering, the amount of outcrop on the valley slopes, the occurrence of limestone or dolomitic rocks that may contain solution features that could cause leakage, and the width and depth of alluvium in the base of the valley are fundamental controls on the suitability of the potential dam sites. Where the rocks are relatively unweathered and outcrop on the abutments of the potential dam site, less stripping will be required to achieve a satisfactory founding level for the dam. In general, where stripping removes the more weathered rock, it is anticipated that the Proterozoic sandstones, siltstones, mudstones and conglomerates will form a reasonably watertight dam foundation requiring conventional grout curtains and foundation preparation. However, because dolostones are soluble over a geological timescale, it is possible that, where they occur within the Proterozoic sequences, potentially leaky dam abutments and reservoir rims may be present, which would require specialised and costly foundation treatment such as extensive grouting. The extent and depth of the Cenozoic or Quaternary alluvial sands and gravels in the floor of the valley are also important geological controls on dam feasibility, as these materials will have to be removed to achieve a satisfactory founding level for the dam. 2.2.3 Major hydrogeological provinces of the Southern Gulf catchments In terms of groundwater, five major hydrogeological provinces exist in the Southern Gulf catchments: (i) the McArthur Basin, which underlies the north-west of the catchment, (ii) the Georgina Basin, which overlies the McArthur Basin in the south to south-west of the catchment, (iii) the Isa Highland, (iv) the geological Carpentaria Basin, which rests on an erosional surface of deformed Proterozoic rocks of the Isa Superbasin and South Nicholson Basin, and (v) the Cenozoic Karumba Basin, which unconformably overlies the geological Carpentaria Basin (Figure 2-4) and alluvial plains. The broad major rock types associated with each geological province include igneous and sedimentary rocks (McArthur, South Nicholson, Georgina and geological Carpentaria basins and Isa Superbasin) and unconsolidated (surficial regolith) to consolidated sediments (Karumba Basin) (Figure 2-4). The McArthur Basin is a geological province underlain by an approximately 10 km thick sequence of sedimentary rocks that in places are intruded by minor igneous rocks of Precambrian age (Paleoproterozoic to Mesoproterozoic). The McArthur Basin extends well beyond the Southern Gulf catchments. In the Southern Gulf catchments, the McArthur Basin is undulating with isolated ranges of quartzite and igneous rocks dissected by river valleys. The rocks of the McArthur Basin have been intruded with dolerite, folded, faulted and uplifted, and subjected to long periods of erosion (both physical and chemical weathering) since they were formed. Most of the sedimentary and igneous rocks of the McArthur Basin have very low primary porosity (<2%), and their pores are very small and not interconnected. Consequently, they do not hold or yield much groundwater and can be impermeable across large areas. Where the upper parts of the sedimentary and igneous rocks are weathered and fractured, they can contain volumes of water that, while not large, can have local importance for stock and domestic use and community water supplies. Geological basins and provinces map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-513_Geological_provinces_v04_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-4 Major geological provinces of the Southern Gulf catchments Source: adapted from Raymond (2018) The most well-known and significant intermediate- to regional-scale groundwater systems within the Southern Gulf catchments are hosted by the Georgina and Carpentaria basins, which underlie approximately 12,400 km2 of the Southern Gulf catchments. There may also be other hydrogeological units that are currently poorly characterised but host local- to intermediate-scale groundwater systems with potential to support small-scale localised irrigated agriculture opportunities (i.e. the Proterozoic rocks and the Bulimba Formation). However, data and information for these systems are very limited. The major hydrostratigraphic units hosted by the Georgina Basin within the Southern Gulf catchments are the Camooweal Dolostone, Wonarah Formation and Thorntonia Limestone. Recent stratigraphic drilling conducted as part of the National Drilling Initiative (Exploring for the future, n.d.) identified a thickness of more than 600 m of Georgina Basin sedimentary rocks (mostly limestone and dolostone) within the Southern Gulf catchments. The Camooweal Dolostone is the shallowest of the Georgina Basin’s more prospective hydrogeological units in the Southern Gulf catchments. It is composed of dolostone and dolomitic limestone. The Wonarah Formation of the Georgina Basin also occurs in the south-west portion of the Southern Gulf catchments and merges laterally into the Anthony Lagoon Formation (the Wiso Basin stratigraphic equivalent). It is predominantly composed of silty dolostone with interbedded (minor) dolomitic and calcareous mudstone and siliciclastic mudstone. Very little information exists for the Wonarah Formation in the Southern Gulf catchments, and it therefore remains unclear if it exhibits the same hydraulic characteristics as its stratigraphic equivalents in adjacent areas. The Thorntonia Limestone is composed mostly of dolostone and dolomitic limestone. The limestone aquifers occur mostly in the NT, but within the Southern Gulf catchments they also outcrop or subcrop in far western Queensland and host a significant regional groundwater system. The sedimentary sequences of the geological Carpentaria Basin are mostly sandstone, siltstone, mudstone and claystone. They have a combined thickness of up to approximately 600 m in the Southern Gulf catchments and thicken significantly offshore. Within the Southern Gulf catchments, the geological Carpentaria Basin rocks overlie the formations of the McArthur Basin in the north-east and the Isa Superbasin and South Nicholson Basin in the east and south-east (Figure 2-4). Most of the rocks of the geological Carpentaria Basin have very low primary porosity and are considered aquitards. However, the extensive and porous sandstones of the GRF are considered a productive aquifer and the Normanton Formation is considered an aquifer or partial aquifer. Groundwater within the GRF is under artesian pressure in some parts of the Southern Gulf catchments (e.g. at the Burketown bore). Consisting of lacustrine, fluvial and, to a lesser extent, shallow-shelf marine sediments, the widespread Cenozoic sediments of the Karumba Basin unconformably overlie the Carpentaria Sub-basin of the GAB. Within the Southern Gulf catchments, Cenozoic sediments are expected to reach a maximum thickness of 40 m (north of Doomadgee), and the most productive aquifer is expected to correspond to the basal section of the Bulimba Formation, which is composed mostly of fine-grained quartzose sediments. Overlying the Karumba Basin are the youngest sediments in the catchments: the alluvial sands, silts and gravels associated with the beds, channels and floodplains of the Leichhardt, Nicholson and Gregory rivers, and Settlement Creek and their tributaries. Only a few groundwater bores intersect the youngest alluvial aquifers; these existing bores suggest that the alluvial sediments are generally relatively shallow (drilled to depths of up to approximately 25 m). 30 | Water resource assessment for the Southern Gulf catchments 2.3 Soils of the Southern Gulf catchments 2.3.1 Introduction Soils in a landscape occur as complex patterns resulting from the interplay of five key factors: parent material, climate, organisms, topography and time (Fitzpatrick, 1986; Jenny, 1941). Consequently, soils can be highly variable across a landscape. Different soils have different attributes that determine their suitability for growing different crops and guide how they need to be managed. The distribution of these soils and their attributes closely reflects the geology and landforms of the catchments. Hence, data and maps of soil and soil attributes that provide a spatial representation of how soils vary across a landscape are fundamental to regional-scale land use planning. This section briefly describes the spatial distribution of soil generic groups (Section 2.3.2) and soil attributes (Section 2.3.3) in the Southern Gulf catchments. The management considerations for irrigated agriculture are also summarised in Table 2-1. Maps showing the suitability of different areas for different crops under different irrigation types in different seasons are presented in Chapter 4. Unless otherwise stated, the material in Section 2.3 is based on findings described in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Soils and their attributes were collected and described according to Australian soil survey standards (National Committee on Soil and Terrain, 2009). 2.3.2 Soil characteristics The soils of the Southern Gulf catchments are presented in a soil generic group (SGG) classification (Figure 2-5; Table 2-1; Table 2-2). These groupings provide a means of aggregating soils with broadly similar properties and management considerations. The distinctive groupings have different potential for agriculture: some have almost no potential (e.g. the shallow and/or rocky soils – SGG 7) and some have moderate to high potential (e.g. the cracking clay soils – SGG 9), assuming other factors such as flooding and the amount of salt in the profile are not limiting. The SGGs were designed to simultaneously cover a number of purposes: (i) to be descriptive so as to assist non-expert communication regarding soil and resources, (ii) to be relatable to agricultural potential, and (iii) to align, where practical, to the Australian Soil Classification (ASC) (Isbell and CSIRO, 2016). Soil generic groups were first used in Queensland to facilitate extension in the sugar industry, and they have been modified to suit the range of soils encountered in the Assessment area. The soil groups and soil characteristics presented below can be viewed in the context of their relationship to physiographic units within the catchment (Figure 2-3; Table 2-1). These physiographic units serve as a useful framework to understand the extent of different SGGs because each unit is derived from a distinct group of lithologies and landforms that give rise to a particular set of soil types and geomorphic patterns. Soil generic group map and locations \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\CR-S-512-Ch4_SGG_v1_Arc10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-5 The soil generic groups (SGGs) of the Southern Gulf catchments produced by digital soil mapping The inset map shows the data reliability, which for SGG mapping is based on the confusion index as described in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Labels on the map relate to the locational description of soils later in this Section (2.3.2). Table 2-1 Soil generic groups (SGGs), descriptions, management considerations and correlations to Australian Soil Classification (ASC) for the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au The Southern Gulf catchments contain soils from nine of the ten SGGs (Figure 2-5); peaty soils (SGG 5) are not found. Only two of the nine SGGs occupy more than 10% of the area, and together these soils represent 78% of the study area (Table 2-2). They are the shallow and/or rocky soils, which are associated with uplands and plateaux (SGG 7, 55.9%), and the cracking clay soils found on the Barkly Tableland and the Armraynald Plain (SGG 9, 22.5%). Brown, yellow and grey sandy soils (SGG 6.2) make up 7.6% of the catchment and seasonally or permanently wet soils (SGG 3) another 6.0%, both being common on the Doomadgee Plain. 34 | Water resource assessment for the Southern Gulf catchments Table 2-2 Area and proportions covered by each soil generic group (SGG) in the Southern Gulf catchments SGG SOIL DESCRIPTION AREA (ha) % OF ASSESSMENT AREA (ROUNDED) 1.1 Sand or loam over relatively friable red clay subsoils 49,000 <1 1.2 Sand or loam over relatively friable brown, yellow and grey clay subsoils 59,000 <1 2 Friable non-cracking clay or clay loam soils 378,000 3 3 Seasonally or permanently wet soils 675,500 6 4.1 Red loamy soils 58,300 <1 4.2 Brown, yellow and grey loamy soils 202,100 2 5 Peaty soils 0 0 6.1 Red sandy soils 168,000 2 6.2 Brown, yellow and grey sandy soils 830,300 8 7 Shallow and/or rocky soils 6,057,700 56 8 Sand or loam over sodic clay subsoils 25,200 <1 9 Cracking clay soils 2,434,800 23 10 Highly calcareous soils 1,700 <1 Total 10,820,000 The soils with some of the greatest agricultural potential in the Assessment area are the cracking clays or Vertosols (SGG 9), which cover 2,434,800 ha. Both the Armraynald Plain (C1 in Figure 2-5) and Barkly Tableland (C2) are extensive natural grasslands with few trees, reflecting the cracking nature of the soils (Figure 2-6). These soils are medium to heavy clays that crack when dry and swell when wet, reducing water permeability. They have a self-mulching clay surface, deep to very deep (1.2 to 1.5 m) effective rooting depth, and the clay texture means the soils have a very high (>220 mm) soil available water capacity (AWC). The clays soils of the Armraynald Plain (C1) are imperfectly drained and can have high salt levels within the profile, whereas the Barkly Tableland soils on the uplands (C2) are moderately well drained and gravel is common. On the Armraynald Plain, soils are suited to a variety of vegetables (but not root crops), rice (Oryza spp.), sugarcane (Saccharum officinarum) and dry-season grain, forage, pulse crops, sweetcorn (Zea mays convar. saccharata var. rugosa) and cotton (Gossypium spp.). On the Barkly Tableland, soils are suited to trickle-irrigated mangoes (Mangifera indica) and vegetables as well as wet-season cotton, grain and forage crops. Along the middle reaches of the Leichhardt River (F1 in Figure 2-5), a moderately well-drained, friable non-cracking clay or clay loam soil (SGG 2; 295,460 ha on the mainland) has formed on the floodplains (Figure 2-7). The soils are moderately well to well drained and have a weakly structured, fine sandy to loam surface soil over a structured sandy clay loam or silty clay subsoil. Effective rooting depth is very deep (>1.5 m) and AWC is moderate (>160 mm). On Mornington Island (F2), this soil is shallower with varying amounts of ironstone gravel reducing the AWC. Soil or landscape photo \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\2_Reporting\SGG_Photos For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-6 Cracking clay (brown Vertosol; SGG 9) Mitchell grass (Astrebla spp.) downs with whitewood (Elaeocarpus sp.) and gutta percha (Palaquium spp.) on the Armraynald Plain physiographic unit, east of the Leichhardt River Photo: CSIRO Soil or landscape photo \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\2_Reporting\SGG_Photos For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-7 Brown Dermosol (SGG 2) buffel grass (Cenchrus ciliaris) open woodland with silver leaf box (Eucalyptus pruinosa) on the Armraynald Plain physiographic unit in the middle reaches of the Leichhardt River Photo: CSIRO On the floodplains of the upper lowlands of the Leichhardt River (U1 in Figure 2-5), a well-drained, sandy loam over a structured red clay subsoil (SGG 1.1; 49,100 ha) has formed. This soil also has a very deep (>1.5 m) effective rooting depth and a high (>180 mm) AWC. Sandy soils (SGG 6) are found on the Doomadgee Plain (S1 in Figure 2-5), on Donors Plateau (S2) (Figure 2-8), in the Gulf Fall (S3) and on the elevated terraces north of the Nicholson River (S4) in the Armraynald Plain, covering 996,660 ha on the mainland. The Doomadgee Plain soils (S1) are brown sands (SGG 6.2) that are well drained and commonly limited by ferricrete rock within 0.6 to 1 m of the soil surface. The soil has a very low AWC (25 to 60 mm), depending on soil depth. On Donors Plateau (S2), the yellow sandy soils (SGG 6.2) are deeper and have a high ironstone gravel content throughout the profile reducing AWC. In the Gulf Fall (S3), the brown (SGG 6.2) and red (SGG 6.1) deep sands occur in the subcatchments of Buddycurrawa, Breakfast, Running and Sandy creeks, and the upper parts of the catchment of the Nicholson River. Near the Doomadgee township (S4), soils are predominantly red sands (SGG 6.1) that are well drained but have a low AWC (60 to 100 mm). Soil or landscape photo \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\2_Reporting\SGG_Photos For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-8 Red sandy soil (SGG 6.1) open woodland of Darwin box (Eucalyptus tectifica), bauhinia (Bauhinia spp.) and Cooktown ironwood (Erythrophleum chlorostachys) near Doomadgee on the Armraynald Plain physiographic unit north of the Nicholson River Photo: CSIRO Loamy soils (SGG 4) have formed along the Nicholson River (K1 in Figure 2-5) and on the Doomadgee (K2) and Cloncurry (K3) plains, and in other isolated areas. They are either red (SGG 4.1; 58,300 ha) well drained or brown (SGG 4.2; 202,100 ha) moderately well drained soils, and they have a high AWC (>170 mm). On the Doomadgee Plain along Westmoreland Road (K2) are brown, yellow and grey sandy clay loams over sandy clays (SGG 4.2) that are poorly to moderately well drained. Soils vary in depth depending on the depth of the underlying rock and have a low to moderate AWC (70 to 150 mm). The red loamy soils (SGG 4.1) on the Cloncurry Plain (K3) are shallower and sandier and commonly have gravel and ironstone throughout the profile; accordingly, these soils have a lower AWC. Irrigation potential is limited to spray- and trickle-irrigated crops on the moderately deep to deep (>1 m) soils. Seasonally or permanently wet soils (SGG 3; 648,510 ha on the mainland) occur on local alluvia along creeks and in swamps, particularly between Lilly and Moonlight creeks on the Doomadgee Plain (W1 in Figure 2-5), and the tidal flats and wetlands of the Karumba Plain (W2). The soils are very poorly drained. Texture contrast soils that have sodic clay subsoils (SGG 8; 25,200 ha on the mainland) are minor areas occurring on Nineteen Mile Creek (D1 in Figure 2-5), a tributary of the Leichhardt River, and on a meander plain just upstream of the Gregory–Nicholson River junction (D2). Shallow soils (<0.25 m depth) or rocky soils (>50% rock) (SGG 7; 6,049,320 ha on the mainland) occur extensively in more than half of the Assessment area (Table 2-2), particularly in the mountainous Isa Highland, Gulf Fall, Dissected Barkly Tableland and Donors Plateau physiographic units. More detail on the soils and their relationship to the physiographic units shown in Figure 2-3 can be found in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Figure 2-9 Level to gently undulating cracking clay soils of the Armraynald Plain suitable for broadacre irrigation Photo: CSIRO – Nathan Dyer 2.3.3 Soil attribute mapping Using a combination of field sampling (Figure 2-10) and digital soil mapping techniques, the Assessment mapped 18 attributes affecting the agricultural and aquaculture suitability of soil for the Southern Gulf catchments, as described in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Descriptions and maps are presented below for six key attributes: •surface soil pH •soil thickness •soil surface texture •permeability •available water capacity in the upper 100 cm of the soil profile (referred to as AWC 100) •surface rockiness. An important feature of each predicted attributes map is the companion reliability map showing the relative confidence in the accuracy of the attribute predictions. Note that mapping is only provided here for regional-scale assessment. Areas of high reliability allow users to be more confident in the accuracy of mapping, whereas areas of low reliability show where users should be cautious. Soil or landscape photo \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\2_Reporting\SGG_Photos For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-10 Soil profile of the brown Vertosol (SGG 9) sampled on the Armraynald Plain physiographic unit east of the Leichhardt River Topsoil is on the left-hand side of the lower 1 m of core tray and the subsoil in the top core tray. The sample core was 1.5 m deep. Photo: CSIRO Surface soil pH The pH value of a soil reflects the extent to which the soil is alkaline or acidic, which affects the extent to which nutrients are available to plants for growth. Surface soil pH is the pH in the top 10 cm of the soil. For the majority of plant species, most soil nutrients are available when the pH range is 5.5 to 6.5. Nutrient imbalances are common for soils with pH greater than 8.5 or less than 5.5 and can lead to toxicity problems. However, almost all of the study area is in the pH range 5.5 to 8.5 (Figure 2-11), which is within the acceptable agronomic threshold (Peverill et al., 1999). Areas in the acid-to-neutral range (pH 5.5 to 7.0) are more associated with the sandier surface soils (Figure 2-13) dominated by shallow and/or rocky soils (SGG 7), red sandy soils (SGG 6.1) and brown, yellow and grey sandy soils (SGG 6.2). Sandier soils tend to be more acidic because of increased soil permeability (Figure 2-14) and have higher leaching rates of neutralising soil components. Coarse-textured soils also tend to have lower buffering capacity by virtue of lower cation exchange capacity supplied by clay minerals and organic matter. The more alkaline soils (pH 7.0 to 8.5) are associated with soils with higher clay content, especially the cracking clay soils (SGG 9) and some areas of seasonally or permanently wet soils (SGG 3). Mapping reliability is generally low to moderate across the Assessment area, being lowest in the Dissected Barkly Tableland and the Karumba Plain physiographic units and in areas of the Gulf Fall unit where lack of data produces less reliable results. Mapping reliability is stronger for the Doomadgee Plain physiographic unit, indicating good data correlation. Soil surface pH map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-511-516_DSM_2x1_v2_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-11 (a) Surface soil pH (top 10 cm) of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction Soil thickness Soil thickness defines the potential root space and the amount of soil from which plants obtain their water and nutrients. The deeper soils dominate the lowlands and Wellesley Islands and the western margins of the uplands (Figure 2-12). The lowland deeper soils are associated with cracking clay soils (SGG 9) of the Cloncurry Plain and Armraynald Plain, and on the Doomadgee Plain with significant areas of brown, yellow and grey sandy soils (SGG 6.2), and brown, yellow and grey loamy soils (SGG 4.2). Additionally, the Karumba Plain is dominated by seasonally or permanently wet soils (SGG 3) that are also generally deep. In upland areas, there are extensive areas of cracking clay soils (SGG 9) on the Barkly Tableland physiographic unit, and the Gulf Fall unit has large areas of deep red sandy soils (SGG 6.1) and deep brown, yellow and grey sandy soils (SGG 6.2) on the western margins of the Dissected Barkly Tableland and Gulf Fall physiographic units. Deep friable non-cracking clay or clay loam soils (SGG 2) dominate Mornington Island. Shallow and/or rocky soils (SGG 7) dominate the Isa Highland physiographic unit and the eastern areas of the Dissected Barkly Tableland and Gulf Fall units. Soil thickness mapping is least reliable along the coastal fringes and the Wellesley Islands, as these areas have few data points, and most reliable in the uplands, especially the Isa Highland physiographic unit where there is less variation in soil thickness across the landscape. Soil thickness map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-511-516_DSM_2x1_v2_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-12 (a) Soil thickness of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction Soil surface texture Soil texture refers to the proportion of sand, silt and clay particles that make up the mineral fraction of a soil. Surface texture influences soil water-holding capacity, soil permeability, soil drainage, water and wind erosion, workability and soil nutrient levels. Light-textured soils are generally those high in sand, and heavy-textured soils are dominated by clay. The clayey surface soils represent a large proportion of the Assessment area and dominate the Barkly Tableland, the Dissected Barkly Tableland, the Armraynald Plain, the Karumba Plain and the Donors Plateau physiographic units (Figure 2-13). These soils are mainly cracking clay soils (SGG 9) and seasonally or permanently wet soils (SGG 3). Sandy surface soils co-dominate the area and are mainly represented by shallow and/or rocky soils (SGG 7) and the brown, yellow and grey sandy soils (SGG 6.2) of the Isa Highland and Doomadgee Plain physiographic units. Sandy-textured surface soils like SGG 7 are also strongly represented in the Donors Plateau physiographic unit. Loamy soil surfaces are associated mainly with SGG 7 soils located in the Gulf Fall and Isa Highland physiographic units, although their contribution across the Assessment area is minor. Mapping reliability of soil surface texture tends to be lower in the areas of higher relief than in the plains. The upland Isa Highland and Dissected Barkly Tableland physiographic units have particularly low reliability, reflecting the variability of surface textures across the landscape. Reliability is highest in the Doomadgee Plain and Armraynald Plain physiographic units in the lowlands and the Barkly Tableland physiographic unit in the uplands, reflecting more consistent surface textures within these units. Soil surface texture map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-511-516_DSM_2x1_v2_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-13 (a) Soil surface texture of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction Permeability The permeability of the profile is a measure of how easily water moves through a soil. Flood and furrow irrigation is most successful on soils with low and very low permeability, which reduces root zone drainage (i.e. water passing below the root zone of a plant), rising watertables and nutrient leaching. Spray or trickle irrigation is more efficient than flood and furrow irrigation on soils with moderate to high permeability. The Assessment area is dominated by moderately permeable soils and, to a lesser extent, slowly permeable soils (Figure 2-14). The latter correlate with patterns of cracking clay soils (SGG 9) of the Armraynald Plain and Barkly Tableland physiographic units. Highly permeable soils are associated with brown, yellow and grey loamy soils (SGG 6.2) of the Doomadgee Plain unit, areas of the Gulf Fall unit, and a part of the Cloncurry Plain unit on the eastern margin of the Assessment area dominated by shallow and/or rocky soils (SGG 7). Generally, mapping reliability in the Assessment area is low where areas have limited data and on the Doomadgee Plain where permeability is quite variable across the landscape. Higher map reliability occurs on the Armraynald Plain physiographic unit where soils have a more consistent permeability. Soil permeability map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-511_Perm_2x1_v2_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-14 (a) Soil permeability of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction Available water capacity to 100 cm AWC 100 is the maximum volume of water that the top 100 cm of soil can hold for plant use. The higher the AWC 100 value, the greater the capacity of the soil to supply plants with water. For irrigated agriculture, it is one factor that determines irrigation frequency and the volume of water required to wet up the soil profile. Soils with low AWC 100 require more frequent watering and lower volumes of water per irrigation. For dryland agriculture, AWC 100 determines the capacity of crops to grow and prosper during dry spells. Patterns of AWC 100 (Figure 2-15) closely correlate with soil thickness patterns (Figure 2-12) and surface soil textures (Figure 2-13). While these soil attributes do not definitively indicate subsurface soil textures and consequently water-holding capacity, the correlation shows the largest AWC values are found where soils are deep and are clay rich, especially the Armraynald Plain, Cloncurry Plain and Barkly Tableland physiographic units. Mapping reliability for AWC is generally highest in upland areas of the Gulf Fall and northern areas of Isa Highland physiographic units. It is least reliable in the Barkly Tableland physiographic unit having fewer measurements. Soil available water capacity to 1m map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-511-516_DSM_2x1_v2_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-15 (a) Available water capacity in the upper 100 cm of the soil profile (AWC 100) of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction Rockiness The rockiness of the soil affects agricultural management and the growth of some crops, particularly root crops. Coarse fragments (e.g. pebbles, gravel, cobbles, stones and boulders), hard segregations and rock outcrops in the plough zone can damage and/or interfere with the efficient use of agricultural machinery. Surface gravel, stone and rock are particularly important and can interfere significantly with planting, cultivation and harvesting machinery used for root crops, small crops, annual forage crops and sugarcane. The rocky soil surfaces (Figure 2-16) closely coincide with the shallow and/or rocky soils (SGG 7). Overall, the reliability of mapping is high (good correlation of data as areas are either consistently rocky or consistently rock-free), although relatively localised areas of lower reliability are found in SGG 7 soil areas, reflecting a lack of data. Soil rockiness map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-511-516_DSM_2x1_v2_ArcGIS10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-16 (a) Surface rockiness in soils of the Southern Gulf catchments represented by presence or absence as predicted by digital soil mapping and (b) reliability of the prediction 2.4 Climate of the Southern Gulf catchments 2.4.1 Introduction Weather, which is defined as short-term atmospheric conditions, is the key source of uncertainty affecting hydrology and crop yield. It influences the rate and vigour of crop growth, and catastrophic weather events can result in extensive crop losses. Climate is defined as weather of a specific region averaged over a long period of time. Key climate parameters controlling plant growth and crop productivity include rainfall, temperature, radiation, humidity, and wind speed and direction. Of all the climate parameters affecting hydrology and agriculture in water-limited environments, rainfall is usually the most important. Rainfall is the main determinant of runoff and recharge and is a fundamental requirement for plant growth. For this reason, reporting of climate parameters is heavily biased towards rainfall data. Other climate variables affecting crop yield are discussed in the companion technical report on climate (McJannet et al., 2023). Climate data presented in this report were calculated using SILO (Scientific Information for Land Owners) climate data surfaces (Jeffrey et al., 2001) unless stated otherwise. Very few climate data are available in the region before 1890, therefore the 132-year period from 1 September 1890 to 31 August 2022 is used in the analysis presented below. Unless otherwise stated, the material in Section 2.4 is based on findings described in the companion technical report on climate (McJannet et al., 2023). 2.4.2 Weather patterns over the Southern Gulf catchments The Southern Gulf catchments are characterised by distinctive wet and dry seasons (Figure 2-17) due to their location in the Australian summer monsoon belt. During the build-up months (typically September to December), the Southern Gulf catchments typically experience low-level easterly winds, which can carry pockets of dry or humid air and result in short-lived thunderstorm activity under favourable conditions. Over inland areas of the Southern Gulf catchments, storms form more frequently during the afternoon because of increased air temperature, which enhances instability and leads to convective cloud formation. Storms also form more readily near the heat trough that is a semi-permanent feature over inland Queensland (predominantly located south of the catchment) during late spring and summer months. There is also a high incidence of thunderstorms in the Southern Gulf catchments where sea breeze convergence and/or boundaries act as a trigger. Thunderstorms show a strong diurnal variation, with most occurring during the afternoon and early evening. Dynamic forcing can cause thunderstorms to develop or persist well beyond the normal diurnal cycle, and if the dynamics are strong enough, thunderstorms can occur at any time. In the Southern Gulf catchments, convection and rain from sea breeze convergence typically occur just inland from the coast from early afternoon in the warmer months. Rain can continue overnight and into the following morning if conditions are favourable, particularly if there is a moisture feed from a strong easterly nocturnal jet stream. During the day, the air over land warms and rises, resulting in lower pressure at the surface. Air flows from the Gulf of Carpentaria waters towards the land to fill this area of lower pressure, resulting in the sea breeze, which can converge with the usually south-easterly synoptic winds. During the wet season, the convergence of persistent easterly winds and broader synoptic north-westerly winds over the Gulf of Carpentaria may trigger storms over the coast after sunset. This phenomenon is most pronounced in the near-coastal areas of the catchment. Historical climate, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\1_Climate\4_S_Gulf\1_GIS\1_Map_docs\1_Exports\Cl-SG-507-Hist-MedAnnRF-ET-RFdeficit.png For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-17 Historical rainfall, potential evaporation and rainfall deficit Median (a) annual, (b) wet-season and (c) dry-season rainfall; median (d) annual, (e) wet-season and (f) dry-season potential evaporation; and median (g) annual, (h) wet-season and (i) dry-season rainfall deficit in the Southern Gulf catchments. Rainfall deficit is rainfall minus potential evaporation. The mean annual rainfall, averaged over the Southern Gulf catchments for the 132-year historical period (1 September 1890 to 31 August 2022), is 602 mm. Rainfall totals are highest near the coast and decline in a southerly direction (Figure 2-17). This is because the more northerly regions of the catchments receive more wet-season rainfall as a result of active monsoon episodes. The Southern Gulf catchments are relatively flat, so there is no noticeable topographic influence on climate parameters such as rainfall or temperature. Approximately 94% of the rain falls in the Southern Gulf catchments during the wet-season months (1 November to 30 April). Figure 2-17 shows the spatial distribution of rainfall over the year and during the wet and dry seasons. Median wet-season rainfall exhibits a very similar spatial pattern to median annual rainfall. Median dry-season rainfall is highest in the most southern part of the Leichhardt catchment and lowest near the coast. The highest monthly rainfall totals typically occur during January and February (Figure 2-18). Tropical cyclones and tropical lows contribute a considerable proportion of total annual rainfall, but the actual amount is highly variable from one year to the next (see the companion technical report on climate (McJannet et al., 2023)) since tropical cyclones do not affect the Southern Gulf catchments in more than half of years. For the 53 tropical cyclone seasons from 1969–70 to 2021–22, 60% of seasons registered no tropical cyclones tracking over the Southern Gulf catchments, 36% experienced one tropical cyclone and 4% experienced two (BOM, 2023). 2.4.3 Potential evaporation and potential evapotranspiration Evaporation is the process by which water is lost from open water, plants and soils to the atmosphere; it is a ‘drying’ process. It has become common usage to also refer to this as evapotranspiration. Evaporation primarily affects the potential for irrigation by influencing: • runoff and deep drainage and, hence, the ability to fill water storages (Section 2.5) • crop water requirements (Section 4.3) • losses from water storages (Section 5.3). Potential evaporation (PE), or potential evapotranspiration (PET), is defined as the amount of evaporation that would occur if an unlimited source of water was available. The Southern Gulf catchments have a mean annual PE of 1900 mm (over the period 1890 to 2022) (Figure 2-17). Evaporation is high all year round, but exhibits a strong seasonal pattern, ranging from about 200 mm/month during November and December to about 100 mm/month during the middle of the dry season (June) (Figure 2-19). Preliminary estimates of mean annual (or seasonal) irrigation demand and net evaporation from water storages are sometimes calculated by subtracting the mean annual (or seasonal) PE from the mean annual (or seasonal) rainfall. This is commonly referred to as the mean annual (or seasonal) rainfall deficit (Figure 2-17). The mean annual rainfall deficit, or mean annual net evaporative water loss, in the Southern Gulf catchments is about 1305 mm, and the deficit increases with distance from the coast. Two common methods for characterising climates are the United Nations Environment Program aridity index and the Köppen–Geiger classification (Köppen, 1936; Peel et al., 2007). The aridity index classifies the Southern Gulf catchments as mainly ‘Semi-arid’, and the Köppen–Geiger classification classifies it as ‘Arid hot steppe’ (see the companion technical report on climate (McJannet et al., 2023)). 2.4.4 Variability and long-term trends in rainfall and potential evaporation Climate variability is a natural phenomenon that can be observed in many ways, for example, warmer-than-average dry seasons or low- and high-rainfall wet seasons. Climate variability can also operate over long-term cycles of decades or more. Climate trends represent long-term, consistent directional changes such as warming or increasingly higher mean rainfall. Separating climate variability from climate change is difficult, especially when comparing climate on a year-to-year basis. In the Southern Gulf catchments, 94% of the rain falls during the wet season (November to April). The highest monthly rainfall in the Southern Gulf catchments typically occurs in January or February (Figure 2-18). The months with the lowest rainfall are June to September. In Figure 2-18, the blue shading (labelled ‘A range’) represents the range under Scenario A (i.e. the historical climate from 1 September 1890 to 31 August 2022). The upper limit of the A range is the value at which monthly rainfall (or PE in Figure 2-19) is exceeded during only 10% of years (10% exceedance). The lower limit of the A range is the value at which monthly rainfall (or PE) is exceeded during 90% of years (90% exceedance). The difference between the upper and lower limits of the A range provides a measure of the potential variation in monthly values from one year to the next. PE also exhibits a seasonal pattern (Figure 2-19): mean PE is about 210 mm during December and it is at its lowest during June (100 mm). Months in which PE is high correspond to those months where the demand for water by plants is also high. Mean wet-season and dry-season PE in the Southern Gulf catchments are shown in Figure 2-17. Compared to rainfall, the variation in monthly PE from one year to the next is small (Figure 2-19). Relative to other catchments in southern and northern Australia, the Southern Gulf catchments have high variability in rainfall from one year to the next. This variability is demonstrated for four locations in the Southern Gulf catchments in Figure 2-18. The highest annual rainfall at Mount Isa (948 mm) occurred in the 1894–95 wet season; it was more than seven times the lowest annual rainfall (127 mm in 1925–26) and more than twice the median annual rainfall value (396 mm). The 10-year running mean provides an indication of the sequences of wet or dry years (i.e. variability at decadal timescales). For an annual time series, the 10-year running mean is the mean of the past 10 years of data, including the current year. Using Mount Isa as an example, the 10-year running mean varied from 291 to 562 mm. Figure 2-18 illustrates that the period between 2000 and 2010 was wetter than average. Under Scenario A, PE exhibits much less inter-annual variability than rainfall at the four demonstration locations (Figure 2-19). Rainfall, graph \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\1_Climate\4_S_Gulf\2_Reporting\plots\climate_report\annual_monthly_rainfall_range_4_station_short2cm.png Figure 2-18 Historical monthly rainfall (left) and time series of annual rainfall (right) in the Southern Gulf catchments at Mount Isa, Doomadgee, Gregory and Burketown ‘A range’ is the 10% to 90% exceedance values of monthly rainfall. Note: the ‘A mean’ line is directly under the ‘A median’ line in some months in these figures. The solid blue line in the right column is the 10-year running mean. Potential evaporation, graph \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\1_Climate\4_S_Gulf\2_Reporting\plots\climate_report\annual_monthly_pet_range_4_station_short2cm.png Figure 2-19 Historical monthly potential evaporation (PE) (left) and time series of annual PE (right) in the Southern Gulf catchments at Mount Isa, Doomadgee, Gregory and Burketown ‘A range’ is the 10% to 90% exceedance values of monthly rainfall. Note: the ‘A mean’ line is directly under the ‘A median’ line in these figures. The solid blue line in the right column is the 10-year running mean The coefficient of variation (CV) provides a measure of the variability of rainfall from one year to the next. CV is calculated as the standard deviation of mean annual rainfall divided by the mean annual rainfall, and the larger the CV value, the larger the variation in annual rainfall relative to a location’s mean annual rainfall. Figure 2-20a shows the CV of annual rainfall for rainfall stations with a long-term record around Australia. Figure 2-20b shows that the inter-annual variation in rainfall in the Southern Gulf catchments is moderately high for northern Australia catchments and high compared to stations in southern Australia with similar mean annual rainfall. (a) (b) \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\1_Climate\4_S_Gulf\1_GIS\1_Map_docs\1_Exports\Cl-SG-516_Cv_map_of_selected_stations_v1_1031.png For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au \\fs1-cbr\{lw-rowra}\work\1_Climate\1_All\2_Reporting\NAWRA2-TR-Cl-A-WB1-v11.xlsm For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-20 (a) Coefficient of variation (CV) of annual rainfall and (b) the CV of annual rainfall plotted against mean annual rainfall for 99 rainfall stations around Australia (a) The grey polygon indicates the extent of the Southern Gulf catchments. (b) The rainfall station in the Southern Gulf catchments is indicated by a red symbol. The light blue diamonds indicate rainfall stations from the rest of northern Australia (RoNA), and hollow squares indicate rainfall stations from southern Australia (SA). Furthermore, Petheram et al. (2008) observed that the inter-annual variability of rainfall in northern Australia is about 30% higher than that observed at rainfall stations from the rest of the world for the same type of climate as northern Australia. Hence, caution should be exercised before drawing comparisons between the agricultural potential of the Southern Gulf catchments and other parts of the world with a similar climate. Several factors are driving this high inter-annual variation in Australia’s climate, including the El Niño – Southern Oscillation (ENSO), the Indian Ocean Dipole, the Southern Annular Mode, the Madden–Julian Oscillation and the Interdecadal Pacific Oscillation. Of these influences, the ENSO phenomenon is considered to be the primary source of global climate variability over the 2- to 6-year timescale (Rasmusson and Arkin, 1993), and it is reported to be a significant cause of climate variability for much of eastern and northern Australia. One of the modes of ENSO, El Niño, has come to be a term synonymous with drought in the western Pacific and eastern and northern Australia (though El Niño does not necessarily mean a ‘drought’ will occur). Rainfall stations along eastern and northern Australia have been observed to have a strong correlation (0.5 to 0.6) with the Southern Oscillation Index (SOI), a measure of the strength of ENSO, during spring, suggesting that ENSO plays a key role in between-year rainfall variability (McBride and Nicholls, 1983). 52 | Water resource assessment for the Southern Gulf catchments Another known impact of ENSO in northern Australia is the tendency for the onset of useful rains after the dry season to be earlier than normal in La Niña years and later than normal in El Niño years. For all years between 1890 and 2022, the mean rainfall onset date (defined as being the date on which 50 mm of rain has accumulated after the dry season) for the Southern Gulf catchments is the last 10 days of October (see the companion technical report on climate (McJannet et al., 2023)). The mean SOI for the September to December period in each year was used to determine whether given years were in negative (SOI < −8, El Niño), positive (SOI > 8, La Niña), or neutral SOI (−8 < SOI < 8). Using this method, in El Niño, neutral and La Niña years, respectively, the median rainfall onset dates for the Southern Gulf catchments are the middle of January, early January and early December. Trends Previously, CSIRO (2009a) found that rainfall in northern Australia between 1997 and 2007 was statistically different to that between 1930 and 1997. In other work, Evans et al. (2014) found a strong relationship between monsoon active periods and the Madden–Julian Oscillation, and that the increasing rainfall trend observed at Darwin Airport was related to increased frequency of active monsoon days rather than increased intensity during active periods. Runs of wet and dry years The rainfall-generating systems in northern Australia and their modes of variability combine to produce irregular runs of wet and dry years. In particular, length and magnitude (intensity) of dry spells strongly influence the scale, profitability and risk of water-resource-related investments. The Southern Gulf catchments are likely to experience dry periods of similar severity to many centres in the Murray–Darling Basin and on the east coast of Australia. The Southern Gulf catchments are characterised by irregular periods of consistently low rainfall when successive wet seasons fail, in addition to the typical annual dry season. Runs of wet years and dry years occur when consecutive years of rainfall occur that are above or below the median, respectively. These are shown in Figure 2-21 at Mount Isa, Doomadgee, Gregory and Burketown stations as annual differences from the median rainfall. A run of consistently dry years may be associated with drought (though an agreed definition of drought continues to be elusive). Analysis of annual rainfall at stations in the Southern Gulf catchments indicates equally long runs of wet and dry years and nothing unusual about the length of the runs of dry years. A graph of numbers and lines Description automatically generated with medium confidence Figure 2-21 Runs of wet and dry years at Mount Isa, Doomadgee, Gregory and Burketown Wet years are shown by the blue columns and dry years by the red columns. Palaeoclimate records for northern Australia The instrumental record of climate data is very short in a geological sense, particularly in northern Australia, so a brief review of palaeoclimate data is provided. The literature indicates that atmospheric patterns approximating the present climate conditions in northern Australia (e.g. the Pacific circulation responsible for ENSO) have been in place since about 3 to 2.5 Ma (Bowman et al., 2010). This suggests many ecosystems in northern Australia have experienced monsoonal conditions for many millions of years. However, past climates have been both wetter and drier than the instrumental record for northern Australia, and the influence of ENSO has varied considerably over recent geological time. Several authors have found that present levels of tropical cyclone activity (i.e. over the instrumental record) in northern Australia are low (Denniston et al., 2015; Forsyth et al., 2010; Nott and Jagger, 2013) and possibly unprecedented over the past 550 to 1500 years (Haig et al., 2014). Furthermore, the recurrence frequencies of high-intensity tropical cyclones (Category 4 to Category 5 events) may have been an order of magnitude higher than that inferred from the current short instrumental records. 2.4.5 Changes in rainfall and evaporation under a future climate The effects of projected climate change on rainfall and PE are presented in Figure 2-22, Figure 2-23 and Figure 2-24. This analysis used 32 global climate models (GCMs) to represent a world where the global mean surface air temperatures are 1.6 °C higher than approximate 1990 global temperatures. This emission scenario is referred to as SSP2-4.5 (IPCC, 2022) and in this report as Scenario C. SSP2-4.5 is the most likely future climate scenario according to Hausfather and Peters (2020). Because the scale of GCM outputs is too coarse for use in catchment- and point-scale hydrological and agricultural computer models, they were transformed to catchment-scale variables using a simple scaling technique (PS, pattern scaled) and are referred to as GCM-PSs. See the companion technical report on climate (McJannet et al., 2023) for further details. In Figure 2-22 the rainfall and PE projections of the 32 GCM-PSs are spatially averaged across the Southern Gulf catchments, and the GCM-PSs are ranked in order of increasing mean annual rainfall. This figure shows that three (10%) of the projections for GCM-PSs indicate an increase in mean annual rainfall by more than 5%, seven (22%) of the projections indicate a decrease in mean annual rainfall by more than 5%, and about two-thirds of the projections indicate a change in future mean annual rainfall of less than 5% under a 1.6 °C warming scenario. Hence, it can be argued that, based on the selected 32 GCM-PSs, the consensus result is that mean annual rainfall in the Southern Gulf catchments is not likely to change significantly under Scenario C. The spatial distribution of mean annual rainfall under Scenario C is shown in Figure 2-23. In this figure, only the third-wettest GCM-PS (i.e. 10% exceedance or Scenario Cwet), the middle (or 11th-wettest) GCM-PS (i.e. Scenario Cmid), and the third-driest GCM-PS (i.e. 90% exceedance or Scenario Cdry) are shown. Figure 2-24a shows mean monthly rainfall under scenarios A and C. The data suggest that mean monthly rainfall under Scenario Cmid will be similar to mean monthly rainfall under Scenario A. Under scenarios Cwet, Cmid and Cdry, the seasonality of rainfall in northern Australia is similar to that under Scenario A. Change in rainfall and evaporation, graph \\fs1-cbr\{lw-rowra}\work\1_Climate\4_S_Gulf\2_Reporting\plots\future_climate\mean.annual.percentage.change.per.degree.SSP245.png For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-22 Percentage change in rainfall and potential evaporation per degree of global warming for the 32 Scenario C simulations relative to Scenario A values for the Southern Gulf catchments GCM-PS ranked by increasing rainfall for SSP2-4.5. Scenario rainfall, map \\fs1-cbr\{lw-rowra}\work\1_Climate\4_S_Gulf\1_GIS\1_Map_docs\1_Exports\Cl-SG-514-annualRain-Cwet-Cmid-Cdry.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-23 Spatial distribution of mean annual rainfall across the Southern Gulf catchments under scenarios (a) Cwet, (b)Cmid and (c) Cdry "\\fs1-cbr\{lw-rowra}\work\1_Climate\4_S_Gulf\3_future_climate\summary\ssp245\mean_month_rain_pet.png" For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-24 (a) Monthly rainfall and (b) potential evaporation for the Southern Gulf catchments under scenarios A and C ‘C range’ is based on the computation of the 10% and 90% monthly exceedance values, so the lower and upper limits in ‘C range’ are not the same as scenarios Cdry and Cwet. Note: the ‘A mean’ line is directly under the ‘Cmid’ line in (a). Potential evaporation The mean annual change in GCM-PS PE shows projected PE increases of about 2% to 9% (Figure 2-22). Under scenarios Cwet, Cmid and Cdry, PE exhibits a similar seasonality to that under Scenario A (Figure 2-24b). However, different methods of calculating PE give different results. Consequently, there is considerable uncertainty as to how PE may change under a warmer climate. See Petheram et al. (2012) and Petheram and Yang (2013) for more detailed discussions. Sea-level rise and sea-surface temperature projections Global mean sea levels have risen at a rate of 1.7 ± 0.2 mm/year between 1900 and 2010, a rate in the order of ten times faster than the preceding century. Australian tide gauge trends are similar to the global trends (CSIRO and Bureau of Meteorology, 2015). Sea-level projections for the Southern Gulf catchments are summarised in Table 2-3. This information may be considered in coastal aquaculture developments and flood inundation of coastal areas. Table 2-3 Projected sea-level rise for the coast of the Southern Gulf catchments Values are the median of Coupled Model Intercomparison Project (CMIP) Phase 5 global climate models (GCMs). Numbers in parentheses are the 5% to 95% range of the same. Projected sea-level rise values are relative to a mean calculated between 1986 and 2005. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au RCP = Representative Concentration Pathway Source: CoastAdapt (2017) Sea-surface temperature increases around Australia are projected with very high confidence for all emissions scenarios, which show warming of around 0.4 to 1.0 °C in 2030 under Representative Concentration Pathway (RCP) 4.5, and 2 to 4 °C in 2090 under RCP 8.5, relative to a 1986 to 2005 baseline (CSIRO and Bureau of Meteorology, 2015). There will be regional differences in sea-surface temperature warming due to local hydrodynamic responses; however, there is only medium confidence in coastal projections as climate models do not resolve local processes (CSIRO and Bureau of Meteorology, 2015). For the Southern Gulf catchments, the corresponding projected sea-surface temperature increases are 0.7 °C (range across climate models is 0.5 to 1.0 °C) in 2030 under RCP 4.5 and 2.9 °C (2.4 to 3.9 °C) in 2090 under RCP 8.5. These changes are relative to a 1986 to 2005 baseline (CSIRO and Bureau of Meteorology, 2015). Note that the data in Table 2-3 use the CMIP5 dataset to provide estimates of sea-level rise. An updated product that uses CMIP6 was not available at the time of writing. 2.4.6 Establishment of an appropriate hydroclimate baseline The allocation of water and the design and planning of water resources infrastructure and systems require great care and consideration and need to be based on scientific evidence and take a genuine long-term view. A hydroclimate baseline from 1890 to 2022 (i.e. current) was deemed the most suitable baseline for the Assessment. A poorly considered design can result in an unsustainable system or preclude the development of a more suitable and possibly larger system, thus adversely affecting existing and future users, industries and the environment. Once water is overallocated, it is economically, financially, socially and politically difficult to reduce allocations in the future, unless water allocations are only assigned over short time frames (e.g. <15 years) and then reassessed. However, many water resource investments, particularly agricultural investments, require time frames longer than 30 years as there are often large initial infrastructure costs and a long learning period before full production potential is realised. Consequently, investors require certainty that, over their investment time frame (and potentially beyond), their access to water will remain at the level of reliability initially allocated. A key consideration in developing a water resource plan, or in assessing the water resources of a catchment, is the time period over which the water resources will be analysed, also referred to as the hydroclimate ‘baseline’ (e.g. Chiew et al., 2009). If the hydroclimate baseline is too short, it can introduce biases in a water resource assessment for various reasons. First, the transformation of rainfall to runoff and rainfall to groundwater recharge is non-linear. For example, averaged across the catchment of the Flinders River in northern Australia, the mean annual rainfall is only 8% higher than the median annual rainfall, yet the mean annual runoff is 59% higher than the median annual runoff (Lerat et al., 2013). Similarly, the median annual rainfall between 1895 and 1945 was the same as the median annual rainfall between 1948 and 1987 (less than 0.5% difference), yet there was a 21% difference in the median annual runoff between these two time periods (and a 40% difference in the mean annual runoff) (Lerat et al., 2013). Consequently, great care is required if using rainfall data alone to justify the use of short periods over which to analyse the water resources of a catchment. In developing a water resource plan, the volume of water allocated for consumptive purposes is usually constrained by the drier years (referred to as dry spells when consecutive dry years occur) in the historical record (see Section 2.4.4). This is because it is usually during dry spells that water extraction most adversely affects existing industries and the environment. All other factors (e.g. market demand, interest rates) being equal, consecutive dry years are usually also the most limiting time periods for new water resource developments and/or investments, such as irrigated agricultural enterprises, particularly if the dry spells coincide with the start of an investment cycle. Consequently, it is important to ensure a representative range of dry spells (i.e. of different durations, magnitudes and sequencing) are captured over the Assessment time period. For example, two time periods may have very similar median annual runoffs, but the duration, magnitude and sequencing of the dry spells may be sufficiently different that they pose different risks to investors and result in different modelled ecological outcomes. In those instances where there is the potential for a long memory, such as in intermediate- and regional-scale groundwater systems or in river systems with large reservoirs, long periods of record are preferable to minimise the influence of initial starting conditions (e.g. assumptions regarding initial reservoir storage volume), to properly assess the reliability of water supply from large storages and to encapsulate the range of likely conditions (McMahon and Adeloye, 2005). All these arguments favour using as long a time period as practically possible. However, in some circumstances a shorter period may be preferable on the basis that it is a more conservative option. For example, in south-western Australia, water resource assessments to support water resource planning are typically assessed from 1975 onwards (Chiew et al., 2012; McFarlane et al., 2012). This is because there has been a marked reduction in runoff in south-western Australia since the mid-1970s, and this declining trend in rainfall is consistent with the majority of GCM projections, which project reductions of rainfall into the future (McJannet et al., 2023). Although there were few rainfall stations in the study area at the turn of the 20th century relative to 2019 (McJannet et al., 2023), an exploratory analysis of rainfall statistics of the early period of the instrumental record does not appear to be anomalous when compared to the longer-term instrumental record. In deciding upon an appropriate time period over which to analyse the water resources of the Southern Gulf catchments, consideration was given to the above arguments, as well as to palaeoclimate records, observed trends in the historical instrumental rainfall data and future climate projections. For the Southern Gulf catchments, although there is evidence of an increasing trend in rainfall in the recent instrumental record, two-thirds of the GCM-PSs project no change in mean annual rainfall for a 1.6 °C warming scenario. Furthermore, palaeoclimate records indicate multiple wetter and drier periods have occurred in the recent geological past (Northern Australia Water Resource Assessment technical report on climate (Charles et al., 2017)). Very few climate data are available in the Southern Gulf catchments before 1890, so the baseline adopted for this Assessment was 1890 to 2022. Note, however, that as climate is changing on a variety of timescales, detailed scenario modelling and planning (e.g. the design of major water infrastructure) should be broader than just comparing a single hydroclimate baseline to an alternative future. 2.5 Hydrology of the Southern Gulf catchments 2.5.1 Introduction The timing and event-driven nature of rainfall events and high PE rates across the Southern Gulf catchments have important consequences for the catchments’ hydrology. The spatial and temporal patterns of rainfall and PE across the Southern Gulf catchments are discussed in Section 2.4. Rainfall can be broadly broken into evaporated and non-evaporated components (the latter is also referred to as ‘excess water’). The non-evaporated component can be broadly broken into overland flow and recharge (Figure 2-25). Recharge replenishes groundwater systems, which in turn discharge into rivers and the ocean. Overland flow and groundwater discharged into rivers combine to become streamflow. Streamflow in the Assessment is defined as a volume per unit of time. Runoff is defined as the millimetre depth equivalent of streamflow. Flooding is a phenomenon that occurs when the flow in a river exceeds the river channel’s capacity to carry the water, resulting in water spilling onto the land adjacent to the river. Water balance, diagram \\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\10_Reporting\1_All\9_Graphics_artist\2_SoG\C Petheram Southern Gulf 1_7_2024 Waterbalance Chpter 2.jpg For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-25 Simplified schematic diagram of terrestrial water balance in the Southern Gulf catchments Runoff is the millimetre depth equivalent of streamflow. Overland flow includes shallow subsurface flow. Numbers indicate mean annual values spatially averaged across the catchment under Scenario A. Numbers will vary locally. Section 2.5 covers the remaining terms of the terrestrial water balance (accounting for water inputs and outputs) of the Southern Gulf catchments, with particular reference to the processes and terms relevant to irrigation at the catchment scale. Information is provided on groundwater, groundwater recharge and surface water – groundwater connectivity. Runoff, streamflow, flooding and persistent waterholes in the Southern Gulf catchments are then discussed. Figure 2-25 shows a schematic diagram of the water balance of the Southern Gulf catchments, along with estimates of the mean annual value spatially averaged across the catchment and an estimate of the uncertainty for each term. The ‘water balance’ comprises all the water inflows and outflows to and from a particular catchment over a given time period. Unless stated otherwise, the material in sections 2.5.2 to 2.5.4 is based on findings described in the companion technical report on hydrogeological assessment (Raiber et al., 2024). Similarly, the material in Section 2.5.5 draws on the findings of the companion technical report on river modelling (Gibbs et al., 2024), unless stated otherwise. 2.5.2 Groundwater Within the Southern Gulf catchments, the distribution, availability and quality of groundwater resources are heavily influenced by the physical characteristics of the sediments and rocks of the major geological divisions (see Section 2.2). Aquifers are the rocks and sediments in the subsurface that store and transmit groundwater. The catchments have several types of aquifer: • fractured and weathered rocks associated with Proterozoic basins (associated with the Proterozoic McArthur and South Nicholson basins and Isa Superbasin) (mostly covered by younger strata) • fractured, fissured and karstic carbonate rocks of the Georgina Basin (Cambrian limestone and Cambrian dolostone in Figure 2-26) • extensive porous sedimentary sandstones of the Great Artesian Basin (not present at the surface in the Southern Gulf catchments; for extent, see Figure 2-4 and Figure 2-31) • porous sandstones of the Karumba Basin (Cenozoic sediments, Figure 2-4) • surficial unconsolidated to consolidated alluvial sands and gravels (Cenozoic alluvium in Figure 2-26). The sedimentary limestones of the Georgina Basin – in particular, the Camooweal Dolostone and Thorntonia Limestone – and the aquifers of the Carpentaria Sub-basin of the GAB host the largest, most regionally extensive groundwater resource in the Southern Gulf catchments. The upper fractured, fissured and karstic parts of the Cambrian carbonate rocks of the Georgina Basin (Figure 2-4 and Figure 2-31) in the Southern Gulf catchments may be regionally connected to the limestones of the adjacent Daly and Wiso basins, which combined are often referred to as the Cambrian Limestone Aquifer. The Cambrian Limestone Aquifer extends for several hundred thousand square kilometres to the north, west and south of the Southern Gulf catchments. It hosts limestones that form a complex, interconnected and highly productive regional-scale groundwater system. That is, the distance between the recharge areas (where there is inflow of water through the soil, past the root zone and into an aquifer) and discharge areas (where there is outflow of water from an aquifer into a water body or as evaporation from the soil or vegetation) can be tens of kilometres to hundreds of kilometres, and the time taken for groundwater to discharge following recharge can be in the order of thousands to hundreds of thousands of years. Simplified regional hydrogeology map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-515_simplified_regional_hydrogeology_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-26 Simplified regional hydrogeology of the Southern Gulf catchments This map does not represent outcropping areas of all hydrogeological units: the blanket of surficial Cretaceous to Quaternary regolith sediments has been removed to highlight the spatial extent of various regional hydrogeological units in the subsurface. Geology data sources adapted from: Department of Industry, Tourism and Trade (2014) and Department of Environment, Parks and Water Security (2008) and Raymond et al. (2012) Spring data sources: Department of Environment, Parks and Water Security (2013); the Groundwater Dependent Ecosystems Atlas (Bureau of Meteorology, 2017) and Queensland Government (2021) The porous sandstones of the geological Carpentaria Basin (Figure 2-4) – in particular the Gilbert River Formation (GRF) and to a lesser extent the Normanton Formation – host the most regionally extensive aquifer systems within the Southern Gulf catchments. They extend over the eastern part of the catchment in Queensland and are entirely covered by the Cenozoic Karumba Basin. The aquifers of the geological Carpentaria Basin form part of the 1.7 million km2 large GAB that reaches very far to the south and east of the Southern Gulf catchments (Figure 2-31) and also includes the Eromanga and Surat basins and parts of the Clarence–Moreton Basin. Within the Southern Gulf catchments, the geological Carpentaria Basin is entirely covered by the Cenozoic Karumba Basin. Recharge and discharge processes within the geological Carpentaria Basin are relatively poorly characterised due to the relatively limited number of groundwater monitoring bores. The near-absence of environmental tracer data in the geological Carpentaria Basin was identified as a key data gap within the GAB in a recent GAB-wide environmental tracer study (Raiber et al., 2022). Other more local- to intermediate-scale aquifer systems may exist in other geological formations within the Southern Gulf catchments. Previous studies described the Proterozoic rocks trending north-west to south-east across the central part of the Southern Gulf catchments (Figure 2-2) as not a feasible groundwater resource. However, other studies suggested that the Mount Isa town water supply was sourced from fractured shales until the construction of the Lake Moondarra dam. These shales occur at depths of 60 to 80 metres below ground level (mBGL), and bores yielded 5 to 10 L/second. In addition, approximately 50% of the bores with stratigraphic information in the Southern Gulf catchments correspond to Proterozoic units composed of unassigned granites or are attributed to one of the geological units of the South Nicholson Basin and Isa Superbasin. Furthermore, the variability of groundwater chemistry found among the Proterozoic aquifers, with dominant ions varying from magnesium, calcium and bicarbonate (Mg–Ca–HCO3) to sodium and sulfate (Na–SO4), may be an indication of localised recharge and interactions of groundwater with variable mineral-rich rocks. The surficial fluvial and, to a lesser extent, shallow-shelf marine Cenozoic sediments of the Karumba Basin and younger alluvial aquifer systems in the Southern Gulf catchments may form local- to intermediate-scale aquifer systems. The basal Bulimba Formation has been described as the most productive aquifer of the Karumba Basin. Its lithological properties have been described as highly variable, from shale to sandy ferricrete, resulting in variable hydraulic properties and groundwater yields. Generally, the sedimentary sequences of the Karumba Basin and younger alluvial aquifer systems within the Southern Gulf catchments remain poorly characterised. Hydrogeological units Hydrogeological units of the Southern Gulf catchments are shown in Figure 2-26. The rocks and sediments of these geological units host a diverse range of aquifers that vary in extent, storage and productivity. The major and most extensive aquifers in the Southern Gulf catchments are found in the Cambrian limestone of the Georgina Basin and the sandstones of the Jurassic to Cretaceous Carpentaria Sub-basin of the GAB. For this Assessment, major aquifer systems are defined as aquifers that contain regional and intermediate-scale groundwater systems with adequate storage volumes (i.e. gigalitres) that could potentially yield water at a sufficient rate (>10 L/second) and be of a sufficient water quality (<1000 mg/L total dissolved solids (TDS)) for a range of irrigated cropping. Minor aquifers are defined as aquifers that contain local-scale groundwater systems with lower storage (i.e. megalitres) (Figure 2-27). The yields from minor aquifers are variable and often low (<5 L/second), and minor aquifers have variable water quality ranging from fresh (~500 mg/L TDS) to brackish (~3000 mg/L TDS). The distribution and characteristics of these rocks is covered in Section Unless otherwise stated, the material in Section 2.5.2 is based on findings described in the companion technical report on hydrogeological assessment by Raiber et al. (2024). Only the major aquifers relevant to potential opportunities for future groundwater resource development are discussed in detail. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-27 Two-dimensional conceptual schematic of the interconnected aquifer system and its variability Bore yields vary significantly depending on where (which geological unit) a bore is installed and at what depth. Adapted from DENR (2016) Cambrian limestone aquifers The Cambrian limestone aquifers of the Georgina Basin occur in the western part of the Southern Gulf catchments (Figure 2-26). Due to the limited number of groundwater bores within the Southern Gulf catchments, data from groundwater bores outside the boundary of the catchment were also used to provide a more robust overview of the hydrogeological properties of the different formations. Camooweal Dolostone and Thorntonia Limestone – Georgina Basin Based on existing hydrogeological data, the fractured and karstic carbonated rocks hosted in the Georgina Basin also offer potential to support productive bores (i.e. yields greater than 10 L/s), but they remain poorly characterised. The Camooweal Dolostone, which occurs in the south-west of the catchments (Cambrian dolostone in Figure 2-26), is mostly composed of dolostone and dolomitic limestone (Geoscience Australia and Australian Stratigraphy Commission, 2021; Matthews, 1992). It is the shallowest of the Georgina Basin’s more prospective hydrogeological units in the Southern Gulf catchments and has an approximate spatial outcrop extent of approximately 2972 km2 within the Southern Gulf catchments. At a new stratigraphic well (Carrara 1) drilled as part of the National Drilling Initiative within the Southern Gulf catchments, the thickness of the Camooweal Dolostone is more than 200 m. The Thorntonia Limestone occurs in the south-west of the Southern Gulf catchments (Cambrian limestone in Figure 2-26) and is composed mostly of dolostone and dolomitic limestone (Geoscience Australia and Australian Stratigraphy Commission, 2021). It is overlain by the Camooweal Dolostone and is confined in parts by the Wonarah Formation. The limestone aquifers occur mostly in the NT, but also outcrop and subcrop in far western Queensland, and host a significant regional groundwater system (Figure 2-26). The Thorntonia Limestone has an outcropping zone of approximately 2515 km2 within the Southern Gulf catchments and is up to approximately 100 m thick (CSIRO, 2009b; Taylor et al., 2021). For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-28 Gregory River at Riversleigh Road Photo: CSIRO – Russell Crosbie The median depth of groundwater bores screened within the Cambrian limestones (Camooweal Dolostone and Thorntonia Limestone, undifferentiated) in the Southern Gulf catchments is 101 m (with a range from 24 to 537 m). Groundwater within these formations is mostly fresh to brackish with a median salinity of approximately 1300 mg/L TDS (ranging from <100 to approximately 6,000 mg/L TDS). The ionic composition of groundwater within the Cambrian limestones is variable, with a predominance of calcium, magnesium and bicarbonate (Ca–Mg–HCO3) or sodium, bicarbonate and chloride (Na–HCO3–Cl For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-29 Lawn Hill Gorge Photo: CSIRO – Russell Crosbie Wonarah Formation – Georgina Basin The Wonarah Formation of the Georgina Basin occurs in the south-west portion of the Southern Gulf catchments (Cambrian siltstone in Figure 2-26) and merges laterally into the Anthony Lagoon Formation (the Wiso Basin stratigraphic equivalent) (Tickell and Bruwer, 2017). The Wonarah Formation is predominantly composed of silty dolostone with interbedded (minor) dolomitic and calcareous mudstone and siliciclastic mudstone. It has a maximum thickness of 244 m in holes drilled in its equivalent Anthony Lagoon Formation (Kruse and Dunster, 2013). Very little information exists for the Wonarah Formation in the Southern Gulf catchments. It therefore remains unclear if it exhibits the same characteristics as the Anthony Lagoon Formation in the Wiso Basin much further to the west in the NT. These formations are fractured, highly heterogeneous and partially karstic, with estimated transmissivities of 13 to 8200 m2/day (Kruse and Dunster, 2013). Indicative hydrogeological data and anecdotal evidence for aquifers in the Wonarah Formation and equivalent come from Tickell (2003). Tickell (2003) described the hydrogeological unit in the Barkly Tableland as fractured and cavernous rocks assigned to both the Georgina and Wiso basins. Aquifers range in depth from 50 to 125 mBGL, and standing water levels vary from 30 to 100 mBGL. Indicative bore yield data based on airlifting at the time of drilling varied from 0.5 to 5 L/second. However, mud circulation losses during drilling were commonly reported, suggesting the presence of cavernous limestones and potential for high-yielding aquifers. Groundwater quality was also described as fresh to brackish (500 to 1500 mg/L TDS) across most of the area investigated (Tickell, 2003). Gilbert River Formation – Carpentaria Sub-basin of the Great Artesian Basin The Gilbert River Formation (GRF), or Gilbert River Aquifer (GRA), is the major aquifer within the Jurassic to Cretaceous sedimentary rocks of the geological Carpentaria Basin in the north-east of the catchments (Figure 2-31). The extensive sandstone aquifers are intersected by bores ranging in depth from approximately 100 to 750 mBGL. The GRF in the Southern Gulf catchments and adjacent areas has significant potential as an aquifer for groundwater-based irrigation. Hydraulic conductivity values for the GRF in the northern geological Carpentaria Basin range from 0.1 to 10 m/day, and transmissivities range from 4 to 570 m2/day (Horn et al., 1995; Klohn Crippen Berger, 2016). Reported bore yields range from less than 0.5 L/second up to 46 L/second (Figure 2-32). It is anticipated that yields would likely exceed 10 L/second with appropriately constructed large-diameter production bores. The GRF is confined by the thick and laterally continuous aquitard of the Rolling Downs Group. In some places this results in artesian conditions as, for example, reported for a bore in Burketown. Groundwater salinity in the GRF ranges from fresh (<500 mg/L TDS) to brackish (>3000 mg/L TDS) with a median TDS of 500 mg/L (Figure 2-33), which is still suitable for irrigation use. The ionic composition of the groundwater of the GRF is dominated by Na–HCO3–Cl and Na–Cl–HCO3, which is indicative of mature groundwaters evolved along very long flow paths. Low calcium and magnesium and high concentrations of fluoride have also been observed. An assessment of recharge processes and flow dynamics in the GAB by Raiber et al. (2022) identified the lack of environmental tracer data in the geological Carpentaria Basin as a key data gap within the GAB. Figure 2-30 The Gregory River receives groundwater discharge from the Cambrian Limestone Aquifer Photo: CSIRO – Nathan Dyer Full extent of Basins and GAB, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-559_Georgina_GAB_v03_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-31 Full extent of the Georgina Basin and Carpentaria Sub-basin of the Great Artesian Basin. Inset map shows full extent of Great Artesian Basin Geology data sources adapted from: Department of Industry, Tourism and Trade (2014) and Department of Environment, Parks and Water Security (2008) and Raymond et al. (2012) Major aquifer bore yield, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-525_Yield_by_aquifer_Major_v06.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-32 Groundwater bore yields for major aquifers across the Southern Gulf catchments Symbol shapes indicate the aquifer within which the bore is sited; colours indicate bore yield class. Bore yield data source: Department of Environment, Parks and Water Security (2014); DNRME (2023) Major aquifer groundwater salinity, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-527_TDS_by_aquifer_Major_v06.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-33 Groundwater salinity for major aquifers in the Southern Gulf catchments Symbol shapes indicate the aquifer within which the bore is sited; colours indicate total dissolved solids (TDS). Salinity data source: Department of Environment, Parks and Water Security (2014); DNRME (2023) Bulimba Formation and alluvial aquifer systems Consisting of lacustrine, fluvial and, to a lesser extent, shallow-shelf marine sediments, the widespread Cenozoic sediments of the Karumba Basin unconformably overlie the Carpentaria Sub- basin of the GAB (Bradshaw et al., 2009). Within the Southern Gulf catchments, Cenozoic sediments are expected to reach a maximum thickness of 40 m (north of Doomadgee). The most productive aquifer is expected to correspond to the basal section of the Bulimba Formation, which is mostly composed of fine-grained quartzose sediments (Herbert, 2000; Radke et al., 2012). Lithological properties of the Bulimba Formation are highly variable, from shale to sandy ferricrete, resulting in variable hydraulic properties and groundwater yields (i.e. 0.25 to 4.5 L/second) (DNRME, 2021). Nevertheless, at the central area of Nicholson catchment, where the alluvial deposits are wider (downstream of Doomadgee) (Figure 2-26), the Bulimba Formation is the most accessed aquifer from the Karumba Basin (Buchanan et al., 2020). Its hydraulic conductivity ranges from 150 to 300 m/day with specific yield of 0.1 (Smerdon et al., 2012). The Cenozoic alluvial aquifers in the middle to lower reaches of the Nicholson, Gregory and Leichhardt rivers, and Settlement Creek and its tributaries, host local-scale groundwater systems. However, data are very sparse, and these water resources remain poorly understood. The median TDS of shallow alluvial groundwaters in the Southern Gulf catchments is approximately 600 mg/L, and the reported yields are relatively low (median of 2 L/second). However, this is based on very few data points, and additional data on the extent, thickness, internal architecture and hydraulic properties of alluvial aquifers are required to assess their suitability as productive and reliable water supplies. Fractured rock aquifers CSIRO (2009c) described the Proterozoic rocks trending north-west to south-east across the central part of the Southern Gulf catchments (Figure 2-2) as not a feasible groundwater resource. However, McEniery (1980) pointed out that the Mount Isa town water supply was sourced from fractured shales until the construction of the Lake Moondarra dam. These shales occur at depths of 60 to 80 mBGL and yield from 5 to 10 L/second. Approximately 50% of the bores with stratigraphic information in the Southern Gulf catchments correspond to Proterozoic units composed of unassigned granites or are attributed to one of the following geological units: Lawn Hill, Corella, Paradise Creek, Surprise Creek, Toole Creek Volcanics, Lady Loretta and Gunpowder Creek. These geological units are considered to potentially host local-scale fractured and weathered rock aquifers (Raiber et al., 2024). According to the data compiled by Bardwell and Grey (2016), a significant spatial variability in groundwater chemistry is found among the Proterozoic aquifers, with dominant ions varying from Mg–Ca–HCO3 to Na–SO4. Such characteristics may be an indication of localised recharge and interactions of groundwater with variable mineral-rich rocks. 2.5.3 Groundwater recharge Groundwater recharge is an important component of the water balance of an aquifer. It can inform how much an aquifer is replenished on an annual basis and therefore how sustainable a groundwater resource may be in the long term. This is particularly important for aquifers with low storage or aquifers that discharge to rivers, streams, lakes and the ocean or via transpiration from groundwater- dependent vegetation. Recharge is influenced to varying degrees by many factors, including spatial changes in soil type (and their physical properties), the amount of rainfall and evaporation, vegetation type (and transpiration), topography and depth to the watertable. Recharge can also be influenced by changes in land use, such as land clearing and irrigation. Directly measuring recharge can be very difficult as it usually represents only a small component of the water balance, can be highly variable spatially and temporally, and can vary depending on the type of measurement or estimate technique used (Petheram et al., 2002). In the Assessment, several independent approaches were used to estimate annual recharge for all aquifers in the Southern Gulf catchments. Figure 2-34 provides an example of recharge estimates using the upscaled chloride mass balance (CMB) method. For more detail on how these estimates were derived, see the companion technical report on hydrogeological assessment (Raiber et al., 2024). Annual recharge, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-509_CMB_R_percentiles_constrained_SG_v2_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-34 Annual recharge estimates for the Southern Gulf catchments Estimates based on upscaled chloride mass balance (CMB) method for the (a) 50th, (b) 5th and (c) 95th percentiles. Figure 2-35 provides a summary of the range in recharge estimates for the outcropping area of seven key hydrogeological units across the Southern Gulf catchments (Figure 2-26). Recharge estimates are based on the mean of the 5th and 95th percentiles and range from approximately: •3 to 28 mm/year for the Camooweal Dolostone •14 to 46 mm/year for the Cretaceous sediments •14 to 55 mm/year for the fractured rocks • 22 to 89 mm/year for the Proterozoic carbonates • 3 to 30 mm/year for the Thorntonia Limestone (or equivalent) • 3 to 20 mm/year for the Wonarah Formation • 9 to 29 mm/year for the alluvium. The estimates of groundwater recharge in the Assessment represent the spatial variability in recharge across the land surface and are a good starting point for estimating a water balance arithmetically or using a groundwater model. However, none of the methods accounts for aquifer storage (available space in the aquifer), so it is unclear whether the aquifers can accept these rates of recharge on an annual basis. The methods also do not account for potential preferential recharge from streamflow or overbank flooding or through karst features across parts of the Southern Gulf catchments. Therefore, the key features of an aquifer must be carefully conceptualised before simply deriving a recharge volume based on the surface area of an aquifer outcrop and an estimated recharge rate. For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-35 Summary of recharge statistics to outcropping areas of key hydrogeological units across the Southern Gulf catchments Recharge rates are based on upscaled chloride mass balance (CMB) method and calculated as the mean of the 5th and 95th percentiles. Error bars represent the standard deviation from the mean. 2.5.4 Surface water – groundwater connectivity As discussed in Section 2.5.2, groundwater discharge to surface water features occurs from a variety of aquifers across the Southern Gulf catchments. Areas of groundwater discharge are important for sustaining both aquatic and terrestrial groundwater-dependent ecosystems and often have cultural significance. These groundwater discharge areas have been mapped in Figure 2-36 as three categories: perennial, seasonally varying and coastal. Perennial groundwater discharge areas often exhibit springs that occur in a variety of hydrogeological settings; these can involve groundwater flow systems at a variety of scales ranging from hundreds of metres to a few hundred kilometres. Areas of seasonally varying groundwater discharge generally exhibit localised alluvial and fractured and weathered rock aquifer systems, and surface water recharges waterholes and alluvial and fractured rock aquifers adjacent to streams during the wet season. These stores of water may sustain the riparian vegetation through the dry season. Although surface water is thought to be the major source for these systems, groundwater discharge from adjacent aquifers can also occur when river levels recede during the dry season. Coastal discharge areas may have a component of coastal submarine groundwater discharge but also have mangroves that use fresh to saline water within the freshwater–saltwater interface. Groundwater discharge at spring vents connected to stream channels likely occurs in some parts of the Southern Gulf catchments. For example, permanently saturated springs of the Boodjamulla spring complex (Figure 2-36) are mapped along Lawn Hill Creek in the headwaters of the Nicholson River. The occurrence of springs in this area has been linked to groundwater discharge from underlying Proterozoic sedimentary rock aquifers (Buchanan et al., 2020). The drivers of potential groundwater discharge from deeper units in this area are not currently well understood due to a lack of hydrochemistry and environmental tracer data. However, multiple possible mechanisms have been described. This includes, for example, groundwater discharge driven by gas overpressure from the shale-rich sequences in the Isa Superbasin. Surface faults and regional structural features occur near the springs, and the location of these springs also corresponds with a distinct change in elevation of approximately 100 m from west to east, which could indicate groundwater discharge at a break in slope (Buchanan et al., 2020). A recent assessment by Geoscience Australia (Dixon-Jain et al., 2024) used geological and historical and newly acquired geophysical (seismic and airborne electromagnetic) data to study the influence of geological structures on the occurrence of groundwater-dependent ecosystems. The study indicated that some of the springs north of Lawn Hill Creek of the Boodjamulla spring complex (Figure 2-36) appear to be associated with the Constance Sandstone, which is part of the South Nicholson Basin, with faults and fractures potentially forming pathways for groundwater discharge. The study also highlighted that other potential spring-source aquifers may exist, including a possible indirect groundwater source originating from the Thorntonia Limestone (part of the Georgina Basin) and flowing into the Constance Sandstone (with alternative hydrogeological conceptualisations discussed in more detail in the companion hydrogeological technical report by Raiber et al., 2024). Other mapped, permanently active springs in the south-western part of the Southern Gulf catchments in the headwaters of the Gregory River (Figure 2-28) are spatially associated with and likely drain the Thorntonia Limestone. Dixon-Jain et al. (2024) highlighted that further hydrochemical sampling, including environmental tracers, is required to understand the hydrogeology of the springs within the Southern Gulf catchments. Shallow groundwater from the Karumba Basin sediments likely supports terrestrial and aquatic groundwater-dependent ecosystems on the alluvial floodplains. Groundwater discharge, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-512_GW_discharge_v3_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-36 Spatial distribution of groundwater discharge classes including surface water – groundwater connectivity across the Southern Gulf catchments Groundwater discharge classes are inferred from remotely sensed estimates of evapotranspiration and open water persistence (based on findings described in the companion technical report on hydrogeological assessment by Raiber et al., 2024). Note: the size of polygons has been greatly exaggerated to allow them to be seen at this scale. Geology data sources: adapted from Department of Industry, Tourism and Trade (2014) and Department of Environment, Parks and Water Security (2008) and Raymond et al. (2012) Spring data sources: Department of Environment, Parks and Water Security (2013); the Groundwater Dependent Ecosystems Atlas (Bureau of Meteorology, 2017) and Queensland Government (2021) 2.5.5 Surface water Streamflow Approximately 60% of Australia’s runoff is generated in northern Australia (Petheram et al., 2010, 2014). Unlike the large internally draining Murray–Darling Basin, however, northern Australia’s runoff is distributed across many hundreds of smaller externally draining catchments (Figure 2-37). To place the Southern Gulf catchments in a broader context, it is useful to compare its size and the magnitude of its median annual streamflow to other river systems across Australia. Figure 2-37 shows the magnitude of median annual streamflow of major rivers across Australia prior to water resource development. Streamflow Australia, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\Hy-S-501_Aust_accumulated_AnnualMedian_flow_AWRA_SouGulf_rescaled.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-37 Modelled streamflow under natural conditions Streamflow under natural conditions is indicative of median annual streamflow prior to European settlement (i.e. without any large-scale water resource development or extractions) assuming the historical climate (i.e. 1890 to 2015). Source: Petheram et al. (2017) The Southern Gulf catchments consist of the contributing area of various rivers and streams that discharge into the southern Gulf of Carpentaria. The most substantial of these are the Leichhardt, Gregory and Nicholson rivers. The catchments of these rivers, plus those on the Wellesley Islands in the Gulf of Carpentaria, Settlement Creek, Morning Inlet, and numerous small coastal creeks, have a total area of 108,200 km2. The Leichhardt catchment has an area of 33,400 km2, and the river itself extends approximately 550 km from the river mouth to Mount Isa in the south of the catchment. Tidal influence on streamflow is detectable approximately 15 km upstream of Burketown on the Albert River before the water becomes brackish. On the Leichhardt River, brackish salinity is likely to extend as far up as Leichhardt Falls, 113 km from the mouth. The falls are just below station 913007B on Leichhardt River at Floraville Station (Figure 2-38), the only operating streamflow station in the lower plains in the Southern Gulf catchments. The median and mean annual discharges from all Southern Gulf catchments are 4961 and 6759 GL/year, respectively. The pronounced difference between the mean and median is due to the mean being biased by a number of very high flow years (Table 2-4). Current surface water licences across the study area total about 114 GL (2.3% of median annual flow). See Section 3.3 for more information. The cease-to-flow column in Table 2-4 gives the percentage of time that no streamflow was observed at each of the streamflow gauging stations in the catchments. Gauges on the Gregory River indicate perennial flow, with dry-season baseflow resulting from discharge from the Thorntonia Limestone hydrogeological unit. Other rivers ceased to flow for 27% to 79% of the time, driven by the seasonal rainfall that has on average 94% of the rainfall falling during the wet season (Section 2.4). The influence of storage at Lake Julius results in a higher cease-to-flow period, with spills occurring 13% of the time. Lake Julius and Lake Moondarra, two large storages (approximately 107 GL each) in the upper Leichhardt River, are used for town and industrial water supply near Mount Isa. There are 75.1 GL of licences allocated from the storages; however, only a portion of this volume is currently used each year (typically ~40%). Table 2-4 Streamflow metrics at selected gauging stations in the Southern Gulf catchments Annual streamflow data are calculated under Scenario A. The 20th, 50th and 80th refer to 20%, 50% and 80% percentile exceedance, respectively. Cease-to-flow percentage (the percentage of all observation days where no streamflow was recorded) is determined using observed data, where streamflow less than 0.1 ML/day was assumed to be equal to zero. The annual streamflow data are shown schematically in Figure 2-39 and Figure 2-40. STATION ID STATION NAME CATCHMENT AREA (HA) ANNUAL STREAMFLOW (GL) CEASE- TO-FLOW % RUNOFF COEFFICIENT MEAN 80TH 50TH 20TH 913014A Leichhardt River at Doughboy Creek 3,520 115 27 72 153 63 0.07 913015A Leichhardt River at Julius Dam 4,748 122 0 41 163 87 0.06 913004A Leichhardt River at Miranda Creek 5,959 213 27 92 277 62 0.08 913006A Gunpowder Creek at Gunpowder 2,412 128 42 85 171 72 0.11 913010A Fiery Creek at 16 Mile Waterhole 721 41 12 27 55 79 0.11 913007B Leichhardt River at Floraville Homestead 23,679 1157 308 693 1670 49 0.10 912108A O Shannassy River at 17.7 Km 5,591 238 103 179 334 0 0.09 912105A Gregory River at Riversleigh No.2 11,382 475 213 343 681 0 0.09 912101A Gregory River at Gregory Downs 12,567 506 226 363 708 0 0.09 912103A Lawn Hill Creek at Lawn Hill No 2 4,003 164 59 116 219 27 0.09 912107A Nicholson River at Connolly’s Hole 13,875 649 82 268 1092 34 0.09 912116A Nicholson River at Doomadgee Mission 15,117 730 98 325 1186 32 0.09 streamflow gauges "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\Hy-S-508_streamflow_gauges.mxd" For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-38 Streamflow observation data availability in the Southern Gulf catchments Figure 2-39 shows how median annual streamflow increases towards the coast in the Southern Gulf catchment. As an indication of variability, Figure 2-40 shows the 20% and 80% exceedance of annual streamflow compared to the median (50% exceedance) in Figure 2-39. Streamflow exceedence, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\Hy-S-000_SouGulf_accumulated_AnnualMedian_flow_(E50)_rescaled_v07.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-39 Median annual streamflow (50% exceedance) in the Southern Gulf catchments under Scenario A Commencement threshold for blue flow accumulation line is 5 GL/year. Labelled points refer to Figure 2-49 Figure 2-41 illustrates the increase in catchment area and decrease in elevation along the Gregory–Nicholson River (a) and Leichhardt River (b) from one of the most upstream source tributaries to their mouth. The large ‘step’ changes in catchment area are where major tributaries join the river. Scenario streamflow exceedence \\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\1_Export\Hy-S-503_2x1_SG_accumulated_E20_E80_flow_rescaled_v03.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-40 (a) 20% and (b) 80% exceedance of annual streamflow in the Southern Gulf catchments under Scenario A For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-41 Catchment area and elevation profile from the upper tributaries in each catchment to the mouth along the (a)Gregory–Nicholson River and (b) Leichhardt River Catchment runoff The simulated mean annual runoff averaged over the Southern Gulf catchments under Scenario A is 76 mm. Figure 2-42 shows the spatial distribution of mean annual rainfall and runoff under Scenario A (1890 to 2022) across the study area. Mean annual runoff broadly follows the same spatial patterns as mean annual rainfall: highest in the north of the study area and lowest in the south. Monthly and annual runoff data in the Southern Gulf catchments exhibit substantial variation from one year to the next. The annual runoff volumes at 20%, 50% (median) and 80% exceedance averaged across the Southern Gulf catchments are 114, 30 and 10 mm, respectively (Figure 2-43). That is, runoff spatially averaged across the catchments will on average exceed 114 mm 1 year in 5, 30 mm half the time and 10 mm 4 years in 5. Figure 2-43 shows the spatial distribution of annual runoff at 20%, 50% and 80% exceedance under Scenario A. Intra- and inter-annual variability in runoff Rainfall, runoff and streamflow in the Southern Gulf catchments are variable between and within years. Approximately 87% of all runoff in the catchments occurs in the 3 months from January to March, which is a very high concentration of runoff compared to rivers in southern Australia (Petheram et al., 2008). While streamflow is ephemeral at many gauge sites, some rivers in the Gregory catchment are perennial (Table 2-4). Figure 2-44b illustrates that there is a high variation in monthly wet-season runoff from one year to the next. For example, during March, spatial mean runoff exceeded 34 mm in 20% of years and was less than 1 mm in 20% of years. The largest catchment mean annual runoff under Scenario A was 491 mm in 1973–74, and the smallest was 3 mm in 1901–02 (Figure 2-44a). The CV of annual runoff aggregated across the Southern Gulf catchments is 1.2. Based on data from Petheram et al. (2008), this variability in annual runoff is slightly above the middle of the range of the annual variability in runoff of other rivers in northern Australia with a comparable mean annual runoff. \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\Hy-S-504_Rain_Runoff_1x2.mxd For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-42 Mean annual (a) rainfall and (b) runoff across the Southern Gulf catchments under Scenario A Pixel-scale variation in mean annual runoff is due to changes in climate and modelled variation based on physiographic units. \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\Hy-S-505_20_50_80_runoff_1x3.mxd For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-43 Annual runoff at (a) 20%, (b) 50% and (c) 80% exceedance across the Southern Gulf catchments under Scenario A Pixel-scale variation in mean annual runoff is due to changes in climate and modelled variation based on physiographic units. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-44 Total runoff across the Southern Gulf catchments under Scenario A showing (a) time series of annual runoff and (b) monthly runoff averaged across the catchments The solid blue line in (a) is the 10-year running mean. In (b) ‘A range’ represents the 80% to 20% exceedance totals for that month. Flooding Intense seasonal rains from monsoonal bursts and tropical cyclones from November to March create flooding in parts of the Southern Gulf catchments and inundate large areas of floodplains, mostly in the downstream reaches between the Nicholson, Gregory and Leichhardt rivers. The floodplains along the Alexandra and Albert rivers are also heavily flooded (Figure 2-45). Floodplains along the Alexandra and Albert rivers (Figure 2-45) are also regularly flooded. Flooding is common in the Burketown area, which is located on a remnant of the main channel of the Albert River and represents the most eastward extent of a very flat ridgeline that provides the highest ground (~5 m above sea level) on the western bank of the river. Burketown is susceptible to flooding from the Albert River floodplain, as well as from overland flow paths within the town area. Flooding of the Albert River could also occur from floodwater breakout from the Nicholson or Gregory rivers. Rivers in the Southern Gulf catchments are largely unregulated. The notable exception is the Leichhardt River, which has two large dams: Moondarra Dam and Julius Dam. Overbank flow is generally governed by the topography of the floodplain. Since 1980, there have been 41 floods greater than or equal to 1 in 1 annual exceedance probability (AEP) in different parts of the Southern Gulf catchments. While floods occur in any month from November to April, historically the month with the most floods is January (42%). Characterising these flood events is important for a range of reasons. Flooding can be catastrophic to agricultural production in terms of loss of stock, pasture and topsoil, and damage to crops and infrastructure. It can also isolate properties and disrupt vehicle traffic providing goods and services to people in the catchment. However, flood events also provide opportunities for offstream wetlands to connect to the main river channel. The high biodiversity found in many unregulated floodplain systems in northern Australia is thought to largely depend on seasonal flood pulses, which allow biophysical exchanges to occur between rivers and offstream wetlands. Further observations of flood characteristics in the Southern Gulf catchments are as follows: • Flood peaks typically take about 2 days to travel from Gregory to Burketown at a mean speed of 3.2 km/hour. • For flood events of AEP 1 in 2, 1 in 5 and 1 in 10, the peak discharges at Riversleigh Road on the Gregory River gauge are 590, 1850 and 2280 m3/second, respectively. • Between 1980 and 2023 (43 years), 51 streamflow events broke the banks of the Gregory River at Riversleigh Road crossing. All events occurred between November and April (inclusive), and about 84% of these events occurred between January and March (inclusive). Of the ten largest flood peak discharges at Riversleigh on the Gregory River, six occurred in January, three in February and one in December. • The maximum area inundated by a flood event of AEP 1 in 38 that occurred in March 2023 was 5983 km2 (Figure 2-46). Flood inundation, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\HC-S-243_Inundation_MODIS_Flood_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-45 Flood inundation map of the Southern Gulf catchments Data captured using Moderate Resolution Imaging Spectroradiometer (MODIS) satellite imagery. This figure illustrates the maximum percentage of each MODIS pixel inundated between 2000 and 2023. Lowland flood inundation, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\HC-S-244_Flood_inundation.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-46 Flood inundation in the Southern Gulf catchments for a flood event of 1 in 38 annual exceedance probability (AEP) in March 2023 Flood frequency in the Southern Gulf catchments Flood frequency analysis was performed for the Southern Gulf catchments to establish streamflow thresholds above which a flood event would occur. Flood frequencies were estimated for the three major rivers (Nicholson, Gregory and Leichhardt). For the Nicholson River, flood frequencies were estimated using streamflow observations from gauging station 912107A (Nicholson River at Connolly’s Hole) as this gauge has reasonable quality data. Flood frequencies were estimated using streamflow observations from gauging station 912105A (Gregory River at Riversleigh) for the Gregory River. For the Leichhardt River, flood frequencies were estimated using streamflow observations from gauging station 913007B (Leichhardt River at Floraville). Traditionally, flood frequencies are estimated based on maximum discharge for individual events. However, in the Assessment, to help determine the true magnitude of the events, the flood frequency analysis accounted for total flow volume as well as peak discharge for each event. This is motivated by the knowledge that the duration of an event, and not only its maximum discharge, can have a great impact on the inundated area. Figure 2-47 displays the relationship between peak flow, flood volume and AEP for the three gauges: one on the Nicholson River (912107A) one on the Gregory River (912105A) and one on the Leichhardt River (913007B). While flow volume is higher for larger floods, duration of flood is a key factor for volume of flood flow. A diagram of different colors Description automatically generated with medium confidence Figure 2-47 Peak flood discharge and annual exceedance probability (AEP) at (a) gauge 912107A (Nicholson River at Connolly’s Hole), (b) gauge 912105A (Gregory River at Riversleigh) and (c) gauge 913007B (Leichhardt River at Floraville) Colours indicate the total event volume of flood water in gigalitres (GL) for different events. Instream waterholes during the dry season The rivers in the Southern Gulf catchments are largely ephemeral. However, perennial flow is associated with discharge from the dolostone aquifer underlying the Barkly Tableland to the Gregory River. In ephemeral reaches, such as most of the Leichhardt River, once streamflow has ceased the rivers break up into a series of waterholes during the dry season. Waterholes that persist from one year to the next are considered to be key aquatic refugia and are likely to be sustaining ecosystems in the catchments (Section 3.2). In some reaches, waterholes may be partly or wholly sustained by groundwater discharge (Section 2.5.2). However, in other reaches there is little evidence that persistent waterholes receive water from groundwater discharge and are likely to be replenished following wet-season flows. Stream gauge data indicate that there is very little to no late dry-season flow for most locations along the Leichhardt River and most of the Nicholson River, while baseflow is apparent in Gregory River observations (Figure 2-48). These properties are reflected in the river model simulations for the month of October when any streamflow is likely the result of groundwater discharge into streams (Figure 2-49). Substantial baseflow in October in the Gregory system is related to discharge from the dolostone aquifer in the upper Gregory catchment. The ecological importance and functioning of key aquatic refugia are discussed in more detail in the companion technical report on ecological modelling (Ponce Reyes et al., 2024). "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\0_Working\1_justin\5_Catchment_report\min_sept_flow.png" Figure 2-48 Minimum observed September streamflow at two stream gauge locations on the Gregory River "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\2_Reporting\1_Catch_report_river\03_minimumflow_boxplots\minOctoberFlow_v3.png" Figure 2-49 Minimum monthly flow over 132 years of simulation for the month of October Assessed at locations indicated by labels in Figure 2-39. Locations are listed in an upstream to downstream order (left to right). The dashed blue horizontal lines equate to 200 ML/day, and the dotted red horizontal lines equate to 400 ML/day As an example, the formation of waterholes following a cease-to-flow event were captured using satellite imagery for a reach of the Flinders River in northern Australia (Figure 2-50). Figure 2-52 maps 1 km river reaches (or segments) in the Southern Gulf catchments in which water is recorded in greater than 90% of dry-season satellite imagery. This is denoted the water index threshold and provides an indication of the river reaches that contain permanent water, typically waterholes that persist over the dry season. Some of these waterholes are maintained by perennial flow, for example, along the Gregory River. Below Gregory, the canopy along the Gregory River can be closed, obscuring the presence of water, and the river width becomes relatively small compared to the Landsat pixel size, which may influence the ability to detect permanent water using this technique. Maps of instream waterhole evolution. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 2-50 Instream waterhole evolution in a reach of the Flinders River This figure shows the area of waterholes in the reach of the river a given time after flow ceased and the ability of the water index threshold to track the change in waterhole area and distribution. Figure 2-51 The Leichhardt River near Kajabbi looking south towards the Isa highlands. In the highly seasonal climate of the Southern Gulf catchments, springs and persistent waterholes provide important ecological refugia during the dry season Photo: CSIRO – Nathan Dyer Permanent waterholes, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\1_GIS\1_Map_Documents\Hy-S-507_SouGulf_permanent_waterholes_v1.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-52 Location of river reaches containing permanent water in the Southern Gulf catchments Persistent river reaches are defined as 1 km river reaches where water was identified in greater than 90% of the dry-season Landsat (Landsat 5, 7 and 8) imagery between 1989 and 2018. Mapping of persistent river reaches is confounded by riparian vegetation along the watercourses. Surface water quality Since the late 1960s, there has been systematic sampling of the water quality of the surface water system at the gauge network through the Surface Water Ambient Water Quality Network (SWAN) program in Queensland. Three sites remain open in the Southern Gulf Assessment area: Gregory River at Riversleigh (912105A), Gunpowder Creek at Gunpowder (913006A) and Leichhardt River at Floraville Homestead (913007B). The data available for regularly analysed water quality parameters of interest (of the over 60 samples at each site) are presented in Figure 2-54. Minimum, median and maximum statistics are also presented in Table 2-5. Environmental values define the suitable uses of the water by aquatic ecosystems and by humans (e.g. for drinking, irrigation, aquaculture, recreation), and water quality objectives define objectives for the physical, chemical and biological characteristics of the water (e.g. nitrogen content, dissolved oxygen, turbidity, toxicants, fish). Environmental values and water quality objectives are being progressively determined for Queensland waters but have not yet been developed for any Southern Gulf rivers. In the absence of local water quality objectives, the water quality data can be interpreted against national guidelines. Salinity concentrations were typically within safe drinking limits (<1000 μS/cm). The maximum value at the Gunpowder Creek site (Table 2-5) coincided with zero discharge, indicating these higher salinities are likely due to the evaporative concentration of water at the sampling location rather than representing water quality of the source water. Table 2-5 Summary of water quality data for the open Southern Gulf catchments sites, with values of minimum, median and maximum for each site and each water quality parameter For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †NTU = nephelometric turbidity units. ‡ below detection limit represented by zero value. All recorded metal concentrations were below livestock drinking water trigger values from ANZECC and ARMCANZ (2000). Median constituent concentrations were well below long-term (100 year) trigger values for heavy metals and nutrients for irrigation water; however, a small number of samples exceeded the iron (0.2 mg/L) and phosphorus (0.05 mg/L) guidelines. All samples were above the national guideline values for total phosphorous for ecosystems in tropical lowland rivers (0.1 mg/L), and 34% of samples were above the national guideline values for total nitrogen (0.3 mg/L). This does not necessarily mean poor-quality water, as both Queensland’s Environmental Protection Policy (Water) and the national guidelines highlight the importance of using guideline values tailored to the local environment. A number of studies have investigated the water quality of the upper Leichhardt River, associated with mining activity, urban runoff and water supply to Mount Isa. Wilson et al. (2007) sampled disconnected dry-season remnant waterholes of the Leichhardt River upstream of Lake Moondarra. They found guideline values for healthy ecosystems for water soluble metals were exceeded, and a number of sites were found to be unfit for primary human contact. Mount Isa city’s sewerage infrastructure, associated waste water reuse scheme, and mining and smelting activities were identified as contributors to impaired water quality, highlighting the need for a catchment management approach to water quality improvement. The Lead Pathways Study investigated sources and pathways of heavy metals (primarily lead) to land, air and water at Mount Isa, including the Leichhardt River (Noller et al., 2012). Five sampling periods encompassing times before, during and after the wet season, were conducted from 2008 to 2010. A number of sites exceeded the ANZECC and ARMCANZ Water Quality Guidelines’ 90% and 95% trigger values for freshwater species for arsenic, cadmium, copper and lead. The water quality guideline process indicates further investigation should be undertaken to assess potential impacts on ecological health. The drinking water source for Mount Isa (Lake Moondarra) was found to meet drinking water guidelines, and sampled metal concentrations indicated a low risk to human health from recreational activities or eating fish caught in the lake or Leichhardt River. Figure 2-53 Accelerated erosion contributes sediment to streamflow Photo: CSIRO – Nathan Dyer For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 2-54 Water quality samples for selected constituents on the Gregory River at Riversleigh (A912005A), Gunpowder Creek at Gunpowder (913006A) and Leichhardt River at Floraville Homestead (913007B) Values below detection limits are seen as constant values, for example, 0.01 mg/L for zinc and 0.05 mg/L for aluminium. Horizontal dashed line represents national guideline levels where relevant. NTU = nephelometric turbidity units. Data sourced from Queensland Government website 2.6 References ANZECC and ARMCANZ (2000) Australian and New Zealand guidelines for fresh and marine water quality, Volume 1, The guidelines. Australian and New Zealand Environment and Conservation Council, Agriculture and Resource Management Council of Australia and New Zealand. Bardwell N and Grey D (2016) Hydrogeochemistry of Queensland: data release: accompanying notes. CSIRO, Australia. Viewed 23 September 2021, https://researchdata.edu.au/hydrogeochemistry- queensland-data-release/671735. BOM (2023) Tropical cyclone databases. Bureau of Meteorology, Canberra. Viewed 6 February 2023, Hyperlink to: Tropical cyclone databases . Bowman DMJS, Brown GK, Braby MF, Brown JR, Cook LG, Crisp MD, Ford F, Haberle S, Hughes J, Isagi Y, Joseph L, McBride J, Nelson G and Ladiges PY (2010) Biogeography of the Australian monsoon tropics. Journal of Biogeography 37(2), 201–216. Hyperlink to: Biogeography of the Australian monsoon tropics . Bradshaw BE, Lahtinen AC, Khider K, Ryan DJ, Colwell JB, Chirinos A and Bradshaw J (2009) Queensland carbon dioxide geological storage atlas. Greenhouse Gas Storage Solutions, Canberra for the Queensland Department of Employment, Economic Development and Innovation, Brisbane. Buchanan S, Dixon-Jain P, Martinez JL, Raiber M and Kumar P (2020) Hydrogeology and groundwater systems for the Isa GBA region. Technical appendix for the Geological and bioregional assessment: stage 2. Department of the Environment and Energy, Bureau of Meteorology, CSIRO and Geoscience Australia, Australia. Bureau of Meteorology, 2017. Groundwater Dependent Ecosystems Atlas. http://www.bom.gov.au/water/groundwater/gde/map.shtml Charles S, Petheram C, Berthet A, Browning G, Hodgson G, Wheeler M, Yang A, Gallant S, Vaze J, Wang B, Marshall A, Hendon H, Kuleshov Y, Dowdy A, Reid P, Read A, Feikema P, Hapuarachchi P, Smith T, Gregory P and Shi L (2017) Climate data and their characterisation for hydrological and agricultural scenario modelling across the Fitzroy, Darwin and Mitchell catchments. 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. Chiew F, Post D and Moran R (2012) Hydroclimate baseline and future water availability projections for water resources planning. 34th Hydrology and Water Resources Symposium 2012, Sydney, NSW, 19–22 November 2012. Engineers Australia, Barton, 461–468. Chiew FHS, Cai W and Smith IN (2009) Advice on defining climate scenarios for use in Murray–Darling Basin Authority Basin Plan modelling. CSIRO report for the Murray–Darling Basin Authority. Christian CS, Noakes LC, Perry RA, Slatyer RO, Stewart GA, Traves DM and Csiro (1954) Survey of the Barkly Region, Northern Territory and Queensland, 1947–48. Land Research Series No.3. CoastAdapt (2017) Sea-level rise and future climate information for coastal councils: Roper Gulf, NT. National Climate Change Adaptation Research Facility and Australian Government Department of Environment and Energy. Viewed 03 May 2018, Hyperlink to: CoastAdapt Roper Gulf, NT . CSIRO (2009a) Water in the Van Diemen region. In: Water in the Timor Sea Drainage Division. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO Water for a Healthy Country Flagship, Australia, 363–452. CSIRO (2009b) Water in the South-West Gulf region. In: Water in the Gulf of Carpentaria Drainage Division. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO Water for a Healthy Country Flagship, Australia, 123–197. Viewed 23 September 2021, Hyperlink to: Water in the Gulf of Carpentaria Drainage Division . CSIRO (2009c) Water in the Gulf of Carpentaria Drainage Division. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO Water for a Healthy Country Flagship, Australia. Viewed 23 September 2021, Hyperlink to: Water in the Gulf of Carpentaria Drainage Division . CSIRO and Bureau of Meteorology (2015) Climate change in Australia. Information for Australia’s natural resource management regions: technical report. CSIRO and Bureau of Meteorology, Australia. Denniston RF, Villarini G, Gonzales AN, Wyrwoll K-H, Polyak VJ, Ummenhofer CC, Lachniet MS, Wanamaker AD, Humphreys WF, Woods D and Cugley J (2015) Extreme rainfall activity in the Australian tropics reflects changes in the El Niño/Southern Oscillation over the last two millennia. Proceedings of the National Academy of Sciences 112(15), 4576-4581. Hyperlink to: Extreme rainfall activity in the Australian tropics reflects changes in the El Niño/Southern Oscillation over the last two millennia . DENR (2016) Bore construction Darwin rural area most common in dolomite country. Department of Environment and Natural Resources. Data downloaded in September 2022, Hyperlink to: Bore construction Darwin rural area most common in dolomite country (figure) . Department of Environment, Parks and Water Security (2008) Northern Territory groundwater – aquifer types 1:2,000,000. Northern Territory Department of Environment, Parks and Water Security, Darwin. Data downloaded in February 2021, Hyperlink to: Northern Territory groundwater – aquifer types . Department of Environment, Parks and Water Security (2013) Northern Territory springs. Northern Territory Department of Environment, Parks and Water Security, Darwin. Data downloaded in February 2021, Hyperlink to: Northern Territory Springs . Department of Environment, Parks and Water Security (2014) Digital groundwater database for the Northern Territory, supplied by the Department of Environment, Parks and Water Security, Copyright – Northern Territory of Australia. Data downloaded from the Northern Territory Government (NTG) Open Data Portal in September 2021, https://data.nt.gov.au/dataset/nt- water-groundwater-licence-extraction-points. Department of Industry, Tourism and Trade (2014) Northern Territory geological map (interp) 2500K. Northern Territory Department of Industry, Tourism and Trade, Darwin. Data downloaded in February 2023, Hyperlink to: Northern Territory Geological map (interp) 2500K . Dixon-Jain P, Bishop C, Lester J, Orlov C, McPherson A, Pho G, Flower C, Lawson S and Lewis S (2024) Hydrogeology and groundwater systems of the South Nicholson and Georgina basins, Queensland and Northern Territory. Record 2024/37, Geoscience Australia, Canberra. DNRME (Queensland Department of Natural Resources, Mines and Energy) (2023) Groundwater database – Queensland. Open Data Portal. Hyperlink to: Groundwater database – Queensland . DNRME (2021) Groundwater database – Queensland. Database storing registered water bore data from private water bores and Queensland Government groundwater investigation and monitoring bores. State of Queensland, Department of Resources. Licensed under Single Supply, licence sourced in September 2021. Viewed 23 September 2021, Hyperlink to: Groundwater Database - Queensland . Environment Australia (2001) A directory of important wetlands in Australia, 3rd edn. Environment Australia, Canberra. Evans S, Marchand R and Ackerman T (2014) Variability of the Australian monsoon and precipitation trends at Darwin. Journal of Climate 27(22), 8487–8500. Hyperlink to: Variability of the Australian monsoon and precipitation trends at Darwin . Exploring for the future (n.d.) Project activity: South Nicholson National Drilling Initiative. Geoscience Australia website. Viewed January 2024, Hyperlink to: South Nicholson National Drilling Initiative . Fitzpatrick EA (1986) An introduction to soil science. Longman Scientific and Technical Group, Essex, UK. Forsyth AJ, Nott J and Bateman MD (2010) Beach ridge plain evidence of a variable late-Holocene tropical cyclone climate, North Queensland, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 297(3–4), 707-716. Hyperlink to: Beach ridge plain evidence of a variable late-Holocene tropical cyclone climate, North Queensland, Australia . Geoscience Australia and Australian Stratigraphy Commission (2021) Australian stratigraphic units database. Viewed 17 January 2024, https://www.ga.gov.au/data- pubs/datastandards/stratigraphic-units. Gibbs M, Hughes J, Yang A, Wang B, Marvanek S and Petheram C (2024) River model scenario analysis for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Grimes KG (1974) Mesozoic and Cainozoic geology of the Lawn Hill, Westmoreland, Mornington, and Cape Van Diemen 1:250 000 Sheet areas, Queensland. Record 1974/106. Geoscience Australia, Canberra. Viewed 19 December 2023, Hyperlink to: Mesozoic and Cainozoic geology of the Lawn Hill, Westmoreland, Mornington, and Cape Van Diemen 1:250 000 Sheet areas, Queensland . Haig J, Nott J and Reichart G-J (2014) Australian tropical cyclone activity lower than at any time over the past 550–1,500 years. Nature 505(7485), 667-671. Hyperlink to: Australian tropical cyclone activity lower than at any time over the past 550-1,500 years . Hausfather Z and Peters GP (2020) Emissions–the ‘business as usual’ story is misleading. Nature 577, 618-620. Herbert GC (2000) Hydrogeology of the Karumba Basin western Cape York Peninsula. Post graduate diploma project, University of Queensland, School of Natural Resource Sciences, Brisbane. Horn AM, Derrington EA, Herbert GC, Lait RW and Hillier JR (1995) Groundwater resources of Cape York Peninsula. Cape York Peninsula land use strategy, Office of the Co-ordinator General of Queensland, Brisbane, Department of the Environment, Sport and Territories, Canberra, Queensland Department of Primary Industries, Brisbane and Mareeba, and Australian Geological Survey Organisation, Mareeba. Ingram JA (1972) 1:250,000 Geological series – explanatory notes, Donors Hill, Queensland, Sheet SE/54-10 International Index. Canberra. IPCC (2022) Climate Change 2022: impacts, adaptation, and vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press. Cambridge University Press, Cambridge, UK and New York, NY, USA. Isbell, RF and CSIRO (2016) The Australian Soil Classification. Australian Soil and Land Survey Handbook Series (Vol. 4), 2nd ed. CSIRO Publishing, Melbourne. Jeffrey SJ, Carter JO, Moodie KB and Beswick AR (2001) Using spatial interpolation to construct a comprehensive archive of Australian climate data. Environmental Modelling and Software 16, 309–330. Hyperlink to: Using spatial interpolation to construct a comprehensive archive of Australian climate data . Jenny H (1941) Factors of soil formation, a system of quantitative pedology. McGraw-Hill, New York. Jolly P and Tickell S (2011) Tufa dams of Northern Australia. NT branch of International Association of Hydrologists (IAH) Australia. Klohn Crippen Berger (2016) Hydrogeological assessment of the Great Artesian Basin: characterisation of aquifer groups, Carpentaria Basin. Department of Natural Resources and Mines (Queensland), Brisbane. Köppen W (1936) Das geographische System der Klimate. In: Köppen W and Geiger G (eds), Handbuch der Klimatologie, 1. C. Gebr, Borntraeger, 1–44. Kruse P and Dunster J (2013) Chapter 28: Georgina Basin. In: Ahmad M and Munson TJ (compilers) Geology and mineral resources of the Northern Territory. Northern Territory Geological Survey, Special Publication 5. Lerat J, Egan C, Kim S, Gooda M, Loy A, Shao Q and Petheram C (2013) Calibration of river models for 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. Matthews I (1992) Alpurrurulam (Lake Nash) groundwater source and borefield review. Power and Water Authority. Technical Report WRA92008. Viewed 23 September 2021, Hyperlink to: Alpurrurulam groundwater source and borefield review . McBride JL and Nicholls N (1983) Seasonal relationships between Australian rainfall and the Southern Oscillation. Monthly Weather Review 111(10), 1998–2004. Hyperlink to: Seasonal relationships between Australian rainfall and the Southern Oscillation . McEniery M (1980) Groundwater resources – north Queensland. In: Henderson R and Stephenson P (eds) The geology and geophysics of northeastern Australia, Volume 1. Geological Society of Australia, Queensland, Brisbane. McFarlane D, Stone R, Martens S, Thomas J, Silberstein R, Ali R and Hodgson G (2012) Climate change impacts on water yields and demands in south-western Australia. Journal of Hydrology 475, 488–498. Hyperlink to: Climate change impacts on water yields and demands in south-western Australia . McJannet D, Yang A and Seo L (2023) Climate data characterisation for hydrological and agricultural scenario modelling across the Victoria, Roper and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. McMahon TA and Adeloye AJ (2005) Water resources yield. Water Resources Publications LLC, Colorado, USA. Ponce Reyes R, Stratford D, Linke S, Merrin L, Kenyon R, Buckworth R, Deng RA, Hughes J, McGinness H, Pritchard J, Seo L and Waltham N (2024) Assessment of the potential ecological outcomes of water resource development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. National Committee on Soil and Terrain (2009) Australian soil and land survey field handbook. CSIRO Publishing, Collingwood. Noller B, Huynh T, Ng J, Zheng J and Harris H (2012) Lead pathways study – Water: sources and pathways of contaminants to the Leichhardt River. Centre for Mined Land Rehabilitation – Sustainable Minerals Institute, The University of Queensland, Report prepared for Mount Isa Mines Limited. Brisbane, Queensland. Nott JF and Jagger TH (2013) Deriving robust return periods for tropical cyclone inundations from sediments. Geophysical Research Letters 40(2), 370–373. Hyperlink to: Deriving robust return periods for tropical cyclone inundations from sediments . Peel MC, Finlayson BL and McMahon TA (2007) Updated world map of the Köppen–Geiger climate classification. Hydrology and Earth System Science 11(5), 1633–1644. Hyperlink to: Updated world map of the Köppen-Geiger climate classification . Petheram C and Yang A (2013) Climate data and their characterisation for hydrological and agricultural scenario modelling across 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. Viewed 6 February 2018, Hyperlink to: Climate data and their characterisation for hydrological and agricultural scenario modelling across the Flinders and Gilbert catchments . Petheram C, Gallant J, Wilson P, Stone P, Eades G, Roger L, Read A, Tickell S, Commander P, Moon A, McFarlane D and Marvanek S (2014) Northern rivers and dams: a preliminary assessment of surface water storage potential for northern Australia. CSIRO Land and Water Flagship Technical Report. CSIRO, Australia. Petheram C, McMahon TA and Peel MC (2008) Flow characteristics of rivers in northern Australia: implications for development. Journal of Hydrology 357(1–2), 93–111. Hyperlink to: Flow characteristics of rivers in northern Australia: implications for development . Petheram C, McMahon TA, Peel MC and Smith CJ (2010) A continental scale assessment of Australia's potential for irrigation. Water Resources Management 24(9), 1791–1817. Hyperlink to: A continental scale assessment of Australia's potential for irrigation . Petheram C, Rogers L, Read A, Gallant J, Moon A, Yang A, Gonzalez D, Seo L, Marvanek S, Hughes J, Ponce Reyes R, Wilson P, Wang B, Ticehurst C and Barber M (2017) Assessment of surface water storage options in the Fitzroy, Darwin and Mitchell catchments. 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. Petheram C, Rustomji P, McVicar TR, Cai W, Chiew FHS, Vleeshouwer J, Van Niel TG, Li L, Cresswell RG, Donohue RJ, Teng J and Perraud J-M (2012) Estimating the impact of projected climate change on runoff across the tropical savannas and semiarid rangelands of northern Australia. Journal of Hydrometeorology 13(2), 483–503. Hyperlink to: Estimating the impact of projected climate change on runoff across the tropical savannas and semiarid rangelands of northern Australia . Petheram C, Walker G, Grayson R, Thierfelder T and Zhang L (2002) Towards a framework for predicting impacts of land-use on recharge: 1. A review of recharge studies in Australia. Soil Research 40, 397–417. Hyperlink to: Towards a framework for predicting impacts of land-use on recharge: 1. A review of recharge studies in Australia . Peverill KI, Sparrow LA and Reuter DJ (eds) (1999) Soil analysis: an interpretation manual. CSIRO Publishing, Collingwood, Victoria. Queensland Government (2021) Springs database. (https://www.data.qld.gov.au/dataset/springs#:~:text=The%20Queensland%20Springs%20Database%20provides%20a%20comprehensive%20catalogue,and%20any%20associated%20surface%20expression%20groundwater%20dependent%20ecosystems.). Radke BM, Kellett JR, Ransley TR and Bell J (2012) Lexicon of the lithostratigraphic and hydrogeological units of the Great Artesian Basin and its Cenozoic cover. A technical report to the Australian Government from the CSIRO Great Artesian Basin Water Resource Assessment, Australian Government. CSIRO Water for a Healthy Country Flagship, Australia. Raiber M, Suckow A, Deslandes A, Gerber C, Martinez J, Crane P, Ransley T, Evans T, Wallace L and Wu G (2022) Assessing recharge processes and flow dynamics using environmental tracers in the Great Artesian Basin. CSIRO, Australia. Raiber M, Taylor AR, Dupuy M, Priestley S, Crosbie RS, Knapton A, Hodgson G and Barry K (2024) Characterising groundwater resources of the Gilbert River Formation, Camooweal Dolostone and Thorntonia Limestone in the Southern Gulf catchments, Queensland and Northern Territory. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Randal MA (1966) 1:250,000 Geological series – explanatory notes, Rankin, N.T. Sheet SE/53-16. Canberra. Rasmusson EM and Arkin PA (1993) A global view of large-scale precipitation variability. Journal of Climate 6, 1495–1522. Hyperlink to: A global view of large-scale precipitation variability . Raymond OL, Liu S, Gallagher R, Zhang W and Highet LM (2012) Surface Geology of Australia 1:1 million scale dataset 2012 edition. Geoscience Australia, Canberra. http://pid.geoscience.gov.au/dataset/ga/74619. Raymond OL (2018) Australian geological provinces 2018.01 edition. Geoscience Australia, Canberra. Viewed 18 October 2020, Hyperlink to: Australian geological provinces 2018.01 edition (database) . Read R (2003) Avon Downs and Ranken sheets groundwater occurrence. Northern Territory Department of Infrastructure, Planning and Environment, Natural Resources Division, Report 40/2003A. Smerdon BD, Ransley TR, Radke BM and Kellett JR (2012) Water resource assessment for the Great Artesian Basin. A report to the Australian Government from the CSIRO Great Artesian Basin Water Resource Assessment. CSIRO Water for a Healthy Country Flagship, Australia. Viewed 23 September 2021, Hyperlink to: Water resource assessment for the Great Artesian Basin . Smith JW and Roberts HG (1972) Mount Drummond, Northern Territory: 1:250,000 Geological series – explanatory notes. Record SE5312. Canberra. Taylor AR, Barron OV, Harrington GA, Chilcott C, Davies PJ, Wohling DL and Gilfedder M (2021) Rapid appraisal of new groundwater-based irrigation infrastructure opportunities across Australia. A technical report from CSIRO to the National Water Grid Authority. CSIRO, Australia. Thomas M, Philip S, Zund P, Stockmann U, Hill J, Gregory L, Watson I and Thomas E (2024) Soils and land suitability for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Tickell SJ (2003) Water resource mapping Barkly Tablelands. Report 23/2003. Department of Infrastructure Planning and Environment Natural Systems Division, Darwin. Viewed 23 September 2021, Hyperlink to: Water Resource Mapping Barkly Tablelands . Tickell SJ and Bruwer Q (2017) Georgina Basin groundwater assessment: Daly Waters to Tennant Creek. Technical Report 17/2017. Department of Environment and Natural Resources, Palmerston. Twidale CF (1956) The physiography of northwest Queensland. Geographic Studies 3(1). Wilson AL, Dehaan RL, Watts RJ, Page KJ, Bowmer KH and Curtis A (eds) (2007) Proceedings of the 5th Australian Stream Management Conference. Australian rivers: making a difference. Charles Sturt University, Thurgoona, New South Wales. 3 Living and built environment of the Southern Gulf catchments Authors: Pethie Lyons, Linda Merrin, Seonaid Philip, Diane Jarvis, Thomas Vanderbyl, Danial Stratford, Rob Kenyon, Simon Linke, Rocio Ponce Reyes, Heather McGinness, Caroline Bruce, Kaylene Camuti, Andrew R Taylor, Nathan Waltham, Jodie Pritchard Chapter 3 discusses a wide range of considerations relating to the living components of the catchments of the Southern Gulf rivers, that is Settlement Creek, Gregory–Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groups.1 This includes the environments that support these components, the people who live in the catchments or have strong ties to it, and the existing transport, power and water infrastructure. 1 Only those islands greater than 1000 ha are mapped 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 establishing a greenfield irrigation development Numbers refer to sections in this chapter. Block diagram of chapter sections \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\1_All\9_Graphics_artist\3_Vic and SoG\C Bruce Vic CR Chp3_8_2024.jpg For more information on this figure please contact CSIRO on enquiries@csiro.au 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 Southern Gulf catchments. It also examines the values, rights, interests and development objectives of Indigenous Peoples. 3.1.1 Key findings Ecology The largely intact habitats and landscapes of the Southern Gulf catchments provide ecosystem services that support high biodiversity, recreational activities, tourism, traditional and commercial fisheries, mining and areas of agricultural production. Within the freshwater sections of the Southern Gulf catchments are extensive areas with high habitat values, including ephemeral and persistent rivers, wetlands, floodplains and groundwater-dependent ecosystems (GDEs), including 13 sites listed in the Directory of Important Wetlands in Australia. Flows from the rivers and creeks in the Southern Gulf catchments into the Gulf of Carpentaria support recreational and commercial fisheries including barramundi (Lates calcarifer) and prawn fisheries. The Southern Gulf catchments support some of northern Australia’s most iconic wildlife species, including freshwater sawfish (Pristis pristis; listed as Vulnerable under the Commonwealth Environment Protection and Biodiversity Conservation Act 1999 (EPBC Act)), marine turtles and dugong (Dugong dugon) that occur in the coastal waters of the Gulf of Carpentaria. A wide variety of waterbirds can be found throughout the catchments as well as many lesser-known plants and animals that are also of great conservation significance (Atlas of Living Australia, 2021). Changes in land and water resources can have serious consequences for the ecology of rivers. Water resource development that changes the magnitude, timing or duration of either low or 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 anthropogenic pressures, and changes to water quality, including the availability and distribution of nutrients. Demographics, industries and infrastructure The Southern Gulf catchments have a population of about 22,500, with a population density one-sixteenth that of Australia as a whole. The study area contains one significant urban area (population >10,000 people), namely, the city of Mount Isa (over 18,000 people). There are several small towns and communities within the catchments, including Burketown, Doomadgee and Kajabbi and Gununa on the Wellesley Islands. Of these settlements, only Doomadgee (population 1387 as at the 2021 Census) has a population greater than 1000. Residents of the Southern Gulf catchments tend to be younger and more likely to identify as Indigenous than the typical resident of Queensland and of Australia as a whole. Incomes differ from the national and state mean, but the direction of the difference depends on the part of the population in question: incomes for the people of Mount Isa are higher than the state and national average while incomes of residents of the remainder of the study area are lower on average. 102 | Water resource assessment for the Southern Gulf catchments Excluding Mount Isa data from the analysis because it is heavily skewed to mining, the data reveal that agriculture provides 17.5% of employment. The dominant land use in the Southern Gulf catchments is grazing (77%). Conservation and protected land occupy 16% of the catchments and water and wetlands occupy another 3.8%. In terms of land tenure, 12% of the Southern Gulf catchments is held as Aboriginal freehold. The gross value of agricultural production (GVAP) in the Southern Gulf catchments is approximately $243.6 million, of which beef cattle contribute around $242.7 million and cropping accounts for the remaining $0.9 million. The Southern Gulf catchments are serviced by a modest network of major roads. The Barkly Highway is the only sealed road between the NT and Queensland. Type 2 road trains (vehicles up to 53 m in length) have access to a large percentage of the pastoral enterprises, providing onward access east to the Port of Townsville, south to markets and west into the NT. A good-quality rail line, which provides a link to Townsville and its port, can be accessed at Mount Isa. The North West Power System (NWPS) is centred in Mount Isa. The only other location receiving high voltage power in the Southern Gulf catchments is the Century Mine (zinc) near Lawn Hill from an approximately 250 km branch. Doomadgee, Burketown and Gununa have their own power stations. The smaller communities are very remote and serviced by stand-alone diesel-powered off-grid electricity generation facilities. Large water storages are found in the North West Minerals Province of the Southern Gulf catchments. The two major storages, Lake Moondarra (107 GL capacity) and Lake Julius (107 GL capacity), are both on the Leichhardt River. They provide a reliable potable water supply to Mount Isa and through transmission pipelines to mining enterprises in the study area. The North West Queensland Water Pipeline from Lake Julius also supplies water to Ernest Henry Mine and Cloncurry, both in the neighbouring Flinders catchment. Other large water storages are directly linked to mining, with two large on-property water storage facilities supporting small-scale irrigation development with surface water allocations of between 1000 and 8000 ML/year. Some moderate surface water licences drawing water from permanent rivers (between 400 and 1000 ML/year) have been granted for town and community water supply. Much smaller surface water licences (<50 ML/year) are associated with stock use. Currently 13 groundwater licences have been granted for a variety of applications. The largest entitlements (150 to 1400 ML/year) are associated with industrial use in mining. Two licensed entitlements of approximately 100 ML/year have been granted for town and community water supplies, and the smallest groundwater licences (<100 ML/year) have been granted for a variety of industrial and agricultural uses. Groundwater is sourced from a variety of aquifers hosted in different geological units. Indigenous values and development objectives This section gives an overview of the information needed on Indigenous water issues in the Assessment area to provide foundations for community consultations and involvement in further research, and planning and decision making with government and industry. The sub-project focuses on publicly available literature with additional material to be provided through discussions with Indigenous land-owning groups within the catchments. The literature review includes the previous Northern Australia Water Resource Assessment reports on Indigenous water rights, values, interests and development goals; publications on Indigenous water values and the colonisation of northern Australia; and reports related to development in the Southern Gulf catchments. The language groups of this Assessment area are: •Leichhardt catchment – Kalkadoon, Mitakoodi, Wakabunga, Mayi-Kutuna, Mayi-Thankurti, Mayi-Yapi, Mayi-Yali and Kukatj•Nicholson catchment – Waanyi, Garawa, Wakabunga, Nguburinji, and Gangalidda•Wellesley Islands – Lardil, Yangkaal, Kaiadilt•Settlement catchment – Garawa, Gangalidda•Morning Inlet catchment – Kukatj. This sub-project builds on findings from the consultations with Indigenous Peoples from the previously conducted Northern Australia Water Resource Assessment reports to offer a regionally specific assessment designed to help non-Indigenous decision makers and others understand general Indigenous valuations of water and relationships to Country and the rights and interests attached to those. It highlights Indigenous perspectives on water-related development, development objectives, and cultural heritage concerns related to access to pastoral and mining leases and Indigenous land and water management and responsibilities. Studies in the mining and community sectors demonstrate increasing concerns about the cumulative impact of development on groundwater and surface water supply and quality. Indigenous Peoples and the groups they belong to have significant land holdings and rights in Country through: •the Commonwealth Aboriginal Land Rights (Northern Territory) Act 1976 (ALRA) •the Northern Territory Aboriginal Sacred Sites Act 1989 (NT) •Queensland Aboriginal Land Act 1991 (ALA) deeds of grant in trust (DOGIT) •native title determinations exclusive and non-exclusive (Queensland) •the Queensland Aboriginal Cultural Heritage Act 2003 •Indigenous Land Use Agreements (ILUAs) (Queensland). Indigenous Peoples within the catchments have a long history of engaging with development interests in pastoralism and mining, two sectors that influence contemporary water values. Indigenous owners in the catchments have actively negotiated with sector development interests to facilitate aligning their economic viability and sustainability objectives with cultural and environmental goals. There are examples of business development successes with Kalkadoon Native Title Aboriginal Corporation Registered Native Title Body Corporate (RNTBC) and the Waanyi People through business partnerships with mining companies. The Lawn Hill and Riversleigh Pastoral Holding Company is a significant agreement that supports the economic development of Waanyi People. However, broader development outcomes of expanding human capability and establishing new regional opportunities are not yet being realised. Ranger groups are engaged in fee-for-service work and, more broadly, environmental management, including collaborations with Southern Gulf Natural Resource Management (NRM) in Queensland. Indigenous representative agencies and industry more broadly are currently being engaged by the Queensland Government in reviewing the Water Plan (Gulf) 2007 (Queensland Government, 2007). 104 | Water resource assessment for the Southern Gulf catchments Queensland Government in reviewing the Water Plan (Gulf) 2007 (Queensland Government, 2007). Legal and policy environment Proponents must be aware of the complex legal, policy and regulatory landscape when contemplating and planning land and water developments within the Southern Gulf catchments. As part of their due diligence process, proponents must secure appropriate land tenure, secure the necessary authorisations to take water, and obtain a range of government approvals prior to commencing construction and operation of a development. The Southern Gulf catchments straddle the border between the NT and Queensland. This means that government powers and responsibilities for managing land and water resources in the Southern Gulf catchments are shared between the Australian Government, the NT and Queensland governments and local governments. The NT and Queensland governments have primary responsibility for land, water and environmental policy and laws. The Australian Government has powers under the EPBC Act relating to matters of national environmental significance (including those arising from the World Heritage Convention, the Ramsar Convention on Wetlands of International Importance and the Convention on Biological Diversity) and the native title rights of Indigenous Peoples. Local governments are established within the states and territories. In Queensland, these have responsibility for land use planning, which involves establishing local planning schemes that regulate land use and development. However, in the NT, the planning system is administered by the NT Government rather than local government. 3.1.2 Introduction This chapter seeks to address the following questions. In the Southern Gulf catchments, what are the existing: •ecological systems •demographic and economic profiles, land use, industries and infrastructure •values, rights, interests and development objectives of Indigenous Peoples? The chapter is structured as follows: •Section 3.2 examines the ecological systems and assets of the Southern Gulf catchments, including the key habitats and biota and their important interactions and connections. •Section 3.3 examines the socio-economic profile of the Southern Gulf catchments, includingcurrent demographics, existing industries and infrastructure of relevance to water resourcedevelopment. •Section 3.4 examines the Indigenous values, rights, interests and development objectives ofTraditional Owners from the Southern Gulf catchments, generated through a literature review ofjournal articles, grey literature and media articles. 3.2 Southern Gulf catchments and their environmental values This section provides an overview of the environmental values and freshwater, marine and terrestrial ecological assets found in the Southern Gulf catchments. Unless otherwise stated, the material in this section is based on findings described in the companion technical report on ecological assets (Merrin et al., 2024). The Southern Gulf catchments span an area of 108,200 km2, of which 21% lies in the NT and 79% lies in Queensland. They are composed of Settlement (17,600 km2), Nicholson (52,200 km2), Leichhardt (33,400 km2) and Morning Inlet (3,700 km2) catchments, and the Wellesley island groups (1,200 km2). Agricultural production is the largest land use in the catchments, mostly cattle grazing on native pastures (77%) (ABARES, 2022). Other land uses include recreational activities, tourism, traditional and commercial fisheries, mining and Indigenous uses. In addition, these catchments have important ecological and environmental values. Within these catchments and the surrounding marine environment are rich and important ecological assets, including species, ecological communities, habitats, and ecological processes and functions (see the conceptualised summary in Figure 3-2). The ecology of the Southern Gulf catchments is maintained by the flow regime in each catchment, shaped by the region’s wet-dry climate and complex geomorphology and topography and driven by seasonal rainfall, evapotranspiration and groundwater discharge. For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-2 Conceptual diagram of selected ecological assets of the Southern Gulf catchments Ecological assets include species of significance, species groups, important habitats and ecological processes. See Table 3-2 for a complete list of the fresh water–dependent, marine and terrestrial ecological assets considered in the Southern Gulf catchments. Biota icons: adapted from Integration and Application Network (2023) The Southern Gulf catchments have a highly seasonal climate with an extended dry season. Rainfall averages 602 mm/year with 94% of the rainfall falling in the wet season (McJannet et al., 2023). The dominant vegetation types in the catchments are open eucalypt open woodlands, Melaleuca forests and woodlands, and tussock grasslands (Department of Climate Change‚ Energy‚ the Environment and Water, 2020). The Southern Gulf catchments have two major water storages: Lake Julius and Lake Moondarra. Both storages are located on the Leichhardt River and are listed in the Directory of Important Wetlands in Australia (DIWA). While the water in these two storages predominantly supplies urban, mining and industrial demand around Mount Isa and Cloncurry, the permanent water provides important dry-season refuge for waterbirds and supports a variety of freshwater fish species (Department of Agriculture‚ Water and the Environment, 2021a). During the wet season, flooding inundates significant parts of the downstream reaches of the catchments connecting wetlands to the river channel, inundating floodplains and driving a productivity boom. This flooding is particularly evident in the downstream parts of the catchments, including the floodplain wetlands, and extensive intertidal flats on the mainland coastline south of Bentinck and Sweers islands, where it delivers extensive discharges into the marine waters of the south-western Gulf of Carpentaria. While most rivers in the Southern Gulf catchments are ephemeral, the Gregory and O'Shannassy rivers and Lawn Hill Creek are perennial, being fed by groundwater from the Thorntonia Limestone (see Section 2.5.4). These perennial waterways provide critical refuge habitat for many aquatic species in this semi-arid environment. During the dry season, river flows are reduced and the streams in the catchment contract, resulting in series of instream waterholes that also provide critical habitat in the dry season. Protected, listed and significant areas of the Southern Gulf catchments The protected areas located in the Southern Gulf catchments include the UNESCO World Heritage-listed Australian Fossil Mammal Sites (Riversleigh), three Indigenous Protected Areas (IPAs) – Ganalanga-Mindibirrina, Nijinda Durlga and Thuwathu/Bujimulla – and Boodjamulla (encompassing Lawn Hill) and Finucane Island national parks and other conservation parks (Figure 3-4). The Australian Fossil Mammal Sites (Riversleigh) is largely situated within the Boodjamulla National Park (Aboriginal Land). It is considered one of the richest and most extensive fossil deposits in the world. The Ganalanga-Mindibirrina IPA (NT) is located in the upper reach of the Nicholson River and covers over 1 million ha. The Nijinda Durlga IPA covers over 186,850 ha and includes habitat for marine turtles, dugongs, shorebirds and seabirds. The Thuwathu/Bujimulla IPA spans across the Wellesley Islands and includes over 1.6 million ha of marine area and over 120,000 ha of land. The area contains significant habitat for sea turtles, shorebirds and seabirds. It is also culturally significant, including having the largest collection of stone fish traps in the southern hemisphere (Australian Indigenous Australians Agency, 2023). Boodjamulla National Park (Aboriginal Land) includes sandstone ranges, Lawn Hill Gorge and the Australian Fossil Mammal Sites (Riversleigh) UNESCO World Heritage site. Lawn Hill Gorge is formed by Lawn Hill Creek, which is groundwater fed, as are the Gregory and O'Shannassy rivers. Lawn Hill Creek and Gregory River contain wet riverine forest and support a variety of plant species. Lawn Hill Creek supports freshwater turtles, including the Gulf snapping turtle (Elseya lavarackorum; listed as Endangered under the Environment Protection and Biodiversity Conservation Act 1999 (EPBC Act)) and more than 20 species of fish. Across the catchment, permanent water sources provide important refuge habitat for a broad variety of species (Queensland Department of Environment and Science, 2023a). Finucane Island National Park, a 7610 ha island within a river network, is located within the Southern Gulf Aggregation DIWA site. It includes estuarine wetlands, salt flats, mangroves and grasslands. The park provides important habitat for fish and waterbirds (Queensland Department of Environment and Science, 2023b). The Southern Gulf catchments contain 13 nationally significant wetlands listed in the DIWA: Bluebush Swamp, Buffalo Lake Aggregation, Forsyth Island Wetlands, Gregory River, Lake Julius, Lake Moondarra, Lawn Hill Gorge, Marless Lagoon Aggregation, Musselbrook Creek Aggregation, Nicholson Delta Aggregation, Southern Gulf Aggregation, Thorntonia Aggregation and Wentworth Aggregation (Figure 3-4) (Department of Agriculture‚ Water and the Environment, 2021a). These DIWA-listed wetlands include a variety of wetland types, ranging from estuarine wetlands with salt flats and saltmarshes to artificial lakes and spring-fed creeks and rivers. No wetlands listed under the Ramsar Convention on Wetlands of International Importance occur within the Southern Gulf catchments. In Queensland, statewide vegetation mapping classifies remnant vegetation communities as regional ecosystems and includes an assessment of the condition of the remnant vegetation with a biodiversity status (Queensland Department of Environment, Science and Innovation, 2024a). Regional ecosystems with a biodiversity status of endangered or of concern are shown on Figure 3-5. This data underpins the Vegetation Management Act 1999 and Planning Act 2016 (Qld) where applications for clearing for development are submitted. Figure 3-3 Estuarine crocodiles inhabit fresh and saltwater environments Photo: CSIRO – Nathan Dyer Ecology wetlands and protected areas map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\4_S_Gulf\1_GIS\1_Map_docs\Ec-S-505_CR_location_map_detailed_v14.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-4 Location of protected areas and important wetlands within the Southern Gulf catchments Assessment area Includes management areas protected mainly for conservation through management intervention as defined by the International Union for Conservation of Nature (IUCN). Dataset: Department of Agriculture‚ Water and the Environment (2020a); Department of Agriculture‚ Water and the Environment (2020b) Department of the Environment and Energy (2010) Qld Regional ecosystem biodiversity status, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\CR-S-519_RemVeg_Status_v2StandardColor.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-5 Regional ecosystem biodiversity status in the Queensland part of the Southern Gulf catchments Datasets: Department of Agriculture‚ Water and the Environment (2020b); Queensland Department of Environment, Science and Innovation (2024a) Important habitat types and values of the Southern Gulf catchments The freshwater sections of the Southern Gulf catchments provide diverse habitats, including persistent and ephemeral rivers, wetlands, floodplains and groundwater-dependent ecosystems (GDEs). The diversity and complexity of habitats, and the connections between habitats within a catchment, are vital for providing the range of habitats needed to support both aquatic and terrestrial biota (Schofield et al., 2018). In the wet season, flooding connects rivers to floodplains. The exchange of water between the river and the floodplain supports higher levels of primary and secondary productivity than is found in surrounding areas that are less frequently inundated (Pettit et al., 2011). Infiltration of water into the soil during the wet season and along persistent streams often enables riparian habitats to form an important interface between the aquatic and terrestrial environments. While riparian habitats often occupy a relatively small proportion of the catchment, they frequently have a higher species richness and abundance of individuals than surrounding habitats (Pettit et al., 2011; Xiang et al., 2016). In the dry season, biodiversity is supported by the perennial rivers and creeks, permanent lakes and inchannel waterholes. These water sources that persist through the dry season become increasingly important as the season progresses; they provide important refuge habitat for species and enable recolonisation into surrounding habitats upon the return of larger flows (Hermoso et al., 2013). Persistent waterholes provide 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). The terrestrial habitats of northern Australia include a range of diverse 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. Many of these are highly dependent upon fresh water supplied from rivers, and their persistence and condition can also be supported by groundwater discharges. Riparian and floodplain communities from the Queensland remnant vegetation regional ecosystem data with endangered and of concern biodiversity status are well represented in the Southern Gulf catchments (Figure 3-5). Areas where the dominant regional ecosystem is endangered (0.7% of the Queensland section of the Southern Gulf catchments, Table 3-1) are spread across riparian zones of the higher relief areas with associated seasonally wet areas nearby having endangered sub-dominant regional ecosystems (3.2%). Areas of the more frequently flooded floodplains of the larger rivers and creeks have dominant regional ecosystems classified as of concern (10.5%) with larger areas of the floodplain less frequently flooded with a biodiversity ‘of concern’ on one or more of the subdominant regional ecosystems (12%) (Queensland Department of Environment, Science and Innovation, 2024a). Table 3-1 Categories of biodiversity status of the Queensland regional ecosystems CATEGORY DEFINITION SUBCLASS AREA OF THE QUEENSLAND PART OF THE SOUTHERN GULF CATCHMENTS (HA) PERCENTAGE OF THE QUEENSLAND PART OF THE SOUTHERN GULF CATCHMENTS (%) Endangered Remnant vegetation is less than 10% of its pre- clearing extent across the bioregion; or 10–30% of its pre-clearing extent remains and the remnant vegetation is less than 10,000 ha. Dominant† Sub- dominant‡ 59,417 271,745 0.7 3.2 Of concern Remnant vegetation is 10 to 30% of its pre- clearing extent across the bioregion; or more than 30% of its pre-clearing extent remains and the remnant extent is less than 10,000 ha. Dominant† Sub- dominant‡ 898,058 1,025,883 10.5 12 †‘Dominant’ subclass means that greater than 50% of the polygon contains the regional ecosystem mapping. ‡‘Sub-dominant’ subclass means that less than 50% of the polygon contains the regional ecosystem mapping. GDEs occur across many parts of the Southern Gulf catchments and come in different forms, including aquatic, terrestrial and subterranean habitats. Aquatic GDEs 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 they provide critical refuge during periods of the late dry season (James et al., 2013). Vegetation occurring adjacent to the waterways in the Southern Gulf catchments relies on water from a range of sources (surface water, soil water, groundwater) which are seasonally dynamic and highly spatially variable. Water may be sourced 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 has only been infrequently investigated (e.g. Canham et al. (2021)). In some locations, vegetation may be sustained by water available in soils and so never use groundwater. In other locations, vegetation may use groundwater sourced from local alluvial recharge processes; alternatively, regional groundwater may be critical for maintaining vegetation condition. Subterranean aquatic ecosystems in the limestone that occur in the upper south-western region of the Southern Gulf catchments support subterranean fauna that depend on the presence of groundwater (e.g. troglofauna, which live in caves, and stygofauna, which live in groundwater systems). The marine and estuarine environments of the Southern Gulf catchments, including the mainland area adjacent to Mornington and Sweers islands, have 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 (Poiner et al., 1987). Seagrass beds in the nearby coastal Gulf of Carpentaria have high diversity, vigorous stands and provide an important food and habitat 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 (Bayliss et al., 2014). Mud crabs (mainly Scylla serrata) are also harvested (Bayliss et al., 2014). Significant species and ecological communities of the Southern Gulf catchments A number of aquatic and terrestrial species occurring in the Southern Gulf catchments are currently listed as Critically endangered, Endangered or Vulnerable under the EPBC Act and by the NT Government’s wildlife classification system and the Queensland Government’s threatened species conservation system. These include freshwater or large tooth sawfish (Pristis pristis; Vulnerable, EPBC Act) and the Gulf snapping turtle (Elseya lavarackorum; Endangered, EPBC Act). The Southern Gulf catchments are important stopover habitat for migratory shorebird species listed under the EPBC Act, including the eastern curlew (Numenius madagascariensis; Critically endangered) and the Australian painted snipe (Rostratula australis; Endangered) (Atlas of Living Australia, 2021; Department of Agriculture‚ Water and the Environment, 2021b). While there are many plant species that fall into the endangered and of concern regional ecosystems from the Queensland remnant vegetation mapping, the regional ecosystem that is most common in the endangered dominant category in the Southern Gulf catchments is the Eucalyptus camaldulensis woodland on channels and levees (regional ecosystem (RE) class 1.3.7), a regional ecosystem of the riparian zone and an important seasonal water bird habitat also providing regional corridors for fauna. The most common regional ecosystem in the ‘of concern’ dominant category is the woodlands of the creek and river frontage (regional ecosystem (RE) class 2.3.20) providing refuges for flora and fauna (Queensland Department of Environment, Science and Innovation, 2024a). Threatening processes for these two areas are the dry season impacts of high grazing pressure leading to habitat degradation and loss leaving areas prone to weed infestation. 3.2.1 Potential threats in the Southern Gulf catchments Land use practices and ecology 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). 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 prawns, mud crabs and barramundi. Riparian vegetation characteristics of the Southern Gulf catchments are considered to not be affected by extensive clearing or development, although impacts that occur are often associated with stock and pest species (Department of Agriculture‚ Water and the Environment, 2021a). A range of economic enterprises, infrastructure and human impacts occur in the Southern Gulf catchments, and the nature and extent to which these have modified habitats and affected species of the Southern Gulf catchments varies. The study area includes the towns of Burketown and Doomadgee and the city of Mount Isa, which support tourism and mining and act as regional hubs for many of the cattle stations across the catchments. While a proportion of the catchments are under conservation reserves, the study area does face environmental threats, including the potential for further mining and tourism-related impacts at sensitive and vulnerable sites. One of the most significant environmental threats to remote regions across northern Australia is that of introduced plants and animals. In the Southern Gulf catchments, introduced animals include feral horses (Equus caballus), wild pig (Sus scrofa) and cane toads (Rhinella marina) (Department of Agriculture‚ Water and the Environment, 2021a; 2021b). Weeds of national significance in the aquatic systems of northern Australia include salvinia (Salvinia molesta) and rubber vine (Cryptostegia grandiflora) (Close et al., 2012). Weed species of interest in and around the Southern Gulf catchments include prickly acacia (Vachellia nilotica), buffel-grass (Cenchrus ciliaris), rubber vine and water hyacinth (Eichhornia crassipes) (Department of Agriculture‚ Water and the Environment, 2021b). Water resource development and ecology Impacts associated with water resource development include the following, which are described below: • flow regime change • altered longitudinal and lateral connectivity • habitat modification and loss • increased invasive and non-native species • 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 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 and to which life-history strategies are adapted. Flow 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). Through the attenuation of flows, water resource development 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). Altered 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, dams and causeways, or a reduction in the magnitude of flows (and the duration or frequency), can affect 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 and enabling movement of biota into new habitats and their return to refuge habitats during drier conditions (Crook et al., 2020). Habitat modification and loss Water resource development can cause direct loss of habitat. For example, artificially creating a lake (inundated) habitat behind an impoundment results in loss of terrestrial and stream habitat. Agricultural development converts existing habitat to more-intensive agriculture. Infrastructure, including roads and canals, can fragment terrestrial habitat or artificially connect aquatic habitats that had been historically separated. Increased invasive and non-native species Water resource development often homogenises flow and flow-related habitats, for example, through changed patterns resulting from capture and release of flows or creation of impoundments for storage and regulation. Invasive species are often at an advantage in such modified habitats (Bunn and Arthington, 2002). Modified landscapes, such as lakes or perennial streams that were previously ephemeral, can be a pathway for introduction of, and support the incidental, accidental or deliberate establishment of, non-native species, including pest plants and fish (Bunn and Arthington, 2002; Close et al., 2012; Ebner et al., 2020). Increased human activity can increase the 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 increased stock pressure; and 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 or accompanying water resource development. Other pressures 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 Southern Gulf catchments and discuss the likely impacts of future water resource development on this system, a suite of ecological assets has been selected (Table 3-2). Assets are classified as species, species groups or habitats. They can be considered as either partially or fully dependent on fresh water, or terrestrial or marine dependent upon freshwater flows (or services provided by freshwater flows). This chapter considers a key subset of assets, as listed in Table 3-2. More information on the ecological assets of the Southern Gulf catchments and their distribution is available in the companion technical report on ecological assets (Merrin et al., 2024). Chapter 7 presents results of the modelling and analysis to explore the potential of change to these assets as a consequence of water resource development. Table 3-2 Freshwater, marine and terrestrial ecological assets with freshwater flow dependences in the Southern Gulf catchments An asterisk (*) represents an asset outlined in this report. All listed species, species groups and habitat assets are detailed in the companion technical report on ecological assets (Merrin et al., 2024). For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au. 3.2.2 Ecological assets in the Southern Gulf catchments Northern Australia’s rivers, floodplains and coastal regions contain high biodiversity, including at least 170 fish species, 150 waterbird species, 30 aquatic and semi-aquatic reptile species, 60 amphibian species and 100 macroinvertebrate families (van Dam et al., 2008). The ecology of the freshwater systems is supported by, and adapted to, the highly seasonal flow regimes of the wet-dry tropics. Water resource development and climate change threaten to affect these habitats and species. This section provides a synthesis of the prioritised assets relevant to the Southern Gulf catchments for the purpose of understanding potential ecological outcomes of flow regime change. Table 3-2 lists the assets evaluated in the Southern Gulf catchments ecology activity. Floodplain wetlands Wetlands in the wet-dry tropics of Australia have great conservation value (Finlayson et al., 1999) and are 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) (Figure 3-6), spawning grounds and nurseries for floodplain-dependent fish (Ward and Stanford, 1995), and habitat for many other aquatic and riparian species (van Dam et al., 2008). 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). For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au. Figure 3-6 Brolgas flying into the sunset at Lake Moondarra Photo: CSIRO Hydrological regimes are fundamental to sustaining the 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 often highly dependent on the seasonal rainfall–runoff pattern and the associated low and high flows from rivers (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 these 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 have the greatest impact on the ecological values, which include species diversity, productivity and habitat structure (Close et al., 2015). Under the Ramsar Convention (Ramsar Convention Secretariat, 2004), wetlands are defined as: 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 NT Government defines wetlands as including coastal saltmarshes, 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 (NT Government, 2020). For delineation, this Assessment considers areas within the river channel to be inchannel waterholes rather than wetlands (see Merrin et al. (2024)). Similarly, marine or saline habitats including mangroves and coastal saltmarshes (salt flats) are also treated as separate assets in this project (see Saltpans and salt flats below and Merrin et al. (2024)). The Southern Gulf catchments have 13 nationally significant wetlands listed under the DIWA (Table 3-3 and Figure 3-7) (Department of Agriculture‚ Water and the Environment, 2021a). There are no Ramsar-listed wetlands within the Southern Gulf catchments. Table 3-3 Nationally important wetlands in the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au. The Wentworth Aggregation is found in the Settlement Creek catchment (Figure 3-7) and covers an area of 82,306 ha (Department of Agriculture‚ Water and the Environment, 2021a). It contains the full range of wetland types (estuarine, lacustrine (lake), palustrine (marshes and swamps) and riverine) and is considered to have high wilderness value due to its remoteness. The main habitat types are estuarine salt flats and saltmarshes (55.6% of the wetland area); coastal and sub-coastal non-floodplain tree swamp – Melaleuca spp. and Eucalyptus spp. (13.5% of the wetland area); coastal and sub-coastal floodplain grass, sedge, herb swamp (7.2% of the wetland area); and estuarine – mangroves and related tree communities (6.7% of the wetland area) (Queensland Department of Environment and Science, 2022). It is considered an important wetland for waterbirds (Department of Agriculture‚ Water and the Environment, 2021a). The Marless Lagoon Aggregation spans the Settlement Creek and Nicholson River catchments (Figure 3-7) and has an area of 166,948 ha (Department of Agriculture‚ Water and the Environment, 2021a). It has extensive palustrine wetlands (97.6% of the wetland area) in which the dominant habitat type is coastal and sub-coastal non-floodplain tree swamp – Melaleuca spp. and Eucalyptus spp. (97.3% of the wetland area) (Queensland Department of Environment and Science, 2022). For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-7 Land subject to inundation (potential floodplain wetlands) and important wetlands in the Southern Gulf catchments Dataset: Geoscience Australia (2017); Department of the Environment and Energy (2010) Bluebush Swamp is located in the Nicholson catchment (Figure 3-7) and is 879 ha (Department of Agriculture‚ Water and the Environment, 2021a). It is a scrub–shrub wetland with Acacia stenophylla as the dominant species, and it has areas of shallow, open water. Bluebush Swamp provides habitat for waterbirds in the late wet season and in autumn and spring (Department of Agriculture‚ Water and the Environment, 2021a). Lawn Hill Gorge is located within Boodjamulla National Park (Aboriginal Land) in the Nicholson catchment (Figure 3-7) and has an area of 1133 ha (Department of Agriculture‚ Water and the Environment, 2021a). Prior to being gazetted as a national park, it was extensively grazed, and feral pigs and other invasive animals continue to affect the area. Various recreational activities are available within the national park, although visitor numbers are controlled. Indigenous usage of the area dates back between 17,000 and 30,000 years (Department of Agriculture‚ Water and the Environment, 2021a). The gorge itself is located on Lawn Hill Creek, which is fed by groundwater from the Thorntonia Limestone and also receives wet-season flushes (Department of Agriculture‚ Water and the Environment, 2021a; Queensland Department of Environment and Science, 2023a). Musselbrook Creek Aggregation is located in the Nicholson catchment (Figure 3-7) and has an area of 45,114 ha (Department of Agriculture‚ Water and the Environment, 2021a). The dominant habitat types are coastal and sub-coastal floodplain tree swamp – Melaleuca spp. and Eucalyptus spp. (34.6% of the wetland area), arid and semi-arid tree swamp (floodplain) (32.8% of the wetland area) and river (27.6% of the wetland area) (Queensland Department of Environment and Science, 2022). The area has been extensively grazed by cattle and feral horses (Department of Agriculture‚ Water and the Environment, 2021a). Nicholson Delta Aggregation is located in the Nicholson catchment (Figure 3-7) and has an area of 63,646 ha (Department of Agriculture‚ Water and the Environment, 2021a). This aggregation has a fresh water – saltwater gradient, as estuarine waters flood the salt flats and tidal channels, particularly during the dry season. There is also a series of freshwater wetlands (Department of Agriculture‚ Water and the Environment, 2021a). The main habitat is estuarine – salt flats and saltmarshes (40.8% of the wetland area) and river (31.1% of the wetland area) (Queensland Department of Environment and Science, 2022). The wetlands in this aggregation are a mix of permanent, semi-permanent and seasonal wetlands, and they provide refugia habitat for waterbirds in the dry season (Department of Agriculture‚ Water and the Environment, 2021a). Thorntonia Aggregation is located in the catchment of the Gregory River (Figure 3-7) and has an area of 298,629 ha (Department of Agriculture‚ Water and the Environment, 2021a). It is partly within the Boodjamulla National Park (Aboriginal Land) and contains the internationally significant Riversleigh fossil field. It has deep permanent channels which are spring fed and shallower seasonal channels. The perennial channels provide refugia habitat in the dry season (Department of Agriculture‚ Water and the Environment, 2021a). The Gregory River DIWA site spans the Nicholson and Leichhardt catchments (Figure 3-7) and has an area of 26,630 ha (Department of Agriculture‚ Water and the Environment, 2021a). It is the largest perennial river in semi-arid and arid Queensland that is spring fed with additional flow from runoff. The area is used for recreational purposes, including camping, canoeing and fishing, and has had extensive cattle grazing. Freshwater crocodiles (Crocodylus johnstoni) are very common (Department of Agriculture‚ Water and the Environment, 2021a). Lake Julius is an artificial lake located in the Leichhardt catchment (Figure 3-7). Lake Julius is on the Leichhardt River and has a surface area of 1255 ha at full supply. It was created when the Julius Dam was completed in 1976 and provides backup supplies to Lake Moondarra as well as supplying water to Cloncurry Council and mines in the North West Minerals Province (Department of Agriculture‚ Water and the Environment, 2021a). The lake is used for recreational purposes, including boating and fishing, and the land surrounding the lake is used to graze cattle. As a permanent lake, it provides important habitat for waterbirds during the dry season. It also supports a variety of freshwater fish, red-clawed crayfish (Cherax quadricarinatus) and freshwater crocodiles (Department of Agriculture‚ Water and the Environment, 2021a). Lake Moondarra (Figure 3-7) is an artificial lake, completed in 1957 and raised in 1971, of 2190 ha and located just outside the town of Mount Isa in the Leichhardt catchment (Figure 3-7). Situated on the Leichhardt River, it is upstream of Lake Julius. As a permanent water body, it provides refugia habitat for waterbirds and is considered an important recreation area, allowing for both boating and fishing. Cattle grazing is extensive, often resulting in damage at the water’s edge. Seasonal turbidity is an issue (Department of Agriculture‚ Water and the Environment, 2021a). Buffalo Lake Aggregation has an area of 1911 ha and is located in the Morning Inlet catchment (Figure 3-7) (Department of Agriculture‚ Water and the Environment, 2021a). The lake is shallow (<1 m depth) and ephemeral, flooding in extreme wet seasons and drying out completely most dry seasons. It is also occasionally flooded by tidal surges (Department of Agriculture‚ Water and the Environment, 2021a). The dominant habitats are coastal and sub-coastal floodplain lakes (69.0% of the wetland area) and coastal and sub-coastal floodplain grass, sedge, herb swamps (24.3% of the wetland area) (Queensland Department of Environment and Science, 2022). The lake habitat provides important habitat for waterbirds, including migratory species. The area has been extensively grazed (Department of Agriculture‚ Water and the Environment, 2021a). Southern Gulf Aggregation is the largest continuous estuarine wetland in Australia at 545,577 ha, and it spans all four of the Assessment catchments (Figure 3-7) (Department of Agriculture‚ Water and the Environment, 2021a). Dominated by estuarine – salt flats and saltmarshes (77.3% of the wetland area) and estuarine – mangroves and related tree communities (20.4% of the wetland area) (Queensland Department of Environment and Science, 2022), the area is governed by estuarine tides, and during the wet season, freshwater flooding (Department of Agriculture‚ Water and the Environment, 2021a). The wetlands located along the inland edges of the aggregation are all seasonal and are brackish. The area is considered one of the most important shorebirds sites in Australia (Department of Agriculture‚ Water and the Environment, 2021a). Forsyth Island Wetlands are located on Forsyth Island, about 10 km off the coast of Bayley Point. It is an estuarine wetland with important seagrass habitats in Government Bay (Department of Agriculture‚ Water and the Environment, 2021a). The habitat consists of estuarine – salt flats and saltmarshes (60.9% of the wetland area) and estuarine – mangroves and related tree communities (39.1% of the wetland area) (Queensland Department of Environment and Science, 2022). The island itself is an Indigenous reserve, with the area, including the surrounding waters, used for fishing and hunting (Department of Agriculture‚ Water and the Environment, 2021a). Significant parts of the floodplain areas within the Southern Gulf catchments are already incorporated into the existing 13 nationally significant wetlands (see land subject to inundation Figure 3-7). There is additional floodplain on Musselbrook Creek, extending beyond the Musselbrook Creek Aggregation. There is also additional floodplain near the Wentworth Aggregation ( Saltpans and salt flats Saltpans and salt flats are intertidal areas that are devoid of marine plants and are located between mangroves and saltmarsh meadows. Saltmarshes (Figure 3-8) occur in the supra-littoral zones, which are inundated only infrequently by the tide and where subsequent water evaporation leaves behind expanses of minerals and salts (Cotin et al., 2011). Inundation of saltpans mostly occurs during the annual wet season when large tides and rainfall surface runoff ponds as shallow wetted areas within the saltpans and shallow tidal-cut gutters that intersect them. Despite their infrequent inundation, saltpans provide habitat for some estuarine fish, such as barramundi (Russell and Garrett, 1983) and metapenaeid shrimps (Bayliss et al., 2014) during periods when the tide covers these habitats. Photo of saltpan. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au. Figure 3-8 This saltpan area in northern Australia is typical in being located between mangrove and saltmarsh areas Photo: Nathan Waltham In northern Australia, saltpan sediments are infused with dormant algae that remain inactive in a desiccated state during the dry season (most of the year). However, during overbank inundation from flooded rivers or extensive rainfall, the saltpan soil algae become active and photosynthesise, increasing nutrient contribution to the ecosystem (Burford et al., 2016). After several days, active algal growth occurs, and carbon, nitrogen and phosphorous compounds are produced. Estimates suggest that saltpans can contribute up to an extra 13% of ecosystem primary productivity depending on the extend of saltpan inundation during the wet season (Burford et al., 2016). Saltpans would be most productive during high-level overbank flood flows. The inundation of saltpans expands the habitat available to estuarine benthic fish and crustaceans that can tolerate euryhaline conditions. In northern Australia, coastal saltpans can extend tens to hundreds of square kilometres. They provide habitat for a range of benthic fauna (Dias et al., 2014), which are an important food source for high-order consumers, including shorebird species that use saltpans as resting and/or feeding areas during their migration, which can include long flights to Asia (Cotin et al., 2011; Lei et al., 2018; Rocha et al., 2017). The extent of saltpans in Australia is unknown, though they are common and extensive in more arid coastal areas, most notably in northern Australia (Duke et al., 2019). The northern Australian coastline extends for thousands of kilometres and is relatively pristine; low beach profiles backed by extensive saltpans, possibly 5 to 10 km inland, are characteristic of hundreds of kilometres of coastline (Short, 2022). Despite limited tidal exchange, saltpans provide important habitat resources for migratory birds that access these areas for feeding and shelter (Lei et al., 2018). In addition, saltpans provide erosion and sediment accumulation opportunities in estuaries as well as carbon sequestration services. Saltpans in the Southern Gulf estuaries are extensive and located behind tide-dominated beaches (Figure 3-10) (Short, 2020). They are mostly restricted to a tidal inundation area on the landward side of the mangroves that line the main river channel, but they also occur adjacent to Buffalo and Sweet swamps. The spatial data presented illustrate the extent of saltpans in these catchments, but the extent was presumably increased as part of the wide-scale dieback of mangroves in the Gulf of Carpentaria between late 2015 and early 2016 (Duke et al., 2017). There has been no targeted scientific survey of fish and crustacean communities over the saltpans of the Southern Gulf catchments marine region, presumably because they are located so high in the intertidal zone and are only infrequently covered with tidal water. Also, the catchments are very remote, which leads to difficulties with access and sampling. Figure 3-9 Australian bustards are common in grasslands and woodlands across northern Australia Photo: CSIRO – Nathan Dyer Saltpans, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\4_S_Gulf\1_GIS\1_Map_docs\Ec-S-506_CR_Salt_Flats_v10.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-10 Location of saltpans in the Southern Gulf catchments marine region Data source: Geoscience Australia (2017) Sawfish (Pristis and Anoxypristis spp.) Sawfish belong to the order Pristiformes. They are characterised by a distinctive tooth-lined rostrum or ‘saw’. As adults, sawfish can attain very large sizes, ranging from 5 to 7 m in total length. They are widely distributed in northern Australian marine waters, although they are not necessarily abundant (Last and Stevens, 2008; Morgan, 2011; Stevens et al., 2009). These species can migrate at landscape and oceanic scales through their life cycle. Inshore waters, including bays and estuaries, are important nursery grounds for neonates and juvenile sawfish up until about 4 to 6 years of age (Morgan, 2011; Morgan et al., 2017; Peverell, 2005). Sawfish adults primarily inhabit tropical and subtropical coastal marine waters (Dulvy et al., 2016; Last and Stevens, 2008). Globally, sawfish are considered one of the most threatened marine taxa (Dulvy et al., 2016). Four species of sawfish occur in Australian waters, all listed as being of conservation significance at both national and international levels. The freshwater or large tooth sawfish (Pristis pristis), the green sawfish (P. zijsron) and the dwarf sawfish (P. clavata) are all listed as Vulnerable under the EPBC Act. The narrow sawfish (Anoxypristis cuspidata) is listed as Migratory under the EPBC Act, and because it is also listed in Appendix I and II under the Convention on the Conservation of Migratory Species of Wild Animals (Bonn Convention), it has similar protection status under the EPBC Act. Additionally, sawfish hold significant cultural and spiritual relevance to Indigenous Australians (Ebner et al., 2016). Sawfish are vulnerable to multiple threats, partly due to their morphology (the shape of their rostra) and behaviour and partly due to their life-history characteristics: long lives, slow growth and low reproductive rates, late maturation, relatively low abundance and high specificity of different habitats in different life stages (Peverell, 2005; Phillips et al., 2017; Stevens et al., 2009). Their habitats often overlap with coastal fisheries, making them highly susceptible to capture in gill-net and trawl fisheries and recreational fishing (because of the shape of their rostra). Sawfish rostra have been collected as trophies for decades (McDavitt, 1996), and there is a growing demand for live sawfish for display in public aquaria (Buckley et al., 2020; Compagno et al., 2006). Recent decades have seen high fishing mortality (Fry et al., 2021). Other pressures include cumulative impacts from climate change, habitat loss, artificial passage barriers and declining water quality that may have a significant impact on the movements of sawfish between freshwater and estuarine environments. The Gulf of Carpentaria is a haven ecosystem for the four recognised species of sawfish found in Australia, with observations occurring across the study area. Despite regional population declines due to fishing mortality over the past 50 years, northern Australia has viable populations of sawfish in contrast to sawfish populations elsewhere in Indo-Pacific region waters. Catfish (order Siluriformes) Catfish are a highly diverse group that inhabit both inland and coastal waters. The group includes freshwater species, marine species, and some that move between the river and the estuary (Pusey et al., 2020). Catfish in northern Australia belong to two families: Plotosidae (19 species in total) and Ariidae (17 species). Plotosidae are found in the Eastern Pacific and Indian Ocean and their tributary rivers, whereas Ariidae are a global family found in both freshwater and marine habitats. Most catfish are bottom-feeding omnivores. They feed on algae, submerged macrophytes, invertebrates and smaller fish. Species within the Ariidae are slow growing and generally large bodied. The family is notable for its reproductive traits: it has the largest eggs of any teleost group (>10 mm), and males exhibit strong parental care behaviour, incubating the eggs and developing the young in the mouth for up to 5 weeks (Pusey et al., 2004). Because of the tendency to feed opportunistically, ariid catfish can be very competitive, consuming a variety and large volumes of food. Thus, they can make up a lot of biomass in a catchment (Crook et al., 2020). The key plotosid species are reasonably tolerant of high temperature and low dissolved oxygen levels, but fish kills at very low dissolved oxygen levels have been reported (Bishop, 1980). The key threat to the two dominant Neosilurus species is potential flow barriers. Plotosidae need high flows to trigger spawning migration, and they require a barrier-free passage to spawning grounds in the headwater streams. While not as important as barramundi or sooty grunters (Hephaestus fuliginosus), the fork-tailed catfish (Neoarius graeffei) has considerable importance as a subsistence fish for Indigenous communities (Finn and Jackson, 2011; Jackson et al., 2011). The modelled potential distribution of the fork-tailed catfish is provided in Figure 3-12. The Southern Gulf catchments support catfish from both families with seven species of Ariidae and four species of Plotosidae. The ariid catfish include Sciades leptaspis, Amissidens hainesi, Hexanematichthys mastersi, Nemapteryx armiger, Plicofollis argyropleuron, Neoarius berneyi, Neoarius graeffei, Neoarius midgleyi, Neosilurus ater, Neosilurus hyrtlii, Porochilus argenteus and Sciades paucus. The larger-bodied ariid catfish like N. graeffei, N. berneyi and S. paucus are mainly found on the main stems of the Leichhardt, Gregory and Nicholson rivers. The usually smaller- bodied Neosilurus species are mainly found in smaller tributaries. Figure 3-11 The O’Shannassy River – one of the northern Australian rivers where catfish are found Photo: CSIRO – Nathan Dyer Catfish distribution, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\4_S_Gulf\1_GIS\1_Map_docs\Ec-S-538_CR_SDM_catfish_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-12 Modelled potential species distribution for fork-tailed catfish (Neoarius graeffei) in the Southern Gulf catchments Probability of occurrence is based upon a general linear model. Observation locations of other catfish species is provided in Merrin et al. (2024). Swimming, grazing and diving waterbirds The swimmers, grazers and divers group comprises species with a relatively high level of dependence on semi-open, open and deeper water environments. These species commonly swim when foraging (including diving, filtering, dabbling and grazing) or when taking refuge. In northern Australia, this group comprises 49 species from 11 families, including ducks, geese, swans, grebes, pelicans, darters, cormorants, shags, swamphens, gulls, terns, noddies and jacanas. This group can be further broken down into the subgroups: • diving swimmers – for example, cormorants, pelicans, grebes • aerial divers – for example, terns, gulls, noddies • grazing swimmers – for example, swans, coots, swamphens, ducks, geese. These species breed in Australia and may be sedentary, nomadic, migratory or partially migratory. Nesting generally occurs seasonally, usually in dense vegetation such as emergent macrophytes, trees and shrubs (Garnett et al., 2015). Nests are usually independent or semi-colonial, and while breeding is usually seasonal, it can be stimulated by flooding or large rainfall events (Kingsford and Norman, 2002). Species diets may be piscivorous, omnivorous or herbivorous (Barker and Vestjens, 1990). Changes in water depth, water extent or water duration can expose nests to predation or drowning, or reduce food availability, resulting in nest failure (McGinness, 2016; Poiani, 2006). The magpie goose (Anseranas semipalmata) (Figure 3-13) is one example of the swimmers, grazers and divers group, and while it is an iconic species in northern Australia, it is also the source of some conflict with humans when resources are limited (Corriveau et al., 2022; Frith and Davies, 1961; Traill et al., 2010). The magpie goose is an ancient and unique species of particular importance to Indigenous Peoples, providing eggs, meat and feathers. This species feeds on aquatic vegetation and often nests colonially (Marchant and Higgins, 1990). While currently abundant in northern Australia, wild magpie goose populations have largely disappeared from southern Australia due to human-driven change such as habitat destruction and hunting (Nye et al., 2007). Climate change is likely to enhance the impacts of such changes on magpie geese in northern Australia (Poiani, 2006; Traill et al., 2009). Photo of magpie goose. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au. Figure 3-13 Magpie goose perched on a fallen tree branch Magpie geese are a representative species of the swimmers, grazers and divers group. Photo: CSIRO 3.2.3 Environmental protection A number of aquatic and terrestrial species in the Southern Gulf catchments are currently listed as Critically endangered, Endangered or Vulnerable under the EPBC Act, and by the wildlife classification systems of the NT and Queensland governments, which are based on the International Union for Conservation of Nature (IUCN) Red List of Threatened Species (Figure 3-14). 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, communities, migratory species or wetlands of importance), it requires approval to proceed under the EPBC Act (Table 3-4). This approval is required irrespective of local government policies. The Australian Government Protected Matters Search Tool lists approximately 40 Threatened species for the Southern Gulf catchments, four of which are listed as Critically endangered. Also listed are 64 migratory species. EPBC species, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\4_S_Gulf\1_GIS\1_Map_docs\Ec-S-531_CR_Listed species_v7.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-14 Distribution of species listed under the Environment Protection and Biodiversity Conservation Act 1999 and by the NT and Queensland governments in the Southern Gulf catchments Datasets: Department of Environment Parks and Water Security (2019); Department of Environment and Science (2023) WildNet; Department of the Environment and Energy (2010); Atlas of Living Australia (2023 a,b) Table 3-4 Definition of threatened categories under the Commonwealth Environment Protection and Biodiversity Conservation Act 1999, the NT wildlife classification system, and the Queensland Nature Conservation Act 1992 For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 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 section describes the current social and economic characteristics primarily of the Queensland portion of the Southern Gulf catchments in terms of the demographics of local communities (Section 3.3.2), current industries and land use (Section 3.3.3), and 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 Southern Gulf catchments 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 (Webster et al., 2024a). For specific details related to the NT context, readers are directed to consult the Victoria River Water Resource Assessment technical report on agricultural viability and socio-economics (Webster et al., 2024b) and the Roper River Water Resource Assessment technical report on agricultural viability and socio-economics (Stokes et al., 2023). 3.3.2 Demographics The Southern Gulf catchments comprise four mainland river catchments (Settlement, Nicholson, Leichhardt and Morning Inlet) plus the larger islands of the Mornington Island ‘catchment’ in the Gulf of Carpentaria. The study area falls mainly within Queensland (79%), but the western part of the study area falls within the NT (21%). Within Queensland, the catchments comprise the entire Mornington, Doomadgee and Burke local government areas together with around half of the Mount Isa local government area and smaller parts of the adjacent local government areas of Carpentaria and Cloncurry. Within the NT, the catchments comprise parts of Barkly and Roper local government areas. At the state and territory level, the catchments include part of the electoral division of Traeger in Queensland and a small part of the Barkly electoral division within the NT. At the federal level, the catchments form part of the Division of Kennedy in Queensland and part of the Division of Lingiari in the NT. The population density of the Southern Gulf catchments is low at one person per 4.8 km2, which is about one-fourteenth that of Queensland and one-sixteenth that of Australia as a whole. The Assessment area contains one significant urban area (population >10,000 people): Mount Isa, a city of over 18,000 residents, was developed to support the mining of the extensive deposits in the surrounding area (particularly for lead, silver, copper and zinc). There are also a number of small towns and communities within the catchments, including Burketown and Doomadgee, and on the Wellesley Islands. Of these smaller settlements, only Doomadgee (population 1387 as at the 2021 Census) has a population greater than 1000. The nearest major cities and population centres are the cities of Townsville and Cairns, respectively, approximately 1000 and 1100 km from Mount Isa. The Queensland capital city of Brisbane is approximately 1925 km from Mount Isa. The demographic profile of the catchments, based on data from the 2021, 2016, 2011 and 2006 censuses, is shown in Table 3-5. 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 approximates the Southern Gulf catchments. Instead, data are shown for: (i)Carpentaria (ABS Statistical Area Level 2 (SA2) region 315021404), being the single region thatencompasses the largest geographic area within the boundaries of the catchments ( Table 3-5 Major demographic indicators for the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the boundaries of the catchments. Sources: ABS (2006, 2011, 2016, 2021a) Census data ABS Statistical area for analysis map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-505_Map_Australia_SoG_tourism_SA2_v1.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-15 Boundaries of the Australian Bureau of Statistics Statistical Area Level 2 (SA2) regions used for demographic data in this analysis Inset shows the red square on the main map. Residents of the Southern Gulf catchments tend to be younger and more likely to identify as Indigenous than the typical resident of Queensland and of Australia as a whole. Incomes differ from the national and state mean, being higher than the state and national averages in Mount Isa but lower in the remainder of the study area. The people in the Southern Gulf catchments are predominantly younger (median age around 31) than is typical for Queensland and the country as a whole (median age around 38); however, the trend from 2011 and 2016 to 2021 suggests that the median age of the Southern Gulf catchments is slightly increasing. The population contains a much larger proportion of Indigenous Peoples (27.3%) than Queensland (4.6%) and the country overall (3.2%). The proportion of Indigenous Peoples varies across the study area: the Mount Isa SA2 region has a much smaller proportion than the remainder of the catchments, although at 21.4% Mount Isa is higher than that for Queensland and Australia. Beyond Mount Isa, the proportion of Indigenous Peoples within the population exceeds 60%. The heterogeneity of the Assessment area is most clearly demonstrated when focusing on incomes. The median weekly household income for the Mount Isa SA2 region ($2236) is substantially above that for Queensland ($1675) and Australia ($1746). In contrast, incomes for the remainder of the study area are substantially lower with median weekly income in the Carpentaria SA2 region being only $1279. A similar pattern is observed for households on low incomes. Excluding Mount Isa, the proportion of households on low incomes (i.e. less than $650/week) in the Carpentaria SA2 region was far higher, and the proportion on high incomes (more than $3000/week) far lower, than the proportion for Queensland and for the country as a whole. The opposite applies to the Mount Isa SA2 region. The Socio-Economic Indexes for Areas (SEIFA) metrics are presented in Table 3-6. The Southern Gulf catchments fall below the national mean for each metric. The remote Carpentaria SA2 region, which comprises a large geographic area of the study area, is classified within the first decile for each index, indicating the region is scoring below 90% of the rest of the country on each of the measures. However, the Mount Isa SA2 region, where most of the population live, scores in the 2nd to 4th decile, depending on the specific index. Overall, weighted by population, the catchments are in the 2nd decile for the Index of Economic Resources (IER), the 3rd decile for Index of Education and Occupation (IEO), and the 4th decile for the remaining indices. Thus, while the level of disadvantage varies across the catchments, the different components and overall scores are all disadvantaged compared to the mean for Australia as a whole. Table 3-6 Socio-Economic Indexes for Areas (SEIFA) scores of relative socio-economic advantage for the Southern Gulf catchments Scores are relativised to a national mean of 1000, with higher scores indicating greater advantage. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the boundaries of the catchments. §Based on both the incidence of advantage and disadvantage. *Based purely on indicators of disadvantage. Source: ABS (2023) 3.3.3 Current industries and land use Employment The economic structure of the combined Southern Gulf catchments differs from that of Queensland and Australia as a whole in having lower unemployment rates and a higher proportion of the adult population (aged 15 and older) within the labour force (see participation rates in Table 3-7). However, the employment rates within study area are highly heterogeneous, and data for the Mount Isa SA2 region (entirely within the study area) are very different to the data for the rest of the study area. For example, the Carpentaria SA2 region, which constitutes the greatest area of the Southern Gulf catchments, has far higher unemployment rates and far lower participation rates than the Queensland and national averages (Table 3-7). The Mount Isa SA2 region provides a contrasting story with far lower unemployment and far higher participation rates than those of Queensland and Australia as a whole. Table 3-7 Key employment data for the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the boundaries of the catchments. Source: ABS (2006, 2011, 2016, 2021a) Census data There are also noticeable differences in the industries providing the most jobs within the region (Table 3-7), both within the catchments and compared to Queensland and Australia. ‘Healthcare and social assistance’, ‘Education and training’ and ‘Retail trade’ are important employers in the region and nationally. However, ‘Construction’ and ‘Professional, scientific and technical services’ feature within the top five industries by employment across the nation on average but are far less significant in the Southern Gulf catchments. ‘Public administration and safety’ is relatively more important to the employment prospects of workers in the Southern Gulf catchments than the national average. However, ‘Mining’ is the most important industry in the Southern Gulf catchments overall, and in the Mount Isa SA2 region within the catchments. It provides more than double the employment provided by ‘Healthcare and social assistance’ professions, the next most important employer in the region. The heterogeneity of the region is important to note, however, as mining is concentrated in the Mount Isa SA2 region (where most of the population is concentrated) but is of negligible importance across the rest of the catchments. Importantly to this Assessment, ‘Agriculture, forestry and fishing’ does not feature within the top five industries for the Southern Gulf catchments. However, the difference across the catchments is evident when the importance of agriculture in different parts of the catchments is considered. Overall, the proportion of employment in the study area provided by agriculture was only 2.7% in 2021, similar to the rate for Queensland (2.6%) and Australia as a whole (2.3%). However, excluding Mount Isa, agriculture provided 17.5% of employment for the rest of the study area and 17.1% in the Carpentaria SA2 region compared with just 0.3% in the Mount Isa region. Over the last three censuses (2021, 2016 and 2011), the percentage of employment from the agricultural sector nationally has been reported as 2.3%, 2.5% and 2.5%, respectively, and for Queensland, 2.6%, 2.8% and 2.7%, respectively, over the same years. That is, the proportion of employment in the agricultural sector has been small and fairly consistent. A broadly similar pattern (fairly consistent and of similar magnitude) is shown within the Southern Gulf catchments overall, with the sector having provided 2.7% of employment in 2021, 2.3% in 2016 and 2.5% in 2011. The structural differences across the Assessment area are notable, as are the far higher dependencies on mining in Mount Isa and on agriculture across the rest of the catchments than are found in Queensland or across Australia. These differences can significantly affect the regional economic benefits that can result from development projects initiated within the region compared to development projects that may be initiated elsewhere. Land use The Southern Gulf catchments cover an area of about 108,200 km2, much of which is used for grazing (77%) (Figure 3-16) based on ACLUMP data current to 2017 for the NT and 2015 for Queensland. Of the remaining area, nearly all is conservation and protected land (16%). A further 4% is classified as water and wetlands, most of which are marine plains and tidal areas located along the coast of the Gulf of Carpentaria. Intensive agriculture and cropping make up a very small portion of the catchment: rainfed and irrigated agriculture and intensive animal production together comprise less than just 0.04% of the area of the catchments. The other intensive localised land uses are transport, communications, services, utilities and urban infrastructure (0.06% of the area of the catchments) and mining (0.06%). While not considered a land use under the land use mapping (because it is a tenure) it is worth noting that Aboriginal freehold title makes up 12% of the Southern Gulf catchments. This land is held under the Commonwealth Aboriginal Land Rights (Northern Territory) Act 1976 (ALRA) in the NT and the Queensland Aboriginal Land Act 1991, including deeds of grant in trust (DOGIT), in Queensland. The title is inalienable freehold, which cannot be sold and is granted to Aboriginal Land Trusts (NT) or trustees (Queensland), which have the power to grant an interest over the land. Land use ACLUMP map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-514_landuse_v4_new_data.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-16 Land use classification for the Southern Gulf catchments Note: land use data shown for the NT on this map is current to 2017 and 2015 for Queensland. Sources: NT Department of Environment, Parks and Water Security (2022); Queensland Government (2021b) Value of agriculture The estimated values of agricultural production for the Southern Gulf catchments and for Queensland as a whole are given in Table 3-8. The Assessment area provides a small proportion of the agricultural production of the state as a whole. The value of production is almost entirely derived from livestock, with a small amount of revenue from crops (predominantly from hay) (Table 3-8). The most recent annual survey data from the ABS describing the value of agriculture by different types of industries (2021–22 survey) are only available at a much larger scale (state and territory level) than the Southern Gulf catchments, making it difficult to accurately estimate the value of agricultural products within the catchments. Hence estimates have been made using the 2020–21 agricultural Census data (Table 3-8), which were published at a finer spatial scale (SA2 level, as used for the socio-economic and demographic catchments estimates). Table 3-8 Value of agricultural production estimated for the Southern Gulf catchments and the value of agricultural production for Queensland for 2020–21 For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the boundaries of the catchments. Source: ABS (2022) Value of agricultural commodities in 2020–21 Beef cattle production Agricultural production in the Southern Gulf catchments is dominated by extensive grazing of beef cattle, valued at $242.7 million in 2020–21 (Table 3-8) and covering about 77% of the Assessment area. Queensland’s beef cattle industry is the largest in Australia, and many Southern Gulf catchment properties turn off young animals (post weaning) to southern fattening properties for domestic and export beef. Sheep were the favoured stock in the earliest days of grazing. Not reaching profitable wool export expectations due to harsh climate, disease (liver fluke, footrot and lung worm), spear grass (Heteropogon spp.) and blowfly incidence, sheep were abandoned in favour of cattle, which have been more successful. The first stations in the Assessment area were formally taken up in the early 1860s. The catchments in this Assessment do not match any stand-alone socio-economic or biogeographical regionalisations used by the cattle industry. However, published information from the ‘Northern Downs’ in Queensland to the south-east of the Southern Gulf (Bowen et al., 2020), the ‘Northern Gulf’ in Queensland directly to the east of the Southern Gulf (Bowen et al., 2019; Rolfe et al., 2016) and the ‘Gulf’ region, encompassing that part of the Assessment area within the NT (Cowley, 2014) have been used below and in Section 4.3.9 to describe the cattle industry within the Southern Gulf catchments. The typical beef production system in the area is a cow-calf operation with sale animals targeted to suit live export, slaughter and the United States grinding beef markets (Bowen et al., 2019; Bowen et al., 2020). A number of properties send cattle to properties further south for backgrounding and fattening. Some of these properties are owned under the one enterprise, with Rolfe et al. (2016) finding a mean of 2.2 properties per business within a survey of 18 properties in the Northern Gulf (Rolfe et al., 2016). Bowen et al. (2020) report that many businesses operate a breeding property in the Northern Gulf region in association with a growing property in the Northern Downs. However, their conclusion was that it was more profitable to turn off live export steers in the Northern Gulf and feed-on steers from the more productive Northern Downs (450 to 480 kg liveweight). 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 January to March period. During the dry season, the total standing biomass and the nutritive value of the vegetation declines. Changes in cattle liveweight closely follow this pattern with higher growth rates in the wet season than the dry season. Indeed, in many cases cattle lose liveweight and body condition throughout the dry season until the next pulse of growth initiated by wet-season rains. 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 mean annual rainfall and its seasonality, and the consequent native vegetation type. Carrying capacities in the Southern Gulf catchments typically range from about 3.5 to 14.2 AE/km2 (i.e. 7 to 28.6 ha/AE) on lands in ‘B’ condition (from a four point condition scale where ‘A’ is highest and ‘D’ is lowest). 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 provide a source of nitrogen for cattle grazing dry-season vegetation. The phosphorus supplements, mostly provided over the wet season, are used because phosphorus is deficient in many areas yet is required for many of the body’s functions, such as building bones, metabolising food and producing milk (Jackson et al., 2012). Winter (1988), working in the Katherine region, found substantial benefits to phosphorus fertilisation and supplementation, particularly in early and late wet-season periods and when grazing pastures which had been oversown with legumes. Cropping Dryland and irrigated agriculture comprise just 0.03% of the Southern Gulf catchments, with a total value of $0.9 million; it all occurs within the Queensland area of the Southern Gulf catchments. Most production is forage sorghum and hay, which are consumed locally. There is at least one small-scale cotton (Gossypium spp.) venture in the catchments. Recent changes to Queensland water regulation and new water releases in the neighbouring catchment of the Flinders River have spurred interest and expertise from growers in southern states to move into northern Australia, focusing on cotton and cereal crops. Horticulture has been proved suitable for the area by Gregory Farm, a 338-ha irrigated development near the Gregory locality where small cropping of mixed vegetables supplied localities in the area in the 1990s. Gregory Farm is currently baling both irrigated forage crops and some seasonal native grasses. Cropping and horticulture have proved to be agronomically suited to the local environment and soils but have been unable to be established as competitive local industries, partly because of difficulties with access to processing, distance to markets and high transport costs. Aquaculture and fisheries There is currently no active aquaculture in the Southern Gulf catchments. The closest aquaculture industry was a small prawn farm established in Karumba (40 km east of the boundary of the catchments) in 1974 and closed the following year for economic reasons (Australian Fisheries, 1975). A comprehensive situational analysis of the aquaculture industry in northern Australia (Cobcroft et al., 2020) identifies key challenges, opportunities and emerging sectors. The Queensland Department of Agriculture and Fisheries supports the aquaculture industry, and research from the Northern Fisheries Centre in Cairns retains aquaculture infrastructure for aquaculture research at the Walkamin Research Facility. Offshore, the Southern Gulf catchments drain into one of the most valuable fisheries in the country. The Northern Prawn Fishery (NPF) spans the northern Australian coast from Cape Londonderry in WA to Cape York in Queensland (Figure 3-17). Most of the catch is 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 from $65 million to $124 million with a mean of $100 million (Steven et al., 2021). The Southern Gulf catchments flow into the Karumba and West Mornington NPF regions (Figure 3-17), the most productive 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 the region of the catchments (Buckworth et al., 2014; Vance et al., 2003). Fishing activity for banana prawns and tiger prawns (Penaeus spp.), which combined 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. 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, it is generally recognised that development of water resources in the Southern Gulf catchments would need to consider the downstream impacts on prawn breeding grounds and the NPF. Commercial barramundi fishing is also found in the study area (Bayliss et al., 2014), but is hard to quantify for the Southern Gulf catchments partly because of the mobility of the industry. Northern prawn fishery regions map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-501_Portrait_map_Australia_NPF_regions_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-17 Regions in the Northern Prawn Fishery and the North West Minerals Province The regions in alphabetical order are Arnhem-Wessels (AW), Coburg-Melville (CM), Fog Bay (FB), Joseph Bonaparte Gulf (JB), Karumba (KA), Mitchell (ML), North Groote (NG), South Groote (SG), Vanderlins (VL), Weipa (WA) and West Mornington (WM). Source: Dambacher et al. (2015) Mining Mining includes extraction of minerals (including coal), petroleum and gas, and quarrying. Mining is well established in Australia, and records indicate that before European settlement Indigenous Peoples quarried different types of stone (particularly gurabaan) and trading materials. In the Southern Gulf catchments, mineral exploration after European settlement dates back to 1882 when Ernest Henry discovered copper at Mount Oxide, north of Mount Isa. Today, mining is the largest industry in the Southern Gulf catchments and in Queensland, was worth $86.5 billion in nominal gross value added (GVA) terms in 2022–23 (Queensland Treasury, 2023). Most of the mining operations in the Southern Gulf catchments involve the extraction of copper, zinc, lead or silver. The North West Minerals Province within the Southern Gulf catchments (Figure 3-17) is considered to be one of the world’s most significant producing areas for base and precious metals (Queensland Department of Regional Development, Manufacturing and Water, 2021). About 75% of Queensland’s base metal mineral endowment is located in this province. Mines currently operating within the Southern Gulf catchments are mainly in the south and the lower rainfall areas, the largest mines occurring in and near Mount Isa (inset map on Figure 3-18). No mines are currently in operation in the NT portion of the Southern Gulf catchments, though as shown in Mineral occurrence and exploration leases, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\CR-S-517-SG_Mineral_Occurences_and_exploration_with_existing_mines_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-18 Main commodity mineral occurrences and exploration tenements in the Southern Gulf catchments Sources: NT Geological Survey (2024); NT Government (2024b); Geological Survey of Queensland (2024); Queensland Government Open Data Portal (2024b) Mining is the largest employer in the Southern Gulf catchments with most of the labour force based in Mount Isa (Table 3-7). In 2021, the population of the Mount Isa urban area was just over 18,000, and approximately 30% of the labour force was employed directly in mining in a range of occupations, including management, administration, professional and technical roles, and machinery operations (ABS, 2021c). Table 3-9 provides employee numbers for mines of various sizes. Table 3-9 Indicative numbers for employment in different types of mining operations in the Southern Gulf catchments MINE TYPE NUMBER OF EMPLOYEES AND CONTRACTORS SOURCE Very large base metal mine and associated processing and refining plants 1000–2000 Glencore (2023a) Mid-sized to large base metal mine 300–600 Evolution Mining (2023), MMG (2024) Mid-sized phosphate mine 200–300 Incitec Pivot (n.d.), Andre and Waterson (2023) Mount Isa region is currently in a period of transition. In 2005 the underground lead mine at Mount Isa closed due to a lack of ore (ABC, 2005), and the Mount Isa Mines underground copper operations are scheduled to cease in 2025 as the remaining mineral resources are deemed economically unviable. These include the underground copper mines Enterprise, X41 and Black Rock. The Lady Loretta zinc mine, a fly-in, fly-out operation 140 km north-west of Mount Isa, is also slated to close in 2025 (Glencore, 2023a). Despite these closures, the global consumptions of copper, zinc and nickel are projected to increase from about 24, 14 and 2 million tonnes in 2019, respectively, to 31, 15 and 4 million tonnes in 2030. Technological advances such as magnets and super magnets, motors, metal alloys, electronic and computing equipment, batteries, catalytic converters, petroleum refining and medical imaging have increased demand for rare earth elements (REEs), a specific group of 17 metals within the critical minerals family. In recognition of the importance of critical and strategic minerals to Australia’s modern technologies, economy and national security, the Australian Government released updated lists of Australia’s critical minerals and strategic materials, and both the NT and Queensland governments have programs to attract investment in critical mineral exploration and infrastructure. Critical minerals and strategic materials currently mined and/or the target of recent and current exploration programs in the Southern Gulf catchments include copper, zinc, phosphorus, the REEs and graphite. See the companion technical report on agricultural viability and socio-economics (Webster et al., 2024a) for a full list of critical minerals and strategic materials. Mineral occurrences for a wide range of commodities have been identified in the Southern Gulf catchments. As shown in Figure 3-18, approximately 68% of the Southern Gulf catchments is covered by mineral or petroleum exploration licences; the highest proportion occur in the Leichhardt catchment, in which 79% is covered by mineral exploration licences. Phosphate occurrences are present across the Southern Gulf catchments with a concentration around Lawn Hill and the Gunpowder localities (Figure 3-18). Exploration for uranium is currently being conducted in the catchments, and occurrences are in the south and north-west (Figure 3-18). There are no operating uranium mines. In Queensland it is government policy not to grant mining leases for uranium, although applications may still be made for mineral development licences or exploration permits for uranium (Queensland Department of Resources, 2021). Uranium mining is permitted in the NT (NT Government, 2024a). Several petroleum exploration bores have been drilled within the catchments. Twenty petroleum exploration bores were drilled on Mornington Island between 1959 and 1961; no hydrocarbons were reported, and the holes were plugged and abandoned. Thirteen wells were drilled in the Queensland part of the study area between 1959 and 2013. Gas was reported in three of the wells. Eleven of the holes are reported to have been plugged and abandoned or suspended and capped; two wells are reported to be current water bores (Queensland Government Open Data Portal, 2024a). One exploration petroleum well was drilled in 1992 in the NT part of the study area close to the western margin of the catchments and is reported to be dry, plugged and abandoned (NT Government, 2024b). Table 3-10 presents resource and reserve data for several major deposits in and on the margins of the Southern Gulf catchments. Table 3-10 Resource and reserve data for several major deposits in and on the margins of the Southern Gulf catchments MINE MASS (million t) MINERAL % MINERAL % MINERAL g/t LIFE OF MINE SOURCE Zinc Lead Silver George Fisher North Mine† (underground) To 2036 Glencore, 2023b, c Measured and indicated resources 2023 164 8.92 3.34 54 Total ore reserves 2023 45 6.84 3.31 54 George Fisher South Mine† (underground) To 2036 Glencore, 2023b, c Measured and indicated resources 2023 55 8.32 5.02 110 Total ore reserves 2023 12.9 6.11 4.76 110 Lady Loretta† (underground) To 2025 Glencore, 2023b, c Measured and indicated resources 2023 5.1 11.29 2.48 54 Total ore reserves 2023 3.5 10.19 2.22 43 Dugald River†* (underground) ≥20 years MMG, 2023, 2024 Measured, indicated and inferred resources 2023 57 11.7 1.6 23 Proved and probable reserves 2023 20 10.8 1.7 40 New Century‡ (tailings and exploration) To 2027 Sibanye Stillwater, 2023 Measured and indicated resources 2022 2.0 5.6 na†† na Proved and probable reserves 2022 6.8 3.0 na na MINE MASS (million t) MINERAL % MINERAL % MINERAL g/t LIFE OF MINE SOURCE Copper Silver Mount Isa Copper† (open-cut and underground) To 2025 Glencore, 2023b, c Measured and indicated Resources 2023 156 1.70 na Total ore reserves 2023 6.5 1.95 na Capricorn Copper† (underground) ≥2034 29Metals, 2022, 2024 Measured, indicated and inferred resources 2023 64.8 1.8 9 Proved and probable reserves 2023 19 1.7 12 Phosphate Paradise South Phosphate§ (open-cut) ≥20 years Andre and Waterson, 2023 Measured, indicated and inferred resources (estimated total deposit) 436.5 9.4 Proven and probable reserves (estimated total deposit) 198.5 12.7 †Resources and reserves reported in accordance with the 2012 edition of the Australasian code for reporting of exploration results, mineral resources and ore reserves (Glencore, 2023b; Joint Ore Reserves Committee, 2012). ‡Resources and reserves reported in accordance with the 2016 edition of The South African code for the reporting of exploration results, mineral resources and mineral reserves (SAMREC, 2016) and subpart 1300 under Regulation S-K of the US Securities Act of 1933 (Sibanye Stillwater, 2023). §Resources and reserves reported in accordance with the 2004 edition of the JORC Code (Joint Ore Reserves Committee, 2004). *The Dugald River mining operations occur outside the Southern Gulf catchments boundary although the tenements associated with the Dugald River project partially occur in the Southern Gulf catchments and the mine water supply is drawn from a purpose-built dam located inside the Leichhardt River catchment with backup water supplied from Lake Julius. ††na = not applicable. Water use in mining and petroleum industries in the Southern Gulf catchments Water is central to the minerals and petroleum industries. Mining uses water in a variety of ways, including for transporting materials, chemical or physical processing, cooling, disposing and storing waste materials, washing, and suppressing dust. Potable water is used for areas that house mining staff (Prosser et al., 2011). Water is also extracted or ‘used’ during de-watering at mines that extend below the water level, such as Century Zinc Mine. Petroleum companies, which use relatively small volumes of water, produce water as a by-product of extraction. Water extracted during de-watering or as a by-product of petroleum extraction must be safely discharged and may need treatment. Water consumption at mining operations is highly variable (Table 3-11). The variations are due to a range of factors, including different mining methods, ore types, ore grades, processing treatments and definitions of water usage. The overall water balance on a site depends on climate conditions, which affect water availability at the site, and the ability to reuse and recycle water within process facilities (Northey and Haque, 2013). While not mined in the Southern Gulf catchments, coal is by far the largest user of water in the mining sector. The water used by mining enterprises does not need to be of potable quality. Table 3-11 Global water consumption in the mining and refining of selected metals PROCESSING STAGE MEAN WATER CONSUMPTION* (m3/tonne of metal) RANGE OF WATER CONSUMPTION§ (m3/tonne of metal) Copper concentrate† 43.235 9.673–99.550 Gold metal‡ 265,861 79,949–477,000 Lead concentrate† 6.597 0.528–11.754 Manganese concentrate† 1.404 1.390–1.410 Palladium metal‡ 210,713 56,779–327,874 Platinum metal‡ 313,496 169,968–487,876 Uranium concentrate (U3O8)† 2,746 46.2–8,207 Zinc concentrate† 11.93 11.07–24.65 †Metal concentrates are typically produced at the site where the ore is mined. ‡Includes mining, smelting and refining of pure metals, assuming mining and processing are all located within a single region or separate regions but with similar water characteristics. *Mean water consumption values per /tonne of metal equivalent in the concentrates or refined metals. §Minimum and maximum water consumption values per tonne of metal equivalent in the concentrates or refined metals. Source: Meissner (2021) Because water is typically a very small fraction of total input cost, and mining produces high-value products, mining enterprises usually develop their own water supplies, which are often regulated separately to the water entitlement system (Prosser et al., 2011). In the Southern Gulf catchments, however, the concentration of mining and industrial activity resulted in sufficiently high water demand for the construction of large purpose-built reservoirs, including the privately funded Leichhardt Dam (Lake Moondarra) and Julius Dam (Lake Julius, now owned by SunWater). Dams in the Southern Gulf catchments are summarised in Chapter 5. Data on water use by mining in the Southern Gulf catchments are difficult to obtain. As described in Section 3.3.4, water use from Lake Julius between 2017–18 and 2012–22 was estimated to range from 5 to 14 GL/year, and from Lake Moondarra from 14 to 17 GL/year. Because these quantities represent a relatively modest proportion of the total supplemented water entitlements, there is scope for existing water storages, including Lake Mary Kathleen (12 GL capacity), which is only used for recreation, to support the expansion of mining activity in the Mount Isa region. Tourism Tourism in the Southern Gulf catchments is in a moderate state of development. Self-drive tourists are the predominant visitor market type in the Southern Gulf catchments, and they represent 87% of visitors to outback Queensland (Outback Queensland Tourism Association, 2021). A strategically promoted transcontinental ‘adventure’ self-drive traveller route, the Savannah Way, connects Cairns (Queensland) to Katherine (NT) then continues to Broome (WA), traversing the Southern Gulf catchments via Burketown (878 km from Cairns), Doomadgee and Hells Gate Roadhouse. Access to much of the Southern Gulf catchments is on unsealed roads. The largest domestic airport is at Mount Isa, which in offering a gateway to the region provides an advantage over many other parts of northern Australia. Smaller commercial regional airports are located in Burketown, Doomadgee and Gununa on Mornington Island, providing regional connections between Mount Isa and Cairns. All airports cater for light aircraft, and the nearest international airports are in Cairns and Darwin. Multiple accommodation options and types are available within and surrounding Burketown and Mount Isa; wet-season closures apply to some camping and accommodation businesses. Key inland attractions for visitors to the Southern Gulf catchments include the iconic Boodjamulla National Park (Aboriginal Land) and Wugudaji-Adels Grove. The Riversleigh World Heritage Area is renowned for its fossil fields containing Gondwanan specimens dating from 25 million years ago (Ma) (Queensland Department of Environment, Science and Innovation, 2024b). Visitor attractions and activities throughout the catchments include a range of nature-based activities such as bird and wildlife watching. Popular coastal attractions and activities include river cruises, four-wheel- drive cultural tours, hot-air ballooning, fishing charters and astronomy (with one Indigenous- owned tour operator from Burketown providing stargazing tours). The annual World Barramundi Fishing Championships in Burketown, the Gregory River Canoe Marathon and Burketown’s Morning Glory Festival are examples of annual events promoted to attract visitors to the region. As well as economic and employment opportunities, tourism can cause impacts such as native habitat loss, and foot traffic, bikes or vehicles may cause environmental damage such as erosion and a loss of amenity to local residents (Larson and Herr, 2008). Other risks include the spread of weeds (Chapter 7) and root rot fungus (Phytopthora cinnamomi) carried on vehicles and people (Pickering and Hill, 2007). The majority (79%) of the Southern Gulf catchments area is within Queensland, and this area is the most accessible to tourists. Visitor statistics in Queensland are usually reported at the local government area (LGA) scale (Figure 3-19). In the Southern Gulf catchments, visitor statistics are dominated by the Mount Isa LGA. The Mount Isa LGA reported 192 tourism businesses, of which 76 were ‘non-employing’, 61 had four or fewer employees, 46 had between 5 and 19 employees, and 15 had 20 or more employees. Annual visitation to the Mount Isa LGA included 154,000 domestic visitors (100,000 of which were intrastate) and 10,000 international visitors. Visitors stayed for a total of 660,000 visitor nights (a mean stay of four nights for domestic visitors and 11 nights for international visitors), contributing more than $132 million in regional expenditure. However, a large proportion of the visitors to Mount Isa LGA (approximately 95,000) had travelled there for business purposes (Tourism Research Australia, 2019). A 2019 survey of Burke Shire LGA reported 11 tourism businesses, of which six were ‘non-employing’, and domestic overnight visitor expenditure in the shire was estimated to be $9 million/year. Other LGAs in the study area reported few if any tourism enterprises, while others such as the Cloncurry LGA are not considered representative because only small parts of the LGA are within the study area. High summer temperatures and humidity result in most tourist visitation occurring during the drier, cooler months with 80% of visitation to the Gulf Country falling between April and October. Road closures and seasonal business closures have further impacts on the ability of the region to have visitors year-round (Tourism Tropical North Queensland, 2024). This seasonality of visitation has flow-on effects for tourism-connected industries such as accommodation, food services and transport. Local governement area and Tourism regiont map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-518_LGA_tourism regions_v03.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-19 Local government areas and the Tropical North Queensland tourism region that statistics on tourism visitation are extracted from Tourism development opportunities and considerations The state of northern Australia’s tourism economy is closely tied to the state of its ecosystems (Prideaux, 2013). With a large proportion of the Southern Gulf catchments in a ‘natural’ state relative to many parts of south-eastern Australia, there is potential for growth in nature-based tourism. However, like other remote areas of northern Australia, the region’s remoteness and distance from urban centres (Bugno and Polonsky, 2024), lack of supporting infrastructure, limited human capital and financial resources, and low awareness of tourism system characteristics (Summers et al., 2019), considerably constrain its potential. The seasonality of visitation also limits enterprise profitability (Bugno and Polonsky, 2024) and permanent employment opportunities. Also important to consider is that much of the catchment’s appeal to self-drive visitors is likely to be the absence of other people and commercial infrastructure, which presents opportunities for exploration and solitude (Lane and Waitt, 2007; Ooi and Laing, 2010). Hence development that alters the region’s current characteristics could be alienating to some current visitor markets. While water resource development for agriculture has the potential to negatively affect tourism and future opportunities in the Southern Gulf catchments, for example, through declining biodiversity and perceived reduced attractiveness (Pickering and Hill, 2007; Prideaux, 2013), such development may present opportunities to foster tourism growth. For example, Lake Argyle in the East Kimberley region (WA), developed as an irrigation dam to supply the Ord River Irrigation Area, is among northern WA’s must-see attractions, offering a wide range of tourism activities (https://www.australiasnorthwest.com/explore/kimberley/lake-argyle). While visitors to the Kimberley region reportedly perceived Lake Argyle in the same way they perceived some ‘natural’ local attractions such as billabongs, irrigated agriculture of the Ord River Irrigation Area is perceived differently, as being ‘domesticated’ (Waitt et al., 2003). Elsewhere in northern Australia, water resource infrastructure, including Fogg Dam (NT), Tinaroo Dam (Queensland) and Lake Moondarra (Queensland), has resulted in increased visitation by tourists for the enhanced wildlife or recreation opportunities they provide (e.g. Regional Development Australia, n.d.). However, the ongoing contributions of dam to their local economies vary. For example, the value of recreational fishing has been found to vary between dams depending upon whether there are other dams nearby and their proximity to tourism traffic (Rolfe and Prayaga, 2007). Visitation numbers to the Southern Gulf catchments suggests that the recreational fishing value of a new dam in the Southern Gulf catchments would be limited. Agritourism opportunities, for example, through property accommodation and other travel support (fuel), offer an opportunity for revenue diversification, although impediments such as highly variable seasonal demand limit profitability (Bugno and Polonsky, 2024). Tourism has the potential to enable economic development within Indigenous communities because Indigenous tourism enterprises, usually microbusinesses, often have some competitive advantages (Fuller et al., 2005). Successful tourism developments in regional and very remote areas such as the Southern Gulf catchments are highly likely to depend on establishing private and public sector partnerships, ensuring effective engagement and careful planning with Traditional Owners and regional stakeholders, and building interregional network connectivity and support (Greiner, 2010; Lundberg and Fredman, 2012). Given the importance of climate on tourism seasonality, demand and travel patterns in northern Australia (Hadwen et al., 2011; Kulendran and Dwyer, 2010), the increased temperatures and occurrence of extreme weather-related events (e.g. drought, flood, severe fires and cyclones) associated with climate change are likely to be significant threats to the industry in the future. These are likely to negatively affect tourist numbers, the also length and quality of the tourist season, tourism infrastructure including roads, and the appeal of the landscape and its changing biodiversity (Amelung and Nicholls, 2014; Prideaux, 2013). 3.3.4 Current infrastructure Transport A modest road network services the Southern Gulf catchments, from sealed major highways to unsealed minor connection roads, all of which are subject to flooding and wet-season closures. The roads in the catchments have previously benefited from the Northern Australia Beef Roads Program, which has funded upgrades to key roads necessary for transporting cattle to improve the reliability and resilience of cattle supply chains in northern Australia, reducing freight costs and strengthening links to markets. The Barkly Highway (Figure 3-20) is the only sealed road between the NT and Queensland. It runs through the most southern part of the Leichhardt catchment at Mount Isa from the Three Ways Roadhouse just north of Tennant Creek, NT, east to Cloncurry, Queensland. This highway continues as the Flinders Highway to Townsville on the east coast, a key route in the national supply chain network supporting many industries, including agriculture and mining. All road network information in this section is from spatial data layers in the Transport Network Strategic Investment Tool (TraNSIT; Higgins et al., 2015). Trucking volumes calculated from TraNSIT show the largest volume of trailers occur on the Wills Developmental and Burke Developmental roads. Both of these roads are sealed and are predominantly used to support the cattle industry (Figure 3-26). National Highway 1 runs east–west across the Southern Gulf catchments through Burketown. It has sealed and unsealed sections and is also known as the Savannah Way, a popular four-wheel drive tourist route. Sections of this road also have local names (e.g. Doomadgee–Westmoreland Road). Rankings of the road network in the Southern Gulf catchments are shown in Figure 3-20. Heavy vehicle access restrictions for roads, as determined by the National Heavy Vehicle Regulator, show good connectivity for Type 2 road trains (Figure 3-21). These are vehicles up to 53 m in length, typically a prime mover pulling three 40-foot (approximately 12 m) trailers (Figure 3-22). A large proportion of the classified non-residential roads in the study area permit Type 2 road trains despite the poor 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 regularly use smaller vehicle configurations on the minor roads due to the difficult terrain and single lane access, particularly during wet conditions. Figure 3-24 shows the mean speed achieved for freight vehicles for the road network. The road speed limits are usually higher than the mean speed achieved for freight vehicles, particularly on unsealed roads. Heavy vehicles using such unsealed roads would usually achieve mean speeds of no more than 60 km/hour, and often lower when transporting livestock. Rail access in the Southern Gulf catchments is from a single point, Mount Isa, and has both a freight and passenger (the Inlander) service. The Great Northern Railway (also known as the Mount Isa line) is a critical link from the North West Minerals Province at Mount Isa to the Port of Townsville, primarily carrying bulk commodity transport (mostly minerals) for export. Several train operators manage rolling stock of incoming and outgoing freight. Queensland Rail (a statutory authority of the Queensland Government) owns the narrow-gauge (3 feet 6 inches, 1067 mm) line. The narrow gauge was chosen purely for economic reasons: because Queensland distances are large, the narrower gauge was cheaper to construct. Road rankings map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-508_TraNSIT_road rankings_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-20 Road rankings and conditions in the vicinity of the Southern Gulf catchments 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 Barkly Highway, which runs from the Three Ways Roadhouse (north of Tennant Creek) in the NT to Cloncurry, Queensland, east of Mount Isa. Road truck class map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-509_TraNSIT_truck type_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-21 Roads accessible to Type 2 vehicles in the vicinity of the Southern Gulf catchments: minor roads are not classified Type 2 vehicles are illustrated in Figure 3-22. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 3-22 Common configurations of heavy freight vehicles used for transporting agricultural goods in Australia Figure 3-23 Road conditions and distance to market impact the economics of development in the Southern Gulf catchments. Photo: CSIRO – Nathan Dyer Road speed map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-510_TraNSIT_road_speed_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-24 Mean speed achieved for freight vehicles on roads in the vicinity of the Southern Gulf catchments Data source: Spatial dataset of the location and attributes of roads and ferries sourced from HERE Technologies (2021). Supply chains and processing Table 3-12 provides volumes of commodities transported into and out of the Southern Gulf catchments annually by road, and Figure 3-26 shows the location of existing pastoral properties in the catchments. Agricultural production is dominated by beef cattle, which is reflected in the annual freight volumes moving across the road network in the Southern Gulf catchments according to TraNSIT records of truck movements. Live export of cattle accounts for the majority of cattle movements. There are also substantial transfers of cattle between properties and smaller volumes directed to domestic markets via abattoirs and feedlots in southern Queensland; however, the closest abattoir is Townsville. Table 3-12 Overview of commodities (excluding livestock) annually transported into and out of the Southern Gulf catchments Indicative transport costs are means for each commodity and include differences in distances between source and destinations. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Source: 2023 data from TraNSIT (Higgins et al., 2015) NA - data not available There are currently no processing facilities for other agricultural produce within the Southern Gulf catchments, although there have been assessments made for cotton processing (MITEZ, 2021). Rail access links Mount Isa to Townsville and the port, primarily carrying bulk commodity transport (mostly minerals) for export. The Port of Townsville is the closest port for bulk export and operates within the Great Barrier Reef World Heritage Area. This port is northern Australia’s largest container import facility, and exports primarily service the Queensland agricultural and mineral provinces. Port services also support imports and exports of general cargo for a range of domestic and industrial commodities (e.g. fuels, food, vehicles, commercial machinery and manufactured items). The port is managed and operated by a government-owned corporation, the Port of Townsville Limited, and currently operates with 11 berths. The Port of Lucinda (inset map on Figure 3-26), located on the east coast 100 km north of Townsville, is a sugar export facility. Ports North, also a corporation owned by the Queensland Government, manages eight ports in far north Queensland, four with sealed road access from the Southern Gulf catchments. The Port of Burketown, inside the catchments, is a declared port, but no commercial trade takes place. To the east just outside the catchments, the Port of Karumba provides for general cargo, fuel, fisheries products and zinc transhipment and has previously seen export of live cattle. On the east coast, the Port of Cairns, a regional port, services bulk and general cargo, a fishing fleet, cruise liners and passenger ferries to the reef. The export of raw sugar and molasses from the local sugar-growing districts is through the Port of Mourilyan. The only port in the western Gulf of Carpentaria is the Port of Bing Bong (inset map on Figure 3-25 Many roads are gravel in the Southern Gulf catchments, and often impassable in the wet season Photo: CSIRO – Nathan Dyer Truck volume and ports map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-511_TraNSIT_ag enterprises_v5.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-26 Annual amounts of trucking in the Southern Gulf catchments and the locations of pastoral properties and ports Declared ports (shown on the inset map) all carry out commercial operations except the Port of Burketown. The thickness of purple lines indicates the volume of traffic (as number of trailers per year) on regional roads connecting properties and enterprises. Energy The only major electricity network in north-western Queensland is the North West Power System (NWPS), which connects Mount Isa to the Century Mine (zinc) near Lawn Hill in the north and Cloncurry to the east outside the Southern Gulf catchments (Figure 3-27). The NWPS is isolated from the National Electricity Market (NEM), Australia’s largest electricity network, which stretches along Australia’s east coast from northern Queensland to Tasmania and SA. Power generation and transmission map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-507_energy generation distribution_v4.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-27 Electricity generation and transmission network and pipelines in the Southern Gulf catchments Most power stations in the NWPS are gas-fired, which has resulted in high electricity costs compared to the NEM (Queensland Government, 2021a). The NWPS does not operate via an electricity market, but rather has negotiated supply contracts, which suits large industrial and mining operations in the region (APA, 2022). Ergon Energy operates the distribution network in this region and all country areas of Queensland (Figure 3-27). Powerlink operates the NEM transmission network in Queensland (AEMO, 2023). Already one of the world’s longest interconnected power systems, the Queensland Government plans to connect the NEM to the NWPS with the development of the CopperString 2032 project. The connection will consist of a 1100 km transmission line from Townsville in the NEM to Mount Isa in the NWPS. Little benefit will be realised within the Southern Gulf catchments outside the Mount Isa area without additional transmission line infrastructure. The Diamantina Power Station, near Mount Isa, is a combined-cycle gas turbine plant with a capacity of 242 MW. This power station supplies the NWPS. It also has 60 MW of backup from the nearby open-cycle gas turbine Leichhardt Power Station. Doomadgee, Burketown and Gununa are not connected to any grids and have their own community power stations, which are supplied and maintained by Ergon Energy. Burketown and Gununa are isolated diesel power stations and Doomadgee’s 568 kW ground-mounted solar farm and 105 kW of rooftop solar (on four Doomadgee Shire Council buildings) provide power with backup supplied by diesel generator. Gas pipelines are located in the most southerly part of the Southern Gulf catchments where the Northern Gas Pipeline stretches from the NT to Mount Isa, and then the Carpentaria Gas Pipeline connects Mount Isa to Ballera in southern Queensland (Australian Energy Regulator, 2021). Renewable energy potential in the Southern Gulf catchments The Southern Gulf catchments have some of the best solar resources in Australia and a low to modest wind resource relative to other locations in Australia. A convenient metric for comparing renewable energy technologies is using the capacity factor of an energy plant, which is the ratio of electricity generated over one year to the nameplate capacity of the solar or wind farm. For example, for a capacity factor of 0.25, each 1 MW of a solar/wind farm will generate about 2190 MWh of electricity per year. In the Southern Gulf catchments, solar photovoltaic capacity factors are uniformly high, ranging between 0.23 to 0.25. In contrast, in southern Australia and along the east coast, the capacity factor can be as low as 0.12 (Figure 3-28). Wind resources for the Southern Gulf catchments are shown in Figure 3-29 as a capacity factor at a turbine hub height of 150 m, which is a typical height for a commercial wind turbine. Although wind capacity factors in the Southern Gulf catchments are comparable to, and in some cases higher than, solar capacity factors, wind farms have a higher capital cost, which can result in a higher cost of electricity production. This is particularly the case for smaller wind turbines than those whose results are shown in Figure 3-29. The generation capacity of these smaller turbines is more likely to be commensurate with the energy requirements of a farm-scale irrigation enterprise. Furthermore solar is modular and scalable and is easier to maintain in remote locations than wind turbines. Wind energy is also a relatively mature technology, and projections of the levelised cost of wind in 2040 suggest that its cost is plateauing. In contrast, solar photovoltaic is projected to steadily decrease such that by 2040 the levelised cost of solar photovoltaic would be 26% to 34% lower than the cost of wind on average (Graham et al., 2023). With the exception of Kajabbi, one of the few locations with soils potentially suitable for irrigated agriculture and a grid connection, solar photovoltaic is the most viable renewable technology across most of the Southern Gulf catchments. Having a grid connection at Kajabbi enables excess renewable electricity to be sold to the grid, potentially changing the economic viability of wind turbines. Solar photovoltaic capacity map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-517_Solar_resource_V3.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-28 Solar photovoltaic capacity factors in the Southern Gulf catchments Inset shows solar photovoltaic capacity factors across Australia. Note: the inset map uses a different colour ramp Wind capacity map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-516_Wind_resource_V4.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-29 Wind capacity factors in the Southern Gulf catchments Inset shows wind capacity factors across Australia. Note: the inset map uses a different colour ramp. At locations distant from a grid connection, such as Doomadgee, it was found that, based on current capital costs and a diesel cost of $1.50/litre (including any rebate), diesel generators were the most cost-effective technology for supplying power to farm infrastructure requiring electricity 24 hours/day or requiring electricity for 30% or fewer days per year. For farm infrastructure operating more than 50% days of the year, and for 12 hours/day or less, a hybrid diesel – solar photovoltaic farm with the renewable fraction between 50% and 75% is most the cost-effective technology. The exception is for farm infrastructure requiring electricity for 4 and 12 hours/day and 365 days/year, for which a 100% solar photovoltaic farm (with batteries) was most the most cost-effective way to provide power. Under a higher cost of diesel ($2.50/litre including rebates), the results were similar except a 100% renewable system with batteries was most cost-effective when electricity had to be supplied for 80% of days or more. Based on solar photovoltaic and battery technology costs projected to 2040, it was found that hybrid diesel – solar photovoltaic systems (with batteries) were most cost-effective when farm infrastructure was operated for 30% of days/year or higher for 12 to 24 hours/day, or 10% of days/year when only operated for a maximum of 8 hours/day at the higher diesel price (i.e. $2.50/litre). See the companion technical report on techno-economic analysis of electricity supply (Hayward, 2024) for more detail. Hybrid grid electricity – renewable systems without batteries are more viable in all instances where soils suitable for irrigated agriculture are located near an existing electricity transmission line and it is possible to sell modest amounts of surplus energy to the grid (i.e. assuming the size of the renewable farm cannot be greater than 25% of the electrical load and assuming expected revenue was only equal to the levelised cost of electricity of wind; see Hayward (2024) for more detail). Solar technology tends to be more viable when the demand is 4 hours/day or less, or the infrastructure is operated is for 10% or fewer days per year. Water Communities and industries in the Southern Gulf catchments source their water from either surface water or groundwater for a variety of purposes, including: (i) stock and domestic uses, (ii) town and community water supplies, and (iii) industries such as agriculture and mining. For some uses, surface water is pumped from the occasional dam or stream. In the case of Mount Isa’s water supply, a major water transmission pipeline supported by pumping stations is used to transfer water from the dams on Lake Julius and Lake Moondarra for treatment prior to distribution in Mount Isa. Small quantities of groundwater (<5 ML/year) may be pumped from a single bore for stock and domestic use. Where larger amounts of groundwater are required (tens to hundreds of megalitres per year), water may be pumped from a borefield consisting of multiple connected production bores. Such applications include town and community water supplies and irrigated agriculture. A water licence is required for some water uses, such as water use for town and community water supply or applications by industry. Other applications, such as stock and domestic use, may or may not require a water licence. This will depend upon the proposed location and magnitude of water take and whether it occurs within a water plan area or could interfere with a watercourse, lake or spring (Queensland Government, 2016). In some areas of the catchments, water use is under a water plan – each plan covers a different extent and plan areas may overlay another plan area (Figure 3-30). Surface water plans will overlay the Water Plan (Gulf) 2007 area (Figure 3-30). Groundwater plans may overlay the Great Artesian Basin and Other Regional Aquifers (GABORA) Plan area. The GABORA Water Plan manages groundwater sources from multiple aquifers hosted in different geological units and within different groundwater sub-areas (Figure 3-30). In 2018, the Queensland Water Act 2000 was changed to formally recognise the importance of water resources to Aboriginal and Torres Strait Islander Peoples. It required new and replacement water plans to explicitly state ‘cultural outcomes’ as distinct from social, economic and environmental outcomes. Water plans can include strategies for monitoring and reporting on the achievement of cultural outcomes. For example, the Water Plan (Gulf) 2007 (Queensland Government, 2007) includes 30,550 ML as an Indigenous reserve ( Surface water and groundwater licences, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-519_WaterLicences_v07.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-30 Location, type and volume of annual licensed surface water and groundwater entitlements across the Southern Gulf catchments License locations are not precise as deidentified location information was used. Unallocated reserves provided for in the Water Plan (Gulf) 2007 (34.3 GL/y in total), are not shown. Currently there are no active groundwater or surface water licences in the NT portion of the Southern Gulf catchments. Data sources: Spatial dataset of the location and attributes for water licences and permits across Queensland sourced from the Queensland Department of Regional Development, Manufacturing and Water (2023) Table 3-13 Unallocated surface water in the Queensland part of the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Source: Waschka and Macintosh (2023) Surface water entitlements Surface water licences with a volumetric entitlement occur at a variety of locations and from a variety of sources across the Southern Gulf catchments (Figure 3-30). Currently, 27 unsupplemented surface water licences with a volumetric entitlement have been granted across the Southern Gulf catchments. These licenses have been granted for a combination of uses, including agriculture and aquaculture, across various parts of the catchments. They have a combined total of about 38,000 ML/year (38 GL/year) (Figure 3-30). The largest entitlements (i.e. between 1000 and 8000 ML/year) have been granted for use in agriculture. Some moderate entitlements (between 400 and 1000 ML/year) have been granted for town and community water supply at Mount Isa, Gregory and Kajabbi (Figure 3-30). Much smaller surface water entitlements (<50 ML/year), are associated with stock use (Figure 3-30). There are also supplemented water entitlements supplied by Lake Julius (48.85 GL) and Lake Moondarra (26.3 GL). These entitlements are used for Mount Isa town water supply and industrial use, as well supply to Cloncurry and Ernest Henry Mine via the North West Queensland Water Pipeline from Lake Julius. Use of these entitlements historically has been low, over the 2017–18 to 2021–22 water years (1 September to 31 August), 11% to 29% of authorised entitlements from Lake Julius were used, and 52% to 66% from Lake Moondarra (Queensland Department of Regional Development, Manufacturing and Water, 2023). Groundwater entitlements Groundwater licences with a volumetric entitlement also occur at a variety of locations and from a variety of sources for different uses across the catchments. Currently 13 groundwater licences with a volumetric entitlement have been granted for a variety of applications with a combined total of about 3.5 GL/year. The largest entitlements (150 to 1400 ML/year) are associated with industrial use in mining with the water sourced from various aquifers hosted in the Paradise Creek Formation and the Currant Bush and Thorntonia limestones (Figure 3-30). Two licensed entitlements of approximately 100 ML/year have been granted for town and community water supplies at Burketown and Gununa (Mornington Island). Both licences have been granted for groundwater sources from the Gilbert River Formation of the Great Artesian Basin (Figure 3-30). The smallest groundwater licences (<100 ML/year) have been granted for a variety of industrial and agricultural uses with groundwater sourced from a variety of aquifers hosted in different geological units (Figure 3-30). The Century Zinc Mine, located about 15 km to the south-east of Lawn Hill, was Australia’s largest open-pit zinc mine before its closure in 2016. It used to de-water part of the Cambrian Limestone Aquifer (CLA) hosted in the Thorntonia Limestone that overlies zinc deposits hosted in the Proterozoic Lawn Hill Formation. When fully operational, the mine was reported to be extracting about 19 GL/year of groundwater in the early 2000s and about 10 GL/year in the mid-2010s. Currently, less than 1 GL/year of groundwater is being extracted at the site. The cessation in de-watering at the site is likely to have resulted in recovery of groundwater levels and storage in the CLA around the site. This may also include an onset of increased discharge from the aquifer to Lawn Hill Creek. However, the timescales for changes in groundwater flow are likely to be in the order of tens of years or longer, and further investigation would be required to confirm this. For more information on groundwater resources of the Southern Gulf catchments, see the companion technical report on groundwater characterisation (Raiber et al., 2024). Community infrastructure The availability of community services and facilities in remote areas can play an important role in attracting people to or deterring people from living in those areas. Development of remote areas, therefore, 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. Like most remote parts of Australia, there are limited primary health resources in the Southern Gulf catchments apart from in the largest population centre, Mount Isa. Mount Isa Hospital has 80 beds and is a Queensland Level 4 (Clinical Services Capability Framework) Specialist Service Base Hospital delivering moderately complex services. Telehealth and specialist outreach services are provided from Mount Isa to remote hospital and health service facilities in the Assessment area. General practitioners and allied health professionals provide most primary healthcare in Mount Isa. The Southern Gulf catchments are serviced by the national primary health network (PHN). Australia is divided into 31 PHNs: one of these covers the whole of the NT, and in Queensland, the area is covered by the Western Queensland PHN. These PHN regions are divided into districts. In the NT, the Katherine Health Service District (HSD) (also known as the Big Rivers Region) and the Barkly HSD provide health services to remote communities and properties in the NT part of the Southern Gulf catchments. There are no hospitals inside the NT section of the Southern Gulf catchments. The nearest healthcare resource is Robinson River Community Health Centre, a nurse- led clinic approximately 100 km north of the western boundary of the catchments. The Queensland North West Hospital and Health Service at Mount Isa works closely with local hospital and clinic networks in smaller communities to provide remote health services. There are three hospitals inside the boundaries of the catchments: Burketown (Level 1), Doomadgee (Level 2) and Mornington Island (Level 2). They provide low-complexity care services and specialist visiting services, including paediatrics, dietetics, oral health and speech therapy. Gidgee Healing Aboriginal Medical Service provides primary and community healthcare in Doomadgee and on Mornington Island. Three Queensland hospitals located close to but outside the boundaries of the catchments provide health services: Camooweal (Level 1), Cloncurry (Level 1) and Normanton (Level 2). The Royal Flying Doctor Service also covers the region, providing weekly general practitioner and fortnightly child health clinics to some communities and properties. The mining town of Mount Isa has all the facilities of a large rural town. It is the central hub for education in the Southern Gulf catchments, with three public schools totalling 1240 full-time equivalent (FTE) enrolled students and 129.4 teachers (FTE) in 2022 (Table 3-14). Delivering education to the smaller communities are three schools, Doomadgee, Burketown and Mornington Island, which are all in the Queensland part of the catchments. A total of 577.2 (FTE) students are enrolled in these schools with 60.8 (FTE) teachers. A further four schools are just outside the Southern Gulf catchments: Camooweal, Cloncurry and Normanton in Queensland and Robinson River in the north of the NT part of the catchments. Mount Isa School of the Air, with 176 students (FTE), also covers the properties and communities in the study area. Table 3-14 Schools servicing the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †FTE = full-time equivalent Source: ACARA (2023) (data presented with permission) At the time of the 2021 Census, around 19% of private dwellings were unoccupied, which is around double the Queensland and national means for unoccupied dwellings (Table 3-15). This suggests that the current pool of housing may have some capacity to absorb small future increases in population. Table 3-15 Number and percentage of unoccupied dwellings and population for the Southern Gulf catchments For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the boundaries of the catchments. Source: ABS (2021b) Census data 3.4 Indigenous values, rights, interests and development goals 3.4.1 Introduction and research scope This section gives an overview of the information needed on Indigenous water issues in the Assessment area to provide foundations for community consultations and involvement in further research, and planning and decision-making with government and industry. It provides some key background information about the Indigenous Peoples in the Assessment area and their specific values, rights, interests and objectives in relation to water. The material in this section draws largely from the literature and is intended to give an overview on water-related development information needs. Additional information on Indigenous aspirations, interests and water values is presented in the companion technical report by Lyons et al. (2024). Indigenous Peoples represent a substantial and growing proportion of the population across northern Australia, and they have secured rights and interests in over 70% of the land. They control significant natural and cultural resource assets, including land, water and coastlines. Indigenous Peoples are crucial owners and will increasingly become critical partners, co-investors and stakeholders in future development. Understanding the past is essential to understanding present circumstances and forms of organisation to engage with development options and future possibilities. The material provided here begins with historical information and a description of the contemporary ownership of the Assessment area. It then focuses on development in the lower reaches of the Southern Gulf catchments. Section 3.4.2 describes the past habitation by Indigenous Peoples, 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. Information on Indigenous water values and perspectives on general water development and associated agriculture has been given limited consideration within the Assessment area. Significant research has been undertaken on pastoralism and mining, particularly the relations between Indigenous and non-Indigenous societies in these sectors. The work undertaken here, and in the companion technical report on Indigenous aspirations, interests and water values (Lyons et al., 2024), addresses these information needs. Engagement with Indigenous Peoples is a strong aspiration across governments and key industries. Nevertheless, engagement models can vary considerably, and competing understandings of what ‘engagement’ means (e.g. consultation, involvement, partnership) can substantially affect successful outcomes. Standard stakeholder models can also marginalise Indigenous interests, reducing what Indigenous Peoples understand as prior and inalienable ownership rights to a single ‘stake’ equivalent to all others at the table. As significant land rights and interest holders, Indigenous Peoples in northern Australia have expressed a desire to drive alternative engagement practices to strengthen local leadership and knowledge partnerships (NAILSMA, 2020). Guided by advice from the Carpentaria Land Council Aboriginal Corporation (CLCAC) and feedback from other Aboriginal corporations, the Assessment activity on Indigenous aspirations, development goals and water values was redesigned to test and understand how findings from the previous Assessments may be relevant to Traditional Owners in the Southern Gulf catchments. At the time of writing this report, on-ground consultations were being planned under a joint CSIRO–CLCAC proposal. This section summarises Indigenous values, rights, interests and development goals sourced from publicly available literature. It is complemented by a legal analysis of Indigenous water rights and interests in the NT and Queensland. 3.4.2 Pre-colonial and colonial history Pre-colonial Indigenous society Pre-colonial Indigenous society is characterised by long residence times; a detailed knowledge of ecology and food gathering techniques on which survival depended; complex systems of kinship and territorial organisation; communal and inalienable ownership of land; and a sophisticated set of religious beliefs, often known as the Dreaming. These Indigenous religious cosmologies provide a source of spiritual and emotional connection and guidance on identity, language, law, territorial boundaries and economic relationships (Merlan, 1982; Rose, 2004; Strang, 2005). From an Indigenous perspective, ancestral powers are always present in the landscape, intimately connected to People, Country and culture. Mythological creators, collectively referred to as Dreaming, have imbued significance to places through creation, leaving evidence of their actions and presence through features in the landscape (Martin and Trigger, 2015; Merlan, 1982; Rose, 2004). The cosmological belief of the Dreaming is present among many Indigenous groups. Totemic figures can be animals or plants, take human-like or inanimate object forms, or be sentient beings with agency to act (Martin and Trigger, 2015; Merlan, 1982; Peterson, 2013). Those powers must be considered in any action that takes place on Country. Northern Australia contains archaeological evidence of Indigenous habitation stretching back many tens of thousands of years (Taçon, 2008), but gaps remain 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 and contain many sites of cultural importance (Barber and Jackson, 2014; McIntyre-Tamwoy et al., 2013; Taçon, 2008). Colonisation The first confrontations of European exploration, settlement and the processes of colonisation resulted in significant levels of violence towards Indigenous Peoples. Continued conquest of Indigenous Peoples and the control of their lands brought about the breakdown of the structure and function of existing Indigenous societies, including the decimation of language groups. Access to land and water was critical to both enterprising pastoral interests and Indigenous subsistence society (Merlan, 1978; Morphy and Morphy, 1981; Scambary, 2013; Strang, 1997). The damage and ruin to water sources caused by cattle affected Indigenous Peoples’ access to other resources on Indigenous lands, which intensified conflicts and lead to Indigenous Peoples taking cattle for sustenance and retaliation (McGrath, 1987; Merlan, 1978, 1986; Strang, 1997). Avoidance, armed defensiveness, skirmishes and violent clashes occurred in colonial relationships as a result of competition for food and water resources, colonial attitudes and cultural misunderstandings (Merlan, 1978). Historical exploration and settlement were practices that placed a colonial overlay of interests and boundaries on Indigenous territories and resources (Scambary, 2013). The opening of the Gulf Country to the pastoralism frontier occurred with the search parties sent from Victoria, Queensland and SA to find the explorers Burke and Wills after their disappearance (Scambary, 2013). Cattle arrived from 1861 onwards after the expedition parties identified fertile pastoral lands. Within 4 years, three stations were established in the Southern Gulf catchments: Beames Brook, Floraville and Gregory Downs stations (Scambary, 2013). Pastoral stations were first established in more accessible and watered lands. Groups belonging to those areas bore the brunt of the displacement activity (Trigger, 1992). People moved in various directions to find refuge: some moved to the coast and others retreated to the NT. Gregory Downs Station remains one of the larger stations in the catchments. The subsequent development of available pastoral lands in the 1860s led to the establishment of Burketown on Gangalidda Country in 1865 (Martin, 2012). The establishment of the Coast Track in 1872 invigorated a new focus on pastoralism in northern Australia, in the southern region of the Gulf of Carpentaria and into the NT. This attention to developing northern Australia was the start of violent relationships between Indigenous Peoples and settlers in the frontier (Roberts, 2009). The Coast Track remains notorious for the violence perpetrated against Indigenous Peoples, a period called the ‘Wild Times’. Police stations were established along the Coast Track in the 1880s in response to station owner complaints of cattle stealing by Indigenous Peoples and to establish law and order in newly settled areas (Merlan, 1986; Roberts, 2009; Strang, 1997). Later, the Native Mounted Police helped manage the Indigenous labour working on stations (Kidd, 1997). The 1890s drought that pushed stock to the coastal fringes affected the Waanyi People most. The Waanyi People, whose lands in the west had been taken up by pastoral stations, moved to the east to secure food supplies from depots and stations in Queensland, obtain tobacco and gain protection from the violence (Martin and Trigger, 2015; Martin, 2012). Burketown on Gangalidda land was established as a supply and support town in 1865 but was later abandoned due to fever. It has continued to be a community township. The Queensland Aboriginals Protection and Restriction of the Sale of Opium Act 1897 legislated for the protection of Aboriginal people and triggered the establishment of Aboriginal reserve lands (Kidd, 1997). Under the Act, the government could remove and relocate individuals and families and regulate the employment of Indigenous Peoples in the pastoral industry (Kidd, 1997). The legislation was designed to reduce Indigenous resistance to pastoral activities and stock loss due to changed resource conditions from the new settler interests (Roberts, 2009; Scambary, 2013). At the same time, the pastoral industry relied heavily on Indigenous labour and knowledge of Country to sustain itself. As a result of the application of the legislation to Indigenous Peoples in Queensland, the Native Mounted Police directed its attention to preventing the movement of NT-based Indigenous Peoples into Queensland (Roberts, 2005). Consequently, Waanyi People continue to live and manage their affairs across historically established jurisdictional borders (Roberts, 2005). The Act included a provision for establishing minimum wages, payable under agreed individual negotiations. The payment of wages was under the discretion of the Protector of Aborigines, who determined an individual’s access to their wages. Wages were paid into trusts that people could not access, and in many cases, money did not reach the workers (Kidd, 1997). Unknowingly, individuals entered into agreements that indentured them to stations. In 2019, the Queensland Government agreed to pay $190 million in stolen wages to settle a class action by Aboriginal and Torres Strait Islander Peoples for unpaid wages dating back to the late 19th century and up to the 1980s. A Presbyterian mission was established on Mornington Island in 1914 and later in the location of Doomadgee in 1936 (Long, 1970). In the 1930s, many Indigenous Peoples were living on the Mornington Island and Doomadgee missions. The missions remained in government control for 50 years and systematically dismantled the social and political systems and kinship responsibilities (Long, 1970). The missions became important sources of labour for stations and protectors of Aboriginal people (Long, 1970). Mornington Island and Doomadgee missions transitioned to community governments in 1978 and 1983, respectively (Memmott and Channells, 2004). 3.4.3 Contemporary Indigenous ownership, management, residence and representation Despite the tensions, disruptions and trauma that stemmed from colonisation, Indigenous Peoples remain closely tied to their territory associations, ceremonial relationships and genealogical and historical residential ties (Martin and Trigger, 2015; Trigger, 1987, 1997). Some of these connections are formally recognised by the Australian state and confer particular rights and interests. Indigenous Traditional Ownership The Indigenous owners’ language groups of the Southern Gulf catchments are: •Leichhardt catchment – Kalkadoon, Mitakoodi, Wakabunga, Mayi-Kutuna, Mayi-Thankurti, Mayi-Yapi, Mayi-Yali and Kukatj•Nicholson catchment – Waanyi, Garawa, Wakabunga, Nguburinji, and Gangalidda•Wellesley Islands – Lardil, Yangkaal, Kaiadilt •Settlement catchment – Garawa, Gangalidda •Morning Inlet catchment – Kukatj. Patterns of ownership and language affiliation follow features of the landscape and waterways and are reflected in the place names and songs of significant Dreamings and totemic figures. Formal boundaries are also negotiated between groups. In some areas, there can be overlapping claims and sharing of territories. The process of colonisation has also shaped land interests. Residential information regarding the identification of potential owners and interest holders is provided by registered organisations such as the CLCAC, Queensland South Native Title Services and prescribed bodies corporate (PBCs) such as the Kalkadoon Native Title Aboriginal Corporation RNTBC. Across Australia, the primary form of recognition for Indigenous interests is native title and associated Indigenous Land Use Agreements (ILUAs). Native title provides a bundle of rights (such as access, hunting and fishing) determined through a legal process. The Australian legal system formally recognises Indigenous exclusive native title interests in 10.8% of the Southern Gulf catchments. Native title, Aboriginal Land Act deeds of grant in trust (DOGIT), the Aboriginal Land Rights (Northern Territory) Act 1976 (ALRA), national park co-management and ILUAs are the formal ways in which contemporary Indigenous Peoples of the Southern Gulf catchments exercise some degree of management control over large areas of their traditional lands and have ownership. In the NT jurisdiction of the Nicholson River, Waanyi Garawa Country, the ALRA provides a form of collective freehold ownership. The entire Waanyi Garawa Aboriginal Land Trust, an area of some 11,000 km2, is an Indigenous Protected Area (Ganalanga-Mindibirrina Indigenous Protected Area) (Figure 3-4). Within Queensland, Doomadgee and Mornington Island and the surrounding islands have been granted DOGIT (Figure 3-31). DOGIT is a legal framework that enables Aboriginal and Torres Strait Islander Peoples to ‘self-manage’ their communities. Indigenous Peoples of the Southern Gulf catchments in Queensland exercise some degree of management control over large areas of traditional lands through native title determinations (Figure 3-32). The Waanyi Native Title area is the most extensive determination in the state. Native title (non-exclusive) areas in the Southern Gulf catchments include the jointly managed Boodjamulla National Park (Aboriginal Land), several extensive pastoral holdings, the Bidunggu Aboriginal Community on the Gregory River, the Wellesley Islands and seas connecting them to mainland, and parts of the mainland that have been designated Indigenous Protected Areas. By definition, native title is based on patterns of customary ownership of lands and waters. There are a variety of formal Indigenous management systems in the Southern Gulf catchments, including joint management of Lawn Hill Gorge, Boodjamulla National Park (Aboriginal Land) and the Indigenous Protected Areas of Thuwathu/Bujimulla (marine), Nijinda Durlga and Ganalanga- Mindibirrina (see Figure 3-33). Aboriginal Lands, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\6_Indigenous\4_S_Gulf\1_GIS\1_Map_docs\Indig_S_501_NT_QLDTenure_Cadastre_v4.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-31 Exclusive Aboriginal lands and pastoral interests make up the majority of the Southern Gulf catchments Native title map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\6_Indigenous\4_S_Gulf\1_GIS\1_Map_docs\Indig_S_500_NativeTitle_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-32 The extent of native title claims and determinations over the Southern Gulf catchments Protected areas \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\6_Indigenous\4_S_Gulf\1_GIS\1_Map_docs\Indig_S_502_ProtectedAreas_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-33 Indigenous protected areas and other protected areas in the Southern Gulf catchments as of April 2020 Data Source: Collaborative Australian Protected Areas Database (CAPAD) - (2020) from commonwealth, state and territory governments, non- government organisations, Indigenous and other protected area managers. Augmenting the formal native title claims are registered ILUAs. ILUAs are voluntary registered agreements between native title claimants or holders and other interested parties to use and manage land and resources. ILUAs occur over 47.1% of the Southern Gulf catchments, and grazing and mining are the main land uses in these areas (see Figure 3-34). In addition to these formally registered and enduring native title areas are pastoral leases held by Indigenous Peoples and corporations, which confer rights under which land may be held and used. Pastoral leaseholders can include Traditional Owners and Indigenous residents, individually and in collective entities. The ownership of Lawn Hill and Riversleigh Pastoral Holding Company Pty Ltd was transferred to the Waanyi People under the Gulf Communities Agreement, which provided the native title approval for the development of the Century Mine. Indigenous Land Use Agreements, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\6_Indigenous\4_S_Gulf\1_GIS\1_Map_docs\Indig_S_503_Registerd_ILUAs_v2.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-34 Indigenous Land Use Agreements (ILUAs) Total extent of all ILUAs is shown in pink. ILUAs cover a range of activities including the energy sector, industry (fishing) and local government. One single area may be covered by multiple ILUAs, only mining and pastoral ILUAs are highlighted by hatching. Over 50 ILUA arrangements exist in the Southern Gulf catchments; they relate to pastoralism, mining, conservation, commercial fishing, local shire business and energy infrastructure. ILUAs relating to pastoralism are most common, followed by mining. Mining ILUAs cover the greatest area at 24.2% of all ILUA types combined. Figure 3-34 shows the regions and types of practices that are related to the ILUAs. Indigenous population and residence The Indigenous population made up 27% of the total population of the Southern Gulf catchments in 2021 (Table 3-5). The 27% includes Indigenous Peoples who are part of the recognised local ownership groups identified above and residents who identify as Indigenous but have their origins elsewhere. Many Traditional Owners may primarily reside outside 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 data do not account for the complexity of Indigenous Peoples’ social, linguistic and economic relations. Indigenous communities in the Southern Gulf catchments face a range of social and demographic challenges, including significant unemployment, poor health and housing, and structural impediments to economic participation, including remoteness and social and family units under high levels of stress. Approximately 60% of the population of the lower section of the Southern Gulf catchments identify as Indigenous. Mornington Island and Doomadgee residents are predominantly Indigenous Peoples, and they live in some of the Assessment area’s poorest townships (Everingham et al., 2013; see Table 3-5). High proportions of the population identify as Indigenous in several other settlements and outstations across the catchments. Only 21.4% of the population of the Mount Isa SA2 region identify as Indigenous (Table 3-5). A review of annual reports and a social impact assessment of the Century Mine (Carpentaria Land Council Aboriginal Corporation, 2021a; Carpentaria Land Council Aboriginal Corporation, 2022; Everingham et al., 2013) both stress the importance of creating local opportunities, including employment, capacity building for Indigenous owners and building assets on traditional lands. For instance, the Waanyi People acquired the Lawn Hill and Riversleigh cattle stations in 2020 under an arrangement with Century Mine. Indigenous governance and representation CLCAC and Queensland South Native Title Services (QSNTS) are the two major agencies representing Indigenous rights and interests within the Southern Gulf catchments. QSNTS mainly focuses on native title services. CLCAC has an economic development unit supporting PBCs to plan commercial opportunities and has an established land and sea ranger program. CLCAC collaborates with the Northern Land Council on the Waanyi Garawa Rangers working on the Waanyi Garawa Aboriginal Land Trust. Land-owning Indigenous groups form diverse organisational and political structures as part of their native title roles and economic development planning. In the southern region of the catchments, Kalkadoon Native Title Aboriginal Corporation RNTBC offers land and resource management and employment services to its members through its partnerships with businesses and stakeholders operating on its customary lands. PBCs and land councils can act for Traditional Owners with respect to Indigenous access, participation, partnerships and ownerships. Indigenous Peoples are formally involved in land and sea management through the Wellesley Islands Rangers, the Gangalidda-Garawa Rangers and within the two pastoral stations of Lawn Hill and Riversleigh. 3.4.4 Indigenous water values and perspectives on development Indigenous water values Documentation of Indigenous oral histories across Australia demonstrate People’s connections to waters and lands and how water is a significant feature of Indigenous cultural landscapes (Martin and Trigger, 2015; Strang, 2005; Toussaint et al., 2005). Indigenous histories share a conceptualisation of water sources, rivers and lands as having been derived from the activities of mythic beings in Dreaming time (Barber and Jackson, 2011; Keen, 2003; Martin and Trigger, 2015; Morphy, 1991; Taçon, 2008). Dreaming necessitates sociocultural institutions and protocols that govern the use of water, determined by spiritual entities in the landscape (Martin and Trigger, 2015; Morphy, 1991). Language place names are often strongly associated with places of fresh water and ancestral connections (Merlan, 1981; Trigger, 1987). In the Gulf Country, many Indigenous place names relate to freshwater sources (Martin and Trigger, 2015). Important places of water, waterholes and springs can also hold art sites of Rainbow Dreaming (Martin and Trigger, 2015; Taçon, 2008). Dreaming connects People from the sea to inland regions, with recountings that tell of mythic beings adopting languages belonging to each territory and group (Martin and Trigger, 2015). Indigenous cultural heritage Indigenous cultural heritage relates to archaeological sites, places associated with traditional knowledge, and places of historical or contemporary significance. Aquatic places and systems strongly correlate with cultural heritage. Any development interests in these areas will likely affect places of cultural heritage significance and require negotiated conditions between Traditional Owners and leaseholders. Early scoping of cultural heritage conditions and requirements will aid consultations between development proponents and Traditional Owners. By way of illustration, the Lawn Hill and Riversleigh Pastoral Holding properties acquired by Waanyi People under an agreement with Century Mine are of significant cultural value (Everingham et al., 2013). Under the agreement, Century Mine preserves access rights for exploration with the mining company. In the southern part of the catchments, Kalkadoon Native Title Aboriginal Corporation RNTBC provides cultural heritage services in the mining and pastoral sectors. A priority of land-owning Indigenous groups of the Gulf Country is to have access rights to pastoral properties to sustain cultural heritage and customary values and roles (Memmott and Channells, 2014). Contemporary Indigenous water values Contemporary Indigenous water values stress the importance of secure water supply and good- quality water for healthy landscapes and remote community health and livelihoods (Barber, 2013, 2018; Barber and Woodward, 2018; Lyons and Barber, 2018). Indigenous water needs include cultural practices, pastoralism, tourism, hunting and fishing, agriculture and natural resource management. In the Southern Gulf catchments, Indigenous water values have been affected by the loss of language groups, dislocation from ancestral territories and pastoralism (Memmott and Channells, 2014). Indigenous involvement and ownership of pastoral leases across the catchments over the years have shifted how water is valued by groups and their members, with increasing emphasis on water places for pastoralism. The impact of many years of cattle grazing on the environment and Country has reshaped the relationship of Indigenous Peoples with water (Martin and Trigger, 2015; Scambary, 2013). Aboriginal corporations have negotiated access to water places for Traditional Owners to fish and hunt. The literature demonstrates that water supplies to significant Dreaming sites, and Indigenous perspectives and beliefs about underground water, remain critical to decisions about types of water harvesting techniques, their use and perceived impacts from long-term use (Martin and Trigger, 2015). Perspectives on water and irrigated agricultural development Traditional Owners and communities in the Southern Gulf catchments face increasing outside interest in irrigated agriculture and water extraction in their region (Carpentaria Land Council Aboriginal Corporation, 2022). Water monitoring is a priority in order to understand the impact of water-related development and threats that may affect Indigenous Peoples’ relationships with Country and community wellbeing. For example, CLCAC initiated a water quality monitoring program to gather baseline data and identify any impacts to fresh and saltwater systems caused by contamination, altered water flow and saltwater incursion into freshwater environments. The monitoring program was implemented with the Australian Rivers Institute, Griffith University, and included a capacity-building component. Regional land and water planning is an emerging priority for other Indigenous Peoples with rights and interests in the Southern Gulf catchments. They have concerns that water extraction for irrigation may adversely affect existing pastoral enterprises and Indigenous communities, as well as their future development options (Southern Gulf Natural Resource Management, 2016). Indigenous interests in water planning Water planning is understood as one way of managing water development risk and advancing sustainable management, but it also has challenges. The National Water Initiative guides and sets the goal for recognising Indigenous Peoples’ values and interests in water in terms of access and management. It considers Indigenous customary objectives, economic development interests, native title needs and Indigenous representation. The Water Plan (Gulf) 2007 is being renewed by the Queensland Government. Engagements are being facilitated with Indigenous Peoples through land councils and Aboriginal corporations. Recognising and incorporating Indigenous interests in contemporary water planning processes with competing water demands is challenging because of the limited knowledge of Indigenous Peoples’ interests and how to accommodate their perspectives. The involvement of Indigenous Peoples in water monitoring programs and their expressed concerns about different water extraction methods demonstrate their interest in maintaining sustainable supplies of good-quality water for diverse interests and values. 3.4.5 Indigenous development objectives Exclusive native title determinations and large areas of Indigenous-owned land and ILUAs mean that communities in the Southern Gulf catchments have access to capital, land and resources. However, as a group, Indigenous Peoples in the study area remain socially and economically disadvantaged. Economic development is a key objective of Southern Gulf catchments Indigenous groups. This has been articulated in documents produced by the CLCAC, including its Strategic plan 2021–2025 (Carpentaria Land Council Aboriginal Corporation, 2021b), annual reports (Carpentaria Land Council Aboriginal Corporation, 2021, 2022), PBC land and sea Country plans, Land and sea management 2014 (Carpentaria Land Council Aboriginal Corporation, 2014a, 2016a, 2016b), Social investment prospectus 2014 (Carpentaria Land Council Aboriginal Corporation, 2014b), Indigenous economic and business development opportunities in the Gulf of Carpentaria region (Carpentaria Land Council Aboriginal Corporation, 2013a), Carpentaria Land Councial Aboriginal Corporation destination and product development plan (Carpentaria Land Council Aboriginal Corporation, 2013b). These documents provide details on Indigenous economic development goals. A key feature of Indigenous development objectives in the Assessment area is the diversity of development interests. These are broadly categorised into four key sectors (Carpentaria Land Council Aboriginal Corporation, 2021b): •resources –mining rehabilitation and land management activities –employment in mining operations –asset acquisition and development •land and sea management –management of Indigenous Protected Areas –carbon abatement –fisheries compliance and monitoring •pastoralism –Lawn Hill and Riversleigh pastoral stations •tourism –ecotourism and Boodjamulla (Lawn Hill) National Park management (Gangalidda andGarawa Native Title Aboriginal Corporation, 2014; Carpentaria Land Council AboriginalCorporation, 2021; Queensland Parks and Wildlife Service, 2022) –ecotourism and business development opportunities (particularly proposals thatacknowledge the role of the Waanyi People as traditional custodians, respect their loreand culture, and provide opportunities for the Waanyi People to improve social andeconomic outcomes. This is a core objective of the Waanyi Strategic plan 2020–2025(Waanyi Native Title Aboriginal Corporation RNTBC, 2020) and can potentially improve thesocial and economic outcomes for the Waanyi People –GRAC Birri Fishing Lodge on Mornington Island (which could be re-established as a viabletourism entity) •service delivery –Jigija Indigenous Fire Training Program of the Indigenous-owned business Gangalidda andGarawa Services Pty Ltd (Gangalidda and Garawa Native Title Aboriginal Corporation(GGNTAC)), which provides fire management and mitigation training in the traditionalCountry of the Gangalidda People. These sectors represent a continued focus on supporting a strong, sustainable region. Expanding sustainable economic opportunities that mobilise Traditional Owners’ values, interests and rights broadly across the Assessment area requires partnership and collaboration. The CLCAC Economic Development and Business Support Unit has a role in enabling, generating and sustaining a wide range of opportunities. It works with its partners, stakeholders and the community to benefit the Southern Gulf catchment’s People and communities (Carpentaria Land Council Aboriginal Corporation, 2021b). Partnerships and planning Indigenous Peoples in the Southern Gulf catchments possess diverse natural, historical and cultural assets. Indigenous corporations such as CLCAC and Kalkadoon Native Title Aboriginal Corporation RNTBC have established partnerships across the private, government, non- government and research sectors. For example, there are partnerships with: •Century Mine under the Gulf Communities Agreement •universities for marine and water-related ecosystem monitoring and research •Indigenous and non-Indigenous pastoralists •natural resource management agencies, including Southern Gulf NRM. 3.5 Legal and policy environment Proponents must be aware of a range of legal, policy and regulatory requirements and approvals when contemplating land and water developments within the Southern Gulf catchments. As part of their due diligence process, proponents must be prepared to secure appropriate land tenure and authorisations to take water and to obtain the necessary approvals well in advance of commencing construction and operation of a development. This section describes the overarching Australian legal context and summarises the key issues and related legal, regulatory and approval considerations that apply to water-related developments in the Southern Gulf catchments. Detailed information is available in the companion technical reports on water planning arrangements (Vanderbyl, 2024) and regulatory requirements for land and water development (Speed and Vanderbyl, 2024). 3.5.1 Australian legal and policy context Australia is a federal constitutional monarchy consisting of six states and two territories. The Southern Gulf catchments straddle the border between Queensland and the NT, as shown in Figure 3-30. There are three levels of government: the Australian Government, state and territory governments, and local governments. The Australian Government has powers under the EPBC Act relating to matters of national environmental significance (including those arising from the World Heritage Convention, the Ramsar Convention on Wetlands of International Importance, and the Convention on Biological Diversity) and powers relating to the native title rights of Indigenous Peoples. Generally, the states and territories are responsible for land, water and environmental policy and laws. However, the NT is an administrative territory established by the Australian Government rather than a state. This means that, although it has been given similar powers to the states, the Australian Parliament retains a right of veto over all NT laws. Local governments are established within the states and territories and usually have responsibility for land use planning, which involves establishing local planning schemes that regulate land use and development. However, the planning system in the NT is administered by the NT Government rather than by local government. Laws made by the Australian Government will generally apply to activities in both the NT and Queensland. Laws made by the NT and Queensland governments, or local laws passed by a local government, will only be applicable within the relevant jurisdiction. 3.5.2 Key legal and regulatory requirements Land tenure and native title Proponents will need to secure appropriate tenure over the land of the proposed development site. Consideration should be given as to whether land tenure can be granted or transferred to the developer (or converted to a more suitable form of tenure) and whether any approvals will be required beyond those held by the current owner or lessee of the land. If the land is not freehold, native title requirements are likely to apply. In that case, the proponent will need to check if a native title determination has been made (or is underway) for the land, who the relevant parties are and whether the proposed development is consistent with the rights of native title holders. The proponent will then need to negotiate with the relevant native title holders for the area prior to undertaking development activities. If the proposed development is on Aboriginal freehold land in the NT, the proponent will need to obtain the consent of the Traditional Owners and approval from the relevant Aboriginal Land Council. If the proposed development is on pastoral lease land in the NT, the proponent will require approval for non-pastoral uses from the Pastoral Land Board. If the proposed development is on state land (including leasehold land) within Queensland, the proponent will need to comply with the requirements of the Queensland Land Act 1994. Authority to take water Proponents will need to secure an appropriate entitlement to take any water that may be required to construct and operate the development. An entitlement will usually be required to take surface water or groundwater. In the NT, this is likely to be in the form of a water licence, while in Queensland this would usually be either a water licence or a water allocation. Water entitlements may be purchased and transferred from an existing entitlement holder subject to state or territory requirements or constraints relating to water trading and the purpose of the water use. Alternatively, it may be possible to seek the grant of a new water entitlement from unallocated water reserves. In Queensland, water trading requirements and unallocated water reserves provisions are set out in the Water Plan (Gulf) 2007 (Figure 3-30) and its associated water management protocol. In the NT, such requirements are typically set out in water allocation plans. However, no such plans are currently applicable to that part of the Southern Gulf catchment, so options for securing water are limited. Planning requirements Proponents will need to ensure that their development will be consistent with local and state or territory planning requirements. This usually involves a formal application and assessment process. A single planning scheme applies across the NT, under which a proposed development may be categorised as: (i) permitted, (ii) merit assessable, (iii) impact assessable, or (iv) prohibited. NT Government websites provide detailed checklists and criteria for helping a proponent to determine the category applicable to their development proposal. A development permit will be required for developments categorised as merit assessable and impact assessable. In addition, the NT Planning Commission may prepare a significant development report to be considered in the assessment of the development permit where a proponent’s development is over a certain investment threshold. In Queensland, a proponent will require a development permit for any activities that are ‘assessable development’ under a planning scheme or planning instrument. The development assessment process is coordinated in the case of matters of state interest, including vegetation clearing and the granting of environmental authorities. In some instances a development may be able to obtain an ‘infrastructure designation’ under the Queensland Planning Act 2016, which would remove the need for a development permit for some activities. In addition, a proponent will require a regional interests development approval for broadacre cropping or water storage within a strategic environmental area. Environmental approvals Proponents will need to obtain approvals for certain activities that have a potential environmental impact, including any building or construction activities. A proponent may require federal environmental approval under the EPBC Act if a development has the potential to affect matters of national environmental significance. Federal environmental impact assessment requirements can be met through the NT or Queensland governments’ assessment processes, allowing for a more streamlined assessment process. However, the ultimate decision under the EPBC Act remains with the Australian Minister for the Environment and Water. In the NT, a proponent will require environmental approval for any actions that will have a significant impact on the environment or that are captured under a ‘referral trigger’. Where required, the NT Environment Protection Authority will undertake an environmental impact assessment. Such processes can take significant time to complete. Under Queensland law, a proponent will require environmental approval for an ‘environmentally relevant activity’, which may require an environmental impact statement. The application process is under the Queensland Planning Act development assessment framework. Cultural heritage Proponents will need to identify potential cultural heritage sites and/or objects (including Indigenous cultural heritage sites and/or objects) if a proposed development will affect cultural heritage at a federal and state or territory level. The proponent will need to undertake searches of the NT Heritage Register, the NT Aboriginal Areas Protection Authority register of sacred sites, the Queensland Heritage Register and the Queensland Cultural Heritage Register. National heritage values will also need to be considered through an environmental impact assessment process under the EPBC Act. A cultural heritage management plan is advisable (and may be required) for significant developments. Works in a watercourse Proponents will need approval to undertake any activities within a watercourse. In the NT, a proponent will require a permit under the Northern Territory Water Act 1992 to interfere with a watercourse (e.g. extraction of materials, construction within a waterway, or diversion of a watercourse). In Queensland, a proponent will require a permit under the Queensland Water Act 2000 to interfere with a watercourse, including for the construction of a barrier and the removal of vegetation or quarry material. Clearing vegetation Proponents will require approval to clear native vegetation to allow for construction or farming or other agricultural activities. Exemptions may apply for routine maintenance and day-to-day management activities. In the NT, restrictions apply to both freehold land (including Aboriginal freehold land) and pastoral leases. For clearing on pastoral land, permit applications are determined by the Pastoral Land Board. For freehold land, applications are assessed under the NT Planning Act 1999 and must be lodged with the Department of Infrastructure, Planning and Logistics. In Queensland, clearing of vegetation is generally captured as assessable development and will require a development permit under the Queensland Planning Act 2016. 3.6 References 29Metals (2022) Capricorn Copper. Viewed 24 February 2024, Hyperlink to: Capricorn Copper . 29Metals (2024) Mineral resources and ore reserves. Viewed 24 February 2024 Hyperlink to: Mineral resources and ore reserves . ABARES (2022) Land use data download – Land use of Australia 2010–11 to 2015–16, 250 m [dataset]. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra. Viewed 20 December 2023, Hyperlink to: Land use data download . ABC (8 September 2005) Xstrada lead mine to close in December. Australian Broadcasting Corporation. Viewed 17 February 2024, Hyperlink to: Xstrada lead mine to close in December . ABS (2006) Census of population and housing time series profile. Catalogue number 2003.0 for various Statistical Area regions falling partly within Southern Gulf catchments. Australian Bureau of Statistics website. Viewed 29 October 2023, Hyperlink to: ABS Census website . ABS (2011) Census of population and housing time series profile. Catalogue number 2003.0 for various Statistical Area regions falling partly within Southern Gulf catchments. Australian Bureau of Statistics, Canberra. Viewed 29 October 2023, Hyperlink to: ABS Census website . ABS (2016) Census of population and housing time series profile. Catalogue number 2003.0 for various SA regions falling partly within Southern Gulf catchments. Australian Bureau of Statistics website. Viewed 29 October 2023, Hyperlink to: ABS Census website . ABS (2021b) Census of population and housing time series profile. Catalogue number 2003.0 for various SA regions falling partly within the Southern Gulf catchments. Australian Bureau of Statistics, Canberra. Viewed 29 October 2023, Hyperlink to: ABS Census website . ABS (2021a) Census of Carpentaria Aboriginal and/or Torres Strait Islander people. Australian Bureau of Statistics, Canberra. Viewed 7 December 2023, https://www.abs.gov.au/census. ABS (2021c) Mount Isa 2021 Census community profiles – Working population profile [dataset], Australian Bureau of Statistics, Canberra. Viewed 18 February 2024, Hyperlink to: Mount Isa 2021 Census community profiles . ABS (2022) Value of agricultural commodities produced, Australia, 2020–21. Australian Bureau of Statistics website. Viewed 29 October 2023, Hyperlink to: Value of agricultural commodities produced, Australia, 2020–21 . ABS (2023) Socio-Economic Indexes for Area (SEIFA) Australia, 2021 (Released 27 April 2023). Australian Bureau of Statistics, Canberra. Viewed 25 September 2023, Hyperlink to: Socio-Economic Indexes for Area (SEIFA) Australia, 2021 . ACARA (2023) My School. My School website. Australian Curriculum, Assessment and Reporting Authority, Sydney. Viewed 13 September 2023, Hyperlink to: My School website . AEMO (2023) The National Electricity Market. Fact sheet. Australian Energy Market Operator. Viewed 17 January 2024, Hyperlink to: The National Electricity Market . Australian Energy Regulator (2021) Pipeline capacity trading: two year review. AER, Canberra. Hyperlink to: Pipeline capacity trading: two year review . Amelung B and Nicholls S (2014) Implications of climate change for tourism in Australia. Tourism Management 41, 228–244. Hyperlink to: Implications of climate change for tourism in Australia . André J and Waterson L (13 October 2023) North West Phosphate begins mining Paradise South in outback Queensland amid global demand for fertiliser, renewables. ABC North West Queensland. Viewed 23 February 2024, Hyperlink to: North West Phosphate begins mining Paradise South in outback Queensland amid global demand for fertiliser, renewables . APA (2022) Diamantina Power Station. APA Group. Viewed 17 January 2024, Hyperlink to: Diamantina Power Station . Atlas of Living Australia (2023a) Atlas of Living Australia: Mornington Island occurrence download at https://doi.org/10.26197/ala.070e61f4-73a4-437f-a15e-ef9f305c7df8 accessed on 8 November 2023. Atlas of Living Australia (2023b) Atlas of Living Australia: Southern Gulf occurrence download at https://doi.org/10.26197/ala.53e3a9bb-e2f1-428f-87c2-ec30b38e8f09 accessed on 13 September 2023. Australian Fisheries (1975) Prawn farm closed despite successes. Australian Fisheries 34(7), 21–22. Australian Indigenous Australians Agency (2023) Environment. Viewed 18 March 2024, https://www.niaa.gov.au/our-work/environment-and-land. Barber M (2013) Indigenous water values, rights and interests 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. Barber M (2018) Indigenous water values, rights, interests and development objectives in the Darwin catchments. 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, Canberra. Hyperlink to: Indigenous water values, rights, interests and development objectives in the Darwin catchments . Barber M and Jackson S (2011) Water and Indigenous People in the Pilbara, Western Australia: a preliminary study. CSIRO, Darwin. Viewed 18 March 2024, https://www.academia.edu/49389081/Water_and_Indigenous_People_in_the_Pilbara_Western_Australia_A_Preliminary_Study. 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. Barber M and Woodward E (2018). Indigenous water values, rights, interests and development objectives in the Fitzroy catchment. 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, Canberra. Hyperlink to: Indigenous water values, rights, interests and development objectives in the Fitzroy catchment . Barker RD and Vestjens WJM (1990) The food of Australian birds: 1. Non-passerines. CSIRO Division of Wildlife and Ecology, Canberra. 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. Bishop K (1980) Fish kills in relation to physical and chemical changes in Magela Creek (East Alligator River system, Northern Territory) at the beginning of the tropical wet season. Australian Zoologist 20, 485–500. Bowen MK, Chudleigh F and Perry L (2020) Northern Downs beef production systems: preparing for, responding to, and recovering from drought. Queensland Department of Agriculture and Fisheries, Queensland, Brisbane. Bowen MK, Chudleigh F, Rolfe JW and English B (2019) Northern Gulf beef production systems: preparing for, responding to, and recovering from drought. Queensland Department of Agriculture and Fisheries, Queensland, Brisbane. 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. Hyperlink to: Impact of water development on river flows and the catch of a commercial marine fishery . 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. Hyperlink to: Nutrients in Australian tropical rivers: changes with agricultural development and implications for receiving environments . Buckley KA, Smith LDG, Crook DA, Pillans RD and Kyne PM (2020) Conservation impact scores identify shortfalls in demonstrating the benefits of threatened wildlife displays in zoos and aquaria. Journal of Sustainable Tourism 28(7), 978-1002. Hyperlink to: Conservation impact scores identify shortfalls in demonstrating the benefits of threatened wildlife displays in zoos and aquaria . 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. Bugno M and Polonsky M (2024) The tyranny of time and distance: impediments for remote agritourism operators accessing support in the Northern Territory, Australia. Tourism Planning & Development 21(2), 200–222. Hyperlink to: The tyranny of time and distance: impediments for remote agritourism operators accessing support in the Northern Territory . Bunn SE and Arthington AH (2002) Basic principles and ecological consequences of altered flow regimes for aquatic biodiversity. Environmental Management 30(4), 492–507. Hyperlink to: Basic principles and ecological consequences of altered flow regimes for aquatic biodiversity . Burford MA, Valdez D, Curwen G, Faggotter JS, Ward DP and O’Brien KR (2016) Inundation of saline supratidal mudflats provides an important source of carbon and nutrients in an aquatic system. Marine Ecology Progress Series 545, 21–33. Hyperlink to: Inundation of saline supratidal mudflats provides an important source of carbon and nutrients in an aquatic system . 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. Hyperlink to: The use of regional and alluvial groundwater by riparian trees in the wet‐dry tropics of northern Australia . Carpentaria Land Council Aboriginal Corporation (2013a) Indigenous economic and business development opportunities in the Gulf of Carpentaria. CLCAC. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/clcac_economic_development_summary_report_2013.pdf. Carpentaria Land Council Aboriginal Corporation (2013b) Destination and product development plan. CLCAC. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/clcac_destination_and_product_development_plan_201306.pdf. Carpentaria Land Council Aboriginal Corporation (2014a) Land and sea management. CLCAC. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/clcac_land-sea_management.pdf. Carpentaria Land Council Aboriginal Corporation (2014b) Social investment prospectus. CLCAC. Viewed 18 March 2024, https://www.clcac.com.au/publications/2014/51. Carpentaria Land Council Aboriginal Corporation (2016a) Nijinda Durlga (Gangalidda) Indigenous Protected Area management plan. Prepared by Carpentaria Land Council Aboriginal Corporation on behalf of the Gangalidda People. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/gangalidda_ipa_management_plan_web_ready.pdf. Carpentaria Land Council Aboriginal Corporation (2016b) Thuwathu/Bujimulla Indigenous Protected Area management plan. Prepared by Carpentaria Land Council Aboriginal Corporation on behalf of the Traditional Owners of the Wellesley Islands. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/wellesly_islands_ipa_management_plan_web_ready_0.pdf. Carpentaria Land Council Aboriginal Corporation (2021) Annual report 2020–2021. CLCAC. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/clcac_annualreport_2020_21_final_.pdf. Carpentaria Land Council Aboriginal Corporation (2021b) Strategic plan 2021–2025. CLCAC. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/clcac_strategic_plan_2021- 25_final_dp_web.pdf. Carpentaria Land Council Aboriginal Corporation(2022) Annual report 2021–2022. CLCAC. Viewed 18 March 2024, Hyperlink to: Annual report 2021–2022 . 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. 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. Cobcroft J, Bell R, Fitzgerald J, Diedrich A and Jerry D (2020) Northern Australia aquaculture industry situational analysis, Project. A.1.1718119. Cooperative Research Centre for Developing Northern Australia (CRCNA). Compagno LJV, Cook SF, Oetinger MI and Fowler SL (2006) Anoxypristis cuspidata. The IUCN Red List of Threatened Species. Corriveau A, Klaassen M, Garnett ST, Kaestli M, Christian K, Crewe TL, Loewensteiner DA and Campbell HA (2022) Seasonal space use and habitat selection in magpie geese: implications for reducing human-wildlife conflicts. Journal of Wildlife Management 86(7), e22289. Hyperlink to: Seasonal space use and habitat selection in magpie geese: implications for reducing human-wildlife conflicts . Cotin J, García-Tarrasón M, Sanpera C, Jover L and Ruiz X (2011) Sea, freshwater or saltpans? Foraging ecology of terns to assess mercury inputs in a wetland landscape: The Ebro Delta. Estuarine, Coastal and Shelf Science 92(1), 188–194. Hyperlink to: Sea, freshwater or saltpans? Foraging ecology of terns to assess mercury inputs in a wetland landscape: The Ebro Delta . Cowley T (2014) The pastoral industry survey – Katherine region. NT Department of Primary Industry and Fisheries, 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). Hyperlink to: Meeting the challenge of interacting threats in freshwater ecosystems: a call to scientists and managers . Crook DA, Buckle DJ, Morrongiello JR, Allsop QA, Baldwin W, Saunders TM and Douglas MM (2020) Tracking the resource pulse: movement responses of fish to dynamic floodplain habitat in a tropical river. Journal of Animal Ecology 89(3), 795–807. Hyperlink to: Tracking the resource pulse: movement responses of fish to dynamic floodplain habitat in a tropical river . 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. Hyperlink to: Human effects on ecological connectivity in aquatic ecosystems: integrating scientific approaches to support management and mitigation . 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. Hyperlink to: Qualitative mathematical models to support ecosystem-based management of Australia’s Northern Prawn Fishery . Department of Agriculture‚ Water and the Environment (2020a) Collaborative Australian Protected Areas Database (CAPAD) 2020 Marine. Viewed 18 September 2024, https://www.dcceew.gov.au/environment/land/nrs/science/capad/2020. Department of Agriculture‚ Water and the Environment (2020b) Collaborative Australian Protected Areas Database (CAPAD) 2020 Terrestrial. Viewed 18 September 2024, https://www.dcceew.gov.au/environment/land/nrs/science/capad/2020. Department of Agriculture‚ Water and the Environment (2021a) Directory of Important Wetlands in Australia. Viewed 08 December 2023, https://www.environment.gov.au/water/wetlands/australian-wetlands-database/directory- important-wetlands. Department of Agriculture‚ Water and the Environment (2021b) EPBC Act Protected Matters Report: Southern Gulf catchments. Viewed 19 September 2023, https://pmst.awe.gov.au/. Department of Climate Change‚ Energy‚ the Environment and Water (2020) National Vegetation Information System (NVIS) v6.0. Viewed 19 January 2024, Hyperlink to: National Vegetation Information System . Department of the Environment and Energy (2010) Directory of Important Wetlands in Australia (DIWA). Viewed 08 December 2023, https://www.environment.gov.au/water/wetlands/australian-wetlands-database. Department of Environment and Science (2023) WildNet wildlife records - published - Queensland, Viewed 28 November 2023. Department of Environment Parks and Water Security (2019) Fauna Atlas N.T. accessed on 28 November 2023. Dias MP, Lecoq M, Moniz F and Rabaça JE (2014) Can human-made saltpans represent an alternative habitat for shorebirds? Implications for a predictable loss of estuarine sediment flats. Environmental Management 53(1), 163–171. Hyperlink to: Can human-made saltpans represent an alternative habitat for shorebirds? . 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. Hyperlink to: Beyond biological performance measures in management strategy evaluation . 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. Hyperlink to: River and wetland food webs in Australia’s wet–dry tropics . Duke NC, Field C, Mackenzie JR, Meynecke J-O and Wood AL (2019) Rainfall and its possible hysteresis effect on the proportional cover of tropical tidal-wetland mangroves and saltmarsh–saltpans. Marine and Freshwater Research 70(8), 1047–1055. Hyperlink to: Rainfall and its possible hysteresis effect on the proportional cover of tropical tidal-wetland mangroves and saltmarsh–saltpans . 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. Hyperlink to: Large-scale dieback of mangroves in Australia’s Gulf of Carpentaria . Dulvy N, Davidson, LNK, Kyne, PM, Simpfendorfer, CA, Harrison, LR, Carlson, JK, and Fordham, SV, (2016) Ghosts of the coast: global extinction risk and conservation of sawfishes. Aquatic Conservation: Marine and Freshwater Ecosystems 26(1), 134-153. Hyperlink to: Ghosts of the coast: global extinction risk and conservation of sawfishes . 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. Ebner BC, Morgan DL, Kerezsy A, Hardie S, Beatty SJ, Seymour JE, Donaldson JA, Linke S, Peverell S, Roberts D, Espinoza T, Marshall N, Kroon F, Burrows D and McAllister R (2016) Enhancing conservation of Australian freshwater ecosystems: identification of freshwater flagship fishes and relevant target audiences. Fish and Fisheries 17(4), 1134–1151. Hyperlink to: Enhancing conservation of Australian freshwater ecosystems . Everingham J, Barnes R and Brereton D (2013) Gulf Communities Agreement 2008–2013. 15-year Review. Centre for Social Responsibility in Mining, The University of Queensland, Brisbane. Evolution Mining (2023) Fact Sheet FY23 – Ernest Henry operation. Viewed 23 February 2024, Hyperlink to: Fact Sheet FY23 – Ernest Henry operation . Finlayson C, Storrs M, Hall R and Bayliss B (1999) Management issues for northern Australian wetlands. In: Finlayson CM 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. Hyperlink to: Protecting Indigenous values in water management . Frith HJ and Davies SJJF (1961) Ecology of the magpie goose, Anseranas semipalmata Latham (Anatidae). CSIRO Wildlife Research 6(2), 91–141. Hyperlink to: Ecology of the magpie goose, Anseranas semipalmata Latham (Anatidae) . Fry G, Laird, A. and Lawrence, E. (2021) Monitoring interactions with bycatch species using crew- member observer data collected in the Northern Prawn Fishery: 2017–2019. Final Report to Australian Fisheries Management Authority, R2017/0835. CSIRO, Australia. Fuller D, Buultjens J and Cummings E (2005) Ecotourism and indigenous micro-enterprise formation in northern Australia opportunities and constraints. Tourism Management. Hyperlink to: Ecotourism and indigenous micro-enterprise formation in northern Australia opportunities and constraints . Gangalidda and Garawa Native Title Aboriginal Corporation (2014) Garawa land and sea country plan. Viewed 18 March 2024, https://www.clcac.com.au/sites/default/files/downloads/clcac_garawa_land_sea_country_plan_web_version_50dpi.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. Hyperlink to: https://doi.org/10.1038/sdata.2015.61 . Geological Survey of Queensland (2024) Mineral Occurrence Database (MINOCC). Queensland Government. Viewed 8 March 2024, https://geoscience.data.qld.gov.au/dataset/ds000004. Geoscience Australia (2017) GEODATA TOPO 250K Series 3 Topographic Data. Viewed 1 October 2023, http://pid.geoscience.gov.au/dataset/ga/64058. Glencore (2023a) Mount Isa Mines operational changes. Media statement 18 October 2018. Viewed 25 July 2023, Hyperlink to: Mount Isa Mines operational changes. Media statement . Glencore (2023b) Glencore latest publications – financial statements and operational reports. Viewed 21 December 2023, Hyperlink to: Glencore latest publications – financial statements and operational reports . Glencore (2023c) Operational changes – Queensland metals. Viewed 10 January 2024, Hyperlink to: Operational changes – Queensland metals . Graham P, Hayward J and Foster J (2023). GenCost 2023–24: consultation draft. CSIRO, Australia. DOI: 10.25919/xxd6-qz89. Greiner R (2010) Improving the net benefits from tourism for people living in remote northern Australia. Sustainability 2(7) 2197–2218. Hyperlink to: Improving the net benefits from tourism for people living in remote northern Australia . Hadwen WL, Arthington AA, Boon PI, Taylor B and Fellows CS (2011) Do climatic or institutional factors drive seasonal patterns of tourism visitation to protected areas across diverse climate zones in eastern Australia? Tourism Geographies 13(2), 187–208. Hyperlink to: Do climatic or institutional factors drive seasonal patterns of tourism visitation to protected areas across diverse climate zones in eastern Australia? . Hayward J (2024) Potential for farm-scale hybrid renewable energy supply options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. HERE Technologies (2021) Traffic products. NAVIGATE. Viewed 15 February 2023, Hyperlink to: Traffic products web page . 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. Hyperlink to: Prioritizing refugia for freshwater biodiversity conservation in highly seasonal ecosystems . 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. Hyperlink to: TRANSIT– a model for simulating infrastructure and policy interventions in agriculture logistics . Incitec Pivot (n.d.) Phosphate Hill – Working with us. Viewed 23 February 2024, https://www.incitecpivot.com.au/siteassets/document-centre/work-with- us/phosphate_hill_leaflet.pdf Integration and Application Network (2023) Integration and Application Network (ian.umces.edu/media-library). Attribution-ShareAlike 4.0 International (CC BY-SA 4.0). Victoria catchment icons. Bos taurus, Penaeus monodon adult: Jane Hawkey; Corymbia spp, Melaleuca rhaphiophylla, Generic tree rainforest 3, Avicennia germinans, Macropus fuliginosus, Pluvialis fulva, trawler: Tracey Saxby; Pennisetum ciliare, Eucalyptus camaldulensis: Kim Kraeer and Lucy Van Essen-Fishman; Lates calcarifer, Carcharhinus leucas (bull shark) 2: Dieter Tracey; Haematopus longirostris (pied oystercatcher): Jane Thomas. Viewed 29 September 2021, 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) Phosphorus management of beef cattle in northern Australia (reprinted 2015). 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, NT. 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. Hyperlink to: Does flood rhythm drive ecosystem responses in tropical riverscapes? . Joint Ore Reserves Committee (2004) Australasian code for reporting of exploration results, mineral resources and ore reserves. The JORC code (2004 edition). The Australasian Institute of Mining and Metallurgy, Australian Institute of Geoscientists and Minerals Council of Australia. Viewed 24 February 2024, Hyperlink to: Australasian code for reporting of exploration results, mineral resources and ore reserves 2004 . Joint Ore Reserves Committee (2012) Australasian code for reporting of exploration results, mineral resources and ore reserves. The JORC code (2012 edition). The Australasian Institute of Mining and Metallurgy, Australian Institute of Geoscientists and Minerals Council of Australia. Viewed 23 February 2024, Hyperlink to: Australasian code for reporting of exploration results, mineral resources and ore reserves 2012 . Junk W, Bayley P and Sparks R (1989) The flood pulse concept in river-floodplain systems. In: Dodge DP (ed.) Proceedings of the International Large River Symposium. Canadian Special Publication of Fisheries and Aquatic Sciences 106, 110–127. Keen (2003) Aboriginal economy and society: Australia at the threshold of colonisation. Oxford University Press, Melbourne. Kidd R (1997) The way we civilise: Aboriginal affairs – the untold story. University of Queensland Press, St Lucia. 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. Hyperlink to: Implications of water extraction on the low-flow hydrology and ecology of tropical savannah rivers . Kingsford R and Norman F (2002) Australian waterbirds – products of the continent's ecology. Emu 102(1), 47–69. Hyperlink to: Australian waterbirds – products of the continent's ecology . Kulendran N and Dwyer L (2010) Seasonal variation versus climate variation for Australian tourism. CRC for Sustainable Tourism, Australia. Lane R and Waitt G (2007) Inalienable places: self-drive tourists in Northwest Australia. Annals of Tourism Research 34(1), 105–121. Hyperlink to: Inalienable places: self-drive tourists in Northwest Australia . Larson S and Herr A (2008) Sustainable tourism development in remote regions? Questions arising from research in the North Kimberley, Australia. Regional Environmental Change 8, 1–13. Hyperlink to: Sustainable tourism development in remote regions? . Last P and Stevens J (2008) Sharks and rays of Australia. CSIRO, Collingwood. Lear KO, Gleiss AC, Whitty JM, Fazeldean T, Albert JR, 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, 17071. Hyperlink to: Recruitment of a critically endangered sawfish into a riverine nursery depends on natural flow regimes . Lei W, Masero JA, Piersma T, Zhu B, Yang H-Y and Zhang Z (2018) Alternative habitat: the importance of the Nanpu Saltpans for migratory waterbirds in the Chinese Yellow Sea. Bird Conservation International 28(4), 549–566. Hyperlink to: Alternative habitat: the importance of the Nanpu Saltpans for migratory waterbirds in the Chinese Yellow Sea . 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. Hyperlink to: Are mangroves and seagrasses sources of organic carbon for penaeid prawns in a tropical Australian estuary? . Long JPM (1970) Aboriginal settlements: a survey of institutional communities in eastern Australia. Australian National University Press, Canberra. Lundberg C and Fredman P (2012) Success factors and constraints among nature-based tourism entrepreneurs. Current Issues in Tourism 15(7), 649–671. Hyperlink to: Success factors and constraints among nature-based tourism entrepreneurs . Lyons P, and Barber, M (2018) Indigenous water values, rights, interests and development objectives in the Mitchell catchment. 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. Hyperlink to: Indigenous water values, rights, interests and development objectives in the Mitchell catchment . Lyons P, Barber M, Braedon P and Pert P (2024) Indigenous water values, rights, interests and development goals in the Southern Gulf Catchments: A literature review and implications for future research. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. 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. Martin RJ and Trigger D (2015) ‘Nothing never change’: mapping land, water and Aboriginal identity in the changing environments of northern Australia's Gulf Country. Settler Colonial Studies 5(4), 317–333. Hyperlink to: ‘Nothing never change’: mapping land, water and Aboriginal identity in the changing environments of northern Australia's Gulf Country . Martin RJ (2012) Representing 'The Gulf': issues of place, race and identity in a region of Northern Australia. PhD thesis, University of Western Australia. McDavitt M (1996) The cultural and economic importance of sawfishes (family Pristidae). Shark News 8, 10–11. McGinness HM (2016) Waterbird responses to flooding, stressors and threats. A report prepared for the Murray–Darling Freshwater Research Centre as part of the Environmental Water Knowledge and Research Project. CSIRO, Canberra. McGrath A (1987) Born in the cattle: Aborigines in cattle country. Allen and Unwin, Sydney. McIntyre-Tamwoy S, Bird M and Atkinson 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, 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. Hyperlink to: Persistence of in-stream waterholes in ephemeral rivers of tropical northern Australia and potential impacts of climate change . McJannet D, Yang A and Seo L (2023) Climate data characterisation for hydrological and agricultural scenario modelling across the Victoria, Roper and Southern Gulf catchments. A technical report from the CSIRO Victoria, Roper and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Meissner S (2021) The impact of metal mining on global water stress and regional carrying capacities – a GIS-based water impact assessment. Resources 2021,10, 120. Hyperlink to: The impact of metal mining on global water stress and regional carrying capacitie . Memmott, P. and Channells, G. (2004). Southern Gulf of Carpentaria sea Country needs and issues Research. Hobart, National Oceans Office. Memmott P and Channells G (2004) Living on saltwater country: Southern Gulf of Carpentaria sea country management needs and issues. Paul Memmott and Associates in association with Aboriginal Environments Research Centre, University of Queensland, on behalf of the Carpentaria Land Council Aboriginal Corporation. 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. Hyperlink to: Land, language and social identity in Aboriginal Australia . 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. Northern Land Council, Darwin. Merrin L, Stratford D, Kenyon R, Pritchard J, Linke S, Ponce Reyes R, Blamey L, Buckworth R, Castellazzi P, Costin B, Deng R, Gannon R, Gao S, Gilbey S, Lachish S, McGinness H, Plagányi E and Waltham N (2024) Ecological assets of the Southern Gulf catchments to inform water resource assessments. A technical report from the CSIRO Southern Gulf catchments Water Resource Assessment for the National Water Grid. CSIRO, Australia. MITEZ (2021) North Queensland cotton gin assessment and feasibility study. Prepared for Mount Isa to Townsville Economic Development Zone Incorporated (MITEZ). Viewed 13 December 2023, Hyperlink to: North Queensland cotton gin assessment and feasibility study . Mitsch WJ, Bernal B and Hernandez ME (2015) Ecosystem services of wetlands. International Journal of Biodiversity Science, Ecosystem Services and Management, 11(1), 1–4. Hyperlink to: Ecosystem services of wetlands . MMG (2023) Mineral resources and ore reserves statement as at June 30 2023. Viewed 24 February 2024, Hyperlink to: Mineral resources and ore reserves statement as at June 30 2023 . MMG (2024) Dugald River. Viewed 24 February 2024, Hyperlink to: Dugald River . Morgan D, Whitty J, Phillips N, Thorburn D, Chaplin J and McAuley R (2011) North-western Australia as a hotspot for endangered elasmobranchs with particular reference to sawfishes and the northern river shark. Journal of the Royal Society of Western Australia 94(2), 345– 358. Morgan DL, Ebner BC, Allen MG, Gleiss AC, Beatty SJ and Whitty JM (2017) Habitat use and site fidelity of neonate and juvenile green sawfish Pristis zijsron in a nursery area in Western Australia. Endangered Species Research 34, 235-249. Hyperlink to: Habitat use and site fidelity of neonate and juvenile green sawfish Pristis zijsron in a nursery area in Western Australia . Morphy H (1991) Ancestral connections: art and an Aboriginal system of knowledge. University of Chicago Press, Chicago. Morphy H and Morphy F (1981) Yutpundji-Djindiwirritj (Roper Bar) Land Claim Book. Northern Land Council, Darwin. NAILSMA (2020) Business on Country diversification strategy – for enterprise, economic development and health and productive lands and seas. CRC for Developing Northern Australia. 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. Hyperlink to: Carbon and nutrient subsidies to a lowland river following floodplain inundation . Northey SA and Haque N (2013) Life cycle based water footprint of selected metal production – Assessing production processes of copper, gold and nickel. CSIRO, Australia. EP137374. NT Department of Environment, Parks and Water Security (2022) Land use mapping project of the Northern Territory (2016–2022) (LUMP). Open Data Portal, NT Government. Sourced 28 March 2023, Hyperlink to: and use mapping project of the Northern Territory (2016–2022) . NT Geological Survey (2024) Mineral Occurrence Database (MODAT). Version March 2023. Northern Territory Government. Viewed 10 January 2024, Hyperlink to: Mineral Occurrence Database . NT Government (2020) Important wetlands. Northern Territory Government, Darwin. Viewed 08 December 2023, https://nt.gov.au/environment/soil-land-vegetation/important-wetlands. NT Government (2024a) Uranium. Resourcing the Territory factsheet. Viewed 23 February 2024, Hyperlink to: Uranium. Resourcing the Territory factsheet . NT Government (2024b) Strike tenure and geoscience information. NT Government. Viewed 10 January 2024, Hyperlink to: Strike tenure and geoscience information . Nye ER, Dickman CR and Kingsford RT (2007) A wild goose chase-temporal and spatial variation in the distribution of the Magpie Goose (Anseranas semipalmata) in Australia. Emu 107(1), 28– 37. Hyperlink to: A wild goose chase-temporal and spatial variation in the distribution of the Magpie Goose (Anseranas semipalmata) in Australia . Ooi N and Laing JH (2010) Backpacker tourism: sustainable and purposeful? Investigating the overlap between backpacker tourism and volunteer tourism motivations. Journal of Sustainable Tourism 18(2), 191–206. Hyperlink to: Investigating the overlap between backpacker tourism and volunteer tourism motivations . Outback Queensland Tourism Association (2021) Outback Queensland destination management plan 2021–2024. Viewed 10 March 2024, Hyperlink to: Outback Queensland destination management plan 2021–2024 . 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. Hyperlink to: On the persistence of sharing: personhood, asymmetrical reciprocity, and demand sharing in the Indigenous Australian domestic moral economy . 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. Hyperlink to: Seasonal contrasts in carbon resources and ecological processes on a tropical floodplain . 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. Hyperlink to: Seasonal changes in water quality and macrophytes and the impact of cattle on tropical floodplain waterholes . 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. Hyperlink to: Productivity and connectivity in tropical riverscapes of Northern Australia . Peverell S (2005) Distribution of sawfishes (Pristidae) in the Queensland Gulf of Carpentaria, Australia, with notes on sawfish ecology. Environmental Biology of Fishes 73(4), 391–402. Hyperlink to: Distribution of sawfishes (Pristidae) in the Queensland Gulf of Carpentaria, Australia, with notes on sawfish ecology . Phillips N, Chaplin JA, Peverell SC and Morgan DL (2017) Contrasting population structures of three Pristis sawfishes with different patterns of habitat use. Marine and Freshwater Research 68(3). Hyperlink to: Contrasting population structures of three Pristis sawfishes with different patterns of habitat use . Pickering CM and Hill W (2007) Impacts of recreation and tourism on plant biodiversity and vegetation in protected areas in Australia. Journal of Environmental Management 85, 791– 800. Hyperlink to: Impacts of recreation and tourism on plant biodiversity and vegetation in protected areas in Australia . Pidgeon R and Humphrey C (1999) Some important ecological features of wetlands in the wet-dry tropics of northern Australia. In: Finlayson CM 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. Hyperlink to: The natural flow regime: a paradigm for river conservation and restoration . Poiani A (2006) Effects of floods on distribution and reproduction of aquatic birds. Advances in Ecological Research, Vol 39, 63–83. DOI: 10.1016/s0065-2504(06)39004-6. Poiner IR, Staples DJ and Kenyon R (1987) Seagrass communities of the Gulf of Carpentaria, Australia. Marine and Freshwater Research 38(1), 121–131. Hyperlink to: eagrass communities of the Gulf of Carpentaria, Australia . Pollino C, Barber E, Buckworth R, Cadiegues M, Cook G, Deng A, Ebner B, Kenyon R, Liedloff A, Merrin L, Moeseneder C, Morgan D, Nielsen D, O'Sullivan, Ponce Reyes R, Robson B, Stratford D, Stewart-Koster B and Turschwell M (2018) 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. Hyperlink to: Synthesis of knowledge to support the assessment of impacts of water resource development to ecological assets in northern Australia: asset analysis . Prideaux B (2013) An investigation into factors that may affect the long term environmental and economic sustainability of tourism in northern Australia. The Cairns Institute, James Cook University, Cairns. Prosser I, Wolf L and Littleboy A (2011). Water in mining and industry. In: Prosser I (ed.) Water: science and solutions for Australia. CSIRO Publishing, Australia, 135–146. Pusey BJ, Jardine TD, Bunn SE and Douglas MM (2020) Sea catfishes (Ariidae) feeding on freshwater floodplains of northern Australia. Marine and Freshwater Research 71(12), 1628– 1639. DOI: 10.1071/MF20012. Pusey BJ, Kennard MJ and Arthington AH (2004) Freshwater fishes of north-eastern Australia. CSIRO Publishing, Collingwood. Queensland Department of Environment and Science (2022) WetlandSummary – facts and maps. WetlandInfo. Queensland Government. Viewed 23 April 2024, Hyperlink to: WetlandSummary – facts and maps . Queensland Department of Environment and Science (2023a) Boodjamulla (Lawn Hill) National Park. Queensland Government. Viewed 17 February 2023, https://parks.des.qld.gov.au/parks/boodjamulla. Queensland Department of Environment and Science (2023b) Finucane Island National Park. Queensland Government. Viewed 08 December 2023, https://parks.des.qld.gov.au/parks/finucane-island. Queensland Department of Environment, Science and Innovation (2024a) Biodiversity status of 2021 remnant regional ecosystems – Queensland, Queensland Government. Viewed 16 September 2024, https://qldspatial.information.qld.gov.au/catalogue/custom/viewMetadataDetails.page?uuid=%7B8FDF54D2-654C-4822-8295-1D8E8E772373%7D. Queensland Department of Environment, Science and Innovation (2024b) Riversleigh World Heritage Area, Boodjamulla National Park (Aboriginal Land). Queensland Government. Viewed 10 March 2024, Hyperlink to: Riversleigh World Heritage Area, Boodjamulla National Park (Aboriginal Land . Queensland Department of Regional Development, Manufacturing and Water (2021) North West Queensland. Queensland Government. Viewed 18 February 2024, Hyperlink to: North West Queensland . Queensland Department of Regional Development, Manufacturing and Water (2023) Spatial dataset of the location and attributes for water licences and permits across Queensland. Department of Regional Development, Manufacturing and Water. State of Queensland 2021. https://qldspatial.information.qld.gov.au/catalogue/custom/viewMetadataDetails.page?uuid=%7B26BC4FC6-C8A1-4C5D-A79D-7EC0C1D2494F%7D. Queensland Department of Resources (2021) Mineral and coal exploration guideline. Viewed 23 February 2024, Hyperlink to: Mineral and coal exploration guideline . Queensland Government (2007) Water Plan (Gulf) 2007. Office of the Queensland Parliamentary Counsel, State of Queensland, Queensland, Australia. Viewed 11 Nov 2023, Hyperlink to: Water Plan (Gulf) 2007 . Queensland Government (2016) Water Regulation 2016. Office of the Queensland Parliamentary Counsel, State of Queensland, Queensland, Australia. Viewed 15 June 2024, Hyperlink to: Water Regulation 2016 . Queensland Government (2021a) Electricity supply options for the North West Minerals Province: consultation regulatory impact statement. Queensland Government, Brisbane. Queensland Government (2021b) Land use mapping – 1999 to 2019 – Queensland. Land use dataset downloaded 2 May 2023, https://www.qld.gov.au/environment/land/management/mapping/statewide- monitoring/qlump/qlump-datasets. Queensland Government Open Data Portal (2024a) Queensland borehole series. Viewed 25 February 2024, Hyperlink to: Queensland borehole series . Queensland Government Open Data Portal (2024b) Queensland mining and exploration tenure series. Viewed 10 January 2024, Hyperlink to: Queensland mining and exploration tenure series . Queensland Parks and Wildlife Service (2022) Boodjamulla (Lawn Hill) National Park management plan 2022. Queensland Department of Environment and Science, Brisbane. Queensland Treasury (2023) About the Queensland economy. Queensland Government. Viewed 15 July 2023, Hyperlink to: About the Queensland economy . Raiber M, Taylor AR, Dupuy M, Priestley S, Crosbie RS, Knapton A, Hodgson G and Barry K (2024) Characterising groundwater resources of the Gilbert River Formation, Camooweal Dolostone and Thorntonia Limestone in the Southern Gulf catchments, Queensland and Northern Territory. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Ramsar Convention Secretariat (2004) The Ramsar Convention manual: a guide to the convention on wetlands (Ramsar, Iran, 1971) (3rd ed.) Ramsar Convention Secretariat, Gland, Switzerland. Regional Development Australia (n.d.) Tourism industry. Viewed 10 March 2024, Hyperlink to: Publications . Roberts, A. (2005). Frontier justice. A history of Gulf Country to 1900. Brisbane: University of Queensland Press. Roberts T. (2009) The brutal truth: what happened in the Gulf Country. The Monthly Nov. 2009, 42–51. Rocha AR, Ramos JA, Paredes T and Masero JA (2017) Coastal saltpans as foraging grounds for migrating shorebirds: an experimentally drained fish pond in Portugal. Hydrobiologia 790(1), 141–155. Hyperlink to: https://doi.org/10.1007/s10750-016-3025-y . Rolfe J and Prayaga P (2007) Estimating values for recreational fishing at freshwater dams in Queensland. The Australian Journal of Agricultural and Resource Economics 51(2), 157–174. Hyperlink to: Estimating values for recreational fishing at freshwater dams in Queensland . Rolfe JW, Larard AE, English BH, Hegarty ES, McGrath TB, Gobius NR, De Faveri J, Srhoj JR, Digby MJ and Musgrove RJ (2016) Rangeland profitability in the northern Gulf region of Queensland: understanding beef business complexity and the subsequent impact on land resource management and environmental outcomes. The Rangeland Journal 38, 261–272. Hyperlink to: Rangeland profitability in the northern Gulf region of Queensland . Rose D, Bird (2004) Fresh water rights and biophilia: Indigenous Australian perspectives. Dialogue 23(3), 35–43. 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. Hyperlink to: Use by juvenile barramundi, Lates calcarifer (Bloch), and other fishes of temporary supralittoral habitats in a tropical estuary in northern Australia . SAMREC (2016) The South African code for the reporting of exploration results, mineral resources and mineral reserves (The SAMREC code). Academy of Science of South Africa, South Africa. Scambary B (2013) My country, mine country: Indigenous People, mining and development contestation in remote Australia. Centre for Aboriginal Economic Policy Research, Australian National University, 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. Hyperlink to: Biota connect aquatic habitats throughout freshwater ecosystem mosaics . Short AD (2020) Australian coastal systems: beaches, barriers and sediment compartments. Springer Cham. Hyperlink to: Australian coastal systems: beaches, barriers and sediment compartments . Short AD (2022) Australian beach systems: are they at risk to climate change? Ocean and Coastal Management 224, 106180. Hyperlink to: Australian beach systems: are they at risk to climate change? . Sibanye Stillwater (2023) Mineral resources and mineral reserves report for the year ended 2022. Viewed 24 February 2024, Hyperlink to: Mineral resources and mineral reserves report for the year ended 2022 . Southern Gulf Natural Resource Management (2016) Southern Gulf Regional Natural Resource Management plan. Viewed 18 March 2024, https://www.southerngulf.com.au/resources/nrm-plan/. Speed R and Vanderbyl T (2024) Regulatory requirements for land and water development in the Northern Territory and Queensland. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Steven AH, Dylewski M and Curtotti R (2021) Australian fisheries and aquaculture statistics 2020. Fisheries Research and Development Corporation, Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra. Hyperlink to: Australian fisheries and aquaculture statistics 2020 . Stevens J, McAuley R, Simpfendorfer C and Pillans R (2009) Spatial distribution and habitat utilisation of sawfish (Pristis spp) in relation to fishing in northern Australia. A report to Department of the Environment, Water, Heritage and the Arts, CSIRO. 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. Strang V, (1997) Uncommon ground: cultural landscapes and environmental values. Berg, Oxford, New York. Strang V (2005) Common senses: water, sensory experience and the generation of meaning. Journal of Material Culture 10(1), 92–120. Hyperlink to: https://doi.org/10.1177/1359183505050096 . Summers J, Cavaye J and Woolcock G (2019) Enablers and barriers of tourism as a driver of economic and social-cultural growth in remote Queensland. Economic Papers 38(2), 77–94. Hyperlink to: Enablers and barriers of tourism as a driver of economic and social-cultural growth in remote Queensland . Taçon PS,(2008) Rainbow colour and power among the Waanyi of Northwest Queensland. Cambridge Archaeological Journal 18(2), 163–176. Hyperlink to: Rainbow colour and power among the Waanyi of Northwest Queensland . Tourism Research Australia (2019) Local government area profiles. Viewed 10 March 2024, Hyperlink to: Local government area profiles. . Tourism Tropical North Queensland (2024) Savannah Way resilience toolkit – Reducing the impacts of seasonality on the Savannah Way: an operator guide. Tourism Tropical North Queensland website. Viewed 25 July 2024, Hyperlink to: Destination publications and brochures . Toussaint S, Sullivan P and Yu S (2005) Water ways in Aboriginal Australia: an interconnected analysis. Anthropological Forum 15, 61–74. Hyperlink to: Water ways in Aboriginal Australia: an interconnected analysis . Traill LW, Bradshaw CJA, Delean S and Brook BW (2010) Wetland conservation and sustainable use under global change: a tropical Australian case study using magpie geese. Ecography 33(5), 818-825. Hyperlink to: Wetland conservation and sustainable use under global change: a tropical Australian case study using magpie geese . Traill LW, Bradshaw CJA, Field HE and Brook BW (2009) Climate change enhances the potential impact of infectious disease and harvest on tropical waterfowl. Biotropica 41(4), 414-423. Hyperlink to: Climate change enhances the potential impact of infectious disease and harvest on tropical waterfowl . Trigger DS, (1987) Languages, linguistic groups and status relations at Doomadgee, an Aboriginal settlement in north-west Queensland, Australia. Oceania 57(3), 217–237. Hyperlink to: Languages, linguistic groups and status relations at Doomadgee, an Aboriginal settlement in north-west Queensland, Australia . Trigger DS, (1992). Whitefella comin: Aboriginal responses to colonialism in northern Australia. Cambridge University Press, Cambridge. Trigger DS (1997) Land rights and the reproduction of Aboriginal culture in the Gulf country. Social Analysis: The International Journal of Anthropology, 41(3), 84–106. van Dam RA, Bartolo R and Bayliss P (2008) Chapter 2: Identification of ecological assets, pressures and threats. In: Bartolo R, Bayliss P and Van Dam RA (eds) Ecological risk assessment for Australia’s northern tropical rivers. Sub-project 2 of Australia’s Tropical Rivers – an integrated data assessment and analysis (DET18). A report to Land and Water Australia. Environmental Research Institute of the Supervising Scientist, National Centre for Tropical Wetland Research, Darwin NT, 14–161. 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. 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. Vanderbyl T (2024) Technical report on the Northern Territory’s water planning arrangements. A technical report from the CSIRO Victoria River Water Resource Assessment for the National Water Grid. CSIRO, Australia. 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. Hyperlink to: Global threats to human water security and river biodiversity . Waanyi Native Title Aboriginal Corporation RNTBC (2020) Strategic Plan 2020–2025. Viewed 18 March 2024, https://waanyipbc.org/strategic-plan. Waitt G, Lane R and Head L (2003) The boundaries of nature tourism. Annals of Tourism Research 30(3), 523–545. Hyperlink to: The boundaries of nature tourism . 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. Ward JV and Stanford J (1995) Ecological connectivity in alluvial river ecosystems and its disruption by flow regulation. Regulated rivers: research and management 11(1), 105-119. Hyperlink to: Ecological connectivity in alluvial river ecosystems and its disruption by flow regulation . 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. Hyperlink to: The ‘wet-dry’ in the wet-dry tropics drives river ecosystem structure and processes in northern Australia . Waschka M and Macintosh A (2023) CSIRO Water Resource Assessments: Indigenous rights and interests in Queensland and the Northern Territory. A report from Barraband Consulting to the CSIRO to inform the CSIRO Roper River, Victoria Rivers and Southern Gulf Catchments Water Resource Assessments. CSIRO, Australia. Webster A, Jarvis D, Jalilov S, Philip S, Oliver Y, Watson I, Rhebergen T, Bruce C, Prestwidge D, McFallan S, Curnock M and Stokes C (2024a) Financial and socio-economic viability of irrigated agricultural development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Webster A, Jarvis D, Jalilov S, Philip S, Oliver Y, Watson I, Rhebergen T, Bruce C, Prestwidge D, McFallan S, Curnock M and Stokes C (2024b) Financial and socio-economic viability of irrigated agricultural development in the Victoria catchment, Northern Territory. A technical report from the CSIRO Victoria River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Winter WH (1988) Supplementation of steers grazing Stylosanthes hamata pastures at Katherine, Northern Territory. Australian Journal of Experimental Agriculture 28, 669-682. 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. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra, 69–89. 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. Hyperlink to: Importance of riparian zone: effects of resource availability at land-water interface . Part III Opportunities for water resource development Chapters 4 and 5 provide information on opportunities for agriculture and aquaculture in the catchment of the Southern Gulf rivers, that is Settlement Creek, Gregory–Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groups. This information covers: •opportunities for irrigated agriculture and aquaculture (Chapter 4) •opportunities to extract and/or store water for use (Chapter 5). Lake Moondarra on the Leichhardt River is a favoured recreational reserve for residents and tourists of Mount Isa. Photo: CSIRO – Nathan Dyer 4 Opportunities for agriculture in the Southern Gulf catchments Authors: Yvette Oliver, Seonaid Philip, Tiemen Rhebergen, Ian Watson, Tony Webster, Peter Zund, Simon Irvin Chapter 4 presents information about the opportunities for irrigated agriculture and aquaculture in the catchment of the Southern Gulf rivers, that is Settlement Creek, Gregory–Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groups1, describing: 1 Only those islands greater than 1000 ha are mapped • land suitability for a range of crop group × season × irrigation type combinations and for aquaculture, including key soil-related management considerations • cropping and other agricultural opportunities, including crop yields and water use • gross margins at the farm scale • prospects for integration of forages and crops into existing beef enterprises • aquaculture opportunities. The key components and concepts of Chapter 4 are shown in Figure 4-1. Figure 4-1 Schematic of agriculture and aquaculture enterprises as well as crop and/or forage integration with existing beef enterprises to be considered in the establishment of a greenfield irrigation development For more information on this figure please contact CSIRO on enquiries@csiro.au 4.1 Summary This chapter provides information on land suitability and the potential for agriculture and aquaculture in the Southern Gulf catchments. A mixture of field surveys and desktop analysis were used to generate the results presented in this chapter. For example, the land suitability results draw on extensive field visits (to describe, collect and analyse soils) and are integrated with state- of-the-art digital soil mapping. Many of the results are expressed in terms of potential. The area of land suitable for cropping or aquaculture, for example, is estimated by considering the set of relevant soil and landscape biophysical attributes at each location and determining the most limiting attribute among them. It does not include water availability; cyclone or flood risk; legislative, regulatory or tenure considerations; or ecological, social or economic drivers that will inevitably constrain the actual area of land that is developed. Crops, forages and cropping systems results are based on data analysis and simulation models, and assume good agronomic practices producing optimum yields given the soil and climate attributes in the catchments. Likewise, aquaculture is assessed in terms of potential, using a combination of land suitability and the productive capacity of a range of aquaculture species. Information is presented in a manner to enable the comparison of a variety of agricultural and aquaculture options. The results from individual components (land suitability, agriculture, aquaculture) are integrated to provide a sense of what is potentially viable in the catchments. This includes providing specific information on a wide range of crop types for agronomy, water use and land suitability for different irrigation types; analyses of economic performance, such as crop gross margins (GMs); how more-intensive mixed cropping systems might be feasible with irrigation; and analyses of what is required for different aquaculture development options to be financially viable. 4.1.1 Key findings Any agricultural resource assessment must consider two major factors: how much soil is suitable for a particular land use and where that soil is located. Based on a sample of 14 individual combinations of crop group × season of use × irrigation type, the amount of land classified as moderately suitable with considerable limitations or better ranges from 780,000 ha (Crop Group 7, wet-season furrow) to 4.7 million ha (Crop Group 14, perennial species, spray) before constraints such as water availability, environmental and other legislation and regulations, and a range of biophysical risks are considered (crop groups are defined in Section 4.2.3). The largest contiguous areas of soil suitable for broad-scale irrigation are the grey cracking clay soils of the lowland alluvial plains (Section 2.3.2, Figure 2-5), which are well located for small-scale irrigation developments based on water harvesting. Downstream of Doomadgee there are contiguous areas of red sandy and loamy soils suitable for irrigated vegetables and in the Leichhardt catchment downstream of Kajabbi there are opportunities for irrigated horticulture on the friable levee soils, and the adjacent friable clayey soils are suitable for broadacre irrigation. The clay soils on the Barkly Tableland in the south-west are suitable for broadacre cropping and overlie areas of intermediate- to regional-scale groundwater resources. Rainfed cropping Despite the theoretical possibility that rainfed crops could be produced using the considerable rainfall that arrives during the wet season, in practice significant agronomic and market-related challenges to rainfed crop production have prevented its expansion. Extensive areas of heavier clay soils (soil generic group (SGG) 9) across the Armraynald Plain and Barkly Tableland store enough plant available water (PAW) that could support potential high crop yields, particularly if cropped opportunistically in wetter years. However, frequent inundation and waterlogging of clay soils means that access for farming operations could be disrupted, increasing the risk to maximum yields through compromised timing of operations. Despite these challenges, higher-value crops such as pulses or cotton show potential, especially when grown in conjunction with irrigated farming. Loamy soils have low water-holding capacity and are hardsetting, which makes consistently achieving viable rainfed yields difficult. Irrigated cropping Irrigation reduces crop water stress and provides greater control over scheduling of crop operations to optimise production, including the option of growing through the cooler months of the dry season. Analyses of the performance of 19 potential irrigated cropping options in the Southern Gulf catchments indicate that achievable annual GMs could be up to about $4500/ha for broadacre crops, $8000/ha for annual row crop horticulture, $6000/ha for perennial fruit tree horticulture and $3000/ha for silviculture (plantation trees). While GMs are a key partial metric of farm performance, they should not be treated as fixed constants determined by the cropping system alone. They are a product of the farming and business management decisions, input costs and market opportunities. As such there are often niche opportunities to improve farm GMs and profitability, but these usually come at the expense of scalability. Farm financial metrics like GMs greatly amplify any fluctuations in commodity prices and input costs, so the mean GM does not accurately reflect the often substantial cashflow challenges in managing years of losses between those of windfall profits (particularly for horticulture). Crop yields and GMs presented in this chapter indicate what might be attained for each cropping option once it has achieved it’s sustainable agronomic potential. It is unrealistic to assume that these levels of performance would be achieved in the early years of newly established farms, and allowance should be made for an initial period of learning (see Chapter 6). Potential crop species that could be grown as a single crop per year were rated and ranked for their performance in the Southern Gulf catchments. Wet-season crops (planted January to early May) that are rated the most likely to be viable are cotton (Gossypium spp.), forages and peanuts (Arachis hypogaea). Dry-season crops (planted late March to August) that are rated the most likely to be viable are annual horticulture and cotton. Financial viability is determined both by crop options with the highest GMs and by associated capital and fixed costs, which are higher in more- intensive farming like horticulture. The farm-scale measures of crop performance presented in this chapter are intended to be used in conjunction with the scheme-scale analyses of financial viability in Chapter 6 (as part of an integrated multi-scale approach). Sequential cropping systems involve planting more than one crop in the same year in the same field. These systems have the potential to significantly increase farm GMs. Annual broadacre and horticultural crops have been grown sequentially for many decades in tropical northern Australia. A wide range of sequential cropping options are potentially viable in the Southern Gulf catchments. Most suitable crop sequences include wet-season mungbean, grain sorghum or peanut with dry-season annual horticulture, wet season mungbean, peanut, soybean or grain sorghum with dry-season cotton, maize, chickpea or forage, and wet-season cotton with dry- season mungbean, sorghum or forage. Scheduling back-to-back crops could be operationally tight in the Southern Gulf catchments, particularly on clay-rich soils with poor drainage, due to limitations on paddock accessibility. Crop selection is market driven in northern Australian regions like the Southern Gulf catchments. Rotations and crop sequences are therefore dynamic as growers develop an understanding of the benefits, trade-offs and management needs of different crop mixes and adapt to changing opportunities as commodity prices change. Integrating forages and hay into existing beef enterprises There are many theoretical benefits to growing irrigated forages and hay on-farm to enhance existing grazing enterprises. The use of on-farm irrigated forage and hay production would allow graziers greater options for marketing cattle: meeting market liveweight specifications for cattle at a younger age, meeting the specifications required for different markets than those typically targeted by cattle enterprises in the Southern Gulf catchments and providing cattle that meet market specification at a different time of the year. Forages and hay may also allow graziers to implement management strategies, such as early weaning or weaner feeding, which should lead to flow-on benefits throughout the herd, including increased reproductive rates. Some of these strategies are already practised within the Southern Gulf catchments but in almost all incidences are reliant on hay or other supplements purchased on the open market. By growing hay on-farm, the scale of these management interventions might be increased, at reduced net cost. Furthermore, the addition of irrigated feeds may allow graziers to increase the total number of cattle that can be sustainably carried on a property. Analysis of two irrigated hay or two irrigated forage stand-and-graze options compared to two base enterprises (with or without purchased hay, for weaners) suggested that irrigated forages or hay increased the total income and the amount of cattle liveweight sold. GMs were highest for the two irrigated hay options. The two stand-and-graze options returned the lowest GMs. A net present value (NPV) analysis suggested that a decision to irrigate would need to assume that beef prices remain high in comparison to the mean of the previous 10 years. Irrigation enterprises of the scale required involve high capital investment and additional or novel management skills. Aquaculture There are considerable opportunities for aquaculture development in northern Australia given the region’s natural advantages of a climate suited to farming valuable tropical species, large areas identified as suitable for aquaculture, and political stability and proximity to large global markets. The main challenges to developing and operating modern and sustainable aquaculture enterprises are regulatory barriers, global cost competitiveness and the remoteness of much of the suitable land area. The three species with the most aquaculture potential in the Southern Gulf catchments are black tiger prawns (Penaeus monodon), barramundi (Lates calcarifer) and red claw (Cherax quadricarinatus). Suitable land for lined ponds for freshwater species is widespread throughout the catchments due to the extensive distribution of favourable soil and land characteristics (flat land, non-rocky, deep soil). In contrast, options for freshwater species in earthen ponds are restricted to the impermeable alluvial clays to allow retention of water. The range for marine aquaculture is restricted to the tidal zones of the catchments and on the coastal plain within 2000 m of access to marine water. High annual operating costs (which can exceed the initial capital costs of development) mean that managing cashflow in the establishment years is challenging, especially for products that require multi-year grow-out periods. Input costs scale with increasing productivity, so improving production efficiency (such as feed conversion rate or labour-efficient operations) is much more important than increasing yields for aquaculture to be viable in the Southern Gulf catchments. It would be essential for any new aquaculture development to refine the production system and achieve the required levels of operational efficiency (input costs per kilogram of produce) using just a few ponds before scaling the enterprise to a larger number of ponds. 4.1.2 Introduction Aspirations to expand agricultural development in the Southern Gulf catchments are not new, and across northern Australia there have been a number of initiatives to put in place large-scale agricultural developments since World War II (Ash, 2014; Ash and Watson, 2018). Ash and Watson (2018) assessed 11 such agricultural developments, four of which continue to operate at a regionally relevant scale, namely the Ord River Irrigation Area, the lower Burdekin, the Mareeba– Dimbulah Water Supply Scheme and the Katherine mango industry. The Lakeland Downs development also continues, although it could not be categorised as regionally significant. Ash and Watson’s assessment included both irrigated and rainfed developments, and considered natural, human, physical, financial and social capitals. Key points to emerge from these analyses include the following: • The natural environment (climate, soils, pests and diseases) makes agriculture in northern Australia challenging, but these inherent environmental factors are not generally the primary reason for a lack of success. • The speed with which many of the developments were undertaken did not allow for a ‘learning by doing’ approach, leading at times to costly mistakes. • Physical capital, in the form of on-farm infrastructure, supply chain infrastructure and crop varieties, was a significant and ongoing impediment to success. For broadacre commodities that require processing facilities, these facilities need to be within a reasonable distance of production sites and at a scale to make them viable in the long term. • Financial plans tended to over estimate early production and returns on capital, and make overly optimistic expectations of the ability to scale up rapidly. This led to financial pressure on investors and a premature end to some developments. Furthermore, the need to have well- connected and well-paying markets was often not fully appreciated. In more remote regions, higher-value products such as fruit, vegetables and niche crops proved more successful, although high supply chain costs to both domestic and export markets remain as impediments to expansion. • Most of the developments began in areas with no history of agricultural development, and there was no significant community of practitioners who could share experiences. • Management, planning and finances were the most important factors in determining the ongoing viability of agricultural developments. For developments to be successful, all factors relating to climate, soils, agronomy, pests, farm operations, management, planning, supply chains and markets need to be thought through in a comprehensive systems design. Particular attention needs to be paid to scaling up at a considered pace and being prepared for reasonable lags before achieving positive returns on investment. This chapter addresses the following questions for the Southern Gulf catchments: • How much land is suitable for cropping and in which suitability class? • Is irrigated cropping economically viable? • Which crop options perform best and how can they be implemented in viable mixed farming systems? • Can crops and forages be economically integrated with beef enterprises? • What aquaculture production systems might be possible? The chapter is structured as follows: • Section 4.2 describes how the land suitability classes are derived from the attributes provided in Chapter 2, with results given for a set of 14 combinations of individual crop group × season × irrigation type. Versatile agricultural land is described, and a qualitative evaluation of cropping is provided for a set of specific locations within the catchment. • Section 4.3 provides detailed information on crop and forage opportunities, including irrigated crop yields, water use and GMs. Agronomic principles, such as selection of sowing time, are provided, including a cropping calendar for scheduling farm operations. The information is synthesised in an analysis of the cropping systems that could best take advantage of opportunities in the Southern Gulf catchments environments while dealing with farming challenges. • Section 4.4 provides synopses for 11 crop and forage groups, including a focus on specific example species. • Section 4.5 discusses the candidate species and likely production systems for aquaculture enterprises, including the prospects for integrating aquaculture with agriculture. 4.2 Land suitability assessment 4.2.1 Introduction The term ‘suitability’ in the Assessment refers to the potential of the land for a specific land use, such as furrow-irrigated cotton. The term ‘capability’ (not used in the Assessment) refers to the potential of the land for broadly defined land uses, such as cropping or pastoral (DSITI and DNRM, 2015). The overall suitability for a particular land use is determined by a number of environmental and soil attributes. These include, but are not limited to, climate at a given location, slope, drainage, permeability, available water capacity (AWC) of the soil, pH, soil depth, surface condition and texture. Examples of some of these attributes are provided in Section 2.3. From these attributes, a set of limitations to suitability are derived, which are then considered against each potential land use. 4.2.2 Land suitability classes The overall suitability for a particular land use is calculated by considering the set of relevant attributes at each location and determining the most limiting attribute among them. This most limiting attribute then determines the overall land suitability classification. The classification is on a scale of 1 to 5 from ‘Suitable with negligible limitations’ (Class 1) to ‘Unsuitable with extreme limitations’ (Class 5), as shown in Table 4-1 (FAO, 1976, 1985). The companion technical report on digital soil mapping and land suitability (Thomas et al., 2024) provides a complete description of the land suitability assessment method, and the material presented in this section is taken from that report. Note that the land suitability maps and figures presented in this section do not consider flooding, risk of secondary salinisation or availability of water as discussed by Thomas et al. (2024). Consideration of these risks and others, along with further detailed soil physical, chemical and nutrient analyses, would be required to plan development at scheme, enterprise or property scale. Caution should therefore be employed when using these data and maps at fine scales. Table 4-1 Land suitability classes based on FAO (1976, 1985) as used in the Assessment For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 4.2.3 Land suitability for crops, versatile agricultural land and evaluation of specific areas of interest The suitability framework used in this Assessment aggregates individual crops into a set of 21 crop groups (Table 4-2). The groups are based on the framework used by Andrews and Burgess (2021), with some additions considered prospective based on previous CSIRO work in northern Australia (e.g. Thomas et al., 2018), including in Queensland. From this set of crop groups, land suitability has been determined for 58 land use combinations of crop group × season × irrigation type (including rainfed) (Thomas et al., 2024). Table 4-2 Crop groups and individual land uses evaluated for irrigation (and rainfed) potential Crop groups and land uses are based on those used by Andrews and Burgess (2021), amended for the Southern Gulf catchments with the addition of crop groups 18 to 21 based on CSIRO’s previous work in northern Australia. Those used in the Northern Australia Water Resource Assessment (Thomas et al., 2018) are in boldface. MAJOR CROP GROUP CROP GROUP INDIVIDUAL CROPS ASSESSED Tree crops/horticulture (fruit) 1 Monsoonal tropical tree crops (0.5 m root zone) – mango, coconut, dragon fruit, Kakadu plum, bamboo, lychee 2 Tropical citrus – lime, lemon, mandarin, pomelo, lemonade, grapefruit Intensive horticulture (vegetables, row crops) 3 Cucurbits – watermelon, honeydew melon, rockmelon, pumpkin, cucumber, Asian melons, zucchini, squash 4 Fruiting vegetable crops – Solanaceae (capsicum, chilli, eggplant, tomato), okra, snake bean, drumstick tree 5 Leafy vegetables and herbs – kangkong, amaranth, Chinese cabbage, bok choy, pak choy, choy sum, basil, coriander, dill, mint, spearmint, chives, oregano, lemon grass, asparagus Root crops 6 Carrot, onion, sweet potato, shallots, ginger, turmeric, galangal, yam bean, taro, peanut, cassava Grain and fibre crops 7 Cotton, grains – sorghum (grain), maize, millet (forage) 8 Rice (lowland and upland) Small-seeded crops 9 Hemp, chia, quinoa, medicinal poppy Pulse crops (food legumes) 10 Mungbean, soybean, chickpea, navy bean, lentil, guar Industrial 11 Sugarcane Hay and forage (annual) 12 Annual grass hay/forages – sorghum (forage), maize (silage) 13 Legume hay/forages – blue pea, burgundy bean, cowpea, lablab, Cavalcade, forage soybean Hay and forage (perennial) 14 Perennial grass hay/forage – Rhodes grass, panics Silviculture/forestry (plantation) 15 Indian sandalwood 16 African mahogany, Eucalyptus spp., Acacia spp. 17 Teak Intensive horticulture (vegetables, row crops) 18 Sweet corn MAJOR CROP GROUP CROP GROUP INDIVIDUAL CROPS ASSESSED Oilseeds 19 Sunflower, sesame Tree crops/horticulture 20 Banana, coffee 21 Cashew, macadamia, papaya A sample of 14 of these individual land use combinations – that covers a mixture of crops, irrigation types and seasons, grown or trialled in northern Australia – is shown in Figure 4-2. Depending on land use, the amount of land classified as Class 3 or better for these sample land uses ranges from about 360,000 ha (Crop Group 10, wet-season rainfed) to 5.1 million ha (Crop Group 14 under spray irrigation). Much of this land is rated as Class 3, and so has considerable limitations, although nearly 1.7 million ha of Class 2 land is available for Crop Group 14 crops under spray irrigation and between about 340,000 ha and about 970,000 ha of Class 2 land for the other crop groups under spray or trickle irrigation. Ranges of suitability geographic distributions are shown on maps in the crop synopses in Section 4.4. Figure 4-2 Area (ha) of the Southern Gulf catchments mapped in each of the land suitability classes for 14 selected land use combinations (crop group × season × irrigation type) The five land suitability classes are described in Table 4-1 and more detail on the crop groups is given in Table 4-2. Land suitability classes mapped for southern gulf For more information on this figure please contact CSIRO on enquiries@csiro.au In order to provide an aggregated summary of the land suitability products, an index of agricultural versatility was derived for the Southern Gulf catchments (Figure 4-3). Versatile agricultural land was calculated by identifying where the highest number of the 14 selected land use options presented in Figure 4-2 were mapped as being suitable (i.e. suitability classes 1 to 3). Qualitative observations on each of the areas mapped as ‘A’ to ‘F’ in Figure 4-3 are provided in Table 4-3. Figure 4-3 Agricultural versatility index map for the Southern Gulf catchments High index values denote land that is likely to be suitable for more of the 14 selected land use options. The map shows specific areas of interest (A to F) from a land suitability perspective, which are discussed in Table 4-3. Note that the versality index mapped here does not consider flooding, risk of secondary salinisation or availability of water. Versatile agricultural land map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-511_AgVers14_PotDev_v1_10_8.png For more information on this figure please contact CSIRO on enquiries@csiro.au Table 4-3 Qualitative land evaluation observations in Southern Gulf catchments areas A to F shown in Figure 4-3 Further information on each soil generic group (SGG) and a map showing spatial distribution can be found in Section 2.3. AREA SOIL AND LOCATION SOIL DESCRIPTION, POTENTIAL LAND USES AND LIMITATIONS A Cracking clay soils (SGG 9) of the Armraynald Plain formed from the Pleistocene Armraynald Beds and broad alluvial plains of the major rivers, particularly the Gregory and lower Leichhardt rivers Comprises rarely flooded plains between the Gregory and Leichhardt rivers and regularly flooded plains on Lawn Hill Creek and between Gregory River and Running Creek. Soils are mainly moderately well-drained to imperfectly drained grey, brown or black cracking clay soils (SGG 9) with self-mulching to hard- setting structured surfaces. The imperfectly drained grey clay soils occur on the northern part of the Armraynald Plain and along Lawn Hill Creek. The brown cracking clays that occur on the south-eastern part of the plain east of the Leichhardt River can also be imperfectly drained. The soils may be suitable for furrow- or spray-irrigated vegetables (except root crops), rice, sugarcane and dry-season grain, forage, pulse crops, sweet corn and cotton. The main limitations are workability and potential wet-season flooding. Management of wet-season cropping needs to consider crop tolerance to seasonal wetness and flood duration, depth and frequency. How soil salinity will accumulate over time in these soils is currently unknown but must be monitored, especially in the imperfectly drained soils. B Friable non-cracking clays or sandy clay loams (SGG 2) and loam over relatively friable red clay subsoils (SGG 1.1) formed on alluvium along the middle reaches of the Leichhardt River The friable non-cracking clays (SGG 2) are moderately well to well-drained, brown, red or grey, structured sandy clay loam or silty clay soils. The loam over red clay subsoils (SGG 1.1) are well drained with moderately thick (<0.2 m), loamy surface soils over red, structured clay subsoils developed on alluvium adjacent to the Leichhardt River. Soils are suitable for a range of spray- or trickle-irrigated vegetables, sugarcane, oilseed, sweet corn and wet-season and dry-season grain, forage, pulse crops and cotton. Wet-season tree crops are also likely to be suitable. Extents of suitable lands are generally minor, resulting in small and/or narrow areas limiting paddock size and irrigation infrastructure layout. The main limitation is flooding post-cyclone. Soil erosion during flood events and compaction from tillage are land degradation risks. C Red loamy soils (SGG 4.1) and red sandy soils (SGG 6.1) occurring near Doomadgee on elevated narrow alluvial plains along the Nicholson River The red loamy soils (SGG 4.1) occur downstream of Doomadgee on the southern side of the Nicholson River on a Pleistocene elevated floodplain. Soils are well-drained brown, silty loam moderately thick (<0.2 m) surface soils over red silty clay subsoils. Soils may be suitable for irrigated agriculture although the narrow tracts along the Nicholson River may limit infrastructure layout. Compaction from tillage is a land degradation risk. On the northern side of the Nicholson River, red sandy soils (SGG 6.1) have developed on an elevated alluvial plain near Doomadgee. Soils are well-drained red sand to sandy loams and have limited very low to low soil AWC, hence are only suited to irrigated horticulture using trickle or drip systems. There is a risk of deep drainage and nutrification of the adjacent river and groundwater table. D Grey cracking clays (SGG 9) from Cenozoic sediments on the Barkly Tableland Soils are self-mulching, grey or occasionally brown, cracking clays (SGG 9), moderately deep (>1.0 m) to very deep (>1.5 m) and moderately well drained. Localised rockiness/stoniness in the soil profile may affect farming. Surfaces are gilgaied, and soils are formed of structured clay with calcareous nodules and gypsum crystals. Soils are suitable for trickle-irrigated mangoes and vegetables as well as wet-season cotton, grain and forage crops. Soil workability and rockiness are the main limitations, and deep gilgai microrelief may restrict land-levelling operations in some areas. There is a risk of water erosion on bare paddocks late in the dry season due to early rains. E Sandy soils (SGG 6.2) formed in sandy sediments on old lateritic surfaces of the Doomadgee Plains Soils are brown, yellow or grey and sandy (SGG 6.2), highly permeable, well- drained, deep to very deep (1–1.5 m), commonly encountering ferricrete rock within 1 m. There is potential for irrigated horticulture using trickle or drip systems. In the absence of irrigation, agricultural potential of these soils is low. Soil depth and water-holding capacity are the main limitations. F Deep sandy soils (SGG 6) formed on an elevated sand plain in the Buddycurrawa Creek subcatchment of the Gulf Fall physiographic unit Sandy soils are brown or red (SGG 6), highly permeable, moderately well to well-drained, and deep to very deep (>1.2 m). At depth these soils may be mottled and with no coarse fragments or nodules. The soil has a low AWC (<60 mm). There is potential for irrigated horticulture using trickle or drip systems. Agricultural potential of these soils is low without irrigation. Land suitability and its implications for crop management are discussed in more detail for a selection of crops in Section 4.4, where land use suitability of a given crop and irrigation combination are mapped, along with information critical to the consideration of the crop in an irrigated farm enterprise. Land suitability maps for all 58 land use combinations are presented in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). 4.3 Crop and forage opportunities in the Southern Gulf catchments 4.3.1 Introduction This section presents results on the farm ‘performance’ of individual crop options, where performance is quantified specifically as crop yields, the amount of applied irrigation water (accounting for application efficiency) and GMs. Performance is presented with information on agronomic principles and farming practices to help interpret the viability of new (greenfield) farming opportunities in the Southern Gulf catchments. The individual crop options are grouped into rainfed broadacre, irrigated broadacre, irrigated horticulture and plantation tree crops (sections 4.3.3 to 4.3.7), and viability is discussed in a section on cropping systems (Section 4.3.8). That section considers the mix of farming opportunities and practices, for both single and sequential cropping systems, with the greatest potential to be profitably and sustainably integrated within the Southern Gulf catchments environments. Finally, Section 4.3.9 evaluates the viability of integrating irrigated forages into existing beef production. These farm-scale analyses are intended to be used in conjunction with the scheme-scale analyses of viability in Chapter 6 (as part of an integrated multi-scale analysis). Nineteen irrigated crop options were selected to evaluate their potential performance in the Southern Gulf catchments (Table 4-4). The crops were selected to be compatible with the land suitability crop groups (Table 4-2), provided that: (i) they had the potential to be viable in the Southern Gulf catchments (based on knowledge of how well these crops grow in other parts of Australia), (ii) they were of commercial interest for possible development in the region and (iii) there was sufficient information on their agronomy, and farming costs and prices, for quantitative analysis. The analyses used a combination of Agricultural Production Systems sIMulator (APSIM) crop modelling and climate-informed extrapolation to estimate potential yield and water use for each crop. Those values were then used in a farm GM tool specifically designed for greenfield farming developments (like those in the Southern Gulf catchments, where there are very few existing commercial farms or farm financial models). In particular, extrapolations used close similarities in climate and soils between possible cropping locations in the Southern Gulf catchments and established irrigated cropping regions at similar latitudes near the Ord River Irrigation Area (WA) and the Mareeba–Dimbulah Irrigation Area (Queensland) (Figure 4-4). Full details of the approach are described in the companion technical report on agricultural viability and socio-economics (Webster et al., 2024). Section 4.4 provides further details on opportunities and constraints in the Southern Gulf catchments, for example, crops in each of the agronomic crop types listed in Table 4-4. Table 4-4 Crop options for which performance was evaluated in terms of water use, yields and gross margins The methods used for estimating crop yield and irrigation water requirements are coded as: A = APSIM; E = climate- informed extrapolation. ‘A, E’ indicates that A is the primary method and E is used for sensibility testing. ‘E, A’ indicates that E is the primary method and A is used for applying adjustments. ‘Mango (KP)’ is Kensington Pride and ‘Mango (PVR)’ is an indicative new high-yielding variety likely to have plant variety rights (e.g. Calypso). Note that crops that are agronomically similar in terms of the commodities they produce (as categorised in the table) may differ in how they respond to soil constraints. The crop type categories in the table are therefore necessarily different to the crop groups used in the land suitability section (which are grouped according to shared soil requirements and constraints; Table 4-2). CROP TYPE CROP IRRIGATION WATER ESTIMATE METHOD YIELD ESTIMATE METHOD Broadacre Cereal Sorghum (grain) A, E A, E Maize A, E A, E Pulse Mungbean A, E A, E Chickpea A, E A, E Soybean A, E A, E Oilseed Sesame E E Peanut A, E A, E Industrial Cotton (dry season) A, E A, E Cotton (wet season) A, E A, E Hemp E E Forage Rhodes grass A, E A, E Horticulture (row) Rockmelon E E Watermelon E E Onion E E Capsicum E E Horticulture (tree) Mango (PVR) E E Mango (KP) E E Lime E E Plantation tree African mahogany E E (a) Mean monthly rainfall (b) Mean daily maximum temperature (c) Mean daily solar radiation (d) Mean daily minimum temperature Monthly daily solar radiation comparisons \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\1_Climate\SoWRA_climate analysis_v2.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Monthly min temp comparisons \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\1_Climate\SoWRA_climate analysis_v2.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au A graph of the month of the year Description automatically generated Figure 4-4 Climate comparisons of Southern Gulf catchments’ sites with established irrigation areas at Kununurra (WA) and Mareeba (Queensland) Southern Gulf catchments sites are Westmoreland, Gregory, Kamilaroi and Gallipoli. Four locations were selected for the APSIM simulations to represent some of the best potential farming conditions across the varied environments in the Southern Gulf catchments: • A Vertosol with a Gregory (–18.65°S, 139.25°E) climate. This soil represents the farming conditions of the lowland cracking clays (SGG 9; marked ‘A’ in Figure 4-3) and are the most extensive arable areas in the Southern Gulf catchments. During the wet-season, access and limitations from floodplain inundation and workability may constrain cropping. Using grain sorghum as an indicator crop, the plant available water capacity (PAWC) of the modelled soil was 212 mm (noting that PAWC differs between crops with different rooting patterns and physiologies). Daily historical meteorological data used for these simulations was from the Gregory weather station, which has a mean annual rainfall of about 540 mm. • A Chromosol with a Kamilaroi (–19.36°S, 140.04°E) climate. This soil represents some of the better farming conditions among the friable non-cracking clay soils (SGG 1 and Dermosols, SGG 2; marked ‘B’ in Figure 4-3) along the middle reaches of the Leichhardt River. The PAWC of this soil for grain sorghum was 93 mm, and the mean annual rainfall for Kamilaroi is about 577 mm. Monthly rainfall comparisons \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\1_Climate\SoWRA_climate analysis_v2.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Monthly max temperature comparisons \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\1_Climate\SoWRA_climate analysis_v2.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au • A red Kandosol with a Westmoreland (–17.34°S, 138.25°E) climate. This soil represents some of the better farming conditions among the loamy soils (SGG 4; marked ‘C’ in • A Vertosol with a Gallipoli (–19.14°S, 137.87°E) climate. This soil represents some of the better farming conditions among the cracking clay soils (SGG 9; marked ‘D’ in Figure 4-3), having less wet-season issues than the Gregory (lowland) cracking clay soils, although surface and profile rock may limit some areas. The PAWC of this soil for grain sorghum was 146 mm, and the mean annual rainfall for Gallipoli is about 420 mm. To assist with interpreting the later results, some information is first provided on agronomic principles related to the scheduling of critical farm operations such as sowing and irrigation in relation to Southern Gulf catchments environments. 4.3.2 Cropping calendar and time of sowing Time of sowing can have a significant effect on achieving economical crop and forage yields, and on the availability and amount of water for irrigation required to meet crop demand. Cropping calendars identify optimum sowing times of different crops and are essential tools for scheduling farm operations (Figure 4-5) so that crops can be reliably and profitably grown. No cropping calendar existed for the Southern Gulf catchments before the Assessment. Sowing windows vary in both timing and length among crops and regions, and they consider the likely suitability and constraints of weather conditions (e.g. heat and cold stress, radiation, and conditions for flowering, pollination and fruit development) during each subsequent growth stage of the crop. Limited field experience currently exists in the Southern Gulf catchments for most of the crops and forages evaluated. This cropping calendar (Figure 4-5) is therefore based on knowledge of crops derived from past and current agricultural experience in the Ord River Irrigation Area (WA), Katherine and Douglas–Daly regions (NT), Mareeba–Dimbulah Water Supply Scheme and the Burdekin region (Queensland). Some annual crops have both wet-season and dry-season cropping options. Perennial crops are grown throughout the year, so growing seasons and planting windows are less well defined. Generally, perennial tree crops are transplanted as small plants, and in northern Australia this is usually timed towards the beginning of the wet season to take advantage of wet-season rainfall. The cropping calendar presented here considers the optimal climate conditions for crop growth and considers operational constraints specific to the local area. Such constraints include wet- season difficulties in access and trafficability, and limitations on the number of hectares that available farm equipment can sow/plant. For example, clay-rich alluvial Vertosols, such as those found across the Armraynald and Cloncurry plains and Barkly Tableland, are likely to present severe trafficability constraints through much of the wet season in the Southern Gulf catchments, while sandier Kandosols would present far fewer trafficability restrictions in scheduling farming operations (Figure 4-6). Figure 4-5 Annual cropping calendar for irrigated agricultural options in the Southern Gulf catchments WS = wet season; DS = dry season. Many suitable annual crops can be grown at any time of the year with irrigation in the Southern Gulf catchments. Optimising crop yield alone is not the only consideration. Ultimately, sowing date selection must balance the need for the best growing environment (optimising solar radiation and temperature) with water availability, pest avoidance, trafficability during the growing season and at harvest, crop rotation, supply chain requirements, infrastructure development costs, market access considerations and potential commodity price. Many summer crops from temperate regions are suited to the tropical dry season (winter) because temperatures are closer to their Crop planting times \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\2_Crops\Cropping_Windows.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au CROP TYPECROPDECJANFEBMARAPRMAYJUNJULAUGSEPOCTNOVCROP DURATION(days) Cereal cropsSorghum (WS)ssssssgggg110—140Sorghum (DS)gssssssssssggg110—140Maize (WS)ssssssggg110—140Maize (DS)gssssssssssggg110—140Rice (WS)ssssgggg120—160+ Rice (DS)ssssgggg90—135 Pulse crops (food legumes)Mungbean (WS)ssssggg70—85Mungbean (DS)ssssggg70—85Chickpeassssgggg100—120OilseedsSoybean (WS)ssssssgggg110—130Sesamessssssggg110—130Root cropsPeanut (WS)ssssssggggg100—140Peanut (DS)gssssggg100—140Cassavassssssssssssssggggg180—210Industrial cropsCotton (WS)ssssssgggg100—120Cotton (DS)ssssssggggg100—120Hemp (fibre)ssssssssgggg110—150Forage, hay, silageRhodes grassggspspspgggspspspspPerennial (regrows) Forage sorghumssssssssgggssssssgg60—80 (regrows) Forage milletssssssssgggssssssgg60—80 (regrows) Forage maizegssssssgggssssssgg75—90Forage legumesCavalcadessggggggssss150—180Lablabssssssssssggggg130—160Horticulture (row crops)Melonsssssssgggg70—110Oniongssssssssssgggg130—160Capsicum, chilli, tomatossssggggg70—90 from transplantPineapplespspspgggggggPerennialHorticulture (vine)Table grapesspspspgggggggggPerenialHorticulture (tree crops)MangospspspgggggggggPerennialAvocadospspspgggggggggPerennialBananaspspspspggggggggPerennialLimespspspgggggggggPerennialLemonspspspgggggggggPerennialOrangespspspgggggggggPerennialCashewspspspgggggggggPerennialMacadamiaspspspgggggggggPerennialPlantation trees (silviculture)Africian mahoganyspspspgggggggggPerennialIndian sandalwoodspspspgggggggggPerennial optima and/or there is more consistent solar radiation (e.g. maize (Zea mays), chickpea (Cicer arietinum) and rice (Oryza sativa)). For sequential cropping systems (which grow more than a single crop in a year in the same field), growing at least one crop partially outside its optimal growing season can be justified if this increases total farm profit per year and there are no adverse biophysical consequences (e.g. pest build-up). (a) (b) Percent of years soil wetness is less than trafficability threshold for soils \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\4_APSIMmodelling\SOGWRA-Charts_APSIM_v3.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-6 Soil wetness indices that indicate when seasonal trafficability constraints are likely to occur on sands, Kandosols (loamy sands) and Vertosols (high clay) with a Gregory climate for two thresholds (a) 80% and (b) 70% of the maximum plant available water capacity The indices show the proportion of years (for dates at bi-monthly intervals) when plant available water (PAW) in the top 30 cm of the soil is below two threshold proportions (70% and 80%) of the maximum plant available water capacity value. Lower values indicate there would be fewer days at that time of year when fields would be accessible and trafficable. Estimates are from 100-year Agricultural Production Systems Simulator simulations without a crop. In actual farming situations, once a crop canopy is established later in the season, crop water extraction from the soil would assist in alleviating these constraints. Growers also manage time of sowing to optimally use stored soil water and in-season rainfall, and to avoid rain damage at maturity. In the Southern Gulf catchments mean monthly rainfall is highly variable between the wet and dry seasons (Figure 4-4) and irrigation allows growers the flexibility in sowing date and in the choice and timing of crop or forage systems in response to seasonal climate conditions. Depending on the rooting depth of a particular species and the length of growing season, crops established at the end of the wet season may access a full profile of soil water (e.g. ≥200 mm PAWC for some Vertosols). While timing sowing to the end of the wet season to take advantage of soil water may reduce the overall irrigation requirement, it may expose crops to periods of unfavourable solar radiation or temperatures during plant development and flowering. It may also prevent the implementation of a sequential cropping system. 4.3.3 Rainfed cropping Rainfed cropping (crops grown without irrigation, relying only on rain) has been practised by farmers in the NT and Queensland for almost 100 years, yet only small areas of rainfed crop production currently occur each year in the very remote northern regions. This indicates that, despite the theoretical possibility of producing rainfed crops using the significant wet-season Percent of years soil wetness is less than trafficability threshold for soils \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\4_APSIMmodelling\SOGWRA-Charts_APSIM_v3.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au rainfall in the Southern Gulf catchments, in practice significant agronomic and market-related challenges to rainfed crop production have prevented its expansion to date. Without the certainty provided by irrigation, rainfed cropping is opportunistic in nature, relying on favourable conditions in which to establish, grow and harvest a crop. The annual cropping calendar in Figure 4-5 shows that, for many crops, the sowing window includes the month of February. For relatively short-season crops, such as forage sorghum and mungbean (Vigna radiata), this coincides with both the sowing time that provides close to maximum crop yield and the time at which the season’s water supply can be accessed with a high degree of confidence. Table 4-5 shows how plant available soil water content at sowing and subsequent rainfall in the 90 days after each sowing date varies over three different sowing dates for a Vertosol in the Southern Gulf catchments at Gregory. As sowing is delayed from February to April, the amount of stored soil water decreases. However, there is a significant decrease in rainfall in the 3 months after sowing. Combining the median PAW in the soil profile at sowing, and the median rainfall received in the 90 days following sowing, provides totals of 392, 250 and 183 mm for the February, March and April sowing dates, respectively. For drier-than-average years (80% probability of exceedance), the soil water stored at sowing and the expected rainfall in the ensuing 90 days (<260 mm) would result in water stress and comparatively reduced crop yields. In wetter-than-average years (20% probability of exceedance) the amount of soil water at the end of February combined with the rainfall in the following 90 days (527 mm) is sufficient to grow a good short-season crop (noting that the timing of rainfall is also important because some rain is ‘lost’ to runoff, evaporation and deep drainage between rainfall events). Opportunistic rainfed cropping would target those wetter years where PAW at the time of sowing indicated a higher chance of harvesting a profitable crop. Table 4-5 Soil water content at sowing, and rainfall for the 90-day period following sowing for three sowing dates, based on a Gregory climate on a Vertosol The 80%, 50% (median) and 20% probabilities of exceedance values are reported for the 100 years between 1920 and 2020. The lower-bound values (80% exceedance) occur in most years, while the upper-bound values only occur in the most exceptional upper 20% of years. PAW = plant available water stored in soil profile. SOWING DATE PAW AT SOWING DATE (mm) RAINFALL IN 90 DAYS FOLLOWING SOWING DATE (mm) TOTAL STORED SOIL WATER + RAINFALL IN SUBSEQUENT 90 DAYS (mm) 80% 50% 20% 80% 50% 20% 80% 50% 20% 1 February 111 176 220 132 200 308 260 392 527 1 March 133 178 208 24 78 180 171 250 369 1 April 120 173 193 0 6 58 136 183 240 Figure 4-7 highlights the impact on rainfed crop yields of the diminishing water availability from early to late wet-season planting. This constraint is much more severe for sandier soils that have less capacity to store PAW (like Kandosols on the Doomadgee Plain in the Southern Gulf catchments, Figure 4-7a), than finer textured soils (like the alluvial Vertosols in the Southern Gulf catchments, Figure 4-7b). However, the frequent inundation and waterlogging of clay soils, which are often located adjacent to rivers, means that crops cannot always be sown at optimum times; fertiliser can be lost due to runoff, drainage and denitrification; and in-crop management (e.g. for weed, disease and insect control) cannot be undertaken cost-effectively with ground-based equipment in a timely manner, a critical requirement for rainfed crop production to succeed. Those disruptions decrease the chance that high potential yields in the top 20% of the seasons could be achieved in practice. (a) Gregory Kandosol (sandy, PAWC 129 mm) (b) Gregory Vertosol (high clay, PAWC 212 mm) Influence of planting date on yield - vertosol \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\4_APSIMmodelling\SOGWRA-Charts_APSIM_v3.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-7 Influence of planting date on rainfed grain sorghum yield at Gregory for a (a) Kandosol and (b) Vertosol Estimates are from Agricultural Production Systems Simulator simulations with planting dates on the 1st and 15th of each month. The shaded band around the median line indicates the 80% to 20% exceedance probability range in year- to-year variation. PAWC = plant available water capacity of given soil profile. Seldom is soil uniform within a single paddock, let alone across entire districts. Without the homogenising input of irrigation to alleviate water limitations (and associated high inputs of fertilisers to alleviate nutrient limitations), yields from low-input rainfed cropping are typically much more variable (both across years and locations) than yields from irrigated agriculture. Furthermore, the capacity of the soil to supply stored water varies with soil type, and it also depends on crop type and variety because each crop’s root system has a different ability to access water, particularly deep in the profile. This makes it harder to make generalisations about the viability of rainfed cropping in the Southern Gulf catchments as farm performance (e.g. yields and GMs) is much more sensitive to slight variations in local conditions. Rigorous estimates of rainfed crop performance, on which investment decisions could be confidently made, would require detailed localised soil mapping and crop trials. Despite the challenges described above, recent efforts have identified potential opportunities for rainfed farming using higher-value crops, such as pulses or cotton, in northern Australia. A preliminary APSIM assessment of the potential for rainfed cotton in the Katherine region suggested that mean lint yields of 2.5 to 3.5 bales per ha may be possible at a range of locations in the vicinity of the Southern Gulf catchments (Yeates and Poulton, 2019). However, there was very high variability in median yields between farms (1–5 bales/ha), depending on management and soil type. 4.3.4 Irrigated crop response and performance metrics Crops that are fully irrigated can yield substantially more than rainfed crops. Figure 4-8 shows how modelled yields for grain sorghum grown on Vertosols in the Southern Gulf catchments increase as more water becomes available to alleviate water limitations and meet increasing proportions of crop demand. With sufficient irrigation, yields are highest for crops grown over the dry season Influence of planting date on yield - kandosol \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\4_APSIMmodelling\SOGWRA-Charts_APSIM_v3.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au when radiation tends to be less limiting comparing plateau of lines in Figure 4-8a and Figure 4-8b. For wet-season sowing, unirrigated yields can approach fully irrigated yields in good years (yields exceeded in the top 20% of years, marked by the upper shaded range in Figure 4-8a). However, irrigation allows greater flexibility in sowing dates, allows sowing in the dry season too (for crops that would then grow through the wet season) and generates more reliable (and higher median) yields. The simulations did not seek to ‘optimise’ supplemental irrigation strategies in years where available water was insufficient to maximise crop yields; irrigators would need to make those decisions in years where available water was lower than total crop demand. A key advantage of irrigated dry-season cropping in northern Australia is that the availability of water in the soil profile and surface water storages is largely known at the time of planting (in the early wet season; Table 4-5). This means irrigators have good advance knowledge for planning how much area to plant, which crops to grow and which irrigation strategies to use, particularly in years where they have insufficient water to fully irrigate all fields. A mix of irrigation approaches could be used, such as expanding the scale of a core irrigated cropping area with other less intensively farmed areas, opportunistic rainfed cropping, opportunistic supplemental irrigation, opportunistic sequential cropping and/or adjusting the area of fully irrigated crops grown to match available water supplies that year. (a) 1 February sowing (wet season) (b) 1 August sowing (dry season) Figure 4-8 Influence of available irrigation water on grain sorghum yields for planting dates of (a) 1 February and (b) 1 August, for a Vertosol with a Gregory climate Estimates are from 100-year Agricultural Production Systems Simulator simulations. The shaded band around the median line indicates the 80% to 20% exceedance probability range in year-to-year variation. Rainfed production is indicated by the zero point, where no allocation is available for irrigation. Measures of farm performance (in terms of yields, water use and GMs) are presented for the 19 cropping options that were evaluated (Table 4-4). Given the limited commercial irrigated farming currently occurring in the Southern Gulf catchments that can provide real-world data, estimates of crop water use and yields should be considered as indicative, and to have a possible 20% margin of error at the catchment scale (with further variation expected between farms and fields). The measures of performance should be considered as an upper bound of what could be achieved under best-practice management after learning and adapting to location-specific conditions. GMs are a key partial metric of farm performance but should not be treated as fixed constants determined by the cropping system alone. They are a product of the farming and business Influence of available water on yield for sorghum planted 1 Feb \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\4_APSIMmodelling\SOGWRA-Charts_APSIM_v3.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Influence of available water on yield for sorghum planted 1 Aug \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\4_APSIMmodelling\SOGWRA-Charts_APSIM_v3.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au 6 management decisions made by individual farmers, input prices, commodity prices and market opportunities (details on calculation of GMs are in Webster et al., 2024). As such, the GMs presented in Table 4-6 should be treated as indicative of what might be attained for each cropping option once its sustainable agronomic potential has been achieved. Any divergence from assumptions about yields and costs would flow through to GM values, as would the consequences of any underperformance or overperformance in farm management. It is unrealistic to assume that the levels of performance in the results below would be achieved in the early years of newly established farms, and allowance should be made for an initial period of learning when yields and GMs are below their potential (see Chapter 6). Collectively however, the GMs and other performance metrics presented here provide an objective and consistent comparison across a suite of likely cropping options for the Southern Gulf catchments, and indicate a maximum performance that could be achievable for greenfield irrigated development for each of the groupings of crops in sections 4.3.5 to 4.3.7. 4.3.5 Irrigated broadacre crops Table 4-6 shows the farm performance (yields, water use and GMs) for the ten broadacre cropping options that were evaluated. For crops that were simulated with APSIM, estimates are provided for locations with four different soil types associated with climates in the Southern Gulf catchments (Kandosol at Westmoreland, Vertosol at Gregory, Chromosol at Kamilaroi and Vertosol at Gallipoli) and include measures of variability (expressed in terms of years with yield exceedance probabilities of 80%, 50% (median) and 20%). For other crops, yield and water use estimates (and resulting GMs) were estimated based on expert experience and climate-informed extrapolation from the most similar analogue locations in northern Australia where commercial production currently occurs. The broadacre cropping options with the best GMs (>$2000/ha) were cotton (both wet-season and dry-season cropping), forages (Rhodes grass (Chloris gayana)) and peanuts (on a Chromosol). These suggest GMs up to $4500/ha might be achievable for broadacre cropping in the Southern Gulf catchments, although not necessarily at scale. Simulated yields (and consequent GMs) were generally lowest on the Kandosol and highest on the heavy Vertosol because of the increased buffering capacity that a high PAWC clay soil provides against hot weather, which triggers water stress even in irrigated crops. The Chromosol yields and GMs were slightly lower than the Vertosol due to its lower PAWC. With Vertosols in the Southern Gulf catchments there could be drainage challenges (Figure 4-6) that could limit the suitable area for farming and may require more careful management than Vertosols that are currently used for cotton farming in other parts of Australia. A breakdown of the variable costs for growing broadacre crops shows that the largest costs are the costs of inputs (mean 28%), farm operations (mean 33%) and marketing (mean 28%) (Table 4-7). The input and operations cost categories would have similar dollar values when growing the same crop in southern parts of Australia, but the cost category that is higher and thus puts northern growers at a disadvantage is market costs (freight and other costs involved in selling the crop). Total variable costs consume 84% of the gross revenue generated, which leaves margin for profitable farms to be able to temporarily absorb small declines in commodity prices or yields without creating severe cashflow problems. 226 | Water resource assessment for the Southern Gulf catchments Table 4-6 Performance metrics for broadacre cropping options in the Southern Gulf catchments: applied irrigation water, crop yield and gross margin (GM) for four environments Performance metrics indicate the upper bound that could be achieved after best management practices for Southern Gulf catchments environments had been identified and implemented. All options are for dry-season (DS) irrigated crops sown between March and May (end of the wet season (WS)), except for the WS cotton, sown in mid-February and DS cotton sown in mid-June. Our modelled results suggest that dry-season planting of cotton in mid-June at Gallipoli led to a high incidence of crop failure and is not shown. Variance in yield estimates from Agricultural Production Systems sIMulator (APSIM) simulations is indicated by providing 80%, 50% (median) and 20% probability of exceedance values (Y80%, Y50% and Y20%, respectively), together with associated applied irrigation water (including on-farm losses) and GMs in those years. The lower-range yields (Y80% exceedance) occur in most years, while the upper-range Y20% yields only occur in the most exceptional upper 20% of years. Note that applied irrigation water is not always higher in years with higher yields (Y20%). ‘na’ indicates 20% and 80% exceedance estimates that were not applicable because APSIM outputs were not available and expert estimates of just the median yield and water use were used instead. Peanut is omitted for the Vertosol location because of the practical constraints of harvesting root crops on clay soils. Freights costs assume processing near Cloncurry for cotton and Townsville for peanut, and that hay is sold locally. No crop model was available for sesame or hemp, so indicative estimates for the catchments were used. Cotton yields and prices are for lint bales (227 kg after ginning), not tonnes, and account for a lint turnout of 40% and a cotton seed price of $280/t. PAWC = plant available water capacity. CROP APPLIED IRRIGATION WATER CROP YIELD YIELD UNIT PRICE VARIABLE COSTS TOTAL REVENUE GROSS MARGIN (ML/ha/y) (yield units) ($/unit) ($/ha/y) ($/ha/y) ($/ha/y) Y80% Y50% Y20% Y80% Y50% Y20% Y80% Y50% Y20% Red Kandosol (129 mm PAWC), Westmoreland climate (~780 mm annual rainfall) Cotton (WS) 4.5 5.3 5.7 8.6 9.4 10 bales/ha 700 4,159 7,415 2,784 3,256 3,683 Cotton (DS) 5.0 5.4 5.9 3.7 5.0 6 bales/ha 700 3,230 3,939 –24 708 1,554 Sorghum (grain) 5.2 5.6 6.0 8.5 9.0 10 t/ha 350 3,779 3,164 –641 –615 –597 Mungbean 3.6 3.9 4.3 1.5 1.8 2 t/ha 1,200 1,417 1,958 323 540 720 Chickpea 2.2 2.3 2.6 0.3 0.3 0 t/ha 750 1,019 254 –776 –765 –732 Soybean 5.6 5.9 6.4 3.3 3.5 4 t/ha 650 2,139 2,265 107 125 182 Peanut 4.0 4.5 4.9 5.1 5.5 6 t/ha 1,000 4,849 5,455 483 607 810 Rhodes grass (hay) 13.7 15.3 16.9 38.1 39.2 41 t/ha 220 5,672 8,624 3,032 2,952 3,005 Maize 5.5 5.8 6.1 8.9 9.3 10 t/ha 380 3,857 3,530 –291 –328 –308 Heavy Vertosol (212 mm PAWC), Gregory climate (~540 mm annual rainfall) Cotton (WS) 4.1 4.7 5.2 9.7 10.7 11 bales/ha 700 3,857 8,425 3,986 4,536 4,950 CROP APPLIED IRRIGATION WATER CROP YIELD YIELD UNIT PRICE VARIABLE COSTS TOTAL REVENUE GROSS MARGIN (ML/ha/y) (yield units) ($/unit) ($/ha/y) ($/ha/y) ($/ha/y) Y80% Y50% Y20% Y80% Y50% Y20% Y80% Y50% Y20% Cotton (DS) 5.2 5.9 6.3 5.0 6.8 9 bales/ha 700 3,889 5,354 1,066 2,096 3,378 Sorghum (grain) 5.2 5.8 6.5 10.0 10.5 11 t/ha 350 3,259 3,671 358 298 348 Mungbean 3.2 3.8 4.4 1.6 1.9 2 t/ha 1,200 3,373 2,068 559 794 968 Chickpea 3.2 3.8 3.9 1.9 2.2 3 t/ha 750 1,275 1,650 90 227 436 Soybean 7.6 8.6 9.3 4.7 5.0 5 t/ha 650 1,423 3,223 913 998 1,109 Rhodes grass (hay) 17.6 20.5 22.3 44.9 45.9 47 t/ha 220 6,566 10,096 3,648 3,530 3,474 Maize 6.7 7.2 7.9 9.6 10.0 10 t/ha 380 3,299 3,800 445 501 544 Chromosol (92 mm PAWC), Kamilaroi climate (~577 mm) Cotton (WS) 4.7 5.4 5.8 9.3 10.5 12 bales/ha 700 3,697 8,268 3834 4,571 5,195 Cotton (DS) 6.3 6.9 7.4 4.7 6.6 8 bales/ha 700 3,183 5,197 815 2,014 2,892 Sorghum (grain) 5.7 6.4 6.9 10.5 11.1 12 t/ha 350 3,159 3,885 657 726 809 Mungbean 3.5 4.1 4.5 1.6 1.9 2 t/ha 1,200 1,253 2,101 608 848 1,090 Chickpea 2.4 2.7 3.3 0.8 1.1 1 t/ha 750 1,128 788 –445 –340 –218 Soybean 7.3 7.8 8.4 4.0 4.2 4 t/ha 650 1,949 2,703 687 754 826 Peanut 5.3 5.8 6.2 6.3 6.9 7 t/ha 1,000 4,766 6,900 1,793 2,134 2,359 Rhodes grass (hay) 20.0 22.4 24.3 45.2 46.3 47 t/ha 220 6,946 10,195 3,319 3,249 3,175 Maize 6.5 6.9 7.3 9.3 9.8 10 t/ha 380 2,937 3,706 700 769 833 Light Vertosol (146 mm PAWC), Gallipoli climate (~420 mm) Cotton (WS) 4.3 4.9 5.4 5.6 7.6 9 bales/ha 700 3,324 5,984 1414 2,660 3,581 Sorghum (grain) 6.5 7.4 7.9 11.1 11.6 12 t/ha 350 3,521 4,076 499 555 603 Mungbean 4.0 4.5 5.0 1.8 2.0 2 t/ha 1,200 1,306 2,212 700 906 1,104 Chickpea 3.5 4.1 4.5 1.9 2.3 3 t/ha 750 1,416 1,692 117 276 538 Soybean 8.8 9.4 10.1 4.8 5.0 5 t/ha 650 2,213 3,240 947 1,026 1,082 CROP APPLIED IRRIGATION WATER CROP YIELD YIELD UNIT PRICE VARIABLE COSTS TOTAL REVENUE GROSS MARGIN (ML/ha/y) (yield units) ($/unit) ($/ha/y) ($/ha/y) ($/ha/y) Y80% Y50% Y20% Y80% Y50% Y20% Y80% Y50% Y20% Rhodes grass (hay) 24.4 22.4 24.4 45.0 46.3 47 t/ha 220 6,829 10,179 4,575 3,350 3,327 Maize 6.1 8.0 8.5 9.7 10.2 11 t/ha 380 3,256 3,895 203 638 675 General estimate for Southern Gulf catchments (not soil specific) Sesame na 5.3 na na 0.9 na t/ha 1,300 2,041 1,170 na –871 na Hemp (grain seed) na 5.0 na na 1.1 na t/ha 3,150 2,519 3,465 na 946 na Table 4-7 Breakdown of variable costs relative to revenue for broadacre crop options The first eight crops (Cotton (WS) to Rhodes grass) are for the Chromosol, Kamilaroi climate (intermediate performance), and the last three crops are for general catchment estimates. ‘Input’ costs are mainly for fertilisers, herbicides and pesticides; the cost of farm ‘operations’ includes harvesting; ‘labour’ costs are the variable component (mainly seasonal workers) not covered in fixed costs (mainly permanent staff); ‘market’ costs include levies, commission and transport to the point of sale. WS = wet season; DS = dry season. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Risk analyses were conducted for the two broadacre crops with the highest GMs: cotton and forages. The risk analysis used a narrative approach, where variable values with the potential to be different from those used in the GMs were varied and new GMs calculated. The narrative approach allows the impact of those variables to be determined. The cotton analysis explored the sensitivity of GMs to opportunities and challenges created by changes in cotton lint prices, crop yields and distance to the nearest gin (Table 4-8). Results show that high recent cotton prices (about $800/bale through 2022) have created a unique opportunity for those looking to establish new cotton farms in NT locations like the Victoria catchment, since growers could transport cotton to distant gins or produce suboptimal yields and still generate GMs above $3000/ha. At lower cotton lint prices, a local gin becomes more important for farms to remain viable. High cotton prices and the opening of a cotton gin 30 km north of Katherine in December 2023 have reduced some of the risk involved in learning to grow cotton as GMs increase from both these developments. At high yields and prices, the returns per megalitre of irrigation water may favour growing a single cotton crop per year, instead of committing limited water supplies to sequential cropping with a dry-season crop (that would likely provide lower returns per megalitre and be operationally difficult/risky to sequence). Table 4-8 Sensitivity of cotton crop gross margins ($/ha) to variation in yield, lint prices and distance to gin The base case is the Gregory heavy Vertosol (Table 4-6) and is highlighted for comparison. The gin locations considered are a local gin near a new cotton farming region in the Southern Gulf catchments near Gregory, a hypothetical gin in Cloncurry, and the existing gin in Emerald, Queensland. Cotton lint prices include a low price for 2015–2020 ($580/bale), a mean price for 2020–2024 ($700/bale) and a high price for 2015–2020 ($900/bale). Effects of a lower yield are also tested (the base case of 10.7 bales/ha for wet-season cropping versus the 6.8 bales/ha estimated as the dry-season yield for this location). FREIGHT COST/TONNE (DISTANCE TO GIN) COTTON CROP GROSS MARGIN ($/ha) LINT PRICE = $580/bale LINT PRICE = $700/bale LINT PRICE = $900/bale YIELD YIELD YIELD 10.7 bales/ha 6.8 bales/ha 10.7 bales/ha 6.8 bales/ha 10.7 bales/ha 6.8 bales/ha $13 (50 km to local gin) 3732 1613 5016 2429 7156 3789 $92 (330 km to Cloncurry gin) 3532 1308 4536 2124 6676 3484 $243 (1250 km to Emerald gin) 2335 725 3619 1541 5759 2901 The narrative risk analysis for irrigated forages also looked at the sensitivity of farm GMs to variations in hay price and distance to markets, but here focuses on the issues of local supply and demand (Table 4-9). Forages, such as Rhodes grass, are a forgiving first crop to grow on greenfield farms as new farmers gain experience of local cropping conditions and ameliorate virgin soils while producing a crop with a ready local market in cattle. While there are limited supplies of hay in the region, growers may be able to sell hay at a reasonable price, given the large amount of beef production in the Southern Gulf catchments and challenges of maintaining livestock condition through the dry season, when the quality of native pastures is low. The scale of unmet local demand for hay limits opportunities for expansion of hay production without depressing local prices and/or having to sell hay further away, both of which lead to rapid declines in GMs (to below zero in many cases; Table 4-9). Another opportunity for hay is for feeding to cattle during live export, which could be integrated into an existing beef enterprise to supply their own live export livestock; this would require the hay to be pelleted. Section 4.3.9 considers how forages could be integrated into local beef productions systems for direct consumption by livestock within the same enterprise. Table 4-9 Sensitivity of forage (Rhodes grass) crop gross margins ($/ha) to variation in yield and hay price The base case is the Gregory heavy Vertosol (Table 4-6) and is highlighted for comparison. Transporting the hay further distances would increase opportunities for finding counter-seasonal markets paying higher prices, but this would be rapidly offset by higher freight costs. FREIGHT COST/TONNE (DISTANCE TO DELIVER) FORAGE CROP GROSS MARGIN ($/ha) HAY PRICE/TONNE $150 $250 $350 $20 (local) 317 3530 9495 $92 (330 km to Cloncurry) –1708 1505 7471 $243 (1250 km to Emerald) –9917 –6704 –738 4.3.6 Irrigated horticultural crops Table 4-10 shows estimates of potential performance for a range of horticultural crop options in the Southern Gulf catchments. Upper potential GMs for annual and tree horticulture are about $5000 per ha per year). Capital costs of farm establishment and operating costs increase as the intensity of farming increases, so ultimate farm financial viability is not necessarily better for horticulture compared to broadacre crops with lower GMs (see Chapter 6). Note also that perennial horticulture crops typically require more water than annual crops because irrigation occurs for a longer period each year (mean of 8.5 compared to 4.4 ML per ha per year, respectively in Table 4-10); this also, indirectly, affects capital costs of development since perennial crops require a larger investment in water infrastructure compared to annual crops to support the same cropped area. Table 4-10 Performance metrics for horticulture options in the Southern Gulf catchments: annual applied irrigation water, crop yield and gross margin Applied irrigation water includes losses of water during application. Horticulture is most likely to occur on well-drained Kandosols. Product unit prices listed are for the dominant top grade of produce, but total yield was apportioned among lower graded/priced categories of produce as well in calculating total income. Transport costs assume sales of total produce are split among southern capital markets in proportion to their size. Applied irrigation water accounts for application losses assuming efficient pressurised micro irrigation systems. KP = Kensington Pride mangoes; PVR = new high-yielding mangoes varieties with plant variety rights (e.g. Calypso). CROP APPLIED IRRIGATION WATER CROP YIELD PRICE PRICING UNIT VARIABLE COSTS TOTAL REVENUE GROSS MARGIN (ML/ha/y) (t/ha/y) ($/unit) (unit) ($/ha/y) ($/ha/y) ($/ha/y) Row crop fruit and vegetables, annual horticulture (less capital intensive) Rockmelon 5.0 25.0 28 15 kg tray 43,216 44,000 784 Watermelon 5.7 47.0 450 500 kg box 52,321 42,300 –10,021 Capsicum 3.0 32.0 19 8 kg carton 71,158 76,000 8,842 Onion 4.0 30.0 15 10 kg bag 36,906 41,850 4,944 Fruit trees, perennial horticulture (more capital intensive) Mango (KP) 7.4 9.3 24 7 kg tray 22,023 28,398 6,375 Mango (PVR) 7.4 17.5 21 7 kg tray 42,786 47,250 4,464 Lime 10.8 28.5 18 5 kg carton 94,913 100,890 5,977 Crop yields and GMs can vary substantially among varieties, as is demonstrated in Table 4-10 for mangoes (Mangifera indica). Mango production is well established in multiple regions of northern Australia, including in the Darwin, Douglas–Daly and Katherine regions of the NT, Bowen, the lower Burdekin and the Mareeba–Dimbulah Irrigation Area in Queensland. For example, the well- established Kensington Pride mangoes typically produce 5 to 10 t/ha while newer varieties (such as Calypso) can produce 15 to 20 t/ha. New varieties are likely to be released with plant variety rights (PVR) accreditation and are denoted as such. Selection of varieties also needs to consider consumer preferences and timing of harvest relative to seasonal gaps in market supply that can offer premium prices. Prices paid for fresh fruit and vegetables can be extremely volatile (Figure 4-9) because produce is perishable and expensive to store, and because regional weather patterns can disrupt target timing of supply, which can result in unintended overlaps or gaps in combined supply between regions. This creates regular fluctuations between oversupply and undersupply, against inelastic consumer demand, to the extent that prices can fall so low at times that it would cost more to pick, pack and transport produce than farms receive in payment. Within this volatility are some counter-seasonal windows in southern markets (where prices are typically higher) that northern Australian growers can target. Figure 4-9 Fluctuations in seedless watermelon prices at Melbourne wholesale markets from April 2020 to February 2023 Percentage change information available; however, prices are commercially sensitive and not available. Source: ABARES (2023) Horticultural enterprises typically run on very narrow margins, where about 90% of gross revenue would be required just to cover variable costs of growing and marketing a crop grown in the Southern Gulf catchments. This makes crop GMs extremely sensitive to fluctuations in variable costs, yield and produce prices, amplifying the effect of already volatile prices for fresh fruit and vegetables. The majority of the variable costs of horticultural production occur from harvest onwards, mainly in freight, labour and packaging. This affords the opportunity to mitigate losses if market conditions are unfavourable at the time of harvest, since most costs can be avoided (at the expense of foregone revenue) by not picking the crop. The narrative risk analysis for horticulture used the crop with the lowest GM (watermelons (Citrullus lanatus) Table 4-10, to illustrate how opportunities for reducing freight costs and targeting periods of higher produce prices could improve GMs to find niches for profitable farms (Table 4-11). Reducing freight costs by finding backloading opportunities or concentrating on just the smaller closest southern capital city market of Brisbane would substantially improve GMs, but a higher price than average is needed to generate positive GMs. The base case already assumed that growers in the Southern Gulf catchments would target the predictable seasonal component of watermelon price fluctuations (Figure 4-9), but any further opportunity to attain premiums in pricing could help convert an unprofitable baseline case into a profitable one. This example also highlights the issue that while there may be niche opportunities that allow an otherwise Influence of available water on yield for sorghum planted 1 Aug https://www.agriculture.gov.au/abares/data/weekly-commodity-price-update/australian-horticulture-prices#daff-page-main For more information on this figure please contact CSIRO on enquiries@csiro.au unprofitable enterprise to be viable, the scale of those niche opportunities also then limits the scale to which the industry in that location could expand; for example: (i) there is a limit to the volume of backloading capacity at cheaper rates, (ii) supplying produce to only the closest market excludes the largest markets (e.g. accessing the larger Sydney and Melbourne markets remains non-viable except when prices are high; Table 4-11) and (iii) chasing price premiums restricts the seasonal windows into which produce is sold or restricts markets to smaller niches that target specialised product specifications. Niche opportunities are seldom scalable, particularly in horticulture, which is partly why horticulture in any region usually involves a range of different crops (often on the same farm). Table 4-11 Sensitivity of watermelon crop gross margins ($/ha) to variation in melon prices and freight costs The base case (Table 4-10) is highlighted for comparison. FREIGHT COST/TONNE WATERMELON PRICE (PERCENTAGE DIFFERENCE FROM BASE PRICE) (MARKET LOCATION) $210/t $337 (–25%) $450 (BASE PRICE) $675 (+50%) $900 (+100%) $342 (backloading to Brisbane) –10,836 –1,702 16,487 34,676 $429 (close market: Brisbane) –14,925 –5,791 12,398 30,587 $519 (all capital cities) –19,155 –10,021 8,168 26,357 $559 (Melbourne) –21,035 –11,901 6,288 24,477 The risk analysis also illustrates just how much farm financial metrics like GMs amplify fluctuations to input costs and commodity prices to which they are exposed. For horticulture, far more than broadacre agriculture, it is very misleading to look just at a single ‘median’ GM for the crop, because that is a poor reflection of what is going on within an enterprise. For example, a –50% to +100% variation in watermelon prices would result in theoretical annual GMs fluctuating between–$19.155/ha and $26,357/ha (Table 4-11). Although, in practice, potentially negative GMs couldbe greatly mitigated (by not harvesting the crop), this still creates cashflow challenges in managingyears of negative returns between years of windfall profits. This amplified volatility is anotherreason that horticultural farms often grow a mix of produce (as a means of spreading risk). For rowcrop production, another common way of mitigating risk is using staggered planting through theseason, so that subsequent harvesting and marketing are spread out over a longer target windowto smooth out some of the price volatility. 4.3.7 Plantation tree crops Estimates of annual performance for African mahogany (Khaya ivorensis) are provided in Table 4-12. The best available estimates were used in the analyses, but information on plantation treeproduction in northern Australia is often commercially sensitive and/or not independentlyverified. The measures of performance presented, therefore, have a low degree of confidence andshould be treated as broadly indicative, noting that actual commercial performance could beeither lower or higher. Table 4-12 Performance metrics for plantation tree crop options in the Southern Gulf catchments: annual applied irrigation water, crop yield and gross margin Yields are values at final harvest and pricing unit is for an 800 kg cube, with 10% of the African mahogany yield as marketable cubes. Other values are annual averages assuming a 20-year life cycle of the crop (representing the idealised ultimate steady state of an operating farm that was set up with staggered plantings for a steady stream of harvests). No discounting is applied to account for the substantial timing offset between when costs are incurred and revenue is received; any investment decision would need to take that into account. African mahogany performance is for unirrigated production. CROP CROP LIFE CYCLE APPLIED IRRIGATION WATER CROP YIELD AT HARVEST PRICE PRICING UNIT VARIABLE COSTS TOTAL REVENUE GROSS MARGIN (y) (ML/ha/y) (t/ha) ($/unit) ($/ha/y) ($/ha/y) ($/ha/y) African mahogany 20 unirrigated 160 4000 cube 1103 4000 2897 Plantation forestry has long life cycles with low-intensity management during most of the growth cycle, so variable costs typically consume less of the gross revenue (28%) than for broadacre or horticultural farming. However, production systems with long life cycles have additional risks over annual cropping: there is a much longer period between planting and harvest for adverse events to affect the yield quantity and/or quality, prices of inputs and harvested products could change substantially over that period, and market access and arrangements with buyers could change. The long lags from planting to harvest also mean that potential investors need to consider other similar competing pipeline developments (that may not be obvious because they are not yet selling product) and long-term future projections of supply and demand (for when their own plantation will start to be harvested and enter supply chains). The cashflow challenges are also significant given the long-term outlay of capital and operating costs before any revenue is generated. Carbon and other externality credits might be able to assist with some early cashflow (e.g. if the ‘average’ state of the plantation, from planting to harvest, stores more carbon than the vegetation it replaced). 4.3.8 Cropping systems This section evaluates the types of cropping systems (crop species × growing season × resource availability × management options) that are most likely to be profitable in the Southern Gulf catchments based on the above analyses of GMs, information from companion technical reports in this Assessment, and cropping knowledge from climate-analogous regions (relative to local biophysical conditions). Cropping system choices could include growing a single crop during a 12- month period, or growing more than one crop – commonly referred to as sequential, double or rotational cropping. Since many of the issues for single cropping options were covered earlier, this section focuses on sequential cropping systems and the mix of cropping options that might be grown in sequence on a unit of land in the Southern Gulf catchments. Cropping system considerations Selecting two or more crops to grow in sequence increases the complexity, beyond the issues already discussed, in finding and adapting individual cropping options for the Southern Gulf catchments. The rewards from successfully growing crops in sequence (versus single cropping) can be substantial if additional net annual revenue can be generated from the same initial capital investment (to establish the farm). To find viable mixes of cropping options for the Southern Gulf catchments, developers will need to consider each of the following four key factors. Markets Whether growing a single crop or doing sequential cropping, the choice of crop(s) to grow is market driven. As the price received for different crops fluctuates, so too will the crops grown. In the Southern Gulf catchments, freight costs, determined by the distance to selected markets, must also be considered. A critical scale of production may be needed for a new market opportunity or supply chain to be viable (e.g. exporting grains from Townsville would require sufficient economies of scale for the required supporting port infrastructure, and shipping routes to be viable). Crops such as cotton, peanut and sugarcane (Saccharum officinarum) require a processing facility. A consistent and critical scale of production is required for processing facilities to be viable. Transport costs of raw cotton from the Southern Gulf catchments to the closest gin in Emerald would be offset by access to a gin locally and go a long way to improving the viability of cotton production (Table 4-8). Most horticultural production from the Southern Gulf catchments would be sent to capital city markets, often using refrigerated transport. Horticultural production in the Southern Gulf catchments would have to accept a high freight cost compared to the costs faced by producers in southern parts of Australia. The competitive advantage of horticultural production in the Southern Gulf catchments is that higher market prices can be achieved from ‘out-of-season’ production compared to large horticultural production areas in southern Australia. Annual horticultural row crops, such as melons, would be grown sequentially, for example with fortnightly planting over 3 to 4 months, to reduce risk of exposure to low market prices and to make it more likely that very high market prices would be achieved for at least some of the produce. Operations Sequential cropping can require a trade-off against sowing at optimal times to allow crops to be grown in a back-to-back schedule. This trade-off could lead to lower yields from planting at suboptimal times. For annual horticultural crops there would be additional trade-offs in the seasonal window over which produce can be sent to market (affecting opportunities to target seasonal peaks in prices and to use staggered planting dates to mitigate risks from price fluctuations). Growing crops sequentially depends on timely transitions between the crops, and selecting crops that are agronomically and operationally compatible with each other, including growing seasons that reliably fit together in the available cropping windows. In the catchments’ variable and often intense wet season, rainfall increases operational risk because of reduced trafficability and the subsequent limited ability to conduct timely operations. A large investment in machinery (either multiple or larger machines) could increase the area that could be planted per day when fields are trafficable within a planting window. With sequential cropping, additional farm machinery and equipment may be required where there are crop-specific machinery requirements, or to help complete operations on time when there is tight scheduling between crops. Any additional capital expenditure on farm equipment would need to be balanced against the extra net farm revenue generated. Sequential cropping can also lead to a range of cumulative issues that need careful management, for example: (i) build-up of pests, diseases (particularly if the sequential cropping is of the same species or family) and weeds; (ii) pesticide resistance; (iii) increased watertable depth; and (iv) soil chemical and structural decline. Many of these challenges can be anticipated before beginning sequential cropping. Integrated pest, weed and disease management would be essential when multiple crop species are grown in close proximity (adjacent fields or farms). Many of these pests and controls are common to several crop species where pests (e.g. aphids) move between fields. Such situations are exacerbated when the growing seasons of nearby crops partially overlap or when sequential crops are grown, because both scenarios create ‘green bridges’ that facilitate the continuation of pest life cycles. When herbicides are required, it is critical to avoid products that could damage a susceptible crop the following season or sequentially. Water Sequential cropping leads to a higher annual crop water demand (versus single cropping) because: (i)the combined period of cropping is longer, (ii) it includes growing during the dry season in theSouthern Gulf catchments and (iii) PAW at planting will have been depleted by the previous crop. Typically, an additional 1 ML/ha on well-drained soils, and 1.5 ML/ha on clays, is required forsequential cropping relative to the combined water requirements of growing each of those cropsindividually (with the same sowing times). This additional water demand needs to be accountedfor in initial farm planning, particularly where on-farm water storage or dry-season waterextraction is required. Irrigating using surface water in the Southern Gulf catchments would face issues with the reliability and the timing of water supplies. Monitored river flows need to be sufficient to allow pumping into on-farm storages for irrigation (i.e. to meet environmental flow and river height requirements). The timing of water availability is analysed in the companion technical report on river model scenario analysis (Gibbs et al., 2024). The availability of water for extraction each wet season affects the options for sequencing a second crop. Soils The largest arable areas in the Southern Gulf catchments are the cracking clay Vertosols (SGG 9, marked ‘A’ and ‘D’ in Figure 4-3), principally on the floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland. Friable, non-cracking clay soils (SGG 1 and 2, marked ‘B’ and ‘C’) and loamy soils (SGG 4, marked ‘F’) make up substantial areas (Figure 4-3). There are good analogues of these environments in the Southern Gulf catchments in successful irrigated farming areas in other parts of northern Australia. Katherine is indicative of farming systems and potential crops grown on well-drained loamy soils irrigated by pressurised systems, and the Burdekin River Irrigation Area and Ord River Irrigation Area are indicative of furrow irrigation on heavy clay soils. The good wet-season trafficability of the well-drained loamy Kandosols permits timely cropping operations and would enhance the implementation of sequential cropping systems. However, Kandosols also present some constraints for farming. Kandosols are inherently low in organic carbon, nitrogen, potassium, phosphorus, sulfur and zinc, and supplementation with other micronutrients (boron, copper and molybdenum) is often required. Very high fertiliser inputs are therefore required at first cultivation. Due to the high risk of leaching of soluble nutrients (e.g. nitrogen and sulfur) during the wet season, in-crop application (multiple times) of the majority of crop requirement for these nutrients is necessary (Yeates, 2001). In addition, high soil temperatures and surface crusting combined with rapid drying of the soil at seed depth reduce crop establishment and seedling vigour for many broadacre species sown during the wet season and early dry season, for example, maize, soybean (Glycine max) and cotton (Abrecht and Bristow, 1996; Arndt et al., 1963). In contrast, the cracking clay Vertosols have poor trafficability following rainfall (Figure 4-6) inundation or irrigation, disrupting cropping operations. Farm design is a major factor on cracking clay soils and needs to minimise flooding of fields from nearby waterways, ensure prompt runoff from fields after irrigation or rain events, and ensure that farm roads maintain access to fields. Timely in-field bed preparation can reduce delays in planting. Clay soils also have some advantages, particularly in costs of farm development by allowing lower-cost gravity-fed surface irrigation (versus pressurised systems) and on-farm storages (where expensive dam lining can be avoided if soils contain sufficient clay) (see companion technical report on surface water storage, Yang et al., 2024). Clay soils also typically have greater inherent fertility than loamy soils, but initial sorption by clay means that phosphorus requirements can be high for virgin soils in the first 2 years of farming. Potentially suitable cropping systems Potential crop species that could be grown as a single crop per year were identified and rated for the Southern Gulf catchments (Table 4-13) based on indicators of farm performance presented above (yields, water use and GMs), together with considerations of growing season, experiences at climate-analogous locations, past research, and known market and resource limitations and opportunities. Many of these crops currently have small to medium high-value markets, hence they are sensitive to Australian and international supply. Annual horticulture, cotton, peanut and forages are the most likely to generate returns that could exceed farm development and growing costs (Table 4-13). Table 4-13 Likely annual irrigated crop planting windows, suitability and viability in the Southern Gulf catchments Crops are rated on likelihood of being financially viable: *** = likely at low-enough development costs; ** = less likely for single cropping (at current produce prices); * S = marginal but possible in a sequential cropping system. Rating qualifiers are coded as L development limitation, M market constraint, P depends on sufficient scale and distance to local processor, and B depends on distance to and type of beef (livestock production) activity it is supporting. Farm viability depends on the cost at which land and water can be developed and supplied (Chapter 6). na = not applicable. WET-SEASON PLANTING (JANUARY TO EARLY MAY) DRY-SEASON PLANTING (LATE MARCH TO AUGUST) CROP RATING CROP RATING Cotton *** P Annual horticulture *** M Forages *** B Cotton *** P Sugarcane *** LP Niche grains (e.g. chia, quinoa) *** SM Peanut (not on clay) *** LMP na na Mungbean ** Mungbean ** Maize ** na na Chickpea ** na na Rice ** L na na Sorghum (grain) *S Sorghum (grain) *S Soybean *S Soybean *S Sesame *S Sesame *S Due to good wet-season trafficability on loamy soils, there are many sequential cropping options for the Southern Gulf catchments Kandosols (Table 4-14). Given the predominance of broadleaf and legume species in many of the sequences (Table 4-14), a grass species is desirable as an early wet-season cover crop. Although annual horticulture and cotton could individually be profitable (Table 4-13), an annual sequence of the two would be very tight operationally. Cotton would be best grown from late January with the need to pick the crop by early August, then destroy cotton stubble, prepare land and remove volunteer cotton seedlings. That scheduling would make it challenging to fit in a late-season melon crop, which would need to be sown by late August to early September. Similar challenges would occur with cotton followed by mungbean or grain sorghum. Table 4-14 Sequential cropping options for Kandosols WET-SEASON PLANTING, DECEMBER TO EARLY MARCH DRY-SEASON PLANTING, MARCH TO AUGUST CROP GROWING SEASON CROP GROWING SEASON Mungbean Early February to late April Annual horticulture Mid-May to late October Sorghum (grain) January to April Peanut (not on clay) January to April or February to May Cotton Late January to early August Mungbean Mid-August to late October Sorghum (grain) Mid-August to mid-November Forage/silage Mid-August to early November; cut then retained as wet-season cover crop Mungbean Early February to late April Cotton Early May to early November Mungbean Peanut Sesame Soybean Early February to late April Early January to late April Early January to late April Early January to late April Maize May to October Sesame or Sorghum (grain) January to late April Chickpea May to August Mungbean Sesame Soybean Early February to late April January to late April January to late April Grass forage/silage May to early November; cut then retained as wet-season cover crop Fully irrigated sequential cropping on the Southern Gulf catchments Vertosols would likely be opportunistic and favour combinations of short-duration crops that can be grown when irrigation water reliability is greatest (March to October), for example, annual horticulture (melons), mungbean, chickpea and grass forages (growing season 2 to 4 months). Following an unirrigated (rainfed) wet-season grain crop with an irrigated dry-season crop could also be possible. However, seasonally dependent soil wetting and drying would limit timely planting and the area planted, which means that farm yields between years would be very variable. Sorghum (grain), mungbean and sesame (Sesamum indicum) are the species most adapted to rainfed cropping due to favourable growing season length, and their tolerance to water stress, and higher soil and air temperatures. Soil drainage, accessibility and trafficability would limit the scale of farming in the wet season within the Southern Gulf catchments (which would restrict opportunities for establishing local processors). 4.3.9 Integrating forage and hay crops into existing beef cattle enterprises A commonly held view within the northern cattle industry is that the development of water resources would allow irrigated forages and hay to be integrated into existing beef cattle enterprises, thereby improving their production and, potentially, their profitability. Currently, cattle graze on native pastures, which rely solely on rainfall and any consequent overland flow. The quality of these pastures is typically low, and it declines throughout the dry season, so that cattle either gain little weight, or even lose weight, during this period. Theoretically, the use of on-farm irrigated forage and hay production would allow graziers greater options for marketing cattle, such as meeting market liveweight specifications for cattle at a younger age, meeting the specifications required for markets different than those typically targeted by cattle enterprises in the Southern Gulf catchments and providing cattle that meet market specification at a different time of the year. Forages and hay may also allow graziers to implement management strategies, such as early weaning or weaner feeding, which should lead to flow-on benefits throughout the herd, including increased reproductive rates. Some of these strategies are already practised within the Southern Gulf catchments but in almost all incidences are reliant on hay or other supplements purchased on the open market. By growing hay on-farm, the scale of these management interventions might be increased, at reduced net cost. Furthermore, the addition of irrigated feeds may allow graziers to increase the total number of cattle that can be sustainably carried on a property. Very few cattle enterprises in northern Australia are set up to integrate on-farm irrigation, notwithstanding the theoretical benefits. Despite its apparent simplicity, fundamentally altering an existing cattle enterprise in this way brings in considerable complexity, with a range of unknowns about how best to increase productivity and profitability. The most comprehensive guide to what might be possible to achieve by integrating forages into cattle enterprises can be found in the guide by Moore et al. (2021), who have used a combination of industry knowledge, new research and modelling to consider the costs, returns and benefits. Because there are so few on-ground examples, modelling has been used in a number of studies to consider the integration of forages and hay into cattle enterprises, summarised by Watson et al. (2021). Bio-economic modelling was used in the Assessment to consider the impact of growing irrigated forages and hay on a representative beef cattle enterprise on the cracking clays of the ‘Bluegrass Browntop Plains’ land type (Southern Gulf NRM, 2016) (see the companion technical report on agricultural viability and socio-economics (Webster et al., 2024) for more detail). The enterprise was based on a self-replacing cow–calf operation, focused on selling into the live export market. Broadly speaking, these enterprise characteristics can be thought of as an owner–manager small cattle enterprise within the Southern Gulf catchments. Cattle numbers are lower than that of the average property in the Southern Gulf catchments but can be scaled to represent larger herds, notwithstanding that economies of scale will result in reduced costs per head in the larger enterprises. More detail on the beef industry in the Southern Gulf catchments can be found in Section 3.3.3. The modelling considered a number of management options: (i) a base enterprise; (ii) base enterprise plus buying in hay to feed weaners; growing forage sorghum, an annual forage grass species, and feeding either as (iii) stand and graze or (iv) as hay; (v) growing lablab (Lablab purpureus), an annual legume, and feeding as stand and graze; and (vi) growing Rhodes grass, a perennial tropical grass, and feeding as hay. Ideally, production would increase by allowing cattle to reach minimum selling weight at a younger age and allowing for greater weight gain during the dry season when animals on native pasture alone either lose weight or gain very little weight. The addition of forages and hay also allows more cattle to be carried, while still maintaining a utilisation rate of native pastures at around 18%. A GM per adult equivalent (AE) was calculated as the total revenue from cattle sales minus total variable costs (Table 4-15). A profit metric, earnings before interest, taxes, depreciation and amortisation (EBITDA), was also calculated as income minus variable and overhead costs, which allows performance to be compared independently of financing and ownership structure (McLean and Holmes, 2015) and is used in the analysis of NPV. Three sets of beef prices were considered: •LOW beef price. Beef prices were set to 275c/kg for males between 12 and 24 months old, declining across age and sex classes to 134c/kg for cows older than 108 months. •MED beef price. Beef prices were set to 350c/kg for males between 12 and 24 months old, declining across age and sex classes to 170c/kg for cows older than 108 months. •HIGH beef price. Beef prices were set to 425c/kg for males between 12 and 24 months old, declining across age and sex classes to 206c/kg for cows older than 108 months. At all three beef prices, total income was highest for the four irrigated forage or hay scenarios compared to the two baseline scenarios. At MED beef prices, EBITDA was highest for the Rhodes grass hay option at $160,929/year and lowest for forage sorghum stand and graze at –$232,238/year. The Rhodes grass hay option and the forage sorghum hay option produced the most liveweight sold per year, and the two highest incomes. An NPV analysis allows consideration of the capital costs involved in development, which are not captured in the gross margin or EBITDA. The analysis used two costings ($15,000 and $25,000/ha) for the capital costs of development used in the NPV analysis. The NPV analysis (see the companion technical report on agricultural viability and socio-economics (Webster et al., 2024)) showed that only two irrigated combinations had a positive NPV, that of Rhodes grass hay at MED and HIGH beef prices and the lower of the two development costs per hectare. All other combinations gave a negative NPV and even the two positive NPVs were low ($18,444 and $114,386), suggesting that a decision to irrigate would need to assume beef prices remaining strong to be viable. Note that cost of capital theory is complex and investors need to understand their weighted average cost of capital and the relative risk of the project compared to the enterprise’s existing project portfolio before drawing their own conclusion from an NPV analysis. Table 4-15 Production and financial outcomes from the different irrigated forage and beef production options for a representative property in the Southern Gulf catchments Details for LOW, MED and HIGH beef prices are in the text above. Descriptions of the six management options are in the companion technical report on agricultural viability and socio-economics (Webster et al., 2024). AE = adult equivalent; EBITDA = earnings before interest, taxes, depreciation and amortisation. Cattle are sold twice per year for all options. Cattle are sold in May for all options. Cattle are sold in September for the two base enterprises and for lablab stand and graze. Cattle are sold in October for forage sorghum stand and graze and the two hay options. BASE ENTERPRISE BASE ENTERPRISE PLUS HAY FORAGE SORGHUM – STAND AND GRAZE FORAGE SORGHUM – HAY LABLAB – STAND AND GRAZE RHODES GRASS – HAY Forage/hay None Bought hay Forage sorghum Forage sorghum Lablab Rhodes grass Maximum number of breeders 1580 1600 1705 1800 1730 1800 Mean of herd size (AE) across calendar year 1841 1867 2107 2182 2138 2188 Pasture utilisation (%) 18.1 18.2 17.9 18.2 18.1 18.0 Weaning rate (%) 55.5 55.4 55.9 57.6 58.4 58.9 Mortality rate (%) 7.0 7.0 6.7 6.3 6.3 6.4 Percentage of ‘one year old castrate males’ (i.e. 7 to 11 months or 8 to 12 months old) sold in September or October 0.0 0.0 0.5 77.5 57.9 77.9 Percentage of ‘one and a half year old castrate males’ (i.e. 15 to 19 months old) sold in May 48.3 60.5 73.2 17.8 25.2 17.8 Percentage of ‘two year old castrate males’ (i.e. 19 to 23 months or 20 to 24 months old) sold in September or October 11.2 11.0 25.0 4.7 16.9 4.2 Percentage of ‘two and a half year old castrate males’ (i.e. 27 to 31 months old) sold in May 40.5 28.6 1.3 0.0 0.0 0.0 Liveweight sold per year (kg) 212,840 220,123 264,395 303,699 290,798 306,171 Gross margin ($/AE) (LOW beef price) 91 78 -64 103 40 129 Profit (EBITDA) ($) (LOW beef price) –96,612 –119,279 –398,807 -40,193 –178,627 17,028 Gross margin ($/AE) (MED beef price) 163 152 15 168 123 194 Profit (EBITDA) ($) (MED beef price) 35,348 19,399 –232,238 102,546 –1,240 160,929 Gross margin ($/AE) (HIGH beef price) 236 226 94 232 205 260 Profit (EBITDA) ($) (HIGH beef price) 169,437 158,076 –65,670 242,247 173,239 304,829 A significant proportion of the animal production increases due to the irrigated forage options came from the increased number of breeders that could be carried, and the decreased number of young animals being carried over an additional wet season in order to achieve sale weight, while still keeping the utilisation rate of native pastures close to 18% (Table 4-15). The two irrigated hay options allowed the highest number of breeders to be carried (1800) compared with 1580 and 1600 for the two base enterprises. This flowed through to the total number of AE carried being about 17% to 19% higher than the two base enterprises averaged across all years. The total liveweight sold each year was about 38% to 44% higher, using the same comparison of options due to the higher liveweight gains from the feeding options combined with the higher AE. The irrigated options increased the herd’s weaning rate by 0.4% to 3.4% compared to the base enterprise without weaner feeding. Even an increase of several per cent is known to have lifetime benefits throughout a herd. The most obvious biophysical impact of the various feeding strategies was the increase in liveweight compared to that of the base enterprise. This allowed a greater proportion of the animals to be sold earlier. For example, for the two hay options, more than 77% of the ‘one-year- old castrate males’ (8–12 months old) were sold in October at a minimum weight of 280 kg, while no animals from the same cohort under the two base enterprise options met the minimum weight at that time (Table 4-15). These latter animals were retained for an additional wet season, with 48.3% (base enterprise) or 60.3% (base enterprise plus hay) being sold in the following May as ‘one-and-a-half-year olds’ (15 to 19 months old). Keeping the utilisation rate at 18.0% meant that carrying these animals for the extra period lowered the number of breeders that could be carried, and the overall stocking rate (AE). In summary, three patterns of growth to reach sale weight (280 kg) occurred: •For the two base enterprises, no animals reached sale weight in September as ‘one-year olds’. By the following May 48.3% (base enterprise) or 60.5% (base enterprise plus hay) had reached sale weight. About 11% were sold in the next September as ‘two-year olds’. The remaining 40.5% (base enterprise) or 28.6% (base enterprise plus hay) were then sold in the following May as‘two-and-a-half-year olds’. •By contrast, the majority of animals in the forage sorghum hay, lablab stand and graze, and Rhodes grass hay options were sold as ‘one-year olds’ in October. The majority of the rest(17.8%, 25.2% and 17.8%, respectively) were sold in the following May. The remainder were sold in the next October. None of this cohort remained for sale in the following May as ‘two-and-a- half-year olds’. •The forage sorghum graze option sat between these two extremes. Very few were sold as ‘one- year olds’ in October, most were sold as ‘one-and-a-half-year olds’ in the following May with almost all of the remainder sold in the following September. Only 1.3% remained to be sold as ‘two-and-a-half-year olds’ in the following May. While there are advantages to some form of irrigated forage or hay production, the introduction of irrigation to an existing cattle enterprise is not for the faint-hearted. The options here range from an area that would require 1.6 pivots of 40 ha each to an area that would require more than five 40 ha pivots. A water allocation of about 1.1 to 2.1 GL would be required to provide sufficient irrigation water. The capital cost of development would range between $975,000 for 65 ha of Rhodes grass hay, at a development cost of $15,000/ha, to $5,125,000 for 205 ha of forage sorghum stand and graze at a development cost of $25,000/ha. In addition, the grazing enterprise would need to develop the expertise and knowledge required to run a successful irrigation enterprise of that scale, which is quite a different enterprise to one of grazing only. This is a constraint recognised by graziers elsewhere in northern Australia (McKellar et al., 2015) and almost certainly contributes to the lack of uptake of irrigation in the Southern Gulf catchments. 4.4 Crop synopses 4.4.1 Introduction The estimates for land suitability in these synopses represent the total areas of the catchments unconstrained by factors such as water availability, landscape complexity, land tenure, environmental and other legislation and regulations, and a range of biophysical risks such as cyclones, flooding and secondary salinisation. These are addressed elsewhere by the Assessment. The land suitability maps are designed to be used predominantly at the regional scale. Farm-scale planning would require finer-scale, more localised assessment. 4.4.2 Cereal crops Cereal production is well established in Australia. The area of land devoted to producing grass grains (e.g. wheat, barley (Hordeum vulgare), grain sorghum, maize, oats (Avena sativa), triticale (× Triticosecale)) each year has stayed relatively consistent at about 20 million ha over the decade from 2012–13 to 2021–22, yielding over 55 Mt with a value of $19 billion in 2021–22 (ABARES, 2022). Production of cereals greatly exceeds domestic demand, and in 2021–22 the majority (82% by value) was exported (ABARES, 2022). Significant export markets exist for wheat, barley and grain sorghum, with combined exports valued at $15 billion in 2021–22. There are additional niche export markets for grains such as maize and oats. Among the cereals, sorghum (grain) is promising for the Southern Gulf catchments. Sorghum is grown over the summer period, coinciding with the Southern Gulf catchments wet season. Sorghum can be grown opportunistically using rainfed production, although the years in which this could be successfully done will be limited. Cereal crop production is higher and more consistent when irrigation is used. From a land suitability perspective, cereal crops are included in Crop Group 7 (Table 4-2; Figure 4-10). Cracking clay soils (Vertosols) make up 23% of the catchments; they are principally found onfloodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lowerparts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and theclay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable(with moderate or minor limitations) for spray irrigation in the dry season, but inadequatedrainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the GulfFall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of theLeichhardt River make up only about 3% of the area but have potential for agriculture, as do theloamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plainsand other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and areunsuitable by definition. Assuming unconstrained development, approximately 4.7 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated cereal cropping (Crop Group 7; Table 4-2) using spray irrigation in the dry season. For spray irrigation in the wet season, nearly 3.1 million ha is suitable with moderate limitations (Class 3) or better. Land considered suitable with moderate limitations for furrow irrigation is limited to about 1.8 million ha in the dry season and about 780,000 ha in the wet season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the loamy and sandy soils are too permeable. There is potential for rainfed cereal production in the wet season over an area of about 360,000 ha. From a land suitability perspective, Crop Group- 7 contains cereal crops and cotton; the latter is considered under industrial (cotton) in these crop synopses (Section 4.4.6). The ‘winter cereals’ such as wheat and barley are not well adapted to the climate of the Southern Gulf catchments. To grow cereal crops, farmers will require access to tillage, fertilising, planting, spraying and harvesting equipment. Harvesting is often a contract operation, and in larger growing regions other activities can also be performed under contract. Because of the low relative value of cereals, good returns are made through production at a large scale. This requires machinery to be large so that operations can be completed in a timely way. Table 4-16 provides summary information relevant to the cultivation of cereals, using sorghum (grain) (Figure 4-11) as an example. The companion technical report on agricultural viability and socio-economics (Webster et al., 2024) provides greater detail for a wider range of crops. Figure 4-10 Modelled land suitability for Crop Group 7 (e.g. sorghum (grain) or maize) using furrow irrigation in the (a) wet season and (b) dry season These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-501_Suit_SorgGrain_MaizeGrain_v2.png For more information on this figure please contact CSIRO on enquiries@csiro.au Table 4-16 Summary information relevant to the cultivation of cereals, using sorghum (grain) as an example For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 4-11 Sorghum (grain) Photo: CSIRO 4.4.3 Pulse crops (food legume) Pulse production is well established in Australia. The area of land devoted to production of pulses (mainly chickpea, lupin (Lupinus spp.) and field pea (Pisum sativum)) each year has varied from 1.1 to 2.0 million ha over the decade from 2012–13 to 2021–22, yielding over 3.8 Mt with a value of $2.5 billion in 2021–22 (ABARES, 2022). The vast majority of pulses in 2021–22 (93% by value) were exported (ABARES, 2022). Pulses produced in the Southern Gulf catchments would most likely be exported, although there is presently no cleaning or bulk handling facility nearby; however, established export ports are located at Townsville and Karumba. Many pulse crops have a relatively short growing season, meaning they are well suited to opportunistic rainfed production, as well as irrigated production either as a single crop or in rotation with cereals or other non-legume crops. Not all pulse crops are likely to be suited to the Southern Gulf catchments. Those that are ‘tender’, such as field peas and beans, may not be well suited to the highly desiccating environment and periodically high temperatures. Direct field experimentation in the catchment is required to confirm this for these and other species. In the Southern Gulf catchments, mungbean and chickpea are likely to be well suited. From a land suitability perspective, pulse crops are included in Crop Group 10 (Table 4-2; Figure 4-12). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found onfloodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lowerparts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Assuming unconstrained development, approximately 4.7 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated pulse cropping (Crop Group 10; Table 4-2) using spray irrigation in the dry season. Nearly 1.7 million ha of land is considered suitable with moderate limitations for furrow irrigation in the dry season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the loamy soils are too permeable. There is potential for rainfed pulse production in the wet season over an area of about 360,000 ha. From a land suitability perspective, Crop Group 10 includes the pulse crops mungbean and chickpea, while soybean is considered under oilseeds in these crop synopses (Section 4.4.4). Pulses are often advantageous in rotation with other crops because they provide a disease break and, being legumes, can provide nitrogen for subsequent crops. Even where this is not the case, their ability to meet their own nitrogen needs can be beneficial in reducing costs of fertiliser and associated freight. Pulses such as mungbean and chickpea can also be of high value (historical prices have reached >$1000/t), so the freight costs as a percentage of the value of the crop are lower than for cereal grains. To grow pulse crops, farmers will require access to tillage, fertilising, planting, spraying and harvesting equipment. Harvesting is generally a contract operation, and in larger growing regions other activities can also be performed under contract. The equipment required for pulse crops is the same as is required for cereal crops, so farmers intending on a pulse and cereal rotation would not need to purchase extra equipment. Table 4-17 provides summary information relevant to the cultivation of many pulses using mungbean (Figure 4-13) as an example. The companion technical report on agricultural viability and socio-economics (Webster et al., 2024) provides greater detail for a wider range of crops. Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-502_Suit_Mung_Mung_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-12 Modelled land suitability for mungbean (Crop Group 10) in the dry season using (a) furrow irrigation and (b)spray irrigation These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. Themethods used to derive the reliability data in the inset maps are outlined in the companion technical report on digitalsoil mapping and land suitability (Thomas et al., 2024). Figure 4-13 Mungbean Photo: CSIRO Table 4-17 Summary information relevant to the cultivation of pulses, using mungbean as an example For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 4.4.4 Oilseed crops The area of land in Australia devoted to production of oilseeds (predominantly canola, Brassica napus) each year has varied between 2.1 and 3.4 million ha over the decade from 2012–13 to 2021–22, yielding over 8.4 Mt with a value of $6.1 billion in 2021–22 (ABARES, 2022). The majority of oilseed produced in 2021–22 (98% by value) was exported (ABARES, 2022). Canola dominates Australian oilseed production, accounting for 98% of the gross value of oilseeds in 2021–22. Soybean, sunflower (Helianthus annuus) and other oilseeds (including peanuts) each accounted for less than 1%. Soybean, canola and sunflower are oilseed crops used to produce vegetable oils and biodiesel, and high-protein meals for intensive animal production. Soybean is also used in processed foods such as tofu. It can provide both green manure and soil benefits in crop rotations, with symbiotic nitrogen fixation adding to soil fertility and sustainability in an overall cropping system. Soybean is used commonly as a rotation crop with sugarcane in northern Queensland, although often as a green manure crop. Summer oilseed crops such as soybean and sunflower are more suited to tropical environments than are winter-grown oilseed crops such as canola. Cottonseed, a by- product of cotton farming separated from the lint during ginning, is also classified as an oilseed. Cottonseed is used for animal feed and oil extraction. Soybean is sensitive to photoperiod (day length) and requires careful consideration in selection of the appropriate variety for a particular sowing window. From a land suitability perspective, soybean is included in Crop Group 10 (Table 4-2; Figure 4-14). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Assuming unconstrained development, approximately 4.7 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated soybean cropping (Crop Group 10; Table 4-2) using spray irrigation in the dry season. Nearly 1.7 million ha of land is considered suitable with moderate limitations for furrow irrigation in the dry season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the sandy and loamy soils are too permeable. There is potential for rainfed pulse production in the wet season over an area of about 360,000 ha. From a land suitability perspective, Crop Group 10 contains the pulse crops mungbean and chickpea, while soybean is considered under oilseeds in these crop synopses. To grow oilseed crops, farmers will require access to tillage, fertilising, planting, spraying and harvesting equipment. Harvesting is generally a contract operation and in larger growing regions other activities can also be performed under contract. The equipment required for oilseed crops is the same as is required for cereal crops, so farmers intending on an oilseed and cereal rotation would not need to purchase oilseed-specific equipment. With no oilseed processing facility in the north, soybean and sunflowers would need to be transported a significant distance until sufficient scales of production are achieved to justify the investment in processing facilities. Given both the modest yield and price, transport costs are likely to be a major constraint on profitability unless there is a well-developed supply chain into Asia. Table 4-18 provides summary information relevant to the cultivation of oilseed crops using soybean (Figure 4-15) as an example. The companion technical report on agricultural viability and socio-economics (Webster et al., 2024) provides greater detail for a wider range of crops. Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-503_Suit_Soy_Soy_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-14 Modelled land suitability for soybean (Crop Group 10) in the dry season using (a) furrow irrigation and (b)spray irrigation These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. Themethods used to derive the reliability data in the inset maps are outlined in the companion technical report on digitalsoil mapping and land suitability (Thomas et al., 2024). Figure 4-15 Soybean Photo: CSIRO Table 4-18 Summary information relevant to the cultivation of oilseed crops, using soybean as an example For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 4.4.5 Root crops, including peanut Root crops, including peanut, sweet potato (Ipomoea batatas) and cassava (Manihot esculenta), are potentially well suited to the lighter soils across much of the Doomadgee Plain. Root crops such as these are not suited to growing on heavier clay soils because they need to be pulled from the ground for harvest, and the heavy clay soils, such as cracking clays, are not conducive to mechanical pulling. While peanut is technically an oilseed crop, it has been included in the root crop category due to its similar land suitability and management requirements (i.e. the need for it to be pulled from the ground as part of the harvest operation). The most widely grown root crop in Australia, peanut is a legume crop that requires little or no nitrogen fertiliser and is very well suited to growing in rotation with cereal crops, as it is frequently able to fix atmospheric nitrogen in the soil for following crops. The Australian peanut industry currently produces approximately 15,000 to 20,000 t/year from around 11,000 ha, which is too small an industry to be reported separately in Australian Bureau of Agricultural and Resource Economics and Sciences statistics (ABARES, 2022). The Australian peanut industry is concentrated in Queensland. In northern Australia, a production area is present on the Atherton Tablelands, and peanuts could likely be grown in the Southern Gulf catchments. The Peanut Company of Australia established a peanut-growing operation at Katherine in 2007 and examined the potential of both wet- and dry-season peanut crops, mostly in rotation with maize. Due to changing priorities within the company, coupled with some agronomic challenges (Jakku et al., 2016), the company sold its land holdings in Katherine in 2012 (and Bega bought the rest of the company in 2018). For peanuts to be successful, considerable planning would be needed in determining the best season for production and practical options for crop rotations. The nearest peanut-processing facilities to the Southern Gulf catchments are at Tolga on the Atherton Tablelands and Kingaroy in southern Queensland. From a land suitability perspective, peanut is included in Crop Group 6 (Table 4-2; Figure 4-16). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Assuming unconstrained development, approximately 3.7 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated root crops (Crop Group 6; Table 4-2) using spray irrigation in the dry season. For spray irrigation in the wet season, about 2.0 million ha is suitable with moderate limitations (Class 3) or better. Furrow irrigation was not considered in the land suitability analysis as root crops prefer lighter-textured soils too permeable for furrow irrigation. To grow root crops, farmers will require access to tillage, fertilising, planting, spraying and harvesting equipment. The harvesting operation requires specialised equipment to ‘pull’ the crop from the ground, and then to pick it up after a drying period. Peanuts are usually dried soon after harvest, in industrial driers. Table 4-19 provides summary information relevant to the cultivation of root crops using peanut (Figure 4-17) as an example. The companion technical report on agricultural viability and socio- economics (Webster et al., 2024) provides greater detail for a wider range of crops. Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-504_Suit_Peanut_Peanut_v2.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-16 Modelled land suitability for peanut (Crop Group 6) using spray irrigation in the (a) wet season and (b) dry season These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Figure 4-17 Peanut Photo: Shutterstock Table 4-19 Summary information relevant to the cultivation of root crops, using peanut as an example For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 4.4.6 Industrial (cotton) Rainfed and irrigated cotton production are well established in Australia. The area of land devoted to cotton production varies widely from year to year, largely in response to availability of water. It varied from 70,000 to 600,000 ha between 2012–13 and 2021–22; a mean of 400,000 ha/year has been grown over the decade (ABARES, 2022). Likewise, the gross value of cotton lint production varied greatly between 2012–13 and 2021–22, from $0.3 billion in 2019–20 to $5.2 billion in 2021–22. Genetically modified cotton varieties were introduced in 1996 and now account for almost all cotton produced in Australia (over 99%). Australia was the fourth largest exporter of cotton in 2022, behind the United States, India and Brazil. Cottonseed is a by-product of cotton processing and is a valuable cattle feed. Mean lint production in Australia in 2015–16 was 8.8 bales/ha (ABARES, 2022). Commercial cotton has a long but discontinuous history of production in northern Australia, including in Broome, the Fitzroy River and the Ord River Irrigation Area in WA; in Katherine and Douglas–Daly in the NT; and near Richmond and Bowen in northern Queensland. An extensive study undertaken by the Australian Cotton Cooperative Research Centre in 2001 (Yeates, 2001) noted that past ventures suffered from: •a lack of capital investment •too rapid a movement to commercial production •a failure to adopt a systems approach to development •climate variability. Mistakes in pest control were also a major issue in early projects. Since the introduction of genetically modified cotton in 1996, yields and incomes from cotton crops have increased in most regions of Australia. The key benefits of genetically modified cotton over conventional cotton are savings in insecticide and herbicide use, and improved tillage management. In addition, farmers can now forward-sell their crop as part of a risk management strategy. Growers of genetically modified cotton are required to comply with the approved practices for growing the genetically modified varieties, including preventative resistance management. Research and commercial test farming have demonstrated that the biophysical challenges are manageable if the growing of cotton is tailored to the climate and biotic conditions of northern Australia (Yeates et al., 2013). In recent years, irrigated cotton crops achieving more than 10 bales/ha have been grown successfully in the Burdekin irrigation region and experimentally in the Gilbert catchment of northern Queensland. Expansion of cotton through private investment is occurring in the catchments of the Leichhardt, Flinders and upper Mitchell rivers, Queensland. Cotton will be processed near Katherine, NT, at a gin commissioned in 2024. New genetically modified cotton using CSIRO varieties that are both pest- and herbicide-resistant are an important component of these northern cotton production systems. Climate constraints will continue to limit production potential of northern cotton crops when compared to cotton grown in more favourable climate regions of NSW and Queensland. On the other hand, the low risk of rainfall occurring during late crop development favours production in northern Australia, as it minimises the likelihood of late-season rainfall, which can downgrade fibre quality and price. Demand for Australian cotton exhibiting long and fine attributes is expected to increase by 10% to 20% during the next decade and presents local producers with an opportunity to target production of high-quality fibre. From a land suitability perspective, cotton is included in Crop Group 7 (Table 4-2; Figure 4-18). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Assuming unconstrained development, approximately 4.7 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated cotton (Crop Group 7; Table 4-2) using spray irrigation in the dry season. For spray irrigation in the wet season, nearly 3.1 million ha is suitable with moderate limitations (Class 3) or better. Land considered suitable with moderate limitations for furrow irrigation is limited to about 1.8 million ha in the dry season and about 780,000 ha in the wet season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the loamy soils are too permeable. There is potential for rainfed cotton production in the wet season over an area of about 360,000 ha. From a land suitability perspective, Crop Group 7 contains both cotton and cereal crops; the latter are considered elsewhere in these crop synopses (Section 4.4.2). In addition to a normal row planter and spray rig equipment used in cereal production, cotton requires access to suitable picking and module or baling equipment, as well as transport to processing facilities. Decisions on initial development costs and scale of establishing cotton production in the catchments would need to consider the need to source external contractors; this could provide an opportunity to develop local contract services to support a growing industry. Cotton production is also highly dependent on access to processing plants (cotton gins). The closest processing facility for cotton grown in the Southern Gulf catchments is Emerald, Queensland. The first cotton gin in northern Australia will be processing in 2024 and is near Katherine in the NT. Niche industrial crops, such as guar (Cyamopsis tetragonoloba) and chia (Salvia hispanica), may be feasible for the Southern Gulf catchments, but verified agronomic and market data on these crops are limited. Past research on guar has been conducted in the NT, and trials are currently underway. Hemp is a photoperiod-sensitive summer annual with a growing season between 70 and 120 days depending on variety and temperature. Hemp is well suited to growing in rotation with legumes, as hemp can use the nitrogen fixed by the legume crop. Industrial hemp can be harvested for grain with modifications to conventional headers, otherwise all other farming machinery for ground preparation, fertilising and spraying can be used. There are legislative restrictions to growing hemp in Australia, and jurisdictions including the NT are implementing industrial hemp legislation to license growing of industrial hemp to facilitate development of the industry. The companion technical report on agricultural viability and socio-economics (Webster et al., 2024) provides greater detail for a wider range of industrial crops. Table 4-20 describes some key considerations relating to cotton production (Figure 4-19). Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-507_Suit_Cotton_Cotton_v2.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-18 Modelled land suitability for cotton (Crop Group 7) using furrow irrigation in the (a) wet season and (b) dry season These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Figure 4-19 Cotton Photo: CSIRO Table 4-20 Summary information relevant to the cultivation of cotton For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 4.4.7 Forages Forage, hay and silage are crops that are grown for consumption by animals. Forage is consumed in the paddock in which it is grown and is often referred to as ‘stand and graze’. Hay is cut, dried, baled and stored before being fed to animals, usually in yards for weaning or when animals are being held for sale. Silage production resembles that for hay, but harvested forage is stored wet in wrapped bales or covered ground pits, where anaerobic fermentation occurs, to preserve the feed’s nutritional value. Silage is often used as a production feed to grow animals to meet the specifications of premium markets. Rainfed and irrigated production of forage crops is well established throughout Australia, with over 20,000 producers, most of whom are not specialist producers. Approximately 85% of forage production is consumed domestically, with the rest primarily used on live export ships, often in a pelleted form. The largest consumers are the horse, dairy and beef feedlot industries. Forage crops are also widely used in horticulture for mulches and for erosion control. There is a significant fodder trade in support of the northern beef industry, with further room for expansion since fodder costs constitute less than 5% of beef production costs (Gleeson et al., 2012). The Southern Gulf catchments are suited to rainfed or irrigated production of forage, hay and silage. Rainfed and irrigated hay production currently occurs in the north-west Queensland region. Non-leguminous forage, hay and silage Forage crops, both annual and perennial, include sorghum, Rhodes grass, maize and Jarra grass (Digitaria milanjiana ‘Jarra’), with specific forage cultivars. If irrigated, these grass forages require considerable amounts of water and nitrogen as they can be high yielding (20 to 40 t dry matter per ha per year). Given the rapid growth of grass forages, crude protein levels can decrease quickly to less than 7%, reducing their value as a feed. To maintain high nutritive value (10% to 15% crude protein), high levels of nitrogen fertiliser need to be applied, and in the case of hay the crop needs to be cut every 45 to 60 days. After cutting, the crop grows back without the need for resowing. The rapid growth of forage during the wet season can make it challenging to match animal numbers to forage growth so that it is kept leafy and nutritious, and does not become rank and of low quality. Producing rainfed hay from perennials gives producers the option of irrigating when required or, if water becomes limiting, allowing the pasture to remain dormant before water again becomes available. Silage can be made from a number of crops, such as grasses, maize and forage sorghum. From a land suitability perspective, Rhodes grass is included in Crop Group 14 (Table 4-2; Figure 4-20). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Assuming unconstrained development, approximately 4.7 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated cropping of annual forages (Crop Group 12; Table 4-2) using spray irrigation in the dry season. For spray irrigation in the wet season, nearly 3.1 million ha is suitable with moderate limitations (Class 3) or better. Land considered suitable with moderate limitations for furrow irrigation is limited to about 1.8 million ha in the dry season and about 780,000 ha in the wet season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the loamy soils are too permeable. There is potential for rainfed production of annual forages in the wet season over an area of about 620,000 ha. For perennial Rhodes grass, about 5.1 million ha is suitable with moderate or minor limitations under spray irrigation and about 1.8 million ha under furrow irrigation. Apart from irrigation infrastructure, the equipment needed for forage production is machinery for planting and fertilising. Spraying equipment is also desirable but not necessary. Cutting crops for hay or silage requires more-specialised harvesting, cutting, baling and storage equipment. Table 4-21 describes Rhodes grass production (Figure 4-21) for hay over 1 year of a 6-year cycle. Information similar to that in Table 4-21 for grazed forage crops is presented in the companion technical report on agricultural viability and socio-economics (Webster et al., 2024). Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-505_Suit_Rhodes_Rhodes_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-20 Modelled land suitability for Rhodes grass (Crop Group 14) using (a) spray irrigation and (b) furrow irrigation These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Figure 4-21 Rhodes grass Photo: CSIRO Table 4-21 Rhodes grass production for hay over 1 year of a 6-year cycle For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Forage legume The use of forage legumes is similar to that of forage grasses. They are generally grazed by animals but can also be cut for silage or hay. Some forage legumes are well suited to the Southern Gulf catchments and would be considered among the more promising opportunities for irrigated agriculture (Figure 4-22). Forage legumes are desirable because of their high protein content and their ability to fix atmospheric nitrogen in the soil. The nitrogen fixed during a forage legume phase is often in excess of requirements and remains in the soil as additional nitrogen available to subsequent crops. Forage legumes are being used by the northern cattle industry, and farmers primarily engaged in extensive cattle production could use irrigated forage legumes to increase the capacity of their enterprise, turning out more cattle from the same area. Cavalcade (Centrosema pascuorum ‘Cavalcade’) and lablab are currently grown in northern Australia and would be well suited to the Southern Gulf catchments. Hay crops are commonly used as a component of forage pellets that are used to feed live export cattle in holding yards and on boats during transport. From a land suitability perspective, forage legumes such as Cavalcade and lablab are included in Crop Group 13 (Table 4-2; Figure 4-22). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet-season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Assuming unconstrained development, approximately 4.8 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) for irrigated forage legumes (Crop Group 13; Table 4-2) using spray irrigation in the dry season. For spray irrigation in the wet season, nearly 2.7 million ha is suitable with moderate limitations (Class 3) or better. Land considered suitable with moderate or minor limitations for furrow irrigation is limited to about 4.8 million ha in the dry season and about 580,000 ha in the wet season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the loamy soils are too permeable. There is potential for rainfed forage legume production in the wet season over an area of about 220,000 ha. The equipment needed for grazed forage legume production is similar to that for forage grasses: a planting method, with fertilising and spraying equipment, is desirable but not essential. Cutting crops for hay or silage requires more-specialised harvesting, cutting, baling and storage equipment. Table 4-22 describes Cavalcade production over a 1-year cycle. The comments could be applied equally to lablab production (Figure 4-23). Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-506_Suit_Cavalcade_Cavalcade_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-22 Modelled land suitability for Cavalcade (Crop Group 13) in the wet season using (a) spray irrigation and (b) furrow irrigation These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Table 4-22 Cavalcade production over a 1-year cycle For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Figure 4-23 Lablab Photo: CSIRO 4.4.8 Horticulture Intensive horticulture is an important and widespread industry in Australia, occurring in every state, particularly close to capital city markets. Horticultural production varied between 2.9 and 3.3 Mt/year between 2012–13 and 2021–22, of which 65% to 70% was vegetables (ABARES, 2022). Unlike broadacre crops, most horticultural production in Australia is consumed domestically. The total gross value of horticultural production was $13.2 billion in 2021–22 (up from $9.3 billion in 2012–13), of which 24% was from exports (ABARES, 2022). Horticulture is also an important source of jobs, employing approximately a third of all people working in agriculture. Horticultural production is more intensive than broadacre production and has a higher degree of risk, such as a short season of supply and highly volatile prices as a result of highly inelastic supply and demand. Managing these issues requires a heightened understanding of risks, markets, transport and supply chain issues (including associated interactions with other horticultural production regions). Production is highly seasonal and can involve multiple crops produced on individual farms to manage labour resources. The importance of freshness in many horticultural products means seasonality of supply is important in the market. Farms in the Southern Gulf catchments have the advantage of being able to produce out-of-season supplies to southern markets. However, they must also compete with production regions in the NT and northern WA, which are already established production areas with associated infrastructure. Southern Gulf catchments may have an advantage over these regions in being geographically closer to most of the urban consumer centres of south-eastern Australia. Horticulture (row crops) Horticultural row crops are generally short-lived, annual crops, grown in the ground, such as seedless watermelons (Citrullus lanatus), rockmelon and honeydew melon (Cucumis melo), as well as sweet corn (Zea mays). Almost all produce is shipped to capital cities where major central markets are located. Row crops such as watermelon and rockmelon use staggered plantings over a season (e.g. planted every 2 to 3 weeks) to extend the period over which harvested produce is sold. This strategy allows better use of labour and better management for risks of price fluctuations. Often only a short period of time with very high prices is enough to make melon production a profitable enterprise. From a land suitability perspective, intensive horticulture row crops such as rockmelon are included in Crop Group 3 (Table 4-2). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet-season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. A wide range of horticultural row crops are considered in the land suitability analysis (crop groups 3, 4, 5, 6 and 18; Table 4-2; Figure 4-24). Assuming unconstrained development, between about 3.4 million ha and 4.9 million ha of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) using spray or trickle irrigation in the dry season. Land considered suitable with moderate limitations for furrow irrigation of sweet corn (Crop Group 18) is limited to about 1.7 million ha in the dry season and only 780,000 ha in the wet season, due to inadequate soil drainage in clay soils (and/or because gilgais are too deep) and because the loamy soils are too permeable. Horticultural row crops are well established throughout the NT, Burdekin and Mareeba–Dimbulah Water Supply Scheme region in Queensland. The NT melon industry, consisting of watermelon (seedless), rockmelon and honeydew, produces approximately 25% of Australia’s melons. Melon production would be well suited to the Southern Gulf catchments, which could compete with NT production. Horticulture typically requires specialised equipment and a large labour force. Therefore, a system for attracting, managing and retaining sufficient staff is also required. Harvesting is often by hand, but packing equipment is highly specialised. Irrigation is generally with micro or trickle equipment, but overhead spray is also feasible. Leaf fungal diseases need to be carefully managed when using spray irrigation. Micro spray equipment has the advantage of being able to deliver fertiliser along with irrigation. Table 4-23 describes some key considerations relating to row crop horticulture production, with rockmelon (Figure 4-25) as an example. Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-508_Suit_Rockmelon_Onion_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-24 Modelled land suitability for (a) cucurbits (e.g. rockmelon, Crop Group 3) using trickle irrigation in the dry season and (b) root crops such as onion (Crop Group 6) using spray irrigation in the wet season These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Rockmelon For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-25 Rockmelon Photo: Shutterstock Table 4-23 Summary information relevant to row crop horticulture production, with rockmelon as an example For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Horticulture (tree crops) Some fruit and tree crops, such as mangoes and citrus (Citrus spp.), are well suited to the climate of the Southern Gulf catchments. Other species, such as avocado (Persea americana) and lychee (Litchi chinensis), are not likely to be as well adapted to the climate due to high temperatures and low humidity. Tree crops are generally not well suited to cracking clays, which make up some of the arable soils for irrigated agriculture in the Southern Gulf catchments. Horticultural tree production is more feasible on the lighter, well-drained soils in the north-west of the Southern Gulf catchments. Fruit production shares many of the marketing and risk features of horticultural row crops, such as a short season of supply and highly volatile prices as a result of highly inelastic supply and demand. Managing these issues requires a heightened understanding of risks, markets, transport and supply chain issues. The added disadvantage of fruit tree production is the time lag between planting and production, meaning decisions to plant need to be made with a long time frame for production and return in mind. Mango production in the NT is buffered somewhat against large- scale competition as its crop matures earlier than the main production areas in Queensland, and it can achieve high returns. Mango production in the NT had a gross value of $129 million in 2020, accounting for 38% of the $341 million total value of horticultural production in the NT and half of all mangoes produced in Australia (Sangha et al., 2022). The perennial nature of tree crops makes a reliable year-round supply of water essential. Some species, such as mango and cashew (Anacardium occidentale), can survive well under mild water stress until flowering. It is critical for optimum fruit and nut production that trees are not water stressed from flowering through to harvest, approximately from June to between November and February, depending on plant species and variety. This is a period in the Southern Gulf catchments when very little rain falls, and farmers would need to have a system in place to access reliable irrigation water during this time. High night-time minimum temperatures can reduce flowering in mangoes, although potential production regions in Southern Gulf catchments should not experience these temperatures extremes. From a land suitability perspective, intensive horticultural tree crops such as mango are included in Crop Group 1, the monsoonal tropical tree crops (Table 4-2). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet-season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. A wide range of horticultural tree crops are considered in the land suitability analysis (crop groups 1, 2, 20 and 21; Table 4-2; Figure 4-26). Assuming unconstrained development, between about 860,000 ha (papaya/cashew/macadamia) and 3.9 million ha (e.g. mango) of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) using spray or trickle irrigation. Furrow irrigation was not considered for horticultural tree crops. Specialised equipment is required for fruit and nut tree production. The requirement for a timely and significant labour force necessitates a system for attracting, managing and retaining sufficient staff. In a remote location the cost of providing accommodation to such staff may be significant. Tree-pruning and packing equipment is highly specialised for the fruit industry, as are the micro irrigation systems typically used in horticulture. Table 4-24 describes some key considerations relating to mango production (Figure 4-27) in the Southern Gulf catchments, as an exemplar of the considerations relating to tree crop production more broadly. Similar information for other fruit tree crops is described in the companion technical report on agricultural viability and socio-economics (Webster et al., 2024). Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-509_Suit_Mango_Lime_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-26 Modelled land suitability for (a) mango (Crop Group 1) and (b) lime (Crop Group 2), both grown using trickle irrigation These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Figure 4-27 Mango Photo: Shutterstock Table 4-24 Summary information relevant to tree crop horticulture production, with mango as an example For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au PVR = plant variety rights. 4.4.9 Plantation tree crops (silviculture) Of the plantation tree crops that could be grown in the Southern Gulf catchments, Indian sandalwood (Santalum album) and African mahogany (Khaya spp.) are likely to be the most economically feasible. Many other plantation species could be grown but returns are much lower than for sandalwood or African mahogany. African mahogany is well established in plantations near Katherine and in north Queensland. Indian sandalwood is grown in the Ord River Irrigation Area (WA), around Katherine (NT) and in northern Queensland. Plantation tree crops require over 15 years to mature, but once established they can tolerate prolonged dry periods. Irrigation water is critical in the establishment and in the first 2 years of a plantation for a number of species. In the case of Indian sandalwood (which is a hemi root parasite), the provision of water is for not only the trees themselves but also the leguminous host plant. From a land suitability perspective, plantation tree crops such as Indian sandalwood, African mahogany and teak (Tectona grandis) are included in crop groups 15, 16 and 17 (Table 4-2). Cracking clay soils (Vertosols) make up 23% of the catchment; they are principally found on floodplains and alluvial plains of the Armraynald Plain and Barkly Tableland physiographic units. Flooding, access and trafficability in the wet season are common constraints across the lower parts of the Armraynald Plain and crop tolerance to poor soil drainage conditions restricts wet- season cropping in these areas. Effective rooting depth is deep to very deep (1.2 to 1.5 m) and the clay texture means the soils have a very high (>220 mm) soil AWC. Much of this area is suitable (with moderate or minor limitations) for spray irrigation in the dry season, but inadequate drainage in the wet season substantially reduces the area suitable for wet-season spray irrigation. Sandy soils have formed on the Doomadgee Plain (marked S1 and S4 on Figure 2-5) and the Gulf Fall (marked S3 on Figure 2-5). In total, the red, brown, yellow and grey sandy soils make up 10% of the area. Friable non-cracking clay or clay loam soils found along the middle reaches of the Leichhardt River make up only about 3% of the area but have potential for agriculture, as do the loamy soils (less than 3% of the area) on the Nicholson River, the Doomadgee and Cloncurry plains and other isolated areas. Shallow and/or rocky soils make up 56% of the catchments and are unsuitable by definition. Depending on the specific tree species being planted and their tolerance to poorly drained soils and waterlogging, the suitable areas vary considerably. A range of silviculture trees were considered in the land suitability analysis (crop groups 15, 16 and 17; Table 4-2). Assuming unconstrained development, between about 3.1 million ha (teak) and 5.1 million ha (African mahogany) of the Southern Gulf catchments is considered to be suitable with moderate limitations (Class 3; Table 4-1) or better (Class 2 or Class 1) using trickle irrigation (Figure 4-28). Furrow irrigation was considered for Indian sandalwood only and about 810,000 ha was assessed as suitable with moderate limitations. Table 4-25 describes Indian sandalwood production (Figure 4-29). Suitability map for crop synopses \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\10_Reporting\4_S_Gulf\1_GIS\1_Map_Docs\1_Export\CR-S-Ch4-510_Suit_IndSand_IndSand_v1.png For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-28 Modelled land suitability for Indian sandalwood (Crop Group 15) grown using (a) trickle or (b) furrow irrigation These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the reliability data in the inset maps are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Indian sandalwood and host plants For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-29 Indian sandalwood and host plants Indian sandalwood trees are those with a darker trunk and leaves, in a line left of centre in the image. Photo: CSIRO Table 4-25 Summary information for Indian sandalwood production For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au 4.4.10 Niche crops Niche crops such as guar, chia, quinoa (Chenopodium quinoa), bush products and others may be feasible in the Southern Gulf catchments, but limited verified agronomic or market data are available for these crops. Niche crops are niche due to the limited demand for their products. As a result, small-scale production can lead to very attractive prices, but only a small increase in productive area can flood the market, leading to greatly reduced prices and making production unsustainable. There is growing interest in bush products but insufficient publicly available information for inclusion with the analyses of irrigated crop options in this report. Bush product production systems could take many forms, from culturally appropriate wild harvesting targeting Indigenous cultural and environmental co-benefits to intensive mechanised farming and processing, resembling something like macadamia (Macadamia integrifolia) farming, with multiple possible combinations and variants in between. The choice of production system would have implications for the extent of Indigenous participation in each stage of the supply chain (farming, processing, marketing and/or consumption), the co-benefits that could be achieved, the scale of the markets that could be accessed (in turn affecting the scale of the industry for that bush product), the price premiums that produce may be able to attract and the viability of those industries. The current publicly available information on bush products mainly focuses on eliciting Indigenous aspirations, biochemical analysis (for safety, nutrition and efficacy of potential health benefits of botanicals), and considerations of safeguarding Indigenous intellectual property (e.g. Woodward et al., 2019). Analysing bush products in a comparable way to other crop options in this report would first require these issues to be resolved, for communities to agree on the preferred type of production systems (and pathways for development), and for agronomic information on yields, production practices and costs to be publicly available. Past research on guar has been conducted in the NT, and trials are underway in northern Queensland, which could prove future feasibility. There is increasing interest in non-leguminous, small-seeded crops such as chia and quinoa, which have high nutritive value. The market size for these niche crops is quite small compared with cereals and pulses, so the scale of production is likely to be small in the short to medium term. There is a small, established chia industry in the Ord River Irrigation Area of WA, but its production and marketing statistics are largely commercial-in-confidence. Nearly all Australian production of chia is contracted to The Chia Company of Australia or is exported to China. In Australia, The Chia Company produces whole chia seeds, chia bran, ground chia seed and chia oil for wholesale and retail sale, and it exports these products to 36 countries. The growing popularity of quinoa in recent years is attached to its marketing as a superfood. It is genetically diverse and has not been the subject of long-term breeding programs. This diversity means it is well suited to a range of environments, including northern Australia, where its greatest opportunity is as a short-season crop in the dry season under irrigation. It is a high-value crop with farm gate prices of about $1000/t. Trials of quinoa production have been conducted at the Katherine Research Station in the NT (approximately 600 km north-west of the Southern Gulf catchments), with reasonable yields being returned. More trials are required in the various northern environments before quinoa could be recommended for commercial production. 4.5 Aquaculture 4.5.1 Introduction There are considerable opportunities for aquaculture development in northern Australia given its natural advantages of a climate suited to farming valuable tropical species, large areas identified as suitable for aquaculture, political stability and proximity to large global markets. The main challenges to developing and operating modern and sustainable aquaculture enterprises are regulatory issues, global cost competitiveness and the remoteness of much of the suitable land area. A comprehensive situational analysis of the aquaculture industry in northern Australia (Cobcroft et al., 2020) identifies key challenges, opportunities and emerging sectors. This section draws on a recent assessment of the opportunities for aquaculture in northern Australia in the Northern Australia Water Resource Assessment technical report on aquaculture (Irvin et al., 2018), summarising the three most likely candidate species (Section 4.5.2), overviewing production systems (Section 4.5.3), land suitability for aquaculture within the Southern Gulf catchments (Section 4.5.4) and the financial viability of different options for aquaculture development (Section 4.5.5). 4.5.2 Candidate species The three species with the most aquaculture potential in the catchments of the Southern Gulf rivers are black tiger prawns (Penaeus monodon), barramundi (Lates calcarifer), and red claw (Cherax quadricarinatus). The first two species are suited to many marine and brackish water environments of northern Australia and have established land-based culture practices and well- established markets for harvested products. Prawns could potentially be cultured in either extensive (low density, low input) or intensive (higher density, higher input) pond-based systems in northern Australia, whereas land-based culture of barramundi would likely be intensive. Red claw is a freshwater crayfish that is currently cultured by a much smaller industry than the other two species. Black tiger prawns Black tiger prawns (Figure 4-30) are found naturally at low abundances across the waters of the western Indo-Pacific region, with wild Australian populations making up the southernmost extent of the species. Within Australia, the species is most common in the tropical north, but does occur at lower latitudes. Figure 4-30 Black tiger prawns Photo: CSIRO Barramundi Barramundi (Figure 4-31) is the most highly produced and valuable tropical fish species in Australian aquaculture. Barramundi inhabit the tropical north of Australia from the Exmouth Gulf in WA through to the Noosa River on Queensland’s east coast. It is also commonly known as the ‘Asian sea bass’ or ‘giant sea perch’ throughout its natural areas of distribution in the Persian Gulf, the western Indo-Pacific region and southern China (Schipp et al., 2007). The attributes that make barramundi an excellent aquaculture candidate are fast growth (reaching 1 kg or more in 12 months), year-round fingerling availability, well-established production methods and hardiness (i.e. they have a tolerance to low oxygen levels, high stocking densities and handling, as well as a wide range of temperatures) (Schipp et al., 2007). In addition, barramundi are euryhaline (able to thrive and be cultured in fresh and marine water), but freshwater barramundi can have an earthy flavour. Figure 4-31 Barramundi Photo: CSIRO Red claw Red claw is a warm-water crayfish species that inhabits still or slow-moving water bodies. The natural distribution of red claw is from the tropical catchments of Queensland and the NT to southern New Guinea. The name ‘red claw’ is derived from the distinctive red markings present on the claws of the male crayfish. The traits of red claw that make them attractive for aquaculture production are a simple life cycle, which is beneficial because complex hatchery technology is not required (Jones et al., 1998); their tolerance of low oxygen levels (<2 mg/L), which is beneficial in terms of handling, grading and transport (Masser and Rouse, 1997); their broad thermal tolerance, with optimal growth achievable between 23 and 31 °C; and their ability to remain alive out of water for extended periods. 4.5.3 Production systems Overview Aquaculture production systems can be broadly classified into extensive, semi-intensive and intensive systems. Intensive systems require high inputs and expect high outputs: they require high capital outlay and have high running costs; they require specially formulated feed and specialised breeding, water quality and biosecurity processes; and they have high production per hectare (in the order of 5000 to 20,000 kg per ha per crop). Semi-intensive systems involve stocking seed from a hatchery, routine provision of a feed, and monitoring and management of water quality. Production is typically 1000 to 5000 kg per ha per crop. Extensive systems are characterised by low inputs and low outputs: they require less-sophisticated management and often require no supplementary feed because the farmed species live on naturally produced feed in open-air ponds. Extensive systems produce about half the volume of global aquaculture production, but there are few commercial operations in Australia. Water salinity and temperature are the key parameters that determine species selection and production potential for any given location. Suboptimal water temperature (even within tolerable limits) will prolong the production season (because of slow growth) and increase the risk of disease, reducing profitability. The primary culture units for land-based farming are purpose-built ponds. Pond structures typically include an intake channel, production pond, discharge channel and a bioremediation pond (Figure 4-32). The function of the pond is as a containment structure – an impermeable layer between the pond water and the local surface water and groundwater. Optimal sites for farms are flat and have sufficient elevation to enable ponds to be completely drained between seasons. It is critical that all ponds and channels can be fully drained during the off (dry-out) season to enable machinery access to sterilise and undertake pond maintenance. Figure 4-32 Schematic of marine aquaculture farm Most production ponds in Australia are earthen. Soils for earthen ponds should have low permeability and high structural stability. Ponds should be lined if the soils are permeable. Synthetic liners have a higher capital cost but are often used in more-intensive operations, which require high levels of aeration – conditions that would lead to significant erosion in earthen ponds. Farms use aerators (typically electric paddlewheels and aspirators) to help maintain optimal water quality in the pond, provide oxygen and create a current that consolidates waste into a central sludge pile (while keeping the rest of the pond floor clear). A medium-sized (50 ha) prawn farm in Australia uses around 4 GWh annually, accounting for most of an enterprise’s energy use (Paterson and Miller, 2013). Backup power capacity sufficient to run all the aerators on the farm, usually with a diesel generator, is essential to be able to cope with power failures. Extensive production systems do not require aeration in most cases. Black tiger prawns A typical pond in the Australian black tiger prawn industry is rectangular, about 1 ha in area and about 1.5 m in depth. The ponds are either wholly earthen, lined on the banks with black plastic and earthen bottoms or (rarely in Australia) fully lined. Pond grow-out of black tiger prawns typically operates at stocking densities of 25 to 50 individuals per square metre (termed ‘intensive’ in this report). These pond systems are fitted with multiple aeration units, which could double from 8 to 16 units as the biomass of the prawn crop increases (Mann, 2012). At the start of each prawn crop, pond bottoms are dried, and unwanted sludge from the previous crop is removed. If needed, additional substrate is added. Before filling the ponds, lime is often added to buffer pH, particularly in areas with acid-sulfate soils. The ponds are then filled with filtered seawater and left for about 1 week prior to postlarval stocking. Algal blooms in the water are encouraged through addition of organic fertiliser to provide shading for prawns, discourage benthic algal growth and stimulate growth of plankton as a source of nutrition (QDPIF, 2006). Postlarvae are purchased from hatcheries and grow rapidly into small prawns in the first month after stocking, relying mainly on the natural productivity (zooplankton, copepods and algae) supported by the algal bloom for their nutrition. Approximately 1 month after the prawns are stocked, pellet feed becomes the primary nutrition source. Feed is a major cost of prawn production: around 1.5 kg of feed is required to produce 1 kg of prawns. Prawns typically reach optimal marketable size (30 g) within 6 months. After harvest, prawns are usually processed immediately, with larger farms having their own production facilities that enable grading, cooking, packaging and freezing. Effective prawn farm management involves maintaining optimal water quality conditions, which becomes progressively complex as prawn biomass and the quantity of feed added to the system increase. As prawn biomass increases, so too does the biological oxygen demand of the microbial population within the pond that is breaking down organic materials. This requires increases in mechanical aeration and water exchanges (either fresh or recycled from a bioremediation pond). In most cases water salinity is not managed, except through seawater exchange, and will increase naturally with evaporation and decrease with rainfall and flooding. Strict regulation of the quality and volume of water that can be discharged means efficient use of water is standard industry practice. Most Australian prawn farms allocate up to 30% of their productive land for water treatment by pre-release containment in settlement systems. Barramundi The main factors that determine productivity of barramundi farms are water temperature, dissolved oxygen levels, effectiveness of waste removal, expertise of farm staff and the overall health of the stock. Barramundi are susceptible to a variety of bacterial, fungal and parasitic organisms. They are at highest risk of disease when exposed to suboptimal water quality conditions (e.g. low oxygen or extreme temperatures). Due to the cost and infrastructure required, many producers elect to purchase barramundi fingerlings from independent hatcheries, moving fish straight into their nursery cycle. Regular size grading is essential during the nursery stage to minimise aggressive and cannibalistic behaviour: size grading helps to prevent mortalities and damage from predation on smaller fish, and it assists with consistent growth. Ponds are typically stocked to a biomass of about 3 kg per 1000 L. Under optimal conditions barramundi can grow to over 1 kg in 12 months and to 3 kg within 2 years (Schipp et al., 2007). The two largest Australian aquafeed manufacturers (located in Brisbane and Hobart) each produce a pellet feed that provides a specific diet promoting efficient growth and feed conversion. The industry relies heavily on these mills to provide a regular supply of high-quality feed. Cost of feed transport would be a major cost to barramundi production in the Southern Gulf catchments. As a carnivorous species, high dietary protein levels, with fishmeal as a primary ingredient, are required for optimal growth. Barramundi typically require between 1.2 and 1.5 kg of pelleted feed for each kilogram of body weight produced. Warm water temperatures in northern Australia enable fish to be stocked in ponds year round. Depending on the intended market, harvested product is processed whole or as fillets and delivered fresh (refrigerated or in ice slurry) or frozen. Smaller niche markets for live barramundi are available for Asian restaurants in some capital cities. Red claw Water temperature and feed availability are the variables that most affect crayfish growth. Red claw are a robust species but are most susceptible to disease (including viruses, fungi, protozoa and bacteria) when conditions in the production pond are suboptimal (Jones, 1995). In tropical regions, mature females can be egg-bearing year round. Red claw breed freely in production ponds, so complex hatchery technology (or buying juvenile stock) is not required. However, low fecundity and the associated inability to source high numbers of quality selected broodstock are an impediment to intensive expansion of the industry. Production ponds are earthen, rectangular in design and on average 1 ha in size. They slope in depth from 1.2 to 1.8 m. Sheeting is used on the pond edge to keep the red claw in the pond (they tend to migrate), and netting surrounds the pond to protect stock from predators (Jones et al., 2000). At the start of each crop, ponds are prepared (as for black tiger prawns above), then filled with fresh water and left for about 2 weeks before stocking. During this period, algal blooms in the water are encouraged through addition of organic fertiliser. Ponds are then stocked with about 250 females and 100 males that have reached sexual maturity. Natural mating results in the production of around 20,000 advanced juveniles. Red claw are omnivorous, foraging on natural production such as microbial biomass associated with decaying plants and animals. Early-stage crayfish rely almost solely on natural pond productivity (phytoplankton and zooplankton) for nutrition. As the crayfish progress through the juvenile stages, the greater part of the diet changes to organic particulates (detritus) on the bottom of the pond. Very small quantities of a commercial feed are added daily to assist with the weaning process and provide an energy source for the pond bloom. Providing adequate shelters (net bundles) is essential at this stage to improve survival (Jones, 2007). Approximately 4 months after stocking, the juveniles are harvested and graded by size and sex for stocking in production ponds. Juveniles are stocked in production ponds at 5 to 10 per square metre. Shelters are important during the grow-out stage, with 250/ha recommended. During the grow-out phase, pellet feed becomes an important nutrition source, along with the natural productivity provided by the pond. Current commercial feeds are low cost and provide a nutrition source for natural pond productivity as much as for the crayfish. Most Australian farmers use diets consisting of 25% to 30% protein. Effective farm management involves maintaining water quality conditions within ranges optimal for crayfish growth and survival as pond biomass increases. As with barramundi, management involves increasing aeration and water exchanges, while strictly managing effluent discharges. Red claw are harvested within 6 months of stocking to avoid reproduction in the production pond. At this stage the crayfish will range from 30 to 80 g. Stock are graded by size and sex into groups for market, breeding or further grow-out (Jones, 2007). Estimated water use An average crop of prawns farmed in intensive pond systems (8 t/ha over 150 days) is estimated to require 127 ML of marine water, which equates to 15.9 ML of marine water for each tonne of harvested product (Irvin et al., 2018). For pond culture of barramundi (30 t/ha over 2 years), 562 ML of marine water, or fresh water, is required per crop, equating to 18.7 ML of water for each tonne of harvested fish. For extensive red claw culture (3 t/ha over 300 days), 240 ML of fresh water is required per pond crop, equating to 16 ML of water for each harvested tonne of crayfish (Irvin et al., 2018). 4.5.4 Aquaculture land suitability The suitability of areas for aquaculture development was also assessed from the perspective of soil and land characteristics using the set of five land suitability classes in Table 4-1. The limitations considered include clay content, soil surface pH, soil thickness and rockiness. Limitations mainly relate to geotechnical considerations (e.g. construction and stability of impoundments). Other limitations, including slope, and the likely presence of gilgai microrelief and acid-sulfate soils, are indicative of more difficult, expensive and therefore less suitable development environments, and a greater degree of land preparation effort. More detail can be found in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). Suitability was assessed for lined and earthen ponds, with earthen ponds requiring soil properties that prevent pond leakage. Soil acidity (pH) was also considered for earthen ponds, as some aquaculture species can be affected by unfavourable pH values exchanged into the water column (i.e. biological limitation). Two aquaculture species were selected to represent the environmental needs of marine species (represented by prawns) and freshwater species (red claw). Additionally, barramundi and other euryhaline species, which can tolerate a range of salinity conditions, may be suited to either marine or fresh water, depending on management choices. Except for aquaculture of marine species, which for practical purposes is restricted by proximity to sea water, no consideration was given in the analysis to proximity to suitable water for aquaculture of fresh and euryhaline species. It was not possible to include proximity to fresh water due to the large number of potential locations where water could be captured and stored within the catchments. Note also that the estimates for land suitability presented below represent the total areas of the catchments unconstrained by factors such as water availability, land tenure, environmental and other legislation and regulations, and a range of biophysical risks such as cyclones and flooding. These are addressed elsewhere by the Assessment. The land suitability maps are designed to be used predominantly at the regional scale. Planning at the enterprise scale would demand more localised assessment. Analysis of suitability of land for marine aquaculture has been restricted to locations within 2 km of a marine water source. Marine aquaculture land suitability is shown in Figure 4-33 and presents suitability across the areas under tidal influence and river margins where cracking clay (SGG 9) and seasonally or permanently wet soils (SGG 3) dominate. These soils show the desired land surface characteristics such as no rockiness, suitable slope and sufficient soil thickness, but they have the risk of acid-sulfate soils and must be managed accordingly. Suitable land for marine aquaculture in lined ponds (Figure 4-33a) totals 300,206 ha (2.8% of the catchments) and is restricted to the Karumba Plain physiographic unit where SGG 3 (seasonally or permanently wet) soils dominate, representing largely Class 2 land (86,000 ha, 0.8%). The suitable area extends into some of the most downstream areas of the Armraynald Plain physiographic unit, where tidal influence is still felt, and coincides with the presence of SGG 9 soils (cracking clays). The land suitability patterns for marine species in earthen ponds (Figure 4-33b) closely mirror those of the marine lined ponds, although areas are restricted to slowly permeable cracking clay soils. Approximately 193,600 ha (1.8% of the catchments) is mapped as suitability Class 3, where the possibility for earthen ponds depends on soil factors including sufficient depth, low soil permeability and heavier surface textures. Marine aquaculture suitability map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-535_Aquaculture_marine_v1_Arc10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-33 Land suitability in the Southern Gulf catchments for marine species aquaculture in (a) lined ponds and (b) earthen ponds These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the suitability data are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). The aquaculture land suitability analyses for freshwater species do not consider availability of fresh water for production, only soil and land attributes (Figure 4-34). This shows that a significant proportion of the catchments is suitable for freshwater aquaculture in lined ponds (6,265,400 ha, 57.9%; Figure 4-34a), with the unsuitable areas associated with higher slopes and shallow and/or rocky soils (SGG 7). The suitable area includes the low slope, deep and non-rocky parts of all SGGs except SGG 7, and the majority of the area is Class 2 (suitable with minor limitations) 5,154,300 ha (47.6%), with smaller proportions of Class 1 (40,300 ha, 0.4%) and Class 3 (1,070,800 ha, 9.9%). In comparison, opportunities for freshwater species in earthen ponds in the Assessment area are more restricted: 2,408,273 ha (22.3%) of which only 170 ha is Class 2 (Figure 4-34b). Shallow and/or rocky (SGG 7) and moderately to highly permeable soils are unsuited to earthen water impoundments. The suitable areas match the cracking clay soils (SGG 9) distribution as these soils provide the necessary soil conditions including depth, slower permeability and clay textured surface soils. There are also significant areas on the Karumba Plain of slowly permeable seasonally or permanently wet (SGG 3) and cracking clay (SGG 9) soils. These coastal plains have potential acid-sulfate soils that would require appropriate management. Freshwater aquaculture suitability map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\3_Land_suitability\4_S_Gulf\1_GIS\1_Map_docs\LL-S-534_Aquaculture_fresh_v2_Arc10_8.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 4-34 Land suitability in the Southern Gulf catchments for freshwater species aquaculture in (a) lined ponds and (b) earthen ponds These land suitability maps do not consider flooding, risk of secondary salinisation or availability of water. The methods used to derive the suitability data are outlined in the companion technical report on digital soil mapping and land suitability (Thomas et al., 2024). 4.5.5 Aquaculture viability This section provides a brief, generic analysis of what would be required for new aquaculture developments in the Southern Gulf catchments to be financially viable. First, indicative costs are provided for a range of four possible aquaculture enterprises that differ in species farmed, scale and intensity of production. The cost structure of the enterprises was based on established tools available from the Queensland Government for assessing the performance of existing or proposed aquaculture businesses (Queensland Government, 2024). Based on the ranges of these indicative capital and operating costs, gross revenue targets that a business would need to attain to be commercially viable are then calculated. Enterprise-level costs for aquaculture development Costs of establishing and running a new aquaculture business are divided here into the initial capital costs of development and ongoing operating costs. The four enterprise types analysed were chosen to portray some of the variation in cost structures between potential development options, not as a like-for-like comparison between different types of aquaculture (Table 4-26). Table 4-26 Indicative capital and operating costs for a range of generic aquaculture development options Costs are provided both per hectare of grow-out pond and per kilogram of harvested produce, although capital costs scale mostly with the area developed, and operating costs scale mainly with crop yield at harvest. Capital costs have been converted to an equivalent annualised cost assuming a 10% discount rate and that a quarter of the developed infrastructure was for 15-year life span assets and the remainder for 40-year life span assets. Indicative breakdowns of cost components are provided on a proportional basis. PARAMETER UNIT PRAWN (EXTENSIVE) PRAWN (INTENSIVE) BARRAMUNDI RED CLAW (SMALL SCALE) Scale of development Grow-out pond area ha 20 100 30 4 Total farm area ha 25 150 100 10 Yield at harvest t/y 30 800 600 32 Yield at harvest per pond area t/ha/y 1.5 8.0 20.0 3.0 Capital costs of development (scale with area of grow-out ponds developed) Land and buildings % 56 26 23 30 Vehicles % 5 2 2 11 Pond-related assets % 27 67 70 41 Other infrastructure and equipment % 11 6 5 17 Total capital cost (year 0) $/ha 74,000 142,000 147,000 163,000 Equivalent annualised cost $/kg 5.41 1.94 0.81 5.95 $/ha/y 8,108 15,558 16,106 17,859 Operating costs (vary with yield at harvest, except overheads) Nursery/juvenile costs % 12 9 7 1 Feed costs % 0 26 30 8 Labour costs % 47 13 12 57 Electricity costs % 16 24 30 9 Packing costs % 2 4 3 2 Transport costs % 6 16 16 11 Overhead costs (fixed) % 17 8 1 12 Total annual operating costs $/kg 19.31 12.47 12.46 17.80 $/ha/y 28,966 99,783 249,211 53,402 Total costs of production Total annual cost $/kg 24.72 14.42 13.27 23.75 $/ha/y 37,100 115,300 265,300 71,300 Capital costs include all land development costs, construction, and plant and equipment accounted for in the year production commences. The types of capital development costs are largely similar across the aquaculture options, with costs of constructing ponds and buildings dominating the total initial capital investment. Indicative costs were derived from the case study of Guy et al. (2014), and consultation with experts familiar with the different types of aquaculture, including updating to December 2023 dollar values (Table 4-26). Operating costs cover both overheads (which do not change with output) and variable costs (which increase as the yield of produce increases). Fixed overhead costs in aquaculture are a relatively small component of the total costs of production. Overheads consist of costs relating to licensing, approvals and other administration (Table 4-26). The remaining operating costs are variable (Table 4-26). Feed, labour and electricity typically dominate the variable costs. Aquaculture requires large volumes of feed inputs, and the efficiency with which this feed is converted to marketed produce is a key metric of business performance. Labour costs consist of salaries of permanent staff and casual staff who are employed to cover intensive harvesting and processing activities. Aerators require large amounts of energy, increasing as the biomass of produce in the ponds increases, which accounts for the large costs of electricity. Transport, although a smaller proportional cost, is important because this puts remote locations at a disadvantage relative to aquaculture businesses that are closer to feed suppliers and markets. In addition, transport costs may be higher at times if roads are cut (requiring much more expensive air freight or alternative, longer road routes) or if the closest markets become oversupplied. Packing is the smallest component of variable costs in the breakdown categories used here. Revenue for aquaculture produce typically ranges from $10 to $20 per kg (on a harvested mass basis), but prices vary depending on the quality and size classes of harvested animals and how they are processed (e.g. live, fresh, frozen or filleted). Farms are likely to deliver a mix of products targeted to the specifications of the markets they supply. Note that the mass of sold product may be substantially lower than the harvested product (e.g. fish fillets are about half the mass of harvested fish), so prices of sold product may not be directly comparable to the costs of production in Table 4-26, which are on a harvest mass basis. Commercial viability of new aquaculture developments Capital and operating costs differ between different types of aquaculture enterprises (Table 4-27), but these costs may differ even more between locations (depending on case-specific factors such as remoteness, soil properties, distance to water source and type of power supply). Furthermore, there can be considerable uncertainty in some costs, and prices paid for produce can fluctuate substantially over time. Given this variation among possible aquaculture developments in the Southern Gulf catchments, a generic approach was taken to determine what would be required for new aquaculture enterprises to become commercially viable. The approach used here was to calculate the gross revenue that an enterprise would have to generate each year to achieve a target internal rate of return (IRR) for given operating costs and development costs (both expressed per hectare of grow-out ponds). Capital costs were converted to annualised equivalents on the assumption that developed assets equated to a mix of 25% 15-year assets and 75% assets with a 40-year life span (using a discount rate matching the target IRR). The target gross revenue is the sum of the annual operating costs and the equivalent annualised cost of the infrastructure development (Table 4-27). Table 4-27 Gross revenue targets required to achieve target internal rates of return (IRR) for aquaculture developments with different combinations of capital costs and operating costs All values are expressed per hectare of grow-out ponds in the development. Gross revenue is the yield per hectare of pond multiplied by the price received for produce (averaged across products and on a harvest mass basis). Capital costs were converted to an equivalent annualised cost assuming a quarter of the developed infrastructure was for 15-year life span assets and the remainder for 40-year life span assets. Targets would be higher after taking into account risks such as initial learning and market fluctuations. OPERATING COSTS ($/ha/y) GROSS REVENUE REQUIRED TO ACHIEVE TARGET IRR ($/ha/y) Capital costs of development ($/ha) 60,000 70,000 80,000 90,000 100,000 110,000 125,000 150,000 175,000 7% target IRR 20,000 25,022 25,859 26,696 27,533 28,371 29,208 30,463 32,556 34,648 50,000 55,022 55,859 56,696 57,533 58,371 59,208 60,463 62,556 64,648 100,000 105,022 105,859 106,696 107,533 108,371 109,208 110,463 112,556 114,648 150,000 155,022 155,859 156,696 157,533 158,371 159,208 160,463 162,556 164,648 200,000 205,022 205,859 206,696 207,533 208,371 209,208 210,463 212,556 214,648 250,000 255,022 255,859 256,696 257,533 258,371 259,208 260,463 262,556 264,648 10% target IRR 20,000 26,574 27,669 28,765 29,861 30,956 32,052 33,695 36,434 39,174 50,000 56,574 57,669 58,765 59,861 60,956 62,052 63,695 66,434 69,174 100,000 106,574 107,669 108,765 109,861 110,956 112,052 113,695 116,434 119,174 150,000 156,574 157,669 158,765 159,861 160,956 162,052 163,695 166,434 169,174 200,000 206,574 207,669 208,765 209,861 210,956 212,052 213,695 216,434 219,174 250,000 256,574 257,669 258,765 259,861 260,956 262,052 263,695 266,434 269,174 14% target IRR 20,000 28,776 30,238 31,701 33,163 34,626 36,089 38,283 41,939 45,596 50,000 58,776 60,238 61,701 63,163 64,626 66,089 68,283 71,939 75,596 100,000 108,776 110,238 111,701 113,163 114,626 116,089 118,283 121,939 125,596 150,000 158,776 160,238 161,701 163,163 164,626 166,089 168,283 171,939 175,596 200,000 208,776 210,238 211,701 213,163 214,626 216,089 218,283 221,939 225,596 250,000 258,776 260,238 261,701 263,163 264,626 266,089 268,283 271,939 275,596 In order for an enterprise to be commercially viable, the volume of produce grown each year multiplied by the sales price of that produce would need to match or exceed the target values provided above. For example, a proposed development with capital costs of $125,000/ha and operating costs of $200,000 per ha per year would need to generate gross revenue of $213,695 per ha per year to achieve a target IRR of 10% (Table 4-27). If the enterprise received $12/kg for produce (averaged across product types, on a harvest mass basis), then it would need to sustain mean long-term yields of 18 t/ha (= $213,695 per ha per year ÷ $12/kg × 1 t/1000 kg) from the first harvest. However, if prices were $20/kg, mean long-term yields would require 11 t/ha (= $213,695 per ha per year ÷ $20/kg × 1 t/1000 kg) for the same $125,000 capital costs per hectare, or only 6 t/ha harvests if the capital costs decreased to $100,000 per hectare. Target revenue would be higher after taking into account risks such as learning and adapting to the particular challenges of a new location, and periodic setbacks that could arise from disease, climate variability, changes in market conditions or new legislation. Key messages From this analysis, a number of key points about achieving commercial viability in new aquaculture enterprises are apparent: •Operating costs are very high, and the amount spent each year on inputs can exceed the upfront(year zero) capital cost of development (and the value of the farm assets). This means that thecost of development is a much smaller consideration for achieving profitability than ongoingoperations and costs of inputs. •High operating costs also mean that substantial capital reserves are required, beyond the capitalcosts of development, as there will be large cash outflows for inputs in the start-up years beforerevenue from harvested product starts to be generated. This is particularly the case for largersize classes of product that require multi-year grow-out periods before harvest. Managingcashflows would therefore be an important consideration at establishment and as yields aresubsequently scaled up. •Variable costs dominate the total costs of aquaculture production, so most costs will increase asyield increases. This means that increases in production, by itself, would contribute little toachieving profitability in a new enterprise. What is much more important is increasingproduction efficiency, such as feed conversion rate or labour efficiency, so inputs per unit ofproduce are reduced (and profit margins per kilogram are increased). •Small changes in quantities and prices of inputs and produce would have a relatively largeimpact on net profit margins. These values could differ substantially between different locations(e.g. varying in remoteness, available markets, soils and climate) and depend on the experienceof managers. Even small differences from the indicative values provided in Table 4-27couldrender an enterprise unprofitable. •Enterprise viability would therefore be very dependent on the specifics of each particular caseand how the learning, scaling up and cashflow were managed during the initial establishmentyears of the enterprise. It would be essential for any new aquaculture development in theSouthern Gulf catchments to refine the production system and achieve the required levels ofoperational efficiency (input costs per kilogram of produce) using just a few ponds before scalingany enterprise. 4.6 References ABARES (2022) Agricultural commodities: September quarter 2022. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra. DOI: doi.org/10.25814/zs85- g927. ABARES (2023) Australian horticulture prices. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra. Viewed 10 March 2023, https://www.agriculture.gov.au/abares/data/weekly-commodity-price-update/australian- horticulture-prices#daff-page-main. Abrecht DG and Bristow KL (1996) Coping with climatic hazards during crop establishment in the semi-arid tropics. Australian Journal of Experimental Agriculture 36, 971–983. DOI: 10.1071/EA9960971. Andrews K and Burgess J (2021) Soil and land assessment of the southern part of Flying Fox Station for irrigated agriculture. Part B: Digital soil mapping and crop specific land suitability. Department of Environment, Parks and Water Security, Northern Territory Government, Darwin. Arndt W, Phillips LJ and Norman MJT (1963) Comparative performance of crops on three soils of the Tipperary region NT. Division of Land Research and Regional Survey Technical Paper No. 23. CSIRO, Canberra. Ash AJ (2014) Factors driving the viability of major cropping investments in Northern Australia – a historical analysis. CSIRO, Australia. Ash A and Watson I (2018) Developing the north: learning from the past to guide future plans and policies. The Rangeland Journal 40, 301–314. DOI: 10.1071/RJ18034 Barbour L (2008) Analysis of plant-host relationships in tropical sandalwood (Santalum album). RIRDC Publication No 08/138. Rural Industries Research and Development Corporation, Canberra. Viewed 28 August 2024, https://agrifutures.com.au/product/analysis-of-plant- host-relationships-in-tropical-sandalwood-santalum-album. Cobcroft J, Bell R, Fitzgerald J, Diedrich A and Jerry D (2020) Northern Australia aquaculture industry situational analysis. Project A.1.1718119. Cooperative Research Centre for Developing Northern Australia, Townsville. Gentry J (2010) Mungbean management guide, 2nd edition. Department of Employment, Economic Development and Innovation, Queensland. Viewed 19 October 2017, https://era.daf.qld.gov.au/id/eprint/7070/1/mung-manual2010-LR.pdf. DSITI and DNRM (2015) Guidelines for agricultural land evaluation in Queensland. Queensland Government (Department of Science, Information Technology and Innovation and Department of Natural Resources and Mines), Brisbane. FAO (1976) A framework for land evaluation. Food and Agriculture Organization of the United Nations, Rome. FAO (1985) Guidelines: land evaluation for irrigated agriculture. Food and Agriculture Organization of the United Nations, Rome. Gibbs M, Hughes J, Yang A, Wang B, Marvanek S and Petheram C (2024) River model scenario analysis for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Gleeson T, Martin P and Mifsud C (2012) Northern Australian beef industry: assessment of risks and opportunities. Australian Bureau of Agricultural and Resource Economics and Sciences report to client, prepared for the Northern Australia Ministerial Forum, Canberra. Guy JA, McIlgorm A and Waterman P (2014) Aquaculture in regional Australia: responding to trade externalities. A northern NSW case study. Journal of Economic & Social Policy 16(1), 115. Irvin S, Coman G, Musson D and Doshi A (2018) Aquaculture viability. 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. Jakku E, Thorburn PJ, Marshall NA, Dowd AM, Howden SM, Mendham E, Moon K and Brandon C (2016) Learning the hard way: a case study of an attempt at agricultural transformation in response to climate change. Climatic Change 137, 557–574. DOI: 10.1007/s10584-016-1698- x Jones C (2007) Redclaw package 2007. Introduction to redclaw aquaculture. Queensland Department of Primary Industries and Fisheries, Brisbane. Jones C, Grady J-A and Queensland Department of Primary Industries (2000) Redclaw from harvest to market: a manual of handling procedures. Queensland Department of Primary Industries, Brisbane. Jones C, Mcphee C and Ruscoe I (1998) Breeding redclaw: management and selection of broodstock. QI98016. Queensland Department of Primary Industries, Brisbane. Jones CM (1995) Production of juvenile redclaw crayfish, Cherax quadricarinatus (von Martens) (Decapoda, Parastacidae) III. Managed pond production trials. Aquaculture 138(1), 247–255. DOI: https://doi.org/10.1016/0044-8486(95)00067-4. Mann D (2012) Impact of aerator biofouling on farm management, production costs and aerator performance. Mid project report to farmers. Australian Seafood Cooperative Research Centre Project No. 2011/734. Department of Agriculture, Fisheries and Forestry, Queensland. Masser M and Rouse B (1997) Australian red claw crayfish. Circular ANR-769. The Alabama Cooperative Extension Service, USA. McKellar L, Bark RH and Watson I (2015) Agricultural transition and land-use change: considerations in the development of irrigated enterprises in the rangelands of northern Australia. The Rangeland Journal 37, 445–457. DOI: 10.1071/RJ14129. McLean I and Holmes P (2015) Improving the performance of northern beef enterprises, 2nd edition. Meat and Livestock Australia, Sydney. Moore G, Revell C, Schelfhout C, Ham C and Crouch S (2021) Mosaic agriculture. A guide to irrigated crop and forage production in northern WA. Bulletin 4915. Western Australia Department of Regional Industries and Regional Development, Perth. Paterson B and Miller S (2013) Energy use in shrimp farming, study in Australia keys on aeration and pumping demands. Global Aquaculture Advocate, November/December, pp. 30–32. QDPIF (2006) Australian prawn farming manual: health management for profit. Queensland Department of Primary Industries and Fisheries, Brisbane. Queensland Government (2024) Business Queensland, Aquaculture farms and production systems. Viewed 15 June 2024, https://publications.qld.gov.au/dataset/agbiz-tools-fisheries- aquaculture. Sangha KK, Ahammad R, Mazahar MS, Hall M, Owens G, Kruss L, Verrall G, Moro J and Dickinson G (2022) An integrated assessment of the horticulture sector in northern Australia to inform future development. Sustainability (Switzerland) 14(18), 1–18. DOI: 10.3390/su141811647. Schipp G, Humphrey JD, Bosmans J and Northern Territory Department of Primary Industry, Fisheries and Mines (2007) Northern Territory barramundi farming handbook. Northern Territory Department of Primary Industry, Fisheries and Mines, Darwin. Southern Gulf NRM (2016). Land condition guide. Southern Gulf NRM, Mount Isa. Thomas M, Gregory L, Harms B, Hill JV, Holmes K, Morrison D, Philip S, Searle R, Smolinski H, Van Gool D, Watson I, Wilson PL and Wilson PR (2018) Land suitability analysis. A technical report from the CSIRO Northern Australia Water Resource Assessment to the Government of Australia. CSIRO, Canberra. Thomas M, Philip S, Zund P, Stockmann U, Hill J, Gregory L, Watson I and Thomas E (2024) Soils and land suitability for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Watson I, Austin J and Ibrahimi T (2021) Other potential users of water. In: Petheram C, Read A, Hughes J, Marvanek S, Stokes C, Kim S, Philip S, Peake A, Podger G, Devlin K, Hayward J, Bartley R, Vanderbyl T, Wilson P, Pena Arancibia J, Stratford D, Watson I, Austin J, Yang A, Barber M, Ibrahimi T, Rogers L, Kuhnert P, Wang B, Potter N, Baynes F, Ng S, Cousins A, Jarvis D and Chilcott C (eds) An assessment of contemporary variations of the Bradfield Scheme. A technical report to the National Water Grid Authority from the Bradfield Scheme Assessment. CSIRO, Australia, chapter 8. Webster A, Jarvis D, Jalilov S, Philip S, Oliver Y, Watson I, Rhebergen T, Bruce C, Prestwidge D, McFallan S, Curnock M and Stokes C (2024) Financial and socio-economic viability of irrigated agricultural development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Woodward E, Jarvis D and Maclean K (2019) The traditional owner-led bush products sector: an overview scoping study and literature review. CSIRO, Australia. Yang A, Petheram C, Marvanek S, Baynes F, Rogers L, Ponce Reyes R, Zund P, Seo L, Hughes J, Gibbs M, Wilson PR, Philip S and Barber M (2024) Assessment of surface water storage options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO Australia Yeates SJ (2001) Cotton research and development issues in northern Australia: a review and scoping study. Australian Cotton Cooperative Research Centre, Darwin. Yeates SJ and Poulton PL (2019) Determining dryland cotton yield potential in the NT: Preliminary climate assessment and yield simulation. Report to NT Farmers, Queensland Cotton and the Cotton Research and Development Corporation. CSIRO, Canberra. Yeates SJ, Strickland GR and Grundy PR (2013) Can sustainable cotton production systems be developed for tropical northern Australia? Crop and Pasture Science 64, 1127–1140. DOI: 10.1071/CP13220. 5 Opportunities for water resource development in the Southern Gulf catchments Authors: Matt Gibbs, Andrew R Taylor, Cuan Petheram, Ang Yang, Steve Marvanek, Lee Rogers, Fred Baynes, Geoff Hodgson, Justin Hughes, Anthony Knapton Chapter 5 examines the opportunities, risks and costs for water resource development in the catchments of the Southern Gulf rivers, that is Settlement Creek, Gregory–Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groupsFigure 5-1. 1. Evaluating the possibilities for water resource development and irrigated agriculture requires an understanding of the development-related infrastructure requirements, how much water can be supplied and at what reliability, and the associated costs. The key components and concepts of Chapter 5 are shown in 1 Only those islands greater than 1000 ha are mapped. Figure 5-1 Schematic of key engineering and agricultural components to be considered in the establishment of a water resource and greenfield irrigation development Numbers in blue refer to sections in this report. 5.1 Summary This chapter provides information on a variety of potential options to supply water, primarily for irrigated agriculture. The methods used to generate these involved collating and synthesising existing data. No fieldwork was undertaken as part of this Assessment. The potential water yields reported in this chapter are based largely on physically plausible volumes, and do not consider economic, social, environmental, legislative or regulatory factors, which will inevitably constrain many developments. In some instances, the water yields are combined with land suitability information from Chapter 4 to provide estimates of areas of land that could potentially be irrigated close to the water source or storage. 5.1.1 Key findings Water can be sourced and stored for irrigation in the Southern Gulf catchments in a variety of ways. If the water resources of the Southern Gulf catchments are developed for consumptive purposes it is likely that some of the options below may help to maximise the cost effectiveness of water supply in different parts of the study area. Groundwater extraction Groundwater is already widely used in parts of the Southern Gulf catchments for a variety of purposes (stock, domestic activities, mining, irrigation, community water supplies) and offers year- round niche opportunities that are geographically distinct from surface water development opportunities. The two most productive groundwater systems in the catchments are the regional- scale Cambrian Limestone Aquifer (CLA) and Gilbert River Aquifer (GRA). Currently, existing annual licensed groundwater entitlements for both aquifers are small. These are 1.8 GL/year for the CLA and <0.25 GL/year for the GRA, neither aquifers are currently used for groundwater-based irrigation. Across different parts of the catchments small amounts of unlicensed groundwater use for stock and domestic purposes also occurs from both aquifers. Water plans exist for both aquifers in the Queensland portion of the study area. The Gulf Water Plan manages groundwater resources of the CLA, and the groundwater resources of the GRA are managed under the Great Artesian Basin and Other Regional Aquifers water plan. There are currently no water plans for the NT parts of the catchments, where (of the two regional-scale aquifers) only the CLA exists. With appropriately sited borefields it may be possible to extract between 10 and 20 GL/year from the CLA in the south-west of the study area. This is where the aquifer outcrops to the west of where the reaches of Lawn Hill Creek and the Gregory and O'Shannassy rivers receive groundwater discharge from the CLA. This indicative scale of the resource is based on an initial estimate of the components of the water balance for the portion of the CLA underlying these catchments, as well as their potential sensitivity to climate variability. However, data for the CLA is sparse, and a more refined estimate of the scale of potential future groundwater resource development would require more detailed investigations of the water balance. Any potential development would also depend on potential impacts to groundwater-dependent ecosystems (GDEs), such as the prescribed reaches of Lawn Hill Creek and the Gregory River, to existing groundwater users (see Section 3.3.4), and community and government acceptance these changes. The GRA may also offer some opportunities for future groundwater resource development, though the aquifer in the north-east of the catchment is deep (>500 metres below ground level (mBGL)) in places and exhibits variable water quality not always suitable for irrigation. It may be possible with appropriately sited borefields to extract about 5 GL/year near the south-western margin of the GRA in the north-east of the catchments, where the aquifer is less than 500 m deep, and in places hosts fresh water (<1000 mg/L total dissolved solids (TDS)). Similar to the CLA, the actual scale of potential future groundwater resource development would depend upon community and government acceptance of potential impacts to GDEs and existing groundwater users (e.g. active water licences for the communities of Burketown and Gununa). Opportunities for potential groundwater resource development from aquifers hosted in other hydrogeological units (Proterozoic igneous, metasedimentary and metamorphic rocks) is most likely to be limited to use for stock and domestic purposes, and occasional community water supply. Aquifers within the Cenozoic alluvium may provide a source of water, though few data exists for these systems. Major dams Indigenous customary residential and economic sites are usually concentrated along major watercourses and drainage lines. Consequently, potential instream dams are more likely to have an impact on areas of high cultural significance than are most other infrastructure developments of comparable size. Topographically and geologically, large parts of the Southern Gulf catchments have considerable potential for large instream dams, and there are already five large water supply dams in the Southern Gulf catchments with capacities of about 10 GL or greater. These dams are used to supply water for mining, industry and the city of Mount Isa. These users typically use considerably less water than the amounts required for irrigated agriculture. In the Southern Gulf catchments a limitation of large dams is the semi-arid climate and typically small catchment areas of those parts of the catchment with suitable topography. This means that most potential sites for dams in the Southern Gulf catchments have relatively low yields (typically less than 100 GL in 85% of years). One of the most cost-effective potential large instream dam sites in the Southern Gulf catchments is on the Gregory River, immediately upstream of a large contiguous area (1.03 million ha) of soil suitable for irrigated agriculture. However, the site sits within the Thorntonia Aggregation wetland. A previous study found that residents at Mount Isa placed a particularly high value on the Gregory Rivers (Jackson et al., 2008). Considering existing consumptive users this potential dam could yield 133 GL in 85% of years and cost $683 million (−20% to +50%) to construct, assuming favourable geological conditions. With the land adjacent to the east bank of the Gregory River sloping away from the river, the land is highly suitable for the more cost-effective gravity channel reticulation infrastructure. A nominal reticulation scheme for 10,000 ha of irrigated land was estimated to cost an additional $31.8 million or $3180/ha (excluding farm development and infrastructure). Water harvesting and offstream storage Water harvesting, where water is pumped from a major river into an offstream storage such as a ringtank, is a cost-effective option for capturing and storing water from the Nicholson, Gregory and Leichhardt rivers. Approximately 21% of the study area (1,672,000 ha) was modelled as having soil and topography likely or possibly suitable for ringtanks, which is a large percentage relative to other catchments across northern Australia. The most favourable soil and topography for ringtanks in the study area is on the east bank of the Gregory River, where the extensive cracking clay soils of the Armraynald Plain gently slope away from the river, enabling the cost-effective conveyance of water using gravity. Water harvesting along the Nicholson is limited by the lack of soil suitable for the construction of off-stream storages. Downstream of Kajabbi the construction of ringtanks adjacent to the Leichhardt River is possible on heavier alluvial soils where the sandy levees are not so wide that the conveyance of water from the river to the ringtank would be prohibitively expensive. From the Nicholson and Leichhardt catchments it is physically possible to extract 150 GL in 75% of years and irrigate 12,000 ha of broadacre by pumping or diverting water from the Nicholson, Gregory and Leichhardt rivers and storing it in offstream storages such as ringtanks. This results in modelled reductions in mean and median annual discharge from the Gregory– Nicholson and Leichhardt rivers to the Gulf of Carpentaria of about 3% and 5%, respectively. Managed aquifer recharge The most promising aquifers for infiltration-based MAR in the Southern Gulf catchments are aquifers within the Cenozoic alluvium associated with many of the rivers in the study area. Visually the Cenozoic alluvium appears to be extensive relative to alluvium associated with many rivers elsewhere in northern Australia. However, bore logs and water-level information are not available for Cenozoic alluvium, and it is likely that the opportunity for MAR may vary between locations. Approximately 122,700 ha (1.1% of the study area) of the Southern Gulf catchments was identified as having potential for aquifers, groundwater and landscape characteristics suitable for infiltration MAR techniques within 5 km of a source of surface water – defined here as being a river with median annual flow greater than 20 GL. Approximately 37,100 ha (0.3% of the study area) within 1 km of such a river had similarly suitable characteristics; however, 94% of this area is underlain by Cenozoic alluvium. There may be potential to use MAR in the CLA to offset the impacts of groundwater pumping on GDEs near Lawn Hill Creek and the Gregory River. However, this would be contingent on finding soils suitable for the construction of ringtanks (Section 5.4.4), to temporarily detain surface water, close to a location that is suitable for MAR (typically on highly permeable soils). Gully dams and weirs Suitably sited large farm-scale gully dams are a relatively cost-effective method of supplying water. Those areas that are more topographically suitable for large-scale gully dam sites generally do not coincide with areas that have soils suitable for their construction or suitable for irrigated agriculture. There are, however, some locations in the north-west of Doomadgee in the Settlement Creek catchment and Mornington Island where there is some topography suitable for gully dams, soil suitable for their construction and versatile agricultural land nearby. These areas are geographically distinct from those areas that are most suitable for water harvesting and ringtanks. The remaining sources of water and storage options, namely weirs and natural water bodies, are estimated to be capable of reliably supplying considerably smaller volumes of water than major instream dams. Sourcing water from natural water bodies, although the most cost-effective option, is highly contentious. Summary of investigation costs, capital, and operation and maintenance costs of different water supply options and potential scale of unconstrained development Table 5-1 provides a summary of indicative investigative, capital, and operation and maintenance costs of different water supply options and estimates of the potential scale of unconstrained development. The development of any of these options will have an impact on existing uses, including ecological systems, to varying degrees, and will depend on the level of development. Ecological implications of altered flow regimes are examined in Section 7.3. All of the water source options reported in Table 5-1 are considerably cheaper than the cost of desalinisation. The initial cost of constructing four large desalinisation plants (capacity of 90 to 150 GL/year) in Australia between 2010 and 2012 ranged from $19,000/ML to $31,000/ML (AWA, 2018), indexed to 2023. This does not include the cost of ongoing operation (e.g. energy) and maintenance or the cost of conveying water to the demand. Table 5-1 Summary of capital costs, yields and costs per ML supply, including operation and maintenance Costs and yields are indicative. Values are rounded. Capital costs are the cost of construction of the water storage/source infrastructure. They do not include the cost of constructing associated infrastructure for conveying water or irrigation development. Water supply options are not independent of one another, and the maximum yields and areas of irrigation cannot be added together. Equivalent annual cost assumes a 7% discount rate over the service life of the infrastructure. Total yields and areas are indicative and based on physical plausibility unconstrained by economic, social, environmental, legislative or regulatory factors, which will inevitably constrain many developments. WATER SOURCE/STORAGE GROUND- WATER† MANAGED AQUIFER RECHARGE‡ MAJOR DAM WEIR§ LARGE FARM- SCALE RINGTANK LARGE FARM- SCALE GULLY DAM NATURAL WATER BODY Cost and service life of individual representative unit Capital cost ($ million) 4.5 1.1 680 5–40 2.5 1.7 0.02 Operation and maintenance ($ million/y)* 0.1 0.07 2.7 0.1–0.8 0.125 0.045 ~0 Assumed service life (y) 50 50 100 50 40 30 10 Potential yield of individual representative unit at water source Yield at source (GL)†† 2 0.6 133 0.5–10 2.4 3 0.125–0.5 Unit cost ($/ML)‡‡ 2,250 1,830 5,110 6,500 1,040 570 100 Levelised cost ($ million/ML) §§ 215 250 380 600 130 60 10 Potential yield of individual representative unit at paddock Assumed conveyance efficiency to paddock (%)††† 95 90 75 80 90 90 90 Yield at paddock (GL) 1.9 0.54 100 0.4–8 2.2 2.7 0.11–0.45 Unit cost ($/ML)‡‡ 2,370 2,040 8,115 8,125 1,160 630 110 Levelised cost ($ million/ML) ††† 225 280 600 750 145 65 12 WATER SOURCE/STORAGE GROUND- WATER† MANAGED AQUIFER RECHARGE‡ MAJOR DAM WEIR§ LARGE FARM- SCALE RINGTANK LARGE FARM- SCALE GULLY DAM NATURAL WATER BODY Total potential yield and area (unconstrained) Total potential yield (GL/y) at source ≥75% reliability‡‡‡ 30 <30 320 <100 90 <100 <20 Potential area that could be irrigated at ≥75% reliability (ha)§§§ 4,500 <4,500 30,000 <10,000 12,000 <10,000 <2,000 †Value assumes extraction from Cambrian Limestone Aquifer assuming mean bore yield of 25 L/s irrigation of 250 ha to meet mean peak evaporative demand over a 3-day period. Assumes a mean depth of 80 m and a drilling failure rate of 50%. ‡Based on recharge weir. §Sheet piling weir. *Operation and maintenance (O&M) cost is the annual cost of operating and maintaining infrastructure and includes cost of pumping groundwater assuming groundwater is 10–20 m below ground level and the cost of pumping water into ringtank. ††Yield at dam wall (taking into consideration net evaporation from surface water storages prior to release) or at groundwater bore. Value assumes large farm-scale ringtanks do not store water past August. ‡‡Capital cost divided by the yield. §§Equivalent annual cost of storage per bore per ML of yield of water. Includes capital cost and O&M costs. Assumes 7% discount rate. †††Conveyance efficiency between dam wall/groundwater bore and edge of paddock (does not include field application losses). ‡‡‡Actual yield will depend upon government and community acceptance of impacts to water-dependent ecosystems and existing users. Yields are not additive. Likely maximum cumulative yield at the dam wall/groundwater bore. Potential yield of major dams based on yield of dams at Waterhouse River. §§§Likely maximum area that could be irrigated (after conveyance and field application losses) in at least 75% of years. Assumes a single crop. Areas provided for each water source are not independent and hence are not additive. Actual area will depend upon government and community acceptance of impacts to water-dependent ecosystems and existing users. 5.2 Introduction 5.2.1 Contextual information Irrigation during the dry season and other periods when soil water is insufficient for crop growth requires sourcing water from a suitable aquifer or from a surface water body. However, decisions regarding groundwater extraction, river regulation and water storage are complex, and the consequences of decisions can be inter-generational, where even relatively small inappropriate releases of water may preclude the development of other, more appropriate (and possibly larger) developments in the future. Consequently, governments and communities benefit by having a wide range of reliable information available prior to making decisions, including the manner of ways water can be sourced and stored, as this can have long-lasting benefits and facilitate an open and transparent debate. Information is presented in a manner to easily enable the comparison of the variety of options. More detailed information can be found in the companion technical reports. Section 5.5 discusses the conveyance of water from the storage and its application to the crop. Transmission and field application efficiencies, and associated costs and considerations, are examined. All costs presented in this chapter are indexed to December 2023. Concepts The following concepts are used in sections 5.3 and 5.4: • Each of the water source and storage sections are structured around: (i) an opportunity- or reconnaissance-level assessment and (ii) a pre-feasibility-level assessment: – Opportunity-level assessments involved a review of the existing literature and a high-level desktop assessment using methods and datasets that could be consistently applied across the entire Assessment area. The purpose of an opportunity-level assessment is to provide a general indication of the likely scale of opportunity and geographic location of each option. – Pre-feasibility-level assessments involved a more detailed desktop assessment of sites/geographic locations that were considered more promising. This involved a broader and more detailed analysis including the development of bespoke numerical models, site-specific cost estimates and site visits. Considerable field investigations were undertaken for the assessment of groundwater development opportunities (Section 5.3.2). • ‘Yield’ is a term used to report the performance of a water source or storage. It is the amount of water that can be supplied for consumptive use at a given reliability. For dams, an increase in water yield results in a decrease in reliability. For groundwater, an increase in water yield results in an increase in the ‘zone of influence’ and can result in a decrease in reliability over time, particularly in local- and intermediate-scale groundwater systems. • Unit cost is the capital cost of the infrastructure divided by the amount of water that can be supplied at a specified reliability. It is commonly used to provide a comparison of the cost effectiveness of assets of similar life spans and operational costs. • Equivalent annual cost is the annual cost of owning, operating and maintaining an asset over its entire life. Equivalent annual cost allows a comparison of the cost effectiveness of various assets that have unequal service lives/life spans. • Levelised cost is the equivalent annual cost divided by the amount of water that can be supplied at a specified reliability. It allows a comparison of the cost effectiveness of various assets that have unequal service lives/life spans and water supply potential. Other economic concepts reported in this chapter, such as discount rates, are outlined in Chapter 6. 5.3 Groundwater and subsurface water storage opportunities 5.3.1 Introduction Groundwater, where an aquifer is relatively shallow (less than a couple of hundred metres) and of sufficient yield to support irrigation (typically greater than 10 L/s), is often one of the cheapest sources of water available, particularly where individual bore yields can be in the order of tens of litres per second thereby reducing the costs of groundwater infrastructure. Even the cheapest forms of MAR, infiltration-based techniques, are usually considerably more expensive than developing a groundwater resource. Further to this, in northern Australia many unconfined aquifers, which are best suited to infiltration-based MAR, either have large areas with no ‘free’ storage capacity at the end of the wet season (because groundwater levels have risen to near the ground surface) or, where they do have the available storage capacity, are often not at economically viable distances (greater than 5 km) from a reliable source of water to recharge the aquifer. Therefore, MAR will inevitably only be developed following development of a groundwater system, where groundwater extraction may create additional storage capacity within the aquifer (by lowering groundwater levels) to allow additional recharge, and hydrogeological information is more readily available to evaluate the local potential of MAR. However, if developed, MAR can increase the quantity of water available for extraction and help mitigate impacts to the environment. Where water uses have a higher value than irrigation (e.g. in mining, energy operations, town water supply), other more expensive but versatile forms of MAR, such as aquifer storage and recovery, can be economically viable and should be considered. The Assessment involved a catchment-wide reconnaissance assessment of: • the potential for groundwater resource development (Section 5.3.2) • MAR (Section 5.3.3). 5.3.2 Opportunities for groundwater development Introduction Planning future groundwater resource developments and authorising licensed groundwater entitlements require value judgments of what is an acceptable impact to receptors such as environmental assets or existing users at a given location. These decisions can be complex, and they typically require considerable input from a wide range of stakeholders, particularly scientific information, to help inform these decisions, which include: • identifying aquifers that may be potentially suitable for future groundwater resource development • characterising their depth, spatial extent, saturated thickness, hydraulic properties and water quality • conceptualising the nature of their flow systems • estimating aquifer water balances • identifying geographical opportunities within key aquifers and the associated risks for future groundwater development. Opportunities include the identification and location of hydrogeological units and the aquifers they host. Risks include the potential to changes in aquifer storage and therefore reliability of access to water for existing users or the environment (groundwater-dependent ecosystems, GDEs). Unless stated otherwise, the material presented in Section 5.3.2 has been summarised from the companion technical report on groundwater characterisation (Raiber et al., 2024) and the companion technical report on groundwater modelling (Knapton et al., 2024). Opportunity-level assessment of groundwater resource development opportunities in the Southern Gulf catchments The hydrogeological units of the Southern Gulf catchments (Figure 5-2) contain a variety of local-, intermediate- and regional-scale aquifers that host localised to regional-scale groundwater flow systems. The regional-scale CLA and regional-scale GRA are present in the subsurface across large areas, collectively occurring beneath about 45% of the study area. Given their large spatial extent, they also underlie and frequently coincide with larger areas of soil suitable for irrigated agriculture (Section 4.2). They contain fresh water (<1000 mg/L TDS) in places and can yield water at a sufficient rate to support irrigation development (>10 L/s). These aquifers also contain larger volumes of groundwater in storage (gigalitres to teralitres) than local-scale aquifers and their storage and discharge characteristics are often less affected by short-term (yearly) variations in recharge rates caused by inter-annual variability in rainfall. Furthermore, their larger spatial extent provides greater opportunities for groundwater resource development away from existing water users and GDEs at the land surface such as springs, spring-fed vegetation and surface water, which can be ecologically and culturally significant. In contrast, local-scale aquifers in the Southern Gulf catchments, such as fractured and weathered rock and alluvial aquifers, host local-scale groundwater systems that are highly variable in composition, salinity and yield. They also have a small and variable spatial extent and less storage compared to the larger aquifers, limiting groundwater resource development to localised opportunities such as for stock and domestic use, or as a conjunctive water resource (i.e. combined use of surface water, groundwater or rainwater). The Assessment identified five hydrogeological units hosting aquifers that may have potential for future groundwater resource development in the Southern Gulf catchments (Table 5-2): •Cambrian limestone and dolostone •Cretaceous rocks •Cenozoic alluvium •Proterozoic igneous rocks •Proterozoic metasedimentary and metamorphic rocks. Table 5-2 Opportunity-level estimates of the potential scale of groundwater resource development opportunities in the Southern Gulf catchments For locations of the hydrogeological units see Figure 5-2. The indicative scale of the groundwater resource is based on the magnitude of the inputs and outputs of the groundwater balance. The actual scale will depend upon potential impacts to groundwater-dependent ecosystems and existing groundwater users, and government and community acceptance of these changes. For more information on this table please contact CSIRO on enquiries@csiro.au For more information on this table please contact CSIRO on enquiries@csiro.au †Actual scale will depend upon government and community acceptance of impacts to groundwater-dependent ecosystems and existing water users. Figure 5-2 Hydrogeological units with potential for future groundwater resource development Presents the spatial extent of the outcropping and subcropping component of each hydrogeological unit with the majority of the Cenozoic cover removed (except the alluvium). Entire spatial extent of the Cambrian limestone and dolostone, and Cretaceous rocks outside the Southern Gulf catchments is shown in Figure 2-28. Simplified regional hydrogeology and springs map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-515_simplified_regional_hydrogeology_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Groundwater development costs The cost of groundwater development is the cost of the infrastructure plus the cost of the hydrogeological investigations required to understand the resource and risks associated with its development. This section presents information relevant to the cost of further developing the groundwater resources of the CLA, including but not limited to the depth to water-bearing formation (control over cost of drilling) and the depth to groundwater (control over cost of pumping). Information on the spatial extent of drawdown of groundwater levels is also presented. This is relevant to the potential hydraulic impact of future development on receptors such as existing licensed water users and culturally and ecologically important GDEs. Aquifer yield information is presented in Section 2.5.2. At a local development scale, individual proponents will need to undertake sufficient localised investigations to provide confidence around aquifer properties and bore performance. This information will also form part of an on-site hydrogeological assessment required by the regulator in order to grant an authorisation to extract groundwater. Key considerations for an individual proponent include: • determining the location to drill a production bore • testing the production bore • determining the location and number of monitoring bores required • conducting a hydrogeological assessment as part of applying for an authorisation to extract groundwater. Estimates of costs associated with these local-scale investigations are summarised in Table 5-3. Table 5-3 Summary of estimated costs for a 250 ha irrigation development using groundwater Assumes mean bore yield of 25 L/s and with 16 production bores required to meet peak evaporative demands of 250 ha. Does not include operating and maintenance costs. DRILLING, CONSTRUCTION, INSTALLATION AND TESTING OF BORES ESTIMATED COST ($) Production bore 2,550,000† Monitoring bores 285,000‡ Submersible pumps 1,360,000§ Mobilisation/demobilisation 6,000§§ Aquifer testing 170,000* Hydrogeological assessment 100,000†† †Value assumes 16 production bores drilled and constructed at a mean depth of 60 m at a cost per bore of $750/m, constructed with 200 mm steel casing at a cost of $82/m and 18 m stainless steel wire-wound screen at a cost of $150/m. Assumes on average two holes need to be drilled for every cased production bore. ‡Value assumes six PVC monitoring bores drilled and constructed at a mean depth of 60 m at a cost of $500/m, constructed with 150 mm PVC and 5 m machine-slotted screen at a cost of $50/m. §Value assumes a pump that is rated to draw water at a rate of up to 60 L/s, as well as rated to draw water from depths of up to 50 mBGL. Value based on 16 pumps. §§Value assumes a mobilisation/demobilisation rate of $10/km from Darwin to Daly Waters and return (approximately 1200 km round trip). *Value assumes six 72-hour aquifer tests (48 hours pumping, 24 hours recovery) at a cost of $500/h and $4000 mobilisation/demobilisation. ††Indicative cost to proponent. Value assumes a small-scale development away from existing users and groundwater-dependent ecosystems. Assumes the regulator has already characterised the aquifers at the intermediate/regional scale to better understand the resource potential under cumulative extraction scenarios, as well as current and future constraints to development. The Assessment identified the CLA to be the most promising intermediate- and regional-scale aquifer, with potential for future groundwater resource development in the Southern Gulf catchments. The CLA coincides with large, contiguous areas of cracking clay soils suitable for irrigated broadacre cropping (308,000 ha) along the south-western margin of the study area (Section 2.3). However, as discussed in Section 2.5.4, groundwater discharge from the CLA provides the dry-season flow in Lawn Hill Creek, and the Gregory and O’Shannassy rivers and some of their tributaries. The parts of the CLA in the south-western part of the Southern Gulf catchments appear to offer some promising opportunities for potential future groundwater resource development but require further investigation. The aquifer outcrops, subcrops and is mostly unconfined in the south-west of the Southern Gulf catchments (Figure 5-2 and Figure 5-5). This means it outcrops at the surface, or is close to the surface (within tens of metres of it) and is directly recharged by outcrop areas or by infiltration of rainfall through either low-permeability black clay soils or by vertical leakage through the overlying Cambrian siltstone where it is present across the southern part of the aquifer (Figure 5-2). The Georgina Basin, which hosts several hydrogeological units (Camooweal Dolostone, Thorntonia Limestone and Wonarah Formation) that collectively host the CLA, ranges in thickness from less than 100 m along the eastern/south-eastern basin margins to about 1000 m along the south- western boundary of the Southern Gulf catchments (Figure 5-4). Figure 5-3 Groundwater dependent ecosystems along Lawn Hill Creek Photo: Auscape/Universal Images Group via Getty Images Figure 5-4 Thickness of the Georgina Basin in the Southern Gulf catchments Only a partial spatial extent of the Georgina Basin is shown beyond the Southern Gulf catchments boundary. Depths are in metres. Georgina Basin extent data source: Raymond (2018) The CLA occurs within the top few hundred metres of the Georgina Basin and varies in thickness from about 50 m along the eastern margin to about 500 m towards the catchment’s boundary (Figure 5-4 and Figure 5-5). At Carrara 1 (deep stratigraphic drill-hole), the CLA is about 500 m thick (Figure 5-4 and Figure 5-5). Thickness of Georgina Basin, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-531_GeorginaBasin_thickness_v03_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-5 Hydrogeological cross-section through the Cambrian Limestone Aquifer in the Georgina Basin and south- west of the Southern Gulf catchments See Figure 5-4 for the spatial location of the eastern part of the cross-section. AHD = Australian Height Datum. While the current level of knowledge for the aquifer is considered low to medium, the following information indicates the aquifer offers potential for future development: • The moderate spatial extent of the outcropping/subcropping area in the catchment (about 12,400 km2, see Figure 5-2) coincides with areas of cracking clay soils potentially suitable for agricultural intensification (Section 2.3.2). • The aquifer can be intersected by drilling at relatively shallow depths in the outcropping and subcropping areas (mostly <100 mBGL). • Moderate bore yields (up to 20 L/s) indicate that the aquifer may have potential to yield water at a sufficient rate for groundwater-based irrigation. • The depth to pump groundwater to the surface is less than 75 mBGL across most areas of the CLA, though groundwater-level data are sparse, and larger areas of shallow depths to groundwater (<10 mBGL) are likely to occur around the middle to lower reaches of the prescribed watercourses of Lawn Hill Creek, and the Gregory and O’Shannassy rivers, where the aquifer discharges (Figure 5-6). • The aquifer host fresh water suitable for a variety of irrigated crops (mostly <1000 mg/L TDS). For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-6 Depth to standing water level (SWL) of the Cambrian Limestone Aquifer (CLA) Only a partial spatial extent of the CLA is shown beyond the Southern Gulf catchments boundary. Depths are in metres below the land surface. Lower left inset indicates the south-western part of the catchments where the mapped data are shown, as well as its location within the eastern part of the Georgina Basin (light purple). The Assessment undertook an extensive review of the hydrogeological data and information to confirm the conceptual model of the CLA in the eastern Georgina Basin. This conceptual model was used to develop an initial two-dimensional groundwater flow model for the CLA in the south- western part of the Southern Gulf catchments (the Undilla Sub-basin in the eastern part of the Georgina Basin). The purpose of the model was to assess the nature and scale of the groundwater resources of the Undilla Sub-basin, which provides baseflow to parts of Lawn Hill Creek, and the Gregory and O’Shannassy rivers. Initial order of magnitude groundwater balance estimates were derived for the entire groundwater model domain, which extends south-west of the catchment boundaries, and also for the portions of the Lawn Hill Creek and Gregory River subcatchments, and Nicholson Groundwater Management Area (NGMA), that coincide with the CLA inside the Southern Gulf catchments (Figure 5-7). SWL CLA \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-558_Georgina_SWL_surface_v03_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-7 Location of the Undilla Sub-basin groundwater flow model in relation to the Southern Gulf catchments and portions of the model that coincide with the Lawn Hill Creek and Gregory subcatchments and Nicholson Groundwater Management Area Lower left inset indicates the south-western part of the catchments where the mapped data are shown, as well as its location within the eastern part of the Georgina Basin (light purple). GW = groundwater. Table 5-4 provides a summary of the mean annual groundwater balances for the 109-year climate sequence (1910–2019) presented for portions of the CLA that coincide with the Lawn Hill Creek and Gregory subcatchments, the NGMA and the entire model domain. Within the model domain the CLA coincides with 2344 km2 of the Lawn Hill subcatchment and 9727 km2 of the Gregory subcatchment. The CLA coincides with 10,266 km2 of the NGMA, whereas the entire model domain for the CLA is 50,867 km2, most of which extends to the south-west outside the catchments (Figure 5-7). When considering the magnitude in modelled recharge volumes by area, these equate to recharge rates of about 25 mm/year for the portion of the CLA in the Lawn Hill Creek subcatchment, 14 mm/year for the portion of the CLA in the Gregory subcatchment, 15 mm/year for the portion of the CLA in the NGMA and 7 mm/year for the entire model domain. CLA GW flow model WAP, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-552_CLA_GW_flow_model_WAP_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Table 5-4 Mean annual groundwater balance for the Cambrian Limestone Aquifer (CLA) in the Undilla Sub-basin of the Georgina Basin for a 109-year climate sequence (1910–2019) and areas of the CLA that coincide with the Lawn Hill Creek and Gregory subcatchments and Nicholson Groundwater Management Area (NGMA) For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au Initial modelling of the mean annual groundwater balance for the CLA in the south-western Southern Gulf catchments suggests the aquifer potentially offers opportunities for groundwater resource development based on the magnitude of groundwater flows. However, the initial model has limitations due to the data-sparse nature of the CLA in this area, including a lack of (i) spatial and temporal groundwater-level data, (ii) gauging of spring flow at discrete springs and (iii) estimates of mean annual recharge using a variety of independent methods. In addition, recent modelling of the CLA by Knapton et al. (2023) highlighted the potential sensitivity of the CLA groundwater balance to climate variables. Based on mean annual recharge summarised in Table 5-4, an indicate scale of the resource for the CLA in the Southern Gulf catchments is estimated to be between 10 and 20 GL. However, this required further investigation. In addition, the actual scale of development will depend upon government and community acceptance of potential impacts to groundwater-dependent ecosystems and existing groundwater users, as well as the approval of licenses to extract groundwater. Groundwater resource development opportunities and risks associated with the Gilbert River Aquifer The GRA within the Great Artesian Basin, despite being data sparse, appears to offer some opportunities for future groundwater resource development but requires further investigation. The GRA is a sandstone aquifer occurring beneath Cretaceous mudstone rocks in the north-west of the catchment and extends offshore out beneath the Gulf of Carpentaria (Figure 5-2). The depth to the top of the GRA beneath the mudstone rocks is shallowest (<150 mBGL) along the south- west margin of the GRA running in a north-west to south-east orientation, from north of Hells Gate Roadhouse to Burke and Wills Roadhouse (Figure 5-8). Depth to the top of the GRA increases in the subsurface in a north-easterly direction towards the coastline where depths are generally between 400 and 500 mBGL (Figure 5-8 and Figure 5-9). The deepest parts of the GRA near the coastline occur around Burketown, where depth to the top of the GRA is between 500 and 700 mBGL (Figure 5-8 and Figure 5-9). For example, the original Burketown community water supply bore intersected the top of the GRA at about 670 mBGL (Figure 5-8). Figure 5-8 Depth to the top of the Gilbert River Aquifer (GRA) This is a mapped spatial extent of the GRA both within and beyond the Southern Gulf catchments boundaries. Depths are in metres below the land surface. Stratigraphic data represent a bore with stratigraphic data to obtain information about changes in geology with depth. A to A′ represents the location of the cross-section in Figure 5-9. Georgina Basin extent data source: Raymond (2018) Depth to GRA, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-533_Depth_to_GilbertRiver_v04_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au In the Gulf of Carpentaria beneath Gununa, the depth to the top of the GRA is similar to that around Burketown (~500 to 700 mBGL). The original Gununa community water supply bore intersected the top of the GRA at about 700 mBGL (Figure 5-8). The Cretaceous mudstone rocks that overlie the GRA are an aquitard (rocks of low permeability that store and transmit little groundwater), which confines the GRA (sealing it off from the atmosphere and pressurising the groundwater it stores and transmits) and prevents recharge from entering the aquifer. The GRA is recharged to the east of the Southern Gulf catchments, east of Karumba and Normanton. The GRA in the Southern Gulf catchments receives its inflows from groundwater in the GRA to the east of the Southern Gulf catchments, which flows into the catchment from west to north-west. Figure 5-9 South-west to north-east cross-section traversing the Great Artesian Basin in the Southern Gulf catchments. See Figure 5-8 for the spatial location of the cross-section. AHD = Australian Height Datum. Changes in the depth to groundwater across the GRA vary from close to artesian (groundwater rises under natural pressure to the land surface) to artesian, though standing water level data for the aquifer are limited. The south-west margin of the GRA groundwater is close to artesian (<20 mBGL) (Figure 5-10). Further to the north-east, towards the coast, groundwater becomes artesian up to about 20 m above the land surface (Figure 5-10). The original Burketown water supply bore, drilled in the 1900s, had an artesian water level of about 40 m above the land surface and a flow rate of about 7 L/second. For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-10 Depth to standing water level of the Gilbert River Aquifer (GRA) Mapped spatial extent of the GRA both within and beyond the Southern Gulf catchments boundary. Positive values are depth above the land surface representing artesian conditions; negative values are depths below the land surface representing sub-artesian conditions. Aquifer extent data source: Raymond (2018) While the current level of knowledge for the GRA has limited pre-feasibility information, the following information indicates it offers potential for future development: (i) the large spatial extent of the aquifer in the subsurface (about 44,600 km2, or about 40% of the study area; see Figure 5-2) that coincides with soils potentially suitable for agricultural intensification (Section 2.3.2); (ii) the aquifer can be intersected by drilling at depths of mostly less than <400 mBGL (Figure 5-8) along the south-western margin of the GRA in the north-east of the catchments, also an area where groundwater is mostly fresh (<1000 mg/L TDS); and (iii) the aquifer is either artesian or close to artesian, thereby reducing the costs of pumping the groundwater to the surface, though groundwater-level data are sparse (Figure 5-10). Currently, insufficient information exists to quantify the water balance for the GRA in the Southern Gulf catchments. Furthermore, the aquifer: (i) has sparse temporal water-level information; (ii) SWL of GRA, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-543_Glibert_River_SWL_inset_v01_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au dips steeply in the subsurface, indicating it shifts across different areas from semi-confined to confined conditions; and (iii) support active water licences for community water supplies at Burketown and Gununa, as well as potentially supporting ecologically and culturally important flora and fauna in the Gulf of Carpentaria’s marine environment through coastal and submarine groundwater discharge. In the absence of further hydrogeological investigations (drilling and pump testing) and hydrological risk assessment modelling to evaluate groundwater extraction impacts to existing water users and GDEs, a conservative value of ≤5 GL/year has been assumed for the indicative scale of the resource for potential groundwater resource development (Table 5-2). However, this requires further investigation, and would depend upon government and community acceptance of potential impacts to groundwater-dependent ecosystems and existing groundwater users, as well as the approval of licenses to extract groundwater. Groundwater resource development opportunities and risks associated with aquifers hosted in other hydrogeological units Opportunities for potential groundwater resource development from aquifers hosted in other hydrogeological units (Proterozoic igneous, metasedimentary and metamorphic rocks) is most likely to be limited to use for stock and domestic purposes, and occasional community water supply. However, productive local-scale aquifers hosted in the Cenozoic alluvium occurring in patches associated with the streambed, stream channel and floodplain of parts of the Nicholson, Gregory and Leichhardt rivers and their tributaries may offer some opportunities but require further investigation. The largest occurrences of the alluvium occur in the north of the catchment along the middle to lower reaches of the Nicholson, Gregory and Leichhardt rivers (Figure 5-2). Indicative bore yield data indicate bore yields can be as high as 5 L/second, which is too low for commercial irrigated agriculture, but the aquifer is currently sparsely tested and yields are likely to be locally variable. Water quality can vary from fresh to brackish but also remains sparsely tested. However, in places the aquifers may offer potential for small-scale (<1 GL/y) localised developments or as a conjunctive water resource. Opportunities are likely to be limited where the alluvium is (i) storage-limited (thin saturated thickness <15 m), (ii) comprising mostly fine-textured sediments (clay lenses), (iii) regularly flooded and (iv) highly connected to perennial reaches of prescribed watercourses and development may impact water availability to GDEs. 5.3.3 Managed aquifer recharge Introduction MAR is the intentional recharge of water to aquifers for subsequent recovery or environmental benefit (NRMMC-EPHC-NHMRC, 2009). Importantly for northern Australia, which has high intra- annual variability in rainfall, MAR can contribute to planned conjunctive use, whereby excess surface water can be stored in an aquifer in the wet season for subsequent reuse in the dry season (Evans et al., 2013; Lennon et al., 2014). Individual MAR schemes are typically small- to intermediate-scale storages with annual extractable volumes of up to 20 GL/year. In Australia, they currently operate predominantly within the urban and industrial sectors, and in the agricultural sector. This scale of operation can sustain rural to urban centres, contribute to diversified supply options in large urban centres and provide localised water management options, and it is suited to mosaic-type irrigation developments. The basic requirements for a MAR scheme are the presence of a suitable aquifer for storage, availability of an excess water source for recharge and a demand for water. The presence of suitable aquifers is determined from previous regional-scale hydrogeological and surface geological mapping (see companion technical report on groundwater characterisation by Raiber et al., 2024)). Source water availability is considered in terms of presence/absence rather than volumes with respect to any existing water management plans. Pre-feasibility assessment was based on MAR scheme entry-level assessment in the Australian guidelines for water recycling: managed aquifer recharge (NRMMC-EPHC-NHMRC, 2009; referred to as the ‘MAR guidelines’). The MAR guidelines provide a framework to assess feasibility of MAR; they incorporate four stages of assessment and scheme development. The first stage is entry-level assessment (pre-feasibility), the second stage involves investigations and risk assessment, the third stage is MAR scheme construction and commissioning, and the fourth stage is operation of the scheme. There are numerous types of MAR (Figure 5-12) and the selection of MAR type is influenced by the characteristics of the aquifer, the thickness and depth of low-permeability layers, land availability and proximity to the recharge source. Infiltration techniques can be used to recharge unconfined aquifers, with water infiltrating through permeable sediments beneath a dam, river or basin. If infiltration is restricted by superficial clay, the recharge method may involve a pond or sump that penetrates the low-permeability layer. Bores are used to divert water into deep or confined aquifers. Infiltration techniques typically have a lower cost than bore injection (Dillon et al., 2009; Ross and Hasnain, 2018) and are generally favoured in this Assessment. The challenge in northern Australia is to identify a suitable unconfined aquifer with capacity to store more water when water is available for recharge. In the Southern Gulf catchments, suitable unconfined aquifers are typically thought to rapidly recharge to full capacity during the wet season. Figure 5-11 Uncontrolled artesian flow from the Gilbert River Aquifer at the Burketown groundwater bore Photo: Shutterstock – Cam Laird A small rock in a muddy area Description automatically generated with medium confidence Figure 5-12 Types of managed aquifer recharge ASR = aquifer storage and recovery; ASTR = aquifer storage, transfer and recovery. Groundwater level indicated by triangle. Arrows indicate nominal movement of water. Source: Adapted from NRMMC-EPHC-NHMRC (2009) Opportunity-level assessment of infiltration-based MAR in the Southern Gulf catchments The most promising aquifers for infiltration-based MAR in the Southern Gulf catchments are aquifers within the Cenozoic alluvium associated with many of the rivers in the study area. Visually the Cenozoic alluvium appears to be extensive relative to alluvium associated with many rivers elsewhere in northern Australia. However, bore logs and water-level information are not available for Cenozoic alluvium, and it is likely the opportunity for MAR may vary between locations. There may be potential to use MAR in the CLA to offset the impacts of groundwater pumping on GDE near Lawn Hill Creek and the Gregory River. However, this would be contingent on finding soils suitable for the construction of ringtanks (Section 5.4.4), to temporarily detain surface water, close to a location that is suitable for MAR (typically on highly permeable soils). Groundwater use lowers groundwater levels and therefore creates storage capacity in the aquifer, which is required for MAR. However, the challenge remains to target aquifers with storage capacity at the end of the wet season, or to identify an available recharge source near and when there is sufficient storage capacity (i.e. early in the dry season). Infiltration techniques recharging unconfined aquifers are generally favoured for producing cost-effective water supplies, hence the initial focus on recharge techniques and limitations for unconfined aquifers. MAR opportunity maps were developed from the best available data at the catchment scale using the method outlined in the Northern Australia Water Resource Assessment technical report on MAR (Vanderzalm et al., 2018). This method uses four suitability classes for the more promising aquifers for MAR: • Class 1 – highly permeable and low slope (<5%) • Class 2 – highly permeable and moderate slope (5% to 10%) • Class 3 – moderately permeable and low slope (<5%) • Class 4 – moderately permeable and moderate slope (5% to 10%). Class 1 is considered most suitable for MAR and Class 4 least suitable. All areas not classified into one of classes 1, 2, 3 and 4 are considered unsuitable. Figure 5-13 shows the suitability map for MAR in the Southern Gulf catchments, with classes 1 and 2 considered potentially suitable for MAR. The opportunity assessment (Figure 5-14) indicates approximately 480 km2 (0.5%) of the Southern Gulf catchments may have aquifers with potential for MAR within 5 km of a major drainage line (excluding the highly intermittent drainage lines on the Sturt Plateau). Approximately 75 km2 (~0.1%) of the study area is considered Class 1 or Class 2 and is within 1 km2 of a major drainage line. Water-level data for monitoring bores across the Southern Gulf catchments provide some insight into the potential for aquifers to store additional water. A watertable level deeper than 4 m is recommended in order to have sufficient storage capacity for MAR. Sufficient aquifer storage space is indicated where depth to water is either greater than 4 m at the end of the wet season (i.e. available for recharge year round) or greater than 4 m at the end of the dry season (i.e. available for seasonal recharge). Bores recording depth to water of less than 4 m at the end of the dry season could be considered as indicative that no storage space exists at any time of year. No water-level data are available for the Quaternary alluvium aquifers (Figure 5-14) and have been used for MAR elsewhere in Australia. However, the opportunity to use these aquifers for MAR in the Southern Gulf catchments is unknown and may vary between locations. Figure 5-13 Managed aquifer recharge (MAR) opportunities for the Southern Gulf catchments independent of distance from a water source for recharge Analysis based on the permeability (Thomas et al., 2024) and terrain slope (Gallant et al., 2011) datasets and limited to the Cambrian Limestone and Cenozoic alluvium aquifers. MAR map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-550_SG_MAR_suitabilitly_v01_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au The opportunity assessment (Figure 5-14) indicates approximately 122,700 ha (1.1%) of the Southern Gulf catchments may have aquifers (including areas of Quaternary alluvium) with potential for MAR within 5 km of drainage lines that have a median annual flow greater than 20 GL. Approximately 37,100 ha (~0.3%) of the study area is considered Class 1 or Class 2 and is within 1 km of drainage lines and with a median annual flow greater than 20 GL. However, 94% of this latter area is underlain by Quaternary alluvium aquifers for which the storage capacity and water level are unknown. See the Northern Australia Water Resource Assessment technical report on MAR schemes in northern Australia (Vanderzalm et al., 2018) for detailed costings on ten hypothetical MAR schemes in northern Australia. Figure 5-14 (a) Managed aquifer recharge (MAR) opportunities in the Southern Gulf catchments within 5 km of major rivers and (b) aquifer underlying the MAR opportunity classes Analysis based on the permeability (Thomas et al., 2024) and terrain slope (Gallant et al., 2011) datasets and limited to the Cambrian Limestone and Cenozoic alluvium aquifers. 5km major riv MAR map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\11_Groundwater\4_S_Gulf\1_GIS\1_Map_docs\Gr-S-551_MAR_Suitability__aquifer_5kmMRivers_v01_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au 5.4 Surface water storage opportunities 5.4.1 Introduction In a highly seasonal climate, such as that of the Southern Gulf catchments, and in the absence of suitable groundwater, surface water storages are essential to enable irrigation during the dry season and other periods when soil water is insufficient for crop growth. The Assessment undertook a pre-feasibility-level assessment of three types of surface water storage options. These were: •major dams that could potentially supply water to multiple properties (Section 5.4.2) •re-regulating structures such as weirs (Section 5.4.3) •large farm-scale or on-farm dams, which typically supply water to a single property (Section 5.4.4 and Section 5.4.5). Both major dams and large farm-scale dams can be further classified as instream or offstream water storages. In the Assessment, instream water storages are defined as structures that intercept a drainage line (creek or river) and are not supplemented with water from another drainage line. Offstream water storages are defined as structures that (i) do not intercept a drainage line or (ii) intercept a small drainage line and are largely supplemented with water extracted from a nearby larger drainage line. Ringtanks and turkey nest tanks are examples of offstream storages with a continuous embankment; the former is the focus in the Assessment due to their higher storage-to-excavation ratios, relative to the latter. The performance of a dam is often assessed in terms of water yield. This is the amount of water that can be supplied for consumptive use at a given reliability. For a given dam, an increase in water yield results in a decrease in reliability. Importantly, the Assessment does not seek to provide instruction on the design and construction of farm-scale water storages. Numerous books and online tools provide detailed information on nearly all facets of farm-scale water storage (e.g. IAA, 2007; Lewis, 2002; QWRC, 1984). Siting, design and construction of weirs, large farm-scale ringtanks and gully dams are heavily regulated in most jurisdictions across Australia and should always be undertaken in conjunction with a suitably qualified professional and tailored to the nuances that occur at every site. Major dams are complicated structures and usually involve a consortium of organisations and individuals. Unless otherwise stated, the material in Section 5.4 originates from the companion technical report on surface water storage (Yang et al., 2024). 5.4.2 Major dams Introduction Major dams are usually constructed from earth, rock and/or concrete materials, and typically act as a barrier wall across a river to store water in the reservoir created. They need to be able to safely discharge the largest flood flows likely to enter the reservoir, and the structure has to be designed so that the dam meets its purpose, generally for at least 100 years. Some dams, such as the Kofini Dam in Greece and the Anfengtang Dam in China, have been in continuous operation for over 2000 years, with Schnitter (1994) consequently coining dams as ‘the useful pyramids’. An attraction of major dams over farm-scale dams is that if the reservoir is large enough relative to the demands on the dam (i.e. water supplied for consumptive use and ‘lost’ through evaporation and seepage), when the reservoir is full, water can last 2 or more years. This has the advantage of mitigating against years with low inflows to the reservoir. For this reason, major dams are sometimes referred to as ‘carry-over storages’. Major instream versus offstream dams Offstream water storages were among the first man-made water storages (Nace, 1972; Scarborough and Gallopin, 1991) because people initially lacked the capacity to build structures that could block rivers and withstand large flood events. One of the advantages of offstream storages is that, if properly designed, they can cause less disruption of the natural flow regime than large instream dams. Less disruption occurs if water is extracted from the river using pumps, or if there is a diversion structure with gates that can be raised, to allow water and aquatic species to pass through when not in use. In the very remote environments of northern Australia, the period in which these gates need to be operated is also the period in which it is difficult to move around wet roads and flooded waterways. The primary advantage of large instream dams is that they provide a very efficient way of intercepting the flow in a river, effectively trapping all flow until the full supply level (FSL) is reached. For this reason, however, they also provide a very effective barrier to the movement of fish and other species within a river system, alter downstream flow patterns and can inundate large areas of land upstream of the dam. Types of major dams Two types of major dams are particularly suited to northern Australia: embankment dams and concrete gravity dams. Embankment dams are usually the most economic, provided suitable construction materials can be found locally, and are best suited to smaller catchment areas where the spillway capacity requirement is small. Concrete gravity dams with a central overflow spillway are generally more suitable where a large-capacity spillway is needed to discharge flood inflows, as is the case in most large catchments in northern Australia. Traditionally, concrete gravity dams were constructed by placing conventional concrete in formed ‘lifts’. Since construction of Kidston Dam (officially known as the Copperfield River Gorge Dam) in 1984 in Australia, however, roller compacted concrete (RCC) has been used, where low-cement concrete is placed in continuous thin layers from bank to bank and compacted with vibrating rollers. This approach allows large dams to be constructed in a far shorter time frame than required for conventional concrete construction, often with large cost savings (Doherty, 1999). RCC is best used for high dams where a larger-scale plant can provide significant economies of scale. This is now the favoured type of construction in Australia whenever foundation rock is available within reasonable depth, and where a larger-capacity spillway is required. In those parts of the Southern Gulf catchments with topography and hydrology most suited to large instream dams, RCC was deemed to be the most appropriate type of dam. Opportunity-level assessment of potential major dams in the Southern Gulf catchments A promising dam site requires inflows of sufficient volume and frequency, topography that provides a constriction of the river channel and, critically, favourable foundation geology. With few studies of large dams in the Southern Gulf catchments identified, the opportunity-level assessment of potential major dams in the Southern Gulf catchments was undertaken using a spatial analysis approach. To ensure no potential dam site had been overlooked, the Assessment used a bespoke computer model, the DamSite model (Petheram et al., 2017), to assess over 50 million sites in the Southern Gulf catchments for their potential as major offstream or instream dams. Broad-scale geological considerations The Southern Gulf catchments drain from the uplands in the south-west and south towards the north-east (referred to as the ‘uplands’), where the river systems cross a broad depositional plain several tens of kilometres wide (referred to as the Armraynald Plain) before emptying into the Gulf of Carpentaria. Favourable foundation conditions include a relatively shallow layer of unconsolidated materials, such as alluvium, and rock that is relatively strong, resistant to erosion, non-permeable or capable of being grouted. Geological features that make dam construction challenging include the presence of faults, weak geological units, landslides and deeply weathered zones. Potentially feasible large dam sites in the Southern Gulf catchments occur where resistant ridges of Proterozoic sandstone beds that have been incised by the river systems outcrop on both sides of river valleys. The sandstones are generally weathered to varying degrees and the depth of weathering and the amount of sandstone outcrop on the valley slopes is a fundamental control on the suitability of the potential dam sites. Where the sandstones are relatively unweathered and outcrop on the abutments of the potential dam site, less stripping will be required to achieve a satisfactory founding level for the dam. The other fundamental control on the suitability of the dam is the extent and depth of the Quaternary alluvial sands and gravels in the floor of the valley, as these materials will have to be removed to achieve a satisfactory founding level for the dam. In general, where stripping removes the more weathered rock, it is anticipated that the Proterozoic sandstones will form a reasonably watertight dam foundation requiring conventional grout curtains and foundation preparation. Where potentially soluble dolomites occur within the Proterozoic sequences (soluble over a geological timescale) then it is possible that potentially leaky dam abutments and reservoir rims may be present, requiring specialised and costly foundation treatment such as extensive grouting. Where rivers are tidal the presence of soft estuarine sediments has the potential to make dam design more challenging and construction more expensive, which may compromise the feasibility of a dam. Sites potentially topographically suitable for large storages for water supply Figure 5-15 displays the most promising sites across the Southern Gulf catchments in terms of topography, assessed in terms of approximate cost of construction per storage volume (ML). Favourable locations with a small catchment area and adjacent to a large river may be suitable as major offstream storages. Figure 5-15 Potential storage sites in the Southern Gulf catchments based on minimum cost per megalitre storage capacity This figure can be used to identify locations where topography is suitable for large offstream storages. At each location the minimum cost per megalitre storage capacity is displayed. The smaller the minimum cost per megalitre storage capacity ($/ML) the more suitable the site for a large offstream storage. Analysis does not take into account geological considerations, hydrology or proximity to water. Only sites with a minimum cost-to-storage-volume ratio of less than $5000/ML are shown. Costs are based on unit rates and quantity of material and site establishment for a roller compacted concrete dam. Inset displays height and width of dam wall at full supply level at the minimum cost per megalitre storage capacity. Potential storage sites cost, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS506-S_Storage_Unit_cost_map_v4_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au In Figure 5-15, only those locations with a ratio of cost to storage less than $5000/ML are shown. This is a simple way to display those locations in the Southern Gulf catchments with the most favourable topography for a large reservoir relative to the size (i.e. cost) of the dam wall necessary to construct the reservoir. This figure can be used to identify more promising sites for offstream storage (i.e. where some or all of the water is pumped into the reservoir from an adjacent drainage line). The threshold value of $5000/ML is nominal and was used to minimise the amount of data displayed. This analysis does not consider evaporation, hydrology or geological suitability for dam construction. Figure 5-15 shows that those parts of the Southern Gulf catchments with the most favourable topography for storing water are on the larger drainage lines on the uplands, in particular the Nicholson and Gregory rivers, Gunpowder Creek and the Leichhardt River. Major instream dams for water and irrigation supply In addition to suitable topography (and geology), instream dams require sufficient inflows to meet a potential demand. Potential dams that command smaller catchments with lower runoff have smaller yields. Results concerning this criterion are presented in terms of minimum cost per unit yield, where the smaller the cost per megalitre yield ($/ML) the more favourable the site for a large instream dam. The potential for major instream dams to cost-effectively supply water is presented in Figure 5-16. No values greater than $10,000/ML are shown. The DamSite modelling results shown in Figure 5-16 indicate that the most cost-effective potential dam sites are on the Nicholson and Gregory rivers and Gunpowder Creek. The streamflow inputs to the DamSite modelling undertaken for the Southern Gulf catchments did not consider existing storages on the Leichhardt River, and hence the suitability of those sites shown on the Leichhardt River would be further diminished with reduced inflows. Also shown on this figure is the versatile agricultural land for the Southern Gulf catchments, with the most versatile agricultural land occurring on the Carpentaria Plain (Thomas et al., 2024). Based on this analysis and a broad-scale desktop geological evaluation, seven of the more cost- effective larger-yielding sites in distinct geographical areas that are proximal to soils suitable for irrigated agriculture were selected for pre-feasibility analysis (see companion technical report on surface water storage, Yang et al., 2024) to explore the potential opportunities and risks of water supply dams in the Southern Gulf catchments. The locations of these pre-feasibility potential dam sites are denoted in Figure 5-16 by black circles and the letters ‘A’ to ‘G’. No potential sites upstream of Julius Dam on the Leichhardt River (Figure 5-17) were examined as part of the pre- feasibility analysis as a dam upstream of Julius Dam would reduce the inflows and yield of the latter. Key parameters and performance metrics are summarised in Table 5-5 and an overall summary comment is recorded in Table 5-6. More detailed analysis of the seven pre-feasibility sites is provided in the companion technical report on surface water storage (Yang et al., 2024). Figure 5-16 Potential storage sites in the Southern Gulf catchments based on minimum cost per megalitre yield at the dam wall This figure indicates those sites more suitable for major dams in terms of cost per megalitre yield at the dam wall in 85% of years overlain on versatile land surface (see companion technical report on land suitability, Thomas et al., 2024). At each location the minimum cost per ML storage capacity is displayed. Only sites with a minimum cost to yield ratio less than $10,000/ML are shown. Costs are based on unit rates and quantity of material required for a roller compacted concrete dam with a flood design of 1 in 10,000. Right inset displays height of full supply level (FSL) at the minimum cost per ML yield and left inset displays width of FSL at the minimum cost per ML yield. Letters indicate potential dams listed in Table 5-5 and Table 5-6. Potential storage sites yield cost, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS507-S_Yield_Unit_cost_map_v4_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-17 Julius Dam on the Leichhardt River Photo: CSIRO Hydro-electric power generation potential in the Southern Gulf catchments The potential for major instream dams to generate hydro-electric power is presented in Figure 5-18 following an assessment of more than 50 million potential dam sites in the Southern Gulf catchments (Yang et al., 2024). This figure provides indicative estimates of hydro-electric power generation potential but does not consider the existence of supporting infrastructure or geological suitability for dam construction. No values greater than $45,000/MWh are shown. Although the topography of the Southern Gulf catchments is moderately suitable for large water storage dams (i.e. narrow constrictions downstream of broad valleys), the topography appears to be less suitable for dams for hydro-electric power generation due to the lack of relief that is required to provide potential elevation head. Aside from a specific site on the Gregory River, Gunpowder Creek, a major tributary of the Leichhardt River, is the most favourable drainage line for dams for the purpose of hydro-electric power generation in the Southern Gulf catchments due to the more favourable topography and its proximity to Mount Isa, a major power demand centre. A companion technical report on hydro-electric power generation (Entura, 2024) undertakes a pre-feasibility analysis of a site on Gunpowder Creek to explore the potential for hydro-electric power in the Southern Gulf catchments. A dam with a body of water Description automatically generated Figure 5-18 Southern Gulf catchments hydro-electric power generation opportunity map Costs are based on unit rates and quantity of material required for a roller compacted concrete dam with a flood design of 1 in 10,000. Right inset displays height of full supply level (FSL) at the minimum cost per megalitre yield and left inset displays width of FSL at the minimum cost per megalitre yield. Hydro power opportunity, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS508-S_Hydropower_Unit_cost_map_v4_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Pre-feasibility-level assessment of potential major dams in the Southern Gulf catchments Seven potential dam sites in the Southern Gulf catchments were examined as part of this pre- feasibility assessment. They are summarised in Table 5-5 and Table 5-6. More detailed descriptions of these seven potential dam sites, including impacts to migratory species and ecological impacts of reservoir inundation, are provided in the companion technical report on surface water storage (Yang et al., 2024). Table 5-5 Potential dam sites in the Southern Gulf catchments examined as part of the Assessment All hypothetical dams are assumed to be roller compacted concrete. Locations of potential dams are shown in Figure 5-16. Geology grade is an ordinal scale between 1 and 5 where grade 1 is best, grade 5 is worst. It is based on a holistic assessment based on whether bedrock is exposed at site, likely depth of weathering/stripping on abutments, likely depth of cut-off and presence of deep alluvium, and overall height-to-width assessment. AMTD = adopted middle thread distance; FSL = full supply level; O&M = operation and maintenance. For more information on this table please contact CSIRO on enquiries@csiro.au *The height of the dam abutments and saddle dams will be higher than the spillway height. **Water yield is based on 85% annual time-based reliability using a perennial demand pattern for the baseline river model under Scenario AE. This is yield at the dam wall (i.e. does not take into account distribution losses or downstream transmission losses). With the exception of the Gregory River AMTD 174 km site and the Gunpowder Creek AMTD 66 km site, the yield values do not take into account downstream existing entitlement holders. At none of the sites do the yield estimates take into account of environmental considerations. # Indicates manually derived preliminary cost estimate, which is likely to be –10% to +50% of ‘true cost’.  Indicates modelled preliminary cost estimate, which is likely to be –25% to +100% of ‘true’ cost. Should site geotechnical investigations reveal unknown unfavourable geological conditions, costs could be substantially higher. ##The unit cost of annual water supply and is calculated as the capital cost of the dam divided by the water yield at 85% annual time reliability. ###Assumes a 7% real discount rate and a dam service life of 100 years. Includes operation and maintenance (O&M) costs, assuming operation and maintenance (O&M) costs are 0.4% of the total capital cost. Table 5-6 Summary comments for potential dams in the Southern Gulf catchments Locations of potential dams are shown in Figure 5-16. AMTD = adopted middle thread distance. SITE NAME MAP ID SUMMARY COMMENT South Nicholson River AMTD 9 km A This potential dam site is the most expensive of those on the short-list and is a considerable distance upstream of the soil potentially suitable for irrigated agriculture. Furthermore, the site is very remote and access would require considerable additional infrastructure. At the adopted full supply level (FSL), the inundation extent of this potential dam site overlaps with parts of the Ganalanga–Mindibirrina Indigenous Protected Area (IPA). There is a high likelihood of unrecorded sites of cultural significance in the inundation area. Asset models predicted that 46% of the catchment has habitat suitable for at least one of the ten modelled migratory species, which constitutes zero % to 6.4% of the total habitat modelled as suitable in the Southern Gulf catchments. Gold Creek AMTD 58 km B This small potential instream dam site on a small catchment in the Settlement Creek Australian Water Resources Council river basin has a low yield and high cost per megalitre released from the dam wall. Downstream of the site, water could potentially be supplied to land with soils suitable for agriculture, with minor or moderate limitations, depending on the land use. Approximately 11% of the catchment was estimated as having habitat suitable for 40% or more of the (11) migratory species modelled. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. Nicholson River AMTD 198 km C This potential instream development has the potential to release water for irrigation along the Nicholson River to the Doomadgee and Armraynald plains and Doomadgee. The foundations appear to be suitable for a roller compacted concrete (RCC) dam. Nonetheless, the site is very remote and is one of the more expensive potential dam sites examined in the Southern Gulf catchments. At the adopted FSL, the inundation extent of this potential dam site overlaps with parts of the Ganalanga–Mindibirrina IPA. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. Gregory River AMTD 174 km D The potential Gregory River dam site is an instream development with potential to supply water to the large contiguous areas of black and grey cracking clay soils on the Armraynald Plain immediately downstream of the dam. Given the potential for significant flooding during construction and the spillway capacity required, an RCC gravity dam with a 400 m wide central uncontrolled spillway would be most suitable. The site is one of the largest-yielding and most cost-effective potential dam sites in the Southern Gulf catchments, and the foundations of this site appear to be suitable for an RCC dam. A major limitation of the site is its proximity to the Boodjamulla (Lawn Hill) National Park (upstream of the site), and the Thorntonia Aggregation wetland, which is up and downstream of the potential dam site. Despite being on a major river, the area of potentially suitable habitat for the four modelled migratory species was relatively small. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. Limiting the FSL of the potential dam so that the area inundated would not extend into the Boodjamulla National Park resulted in a modelled yield of 180 GL in 85% of years at the dam wall. However, taking into consideration existing downstream users and assuming a dry-season crop resulted in a modelled yield of 133 GL in 85% of years. Gunpowder Creek AMTD 66 km E This potential dam site is on Gunpowder Creek, a large tributary of the Leichhardt River, and it has the second-lowest cost per megalitre of the short-listed sites in the Southern Gulf catchments. The foundations appear to the suitable for an RCC dam. The site would nominally supply water to a large plain of recent alluvium at the junction of Gunpowder Creek and the Leichhardt River. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. The area of potentially suitable habitat for modelled migratory species upstream of this dam site is relatively small. Previous studies of dams on Gunpowder Creek focused on areas further upstream, predominantly to supply water for mining. Mistake Creek AMTD 60 km F This potential dam site would supply water to large areas of contiguous soils suitable for irrigated agriculture on the Carpentaria Plains. Being a relatively small tributary of the Leichhardt River, the site has a low yield, and it has the highest cost per megalitre released from the dam wall of all the short-listed sites examined in the Southern Gulf catchments. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. For the migratory species modelled as part of the Assessment, the catchment of a dam at this potential site would constitute less than 1% of the total potentially suitable habitat modelled in the Southern Gulf catchments. Ewen Creek AMTD 6 km G This low-yielding and relatively expensive potential dam site is located on the tributary that joins the Leichhardt River downstream of Lake Julius. The foundations appeared to be suitable for a RCC dam; however, the amount of dolomite rock in the foundation may represent a problem, due to cavities within dolomite strata. This would need to be investigated. Given the small catchment area and modest percentage of habitat suitable for migratory species upstream of the potential dam (22% for at least one species), a potential dam at this location would have an effect on zero % to 0.5% of the total potentially suitable habitat modelled in the Southern Gulf catchments. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. The investigation of a potential large dam site generally involves an iterative process of increasingly detailed studies over a period of years, occasionally over as few as 2 or 3 years but often over 10 or more years. It is not unusual for the cost of the geotechnical investigations for a potential dam site alone to exceed several million dollars. For any of the options listed in this report to advance to construction, far more comprehensive studies would be needed, including geotechnical investigations, field measurements of sediment yield, archaeological surveys and ground-based vegetation and fauna surveys, as well as extensive consultations with Traditional Owners and other stakeholders. Studies at that detail are beyond the scope of this regional-scale resource assessment. The companion technical report on surface water storage (Yang et al., 2024) outlines the key stages in investigation of design, costing and construction of large dams. More comprehensive descriptions are provided by Fell et al. (2005), while Indigenous Peoples’ views on large-scale water development in north-eastern Queensland can be found in the companion technical report on Indigenous aspirations, interests and water values (Lyons et al., 2024). Other important considerations Cultural heritage considerations Indigenous Peoples traditionally situated their campsites, and hunting and foraging activities, along major watercourses and drainage lines. Consequently, dams are more likely to affect areas of high cultural significance than are most other infrastructure developments (e.g. irrigation schemes, roads). No field-based cultural heritage investigations of potential dam and reservoir locations were undertaken in the Southern Gulf catchments as part of the Assessment. However, based on existing records and statements from Indigenous participants in the Assessment, it is highly likely such locations will contain heritage sites of cultural, historical and wider scientific significance. Information relating to the cultural heritage values of the potential major dam sites is insufficient to allow full understanding or quantification of the likely impacts of water storages on Indigenous cultural heritage. The cost of cultural heritage investigations associated with large instream dams that could potentially impound large areas is high relative to other development activities. Ecological considerations of the dam wall and reservoir The water impounded by a major dam inundates an area of land, drowning not only instream habitat but surrounding flora and fauna communities. Complex changes in habitat resulting from inundation could create new habitat to benefit some of these species, while other species would be affected by loss of habitat. For instream ecology, the dam wall acts as a barrier to the movements of plants, animals and nutrients, potentially disrupting connectivity of populations and ecological processes. There are many studies linking water flow with nearly all the elements of instream ecology in freshwater systems (e.g. Robins et al., 2005). The impact of major dams on the movement and migration of aquatic species will depend upon the relative location of the dam walls in a catchment. For example, generally a dam wall in a small headwater catchment will have less of an impact on the movement and migration of species than a dam lower in the catchment. A dam also creates a large, deep lake, a habitat that is in stark contrast to the usually shallow and often flowing, or ephemeral, habitats it replaces. This lake-like environment favours some species over others and will function completely differently to natural rivers and streams. The lake-like environment of an impoundment is often used by sports anglers to augment natural fish populations by artificial stocking. Whether fish stocking is a benefit of dam construction is a matter of debate and point of view. Stocked fisheries provide a welcome source of recreation and food for fishers, and no doubt an economic benefit to local businesses, but they have also created a variety of ecological challenges. Numerous reports of disruption of river ecosystems (e.g. Drinkwater and Frank, 1994; Gillanders and Kingsford, 2002) highlight the need for careful study and regulatory management. Impounded waters may be subject to unauthorised stocking of native fish and releases of exotic flora and fauna. Further investigation of any of these potential dam sites would typically involve a thorough field investigation of vegetation and fauna communities. Ecological assets in the Victoria catchment are discussed in Section 3.2 and described in more detail in the companion technical reports on ecological assets (Merrin et al., 2024) and surface water storage (Yang et al., 2024). Potential changes to instream, riparian and near-shore marine species arising from changes in flow are discussed in Section 7.3 . Sedimentation Rivers carry fine and coarse sediment eroded from hill slopes, gullies and banks, and sediment stored within the channel. The delivery of this sediment into a reservoir can be a problem because it can progressively reduce the volume available for active water storage. The deposition of coarser-grained sediments in backwater (upstream) areas of reservoirs can also cause back- flooding beyond the flood limit originally determined for the reservoir. Although infilling of the storage capacity of smaller dams has occurred in Australia (Chanson, 1998), these dams had small storage capacities, and infilling of a reservoir is generally only a potential problem where the volume of the reservoir is small relative to the catchment area. Sediment yield is strongly correlated to catchment area (Tomkins, 2013; Wasson, 1994). Sediment yield to catchment area relationships developed for northern Australia (Tomkins, 2013) were found to predict lower sediment yield values than global relationships. This is not unexpected given the antiquity of the Australian landscape (i.e. it is flat and slowly eroding under ‘natural’ conditions). Using the relationships developed by Tomkins (2013), potential major dams in the Southern Gulf catchments were estimated to have about 2% or less sediment infilling after 30 years and less than 5% sediment infilling after 100 years. The exception is the potential dam at Gregory River AMTD 174 km with an FSL of 138 mEGM96 to avoid incursion of the reservoir into the Boodjamulla National Park upstream. At this FSL the reservoir capacity is small (118 GL) relative to the catchment area of the potential dam (11,381 km2), and consequently it was estimated sediment could potentially infill about 15% and 50% of the reservoir’s capacity after 30 years and 100 years, respectively. Cumulative yield of multiple hypothetical dams in the Southern Gulf catchments This analysis examined the combined or cumulative yield of multiple dams in the Southern Gulf catchments and the resulting impact to end-of-system (EOS) flows. To undertake this analysis, the most promising hypothetical dam sites (in terms of lowest cost per megalitre at the dam wall) were incrementally included in each river model simulation. Cumulative yields are reported at the dam wall and do not include transmission and conveyance losses. Figure 5-19 shows the cumulative yield (left y-axis) from sequential dams as triangles, and the percentage change in median annual EOS volume is represented as circles on the right y-axis. The total yield at the dam wall at an annual time reliability of 85% while also maintaining supply to existing entitlement holders downstream of six of the seven instream dams analysed is 733 GL, with the majority of this yield provided by the first three dams (641 GL). The seventh dam, South Nicholson River AMTD 9 km, is not included in this analysis as it is upstream of the Nicholson River AMTD 198 km dam, resulting in reduced cumulative yield. The results from this analysis were used to investigate the cumulative impacts of multiple dams in the Southern Gulf catchments (Section 7.3). Figure 5-19 Cumulative yield at 85% annual time reliability versus cumulative cost of water in $/ML and change in the end-of-system (EOS) volume in the Southern Gulf catchments Yield is reported at the dam wall under Scenario A. Triangles indicate combined water yield at 85% annual time reliability of one or more dams, and the colour of the dot indicates the most recently included dam in the cumulative yield calculation. Circles indicate change in median annual streamflow at the EOS for all mainland catchments compared to under Scenario A. For more information on this figure please contact CSIRO on enquiries@csiro.au Exploration of two potential dam sites in the Southern Gulf catchments Two potential dam sites on different rivers are summarised here. These sites are described because they are among the more cost-effective sites near relatively large continuous areas of land suitable for irrigated agriculture in the Southern Gulf catchments, and not too remote relative to other potential dam sites in the study area. More detailed descriptions of the seven sites selected for pre-feasibility assessment are provided in the companion technical report on surface water storage (Yang et al., 2024). Potential dam on Gunpowder Creek AMTD 66 km and FSL 186 mEGM96 for water supply The potential Gunpowder Creek dam site is an instream development with the potential to provide irrigation supplies downstream near the junction of Gunpowder Creek and the Leichhardt River. Previous investigations commissioned by the then Queensland Irrigation and Water Supply Commission (QWIS, 1960, 1974) in the 1970s examined a number of potential dam sites along Gunpowder Creek for their potential to supply proposed mining developments. The potential sites investigated by the Queensland Irrigation and Water Supply Commission are upstream of the potential dam site examined in this section. There have been no recent publicly funded investigations. The foundations at the nominated site appear suitable for an RCC dam and it is estimated that both the depth of alluvium and stripping on the abutments would be about 3 to 5 m. However, the extent of any dolomite in the sequence at the potential dam site should be checked in case there are strata with cavities, although the previous investigations at nearby sites did not note any problems. Given the potential for significant flooding during construction and the spillway capacity required, an RCC dam with a 140 m wide central uncontrolled spillway would be most suitable. Releases of water downstream of the dam could be made through two conduits installed in the right abutment of the dam, and would be regulated by two 900 mm diameter fixed-cone regulating valves. A fish-lift transfer facility would be installed in the right abutment of the potential dam. No site-specific evaluation of cultural heritage considerations was possible, as pre-existing Indigenous cultural heritage site records were not made available to the Assessment. Land tenure and native title information were derived from regional land councils and the National Native Title Tribunal. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. Habitat species distribution models developed for the study area estimate that approximately 5% of the potential dam catchment (18,028 ha) has suitable habitat for at least 40% of the ten species modelled (Figure 5-20). The estimated area of suitable habitat for these water- dependent species upstream of the potential dam site is small relative to the total estimated suitable habitat in the Southern Gulf catchments, ranging from zero% to 3% of the study area depending on the species. These species may have their habitat fragmented, and/or their movement may be impeded by a dam. A manual cost estimate undertaken as part of the Assessment for a hypothetical RCC dam on the Gunpowder Creek AMTD 66 km site at an FSL of 186 mEGM96 was approximately $773 million. Access to the site could be along a 75 km long extension of the existing Mount Isa to Gunpowder Road, although this extension would require further investigation should this proposal be considered further. The total distance from Mount Isa to the dam site would be 205 km. Assuming full use of existing licence entitlement holders in the Leichhardt catchment and a dry-season demand pattern, the yield at 85% annual time reliability and FSL of 186 mEGM96 was modelled to be 119 GL at the dam wall (Gibbs et al., 2024). Figure 5-21 shows the modelled dam cost and yield at the dam wall without consideration of existing entitlement holders. Figure 5-20 Location of listed species, water-dependent assets and aggregated modelled habitat in the vicinity of the potential dam site on Gunpowder Creek dam AMTD 66 km and reservoir extent Near the junction of Gunpowder Creek and the Leichhardt River, approximately 35 km downstream of the potential dam site, a large plain of recent alluvium has formed. The nearest location for a potential re-regulating structure appears to be about 5 km upstream of the hypothetical target area. Friable non-cracking clay or clay loam soils (soil generic group (SGG) 2) and sand to loam over relatively friable red clay subsoils (SGG 1.1) have formed on the recent alluvium of the floodplain. Brown cracking clay soils (SGG 9) occur on the older Cloncurry Plain. Gunpowder Ck dam water-dependent assets, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS533-S_Dam024_Ecology_v03_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-21 Potential dam site on Gunpowder Creek AMTD 66 km: cost and yield at the dam wall Dam length and dam cost versus full supply level (FSL), and (b) dam yield and yield/$ million at 75% and 85% annual time reliability. Due to the relatively steep slope of the land adjacent to the re-regulating weir and due to the texture of the soils, the likely irrigation application methods being spray or trickle, piped reticulation from the weir pump station would be likely. It is possible that this potential dam site could service about 11,000 ha, depending on crop type, irrigation method and reticulation arrangement (see companion technical report on irrigation systems in the Victoria and Southern Gulf catchments, Devlin, 2024). Based on a notional layout of pipeline infrastructure outlined by Devlin (2024), the cost of reticulation infrastructure to irrigate this area would be about $320 million or about $29,100 per irrigated hectare. Potential dam on Gregory River AMTD 174 km and FSL 138 mEGM96 for water supply The potential dam site on the Gregory River AMTD 174 km was first identified by the then Queensland Irrigation and Water Supply Commission in the late 1960s (QIWS, 1969) and subsequently identified by the Assessment as one of the largest-yielding and most cost-effective potential dam sites in the Southern Gulf catchments. There has been no recent consideration of this site for a potential water storage development. A major limitation of the site is its proximity to the Boodjamulla (Lawn Hill) National Park (upstream of the site), and parts of the Thorntonia Aggregation wetland, a Directory of Important Wetlands in Australia nationally important wetland, overlaps the potential reservoir and extends downstream from the potential dam wall. This hypothetical instream development has potential to supply water to the Armraynald Plain downstream of the dam. The hypothetical dam site is located on Proterozoic rocks of the Shady Bore Quartzite, which consists of white medium orthoquartzite, siltstone and dolomitic fine sandstone, which appear to be gently folded and dipping upstream. The foundations appear suitable for an RCC dam and the depth of alluvium is estimated to be 7 to 10 m; 5 to 7 m of stripping would be required on the potential dam abutments. Given the potential for significant flooding during construction, and the spillway capacity required, the RCC dam would have a 400 m wide central uncontrolled spillway. A hydraulic jump type \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\3_Plotting_scripts\1_output\figure2 For more information on this figure please contact CSIRO on enquiries@csiro.au spillway basin would be provided to protect the river bed against erosion during spillway overflows. Releases downstream of the dam would be made through pipework installed in a diversion conduit located in the right abutment of the dam. A fish-lift transfer facility would also be installed in the right abutment. Access to the site would be along 27 km of new road built to serve the potential dam branching from the Camooweal–Gregory Road some 58 km south-west of Gregory. The total distance of the site from Mount Isa via Camooweal is 371 km. No site-specific evaluation of cultural heritage considerations was possible, as pre-existing Indigenous cultural heritage site records were not made available to the Assessment. Land tenure and native title information were derived from regional land councils and the National Native Title Tribunal. There is a high likelihood of unrecorded sites of cultural significance in the inundation area. Despite being on a major river, the potentially suitable habitat for the four modelled migratory species was relatively small. It was estimated that approximately 3% of the potential dam catchment (35,144 ha) has suitable habitat for at least 40% of the ten species modelled (Figure 5-22). However, the estimated area of suitable habitat for these water-dependent species upstream of the potential dam site is small relative to the total estimated suitable habitat in the Southern Gulf catchments, ranging from 0.04% to 6.8% of the study area, depending on the species. These species may have fragmented habitat and/or their movement may be impeded by a dam. Figure 5-22 Location of listed species, water-dependent assets and aggregated modelled habitat in the vicinity of the potential dam site on the Gregory River AMTD 174 km The FSL 138 m EGM96 was selected such that the reservoir did not encroach into the Boodjamulla (Lawn Hill) National Park and Lawn Hill Resources Reserve, which are located about 9 km upstream of the potential dam wall. At this reduced FSL the yield is considerably reduced and the reservoir, which is small relative to its catchment area, has an elevated risk of having its capacity notably reduced by sediment infill (Yang et al., 2024). For example, taking into consideration existing licence entitlement holders and assuming a dry-season demand pattern, the yields at 85% annual time reliability at FSLs of 138 mEGM96 and 145 mEGM96 were modelled to be 133 GL and 233 GL at the dam wall, respectively (see companion technical report on river model simulation in the Southern Gulf catchments, Gibbs et al., 2024). Figure 5-23 shows the modelled dam cost and yield at the dam wall without consideration of existing entitlement holders. Gregory River dam water-dependent assets, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS532-S_Dam003_Ecology_v03_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-23 Potential dam site on Gregory River AMTD 174 km: cost and yield at the dam wall (a) Dam length and dam cost versus full supply level (FSL), and (b) dam yield and yield/$ million at 75% and 85% annual time reliability. A manual cost estimate undertaken as part of the Assessment for a hypothetical RCC dam on the Gregory River at AMTD 174 km at an FSL of 138 mEGM96 found the dam would cost approximately $683 million. Although the nearby Boodjamulla (Lawn Hill) National Park (upstream of the site) and Thorntonia Aggregation wetland are major limitations of this potential dam site, a major benefit is its proximity to the Armraynald Plain (approximately 35 km downstream of the potential dam site), which is a large contiguous area of Pleistocene sediments that formed black and grey cracking clay soils (SGG 9) suitable for a range of dry-season crops. Due to the slope of the Gregory River near the potential dam site, water would be released from the dam down the river to a potential re- regulating weir on a natural bar with a 2.5 km long pool upstream. One of the favourable features of the site is that the river levees adjacent to the potential re-lift pump point mean the land naturally slopes away from the river at a gradient favourable for cost-effective open channels. Notionally the hypothetical scheme would be based around using two main channel alignments and constructed from the in-situ cracking clay soils. Under the nominal irrigation scheme outlined by Devlin (2024), a hypothetical dam on the Gregory River at AMTD 174 km and FSL 138 mEMG96 could potentially supply sufficient water to irrigate approximately 10,000 ha at a reticulation scheme cost of $31.8 million or $3180/ha (excluding farm development and infrastructure; Devlin 2024). 5.4.3 Weirs and re-regulating structures Re-regulating structures, such as weirs, are typically located downstream of large dams. They allow for more efficient releases from the storages and for some additional yield from the weir storage itself, thereby reducing the transmission losses normally involved in supplemented river systems. As a rule of thumb, however, weirs are constructed to one-half to two-thirds of the river bank height. This height allows the weirs to achieve maximum capacity, while ensuring the change in downstream hydraulic conditions does not result in excessive erosion of the toe of the structure. It For more information on this figure please contact CSIRO on enquiries@csiro.au also ensures that large flow events can still be passed without causing excessive flooding upstream. Two types of weir structures have been constructed in Queensland: concrete gravity weirs and sheet piling weirs. These are discussed below. For each type of weir, rock-filled mattresses are often used on the stream banks, extending downstream of the weir to protect erodible areas from flood erosion. A brief discussion on sand dams is also provided. Weirs, sand dams and diversion structures obstruct the movement of fish in a similar way to dams during the dry season. Concrete gravity type weirs Where rock bars are exposed at bed level across a stream, concrete gravity type weirs have been built on the rock at numerous locations across Queensland. This type of construction is less vulnerable to flood erosion damage, both during construction and in service. Indicative costs are provided for a small weir structure with only sufficient height (e.g. 0.75 m above river bed) to submerge pumping infrastructure. Assuming exposed bedrock across the river bed, and rock for aggregates and mattresses, are available locally, the cost of a low reinforced concrete slab with upstand (i.e. 0.75 m above river bed, nominally 150 m width along crest) for the purpose of providing pump station submergence is estimated to cost about $13 million. Nominal allowances were made for site access, services and construction camp costs on the basis that more substantial site establishment costs would be incurred by the nearby irrigation development (Yang et al., 2024). Sheet piling weirs Where rock foundations are not available, stepped steel sheet piling weirs have been successfully used in many locations across Queensland. These weirs consist of parallel rows of steel sheet piling, generally about 6 m apart, with a step of about 1.5 to 1.8 m high between each row. Reinforced concrete slabs placed between each row of piling absorb much of the energy as flood flows cascade over each step. The upstream row of piling is the longest, driven to a sufficient depth to cut off the flow of water through the most permeable material (Figure 5-24). Indicative costs are provided in Table 5-7. In recent years the Queensland Department of Agriculture and Fisheries has not approved stepped weirs on the basis that the steps result in fish mortalities. Sheet piling weirs would therefore have to have a sloping face with a more extensive dissipator at bed level. Figure 5-24 Schematic cross-section of sheet piling weir Source: Petheram et al. (2013) Table 5-7 Estimated construction cost of 3 m high sheet piling weir Cost indexed to 2023. WEIR CREST LENGTH (m) ESTIMATED CAPITAL COST ($ million) 100 32 150 42 200 50 Sand dams As many of the large rivers in northern Australia are very wide (e.g. >300 m), weirs are likely to be impractical and expensive at many locations. An alternative structure is a sand dam, which consists of low embankments built of sand and constructed at the start of each dry season during periods of low or no flow, when heavy earth-moving machinery can access the bed of the river. A sand dam is constructed to form a pool of sufficient depth to enable pumping (i.e. typically greater than 4 m depth) and this type of dam is widely used in the Burdekin River near Ayr in Queensland, where the river is too wide to construct a weir. Typically, sand dams take three to four large excavators approximately 2 to 3 weeks to construct, and no further maintenance is required until they need to be reconstructed again after the wet season. Bulldozers can construct a sand dam more quickly than can a team of excavators, but they have greater access difficulties. Because a sand dam only needs to form a pool of sufficient size and depth from which to pump water, it usually only partially spans a river and is typically constructed immediately downstream of large, naturally formed waterholes. The cost of 12 weeks of hire for a 20 t excavator and float (i.e. for transportation) is approximately $100,000. Although sand dams are cheap to construct relative to a concrete or sheet piling weir, they require annual rebuilding and have much larger seepage losses beneath and through the dam wall. No studies are known to have quantified losses from sand dams. For more information on this figure please contact CSIRO on enquiries@csiro.au 5.4.4 Large farm-scale ringtanks Large farm-scale ringtanks are usually fully enclosed circular earthfill embankment structures constructed close to major watercourses/rivers to minimise the cost of pumping infrastructure by ensuring long ‘water harvesting’ windows. For this reason, they are often subject to reasonably frequent inundation, usually by slow-moving flood waters. In some exceptions embankments may not be circular; rather, they may be used to enhance the storage potential of natural features in the landscape such as horseshoe lagoons or cut-off meanders adjacent to a river (see Section 5.4.6 for discussion on extracting water from persistent waterholes). An advantage of ringtanks over gully dams is that the catchment area of the former is usually limited to the land that it impounds, so costs associated with spillways, failure impact assessments and constructing embankments to withstand flood surges are considerably less than those for large farm-scale gully dams. Another advantage of ringtanks is that unless a diversion structure is utilised in a watercourse to help ‘harvest’ water from a river, a ringtank and its pumping station do not impede the movement of aquatic species or transport of sediment in the river. Ringtanks also have to be sited adjacent to major watercourses to ensure there are sufficient days available for pumping. While this limits where they can be sited, it means that because they can be sited adjacent to major watercourses (on which gully dams would be damaged during flooding – large farm-scale gully dams are typically sited in catchments of areas less than 40 km2), they often have a higher reliability of being filled each year than gully dams. However, operational costs of ringtanks are usually higher than those of gully dams because water must be pumped into the structure each year from an adjacent watercourse, typically using diesel-powered pumps (solar and wind energy do not generate sufficient power to operate high-volume axial flow or centrifugal flow pumps). Even where diversion structures are utilised to minimise pumping costs, the annual cost of excavating sediment and debris accumulated in the diversion channel can be in the order of tens of thousands of dollars. For more information on ringtanks in the Southern Gulf catchments, refer to the companion technical reports on surface water storage (Yang et al., 2024) and river model simulation (Gibbs et al., 2024). Also of relevance is the Northern Australia Water Resource Assessment technical report on large farm-scale dams (Benjamin, 2018) and companion technical report on pumping stations in northern Australia (Devlin, 2024). A rectangular ringtank in the catchment of the Flinders River (Queensland) is pictured in Figure 5-25. In this section, the following assessments of ringtanks in the Southern Gulf catchments are reported: • suitability of land for large farm-scale ringtanks • reliability with which water can be extracted from different reaches • indicative evaporative and seepage losses from large farm-scale ringtanks • indicative capital, operation and maintenance costs of large farm-scale ringtanks. Figure 5-25 Rectangular ringtank and 500 ha of cotton in the Flinders catchment (Queensland) The channel along which water is diverted from the Flinders River to the ringtank can be seen in the foreground. Photo: CSIRO Suitability of land for ringtanks in the Southern Gulf catchments Figure 5-26 displays the broad-scale suitability of land for large farm-scale ringtanks in the Southern Gulf catchments. Approximately 21% of the Southern Gulf catchments is classed as being likely suitable. Those areas of the Southern Gulf catchments with soil and topography suitable for ringtanks are mainly restricted to the level, slowly permeable, rock-free cracking clay soils (SGG 9) of the Armraynald Plain, Barkly Tableland and northern parts of Donors Plateau (Figure 5-26). The very poorly drained saline coastal marine plains (Karumba Plain), which are subject to tidal inundation and have very deep, strongly mottled, grey non-cracking and cracking clay soils with potential acid-sulfate deposits in the profile, are likely to be suitable for ringtanks but are subject to storm surge from cyclones. The soils of the Armraynald Plain are very deep (>1.5 m), imperfectly drained, slowly permeable, medium to heavy clays that crack when dry and swell when wet, reducing the rate of deep drainage. Soils have a self-mulching clay surface with gilgai common. On the Barkly Tableland, the cracking clays are deep (1.2 to 1.5 m) and underlain by limestone and dolomite karst, and hence are often moderately well drained and with gravel common. On Donors Plateau, the cracking clay soils are shallower (<0.5 m). Figure 5-26 Suitability of land for large farm-scale ringtanks in the Southern Gulf catchments Soil and subsurface data were only available to a depth of 1.5 m, hence the Assessment does not consider the suitability of subsurface material below this depth. This figure does not consider the availability of water. Data are overlaid on a shaded relief map. The results presented in this figure are only indicative of suitable locations for siting a ringtank; site-specific investigations by a suitably qualified professional should always be undertaken prior to ringtank construction. Ring tank suitability map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS529-S-RingtankSuitability_v02_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Reliability of water extraction The existence of a water entitlement does not mean the full volume will be available in all years. The reliability for extraction of an allocation or volume of water from a river depends upon a range of factors including the: •quantity of discharge and the natural inter- and intra-variability of a river system (Section 2.5.5) •capacity of the pumps or diversion structure (expressed here as the number of days taken topump an allocation) •quantity of water being extracted by other users, and their locations •conditions associated with a licence to extract water, such as: –a minimum threshold (i.e. water height level/discharge) at which pumping can commence(pump start threshold) –an annual diversion commencement flow requirement (ADCFR), the volume that mustpass through the system before pumping can commence each water year (1 September to31 August). The ADCFR was assumed to be the combined outflow from rivers thatincluded hypothetical extractions, the Gregory–Nicholson and Leichhardt River basins. Licence conditions can be imposed on a potential water user to ensure downstream entitlement holders are not affected by new water extractions and to minimise environmental change that may arise from changes to streamflow. In some cases a pump start threshold may be a physical threshold below which it is difficult to pump water from a natural pumping pool, but it can also be a regulatory requirement imposed to minimise impacts to existing downstream users and mitigate changes to existing water-dependent ecosystems. The impact of pump start thresholds and ADCFR flow requirements on extraction reliability are explored because they are the least complex environmental flow provision to regulate and ensure compliance in remote areas. Although more targeted environmental flow provisions may be possible, these are inevitably more complicated for irrigators to adhere to (usually requiring many dozens of pump operations during a season) and more difficult for regulators to enforce. Within each river reach, water could be harvested by one or more hypothetical water harvesters and the water nominally stored in ringtanks adjacent to the river reach. The reliability of water extraction under different conditions and at different locations in the Southern Gulf catchments is detailed in the companion technical report for river model simulation (Gibbs et al., 2024). A selection of plots from that report are provided in Figure 5-27 to Figure 5-31 to illustrate key concepts. The locations of the hypothetical extractions are illustrated in the maps in the bottom right corners, and their relative proportions of the total system allocation (left vertical axis) were assigned based on joint consideration of crop versatility, broad-scale flooding, ringtank suitability and river discharge (see companion technical report on river modelling, Gibbs et al., 2024). Figure 5-27 can be used to explore the reliability of extracting (‘harvesting’) or diverting increasing volumes of water at five locations in the Southern Gulf catchments under varying pump start thresholds. The left vertical axis (y1-axis) indicates the system target volume, which is the maximum volume of water extracted across the Southern Gulf catchments each season (nominal catchment-wide entitlement volume). The right vertical axis (y2-axis) is the maximum volume of water extracted in that reach each season (nominal reach entitlement volume). Figure 5-27 Annual reliability of diverting annual system and reach target volumes for varying pump start thresholds The results assume no annual diversion commencement flow requirement before pumping can commence. Pump capacity is such that system and reach target volumes can be pumped in 20 days. Seven-digit numbers refer to model node locations. Black contour lines indicate conditions that meet a 75% annual reliability of supply, and grey contour lines on node 9139000 indicate proportion of years with a reduction in supply to existing users compared to Scenario A. For more detail see companion technical report on river modelling (Gibbs et al., 2024). "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\6_River_model\SG_AWRAR\4_water_harvest\1_scripts\Catch_report_WH_plots.R" This example assumes a 20-day pump capacity – that is, pump capacities are set to enable the system and reach target volumes to be pumped in 20 days (not necessarily consecutively). This means an irrigator with a 4 GL ringtank has a pump capacity of 200 ML/day to fill their ringtank in 20 days. In this example there is no annual diversion commencement flow requirement based on the volume at the end of the system. At the smallest pump start threshold examined, 200 ML/day (nominally representative of a lower physical pumping limit), approximately 300 GL of water can be extracted in the Southern Gulf catchments in 75% of years; however, this relatively low pump start threshold results in some impacts on existing downstream licence holders along the Leichhardt River (grey contour lines on node 9139000 in Figure 5-27). This figure shows that as the total system and reach targets increase, the extraction reliability for that volume decreases. Similarly, as the pump start threshold increases, reducing the opportunities to extract water, the extraction reliability for the full system and reach targets decreases. At a pump start threshold of 600 ML/day, which does not affect existing users downstream, approximately 150 GL of water can be extracted in the Southern Gulf catchments in 75% of years. The data presented in Figure 5-28 are similar to that presented in Figure 5-27 except in Figure 5-28 an additional extraction condition is imposed: a combined total of 150 GL has to flow past the outlets of the Gregory–Nicholson and Leichhardt rivers each wet season before any water can be extracted. Figure 5-28 shows that increasing the EOS flow requirement reduces the extraction reliability for the system and reach targets for different pump start thresholds. Figure 5-29 models increasing the pumping capacity by modifying the conditions such that the target volume can be extracted in 10 days instead of 20 days, with the result being that reliability of supply increases for a given target and pump start threshold. This relationship is shown in more detail in Figure 5-30, this time with pump rate in days on the x-axis instead of pump start threshold, which has been fixed to 600 ML/day. With a pump start threshold of 600 ML/day and an annual EOS flow requirement of 150 GL, large pump capacities (i.e. to enable pumping in 20 days or less) are required to extract the system and reach target volumes in 75% of years or greater. Figure 5-31 and Figure 5-32 indicate, respectively, the post-extraction 50% and 80% annual flow exceedance combined across the outlets with water harvest extractions upstream as a proportion of change relative to Scenario A. The median annual flow is relatively insensitive to the ADCFR flow requirement (Figure 5-31). However, in drier years (represented by the 80% annual flow exceedance data in Figure 5-32) the ADCFR flow requirement has the effect of ‘protecting’ streamflow, with a higher relative proportion of the Scenario A EOS flow maintained. Figure 5-28 Annual reliability of diverting annual system and reach target volumes for varying pump start thresholds with annual diversion commencement flow requirement of 150 GL Assumes pumping capacity of 20 days (i.e. system and reach target volumes can be pumped in 20 days). Seven-digit numbers refer to model node locations. Black contour lines indicate conditions that meet a 75% annual reliability of supply, and grey contour lines on node 9139000 indicate proportion of years with a reduction in supply to existing users compared to Scenario A. For more detail see companion technical report on river modelling (Gibbs et al., 2024). "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\6_River_model\SG_AWRAR\4_water_harvest\1_scripts\Catch_report_WH_plots.R" Figure 5-29 Annual reliability of diverting annual system and reach target volumes for varying pump start thresholds assuming pumping capacity of 10 days Annual diversion commencement flow requirement of 150 GL before pumping can commence and pumping capacity such that system and reach target volumes can be pumped in 10 days. Seven-digit numbers refer to model node locations. Black contour lines indicate conditions that meet a 75% annual reliability of supply, and grey contour lines on node 9139000 indicate proportion of years with a reduction in supply to existing users compared to Scenario A. For more detail see companion technical report on river modelling (Gibbs et al., 2024). "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\6_River_model\SG_AWRAR\4_water_harvest\1_scripts\Catch_report_WH_plots.R" Figure 5-30 Annual reliability of diverting annual system and reach target volumes for varying pump rates assuming a pump start threshold of 600 ML/day Annual diversion commencement flow requirement of 150 GL before pumping can commence. Seven-digit numbers refer to model node locations. Black contour lines indicate conditions that meet a 75% annual reliability of supply, and no grey contour lines on node 9139000 indicate no impact on reliability of supply to existing users (e.g. Figure 5-29) compared to Scenario A with a pump start threshold of 600 ML/day. For more detail see companion technical report on river modelling (Gibbs et al., 2024). "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\6_River_model\SG_AWRAR\4_water_harvest\1_scripts\Catch_report_WH_plots.R" Figure 5-31 50% annual exceedance (median) streamflow relative to Scenario A in the Southern Gulf catchments for a pump start threshold of 600 ML/day and a pump capacity of 20 days Seven-digit numbers refer to model node locations. For more detail see companion technical report on river modelling (Gibbs et al., 2024). ADCFR = Annual diversion commencement flow requirement, the volume flowing past the end of the system before pumping can commence. "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\6_River_model\SG_AWRAR\4_water_harvest\1_scripts\Catch_report_WH_plots.R" Figure 5-32 80% annual exceedance streamflow relative to Scenario A in the Southern Gulf catchments for a pump start threshold of 600 ML/day and a pump capacity of 20 days Seven-digit numbers refer to model node locations. For more detail see companion technical report on river modelling (Gibbs et al., 2024). ADCFR = Annual diversion commencement flow requirement, the volume flowing past the end of the system before pumping can commence. "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\2_Hydrology\201_SG\6_River_model\SG_AWRAR\4_water_harvest\1_scripts\Catch_report_WH_plots.R" Evaporation and seepage losses Losses from a farm-scale dam occur through seepage and evaporation. A study of 138 farm dams ranging in capacity from 75 to 14,000 ML from southern NSW to central Queensland by the Cotton Catchment Communities CRC (2011) found mean seepage and evaporation rates of 2.3 and 4.2 mm/day, respectively. Of the 138 dams examined, 88% had seepage values of less than 4 mm/day and 64% had seepage values less than 2 mm/day. These results largely concur with those of the Irrigation Association of Australia (IAA, 2007), which states that reservoirs will have seepage losses equal to or less than 1 to 2 mm/day when constructed on suitable soils and greater than 5 mm/day if sited on less suitable (i.e. permeable) soils. When calculating evaporative losses from farm dams it is important to calculate net evaporation (i.e. evaporation minus rainfall) rather than just evaporation. Ringtanks with greater mean water depths lose a lower percentage of their total storage capacity to evaporation and seepage; however, they have a smaller ratio of storage capacity to excavation. In Table 5-8, effective volume refers to the actual volume of water that could be used for consumptive purposes after losses due to evaporation and seepage. For example, if water is stored in a ringtank with mean water depth of 3.5 m from April until January and the mean seepage loss is 2 mm/day, nearly three-quarters of the stored volume (i.e. 70%) would be lost to evaporation and seepage. The example provided in Table 5-8 is for a 4000 ML storage but the effective volume expressed as a percentage of the ringtank capacity is applicable to any storage (e.g. ringtanks or gully dams) of any capacity for mean water depths of 3.5, 6.0 and 8.5 m. Table 5-8 Effective volume after net evaporation and seepage for ringtanks of three mean water depths, under three seepage rates, near Century Mine in the Southern Gulf catchments Effective volume refers to the actual volume of water that could be used for consumptive purposes as a result of losses due to net evaporation and seepage, assuming the storage capacity is 4000 ML. For storages of 4000 ML capacity and mean water depths of 3.5, 6.0 and 8.5 m, reservoir surface areas are 110, 65 and 45 ha, respectively. Effective volumes are calculated based on the 20% exceedance net evaporation. For more detail see companion technical report on surface water storage (Yang et al., 2024). S:E ratio = storage capacity to excavation ratio. MEAN WATER DEPTH† (m) S:E RATIO SEEPAGE LOSS (mm/day) EFFECTIVE VOLUME (ML) EFFECTIVE VOLUME AS PERCENTAGE OF CAPACITY (%) EFFECTIVE VOLUME (ML) EFFECTIVE VOLUME AS PERCENTAGE OF CAPACITY (%) EFFECTIVE VOLUME (ML) EFFECTIVE VOLUME AS PERCENTAGE OF CAPACITY (%) 5 months (April to August) 7 months (April to October) 10 months (April to January) 3.5 14:1 1 2879 72 2327 58 1545 39 14:1 2 2711 68 2093 52 1210 30 14:1 5 2210 55 1390 35 203 5 6 7.5:1 1 3332 83 3004 75 2537 63 7.5:1 2 3233 81 2865 72 2338 58 7.5:1 5 2937 73 2450 61 1744 44 8.5 5:1 1 3535 88 3306 83 2981 75 5:1 2 3467 87 3210 80 2843 71 5:1 5 3262 82 2923 73 2432 61 †Mean water depth above ground surface. Strategies to minimise evaporation include liquid and solid barriers, but these are typically expensive per unit of inundated area (e.g. $12 to $40 per m2). In non-laboratory settings, liquid barriers such as oils are susceptible to being dispersed by wind and have not been shown to reduce evaporation from a water body (Barnes, 2008). Solid barriers can be effective in reducing evaporation but are expensive, at approximately two to four times the cost of constructing a ringtank. Evaporation losses from a ringtank can also be reduced slightly by subdividing the storage into multiple cells and extracting water from each cell in turn to minimise the total surface water area. However, constructing a ringtank with multiple cells requires more earthworks and incurs higher construction costs than outlined in this section. Capital, operation and maintenance costs of ringtanks Construction costs of a ringtank may vary considerably, depending on its size and the way the storage is built. For example, circular storages have a higher ratio of storage volume to excavation cost than rectangular or square storages. As discussed in the section on large farm-scale gully dams (Section 3.5.6), it is also considerably more expensive to double the height of an embankment wall than double its length due to the low angle of the walls of the embankment (often at a 3:1 ratio, horizontal to vertical). Table 5-9 provides a high-level breakdown of the capital and operation and maintenance (O&M) costs of a large farm-scale ringtank, including the cost of the water storage, pumping infrastructure, up to 100 m of pipes, and O&M costs of the scheme. In this example it is assumed that the ringtank is within 100 m of the river and pumping infrastructure. The cost of pumping infrastructure and conveying water from the river to the storage is particularly site specific. In flood-prone areas where flood waters move at moderate to high velocities, riprap (rocky material) protection may be required. The addition of riprap protection may increase the construction costs presented in Table 5-9 and Table 5-10 by 10% to 20% depending upon the volume of rock required and proximity to a quarry with suitable rock. For a more detailed breakdown of ringtank costs see the Northern Australia Water Resource Assessment technical report on large farm-scale dams (Benjamin, 2018), and for more information on pumping infrastructure see the companion technical report on pump stations for flood harvesting and irrigation downstream of storages (Devlin, 2024). 356 | Water resource assessment for the Southern Gulf catchments Table 5-9 Indicative costs for a 4000 ML ringtank Assumes a 4.25 m wall height, 0.75 m freeboard, 3:1 ratio on upstream slope, 3:1 ratio on downstream slope and crest width of 3.1 m, approximately 60% of material can be excavated from within storage, and costs of earthfill and compacted clay are $5.40/m3 and $7/m3, respectively. Earthworks costs include vegetation clearing, mobilisation/demobilisation of machinery and contractor accommodation. For a more detailed costing, see the Northern Australia Water Resource Assessment technical report on large farm-scale dams (Benjamin, 2018). Pump station costs were derived from the companion technical report on pump stations in northern Australia (Devlin, 2023). Costs indexed to 2023. Pump station operation and maintenance (O&M) costs are based on a diesel cost of $1.49/L. SITE DESCRIPTION/ CONFIGURATION EARTHWORKS COSTS ($) GOVERNMENT PERMITS AND FEES ($) INVESTIGATION AND DESIGN FEES ($) PUMP STATION COSTS ($) TOTAL CAPITAL COST ($) O&M COSTS OF RINGTANK ($/y) O&M COSTS OF PUMP STATION ($/y) TOTAL O&M COSTS ($/y) 4000 ML ringtank 2,000,000 43,000 92,000 380,000 2,515,000 21,000 92,000 113,000 The capital costs can be expressed over the service life of the infrastructure (assuming a 7% discount rate) and combined with O&M costs to give an equivalent annual cost for construction and operation. This enables infrastructure with differing capital and O&M costs and service life to be compared. The total equivalent annual costs for the construction and operation of a 1000 ML ringtank with 4.25 m high embankments and 55 ML/day pumping infrastructure are approximately $143,600 (Table 5-10). For a 4000 ML ringtank with 4.25 m high embankments and 160 ML/day pumping infrastructure, the total equivalent annual cost is approximately $301,550. For a 4000 ML ringtank with 6.75 m high embankments and 160 ML/day pumping infrastructure, the total equivalent annual cost is approximately $457,600. Table 5-10 Annualised cost for the construction and operation of three ringtank configurations Assumes freeboard of 0.75 m, pumping infrastructure can fill ringtank in 25 days and assumes a 7% discount rate. Costs based on those provided for 4000 ML provided in Northern Australia Water Resource Assessment technical report on large farm-scale dams (Benjamin, 2018). Costs indexed to 2021. Pump station operation and maintenance (O&M) costs assume cost of diesel of $1.49/L. CAPACITY AND EMBANKMENT HEIGHT ITEM CAPITAL COST ($) LIFE SPAN (y) ANNUALISED CAPITAL COST ($) ANNUAL O&M COST ($) 1000 ML and 4.25 m Ringtank 1,075,000 40 80,480 10,700 Pumping infrastructure† 245,000 15 26,900 4,500 Pumping cost (diesel) NA NA NA 21,000 4000 ML and 4.25 m Ringtank 2,000,000 40 150,000 17,250 Pumping infrastructure† 380,000 15 41,700 7,600 Pumping cost (diesel) NA NA NA 85,000 4000 ML and 6.75 m Ringtank 3,863,000 40 290,000 33,300 Pumping infrastructure† 380,000 15 41,700 7,600 Pumping cost (diesel) NA NA NA 85,000 NA = data not available. †Costs include rising main, large-diameter concrete or multiple strings of high-density polypipe, control valves and fittings, concrete thrust blocks and headwalls, dissipator, civil works and installation. Although ringtanks with a mean water depth of 3.5 m (embankment height of 4.25 m) lose a higher percentage of their capacity to evaporative and seepage losses than ringtanks of equivalent capacity with mean water depth of 6 m (embankment height of 6.75 m) (Table 5-8), their annualised unit costs are lower (Table 5-11) due to the considerably lower cost of constructing embankments with lower walls (Table 5-10). In Table 5-11 the levelised cost, or the equivalent annual cost per unit of water supplied from the ringtank, takes into consideration net evaporation and seepage from the storage, which increase with the length of time water is stored (i.e. crops with longer growing seasons will require water to be stored longer). In this table, the results are presented for the equivalent annual cost of water yield from a ringtank of different seepage rates and lengths of time for storing water. Table 5-11 Equivalent annual cost per megalitre for two different capacity ringtanks under three seepage rates based on a climate station near the Century Zinc Mine in the Southern Gulf catchments Assumes a 0.75 m freeboard, 3:1 ratio on upstream slope, 3:1 ratio on downstream slope. Crest widths are 3.1 and 3.6 m for embankments with heights of 4.25 and 6.75 m, respectively, and assumes earthfill and compacted clay costs of $5.40/m3 and $7/m3, respectively. Earthworks costs include vegetation clearing, mobilisation/demobilisation of machinery and contractor accommodation. 1000 ML ringtank reservoir has surface area of 27 ha and storage volume to excavation ratio of about 7:1. 4000 ML ringtank and 4.25 m embankment height reservoir has surface area of 110 ha and storage volume to excavation ratio of about 14:1. 4000 ML ringtank with 6.75 m embankment height reservoir has surface area of 64 ha and storage volume to excavation ratio of about 7.5:1. Annualised cost indexed to 2023 and assumes a 7% discount rate. For more information on this table please contact CSIRO on enquiries@csiro.au 5.4.5 Large farm-scale gully dams Large farm-scale gully dams are generally constructed of earth, or earth and rockfill embankments with compacted clay cores, and usually to a maximum height of about 20 m, though most gully dams are homogenous fill embankments and their height is typically 8 m or less. Large farm-scale gully dams typically have a maximum catchment area of about 30 km2 due to the challenges in passing peak floods from large catchments (large farm-scale gully dams are generally designed to pass an event with an annual exceedance probability of 1%), unless a site has an exceptionally good spillway option. Like ringtanks, large farm-scale gully dams are a compromise between best-practice engineering and affordability. Designers need to follow accepted engineering principles relating to important aspects of materials classification, compaction of the clay core and selection of an appropriate embankment cross-section. However, costs are often minimised where possible; for example, by employing earth bywashes and grass protection for erosion control rather than more expensive concrete spillways and rock protection as found on major dams. This can compromise the integrity of the structure during extreme events, and the longevity of the structure, as well as increase the ongoing maintenance costs, but can considerably reduce the upfront capital costs. In this section the following assessments are reported: • suitability of the land for large farm-scale gully dams • indicative capital and O&M costs of large farm-scale gully dams. Net evaporation and seepage losses also occur from large farm-scale gully dams. The analysis presented in Section 5.4.4 is also applicable to gully dams. Suitability of land for large farm-scale gully dams Figure 5-33 indicates those locations where it is more topographically and hydrologically favourable to construct large farm-scale gully dams in the Southern Gulf catchments and the likely density of options. This analysis takes considers those sites likely to have more favourable topography. It does not explicitly consider those sites that are underlain by soil suitable for the construction of the embankment and to minimise seepage from the reservoir base. This is shown in Figure 5-34. In reality, dams can be constructed on eroded or skeletal soils provided there is access to a clay borrow pit nearby for the cut-off trench and core zone. However, these sites are likely to be less economically viable. These figures indicate that those parts of the Southern Gulf catchments that are more topographically suitable as large-scale gully dam sites generally do not coincide with areas with soils that are moderately suitable for irrigated agriculture. Furthermore, in many areas topographically suitable for gully dams, dam walls would need to be constructed from rockfill and imported clay soils, increasing the cost of their construction. There are, however, some locations in the north-west of Doomadgee in the Settlement Creek catchment and Mornington Island where there is some topography suitable for gully dams, soil suitable for their construction and versatile agricultural land nearby. Gully dam suitability and agricultural versatility, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS560-S_Gullydam_Damsite_Land_Versatile_v4_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-33 Most economically suitable locations for large farm-scale gully dams in the Southern Gulf catchments Agricultural versatility data (see Section 4.3) indicate those parts of the catchment that are more or less versatile for irrigated agriculture. For the gully dam analysis, soil and subsurface data were only available to a depth of 1.5 m, hence this Assessment does not consider the suitability of subsurface material below this depth. Sites with catchment areas greater than 40 km2 or yield to excavation ratio less than 10 are not displayed. The results presented in this figure are modelled and consequently only indicative of the general locations where siting a gully dam may be most economically suitable. This analysis may be subject to errors in the underlying digital elevation model, such as effects due to the vegetation removal process. An important factor not considered in this analysis was the availability of a natural spillway. Gully dam suit and excavation yield ratio, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\4_Water_storage\3_S_Gulf\1_GIS\1_Map_docs\WS561-S_Gullydam_Suitability_Damsite_v3_CR.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 5-34 Suitability of soils for construction of gully dams in the Southern Gulf catchments Capital, operation and maintenance costs of large farm-scale gully dams The cost of a large farm-scale gully dam will vary depending upon a range of factors, including the suitability of the topography of the site, the size of the catchment area, quantity of runoff, proximity of site to good-quality clay, availability of durable rock in the upper bank for a spillway and the size of the embankment. The height of the embankment, in particular, has a strong influence on cost. An earth dam to a height of 8 m is about 3.3 times more expensive to construct than a 4 m high dam, and a dam to a height of 16 m will require 3.6 times more material than the 8 m high dam, but the cost may be more than five times greater, due to design and construction complexity. As an example of the variability in unit costs of gully dams, actual costs for four large farm-scale gully dams in northern Queensland are presented in Table 5-12. Table 5-12 Actual costs of four gully dams in northern Queensland Sourced from Northern Australia Water Resource Assessment technical report on farm-scale design and costs, (Benjamin, 2018). Costs indexed to 2023. Performance and cost of three hypothetical farm-scale gully dams in northern Australia A summary of the key parameters for three hypothetical 4 GL (i.e. 4000 ML) capacity farm-scale gully dam configurations is provided in Table 5-13 and a high-level breakdown of the major components of the capital costs for each of the three configurations is provided in Table 5-14. Detailed costs for the three hypothetical sites are provided in the Northern Australia Water Resource Assessment technical report on large farm-scale dams (Benjamin, 2018). Table 5-15 presents calculations of the effective volume for three configurations of 4 GL capacity gully dams (varying mean water depth/embankment height) for combinations of three seepage losses and water storage capacities over three time periods in the Southern Gulf catchments. Based on the information presented in Table 5-13, an equivalent annual unit cost including annual O&M cost for a 4 GL gully dam with a mean depth of about 6 m is about $220,000 (Table 5-16 and Table 5-17). Where the topography is suitable for large farm-scale gully dams and a natural spillway is present, large farm-scale gully dams are typically cheaper to construct than ringtanks. Table 5-13 Cost of three hypothetical large farm-scale gully dams of capacity 4 GL Costs include government permits and fees, investigation and design and fish passage. For a complete list of costs and assumptions see the Northern Australia Water Resource Assessment technical report on farm-scale dams (Benjamin, 2018). Costs indexed to 2023. O&M = operation and maintenance; S:E ratio = storage capacity to excavation ratio. Table 5-14 High-level breakdown of capital costs for three hypothetical large farm-scale gully dams of capacity 4 GL Earthworks include vegetation clearing, mobilisation/demobilisation of equipment and contractor accommodation. Investigation and design fees include design and investigation of fish passage device and failure impact assessment (i.e. investigation of possible existence of population at risk downstream of site). Costs indexed to 2023. Table 5-15 Effective volumes and cost per megalitre for a 4 GL storage with different mean depths and seepage loss rates based on a climate station at the Century Zinc Mine in the Southern Gulf catchments Time periods of 4, 6 and 9 months refer to length of time water is stored or required for irrigation. Table 5-16 Cost of construction and operation of three hypothetical 4 GL gully dams Assumes operation and maintenance (O&M) cost of 3% of capital cost and a 7% discount rate. Figures have been rounded. Table 5-17 Equivalent annualised cost and effective volume for three hypothetical 4 GL gully dams with various mean depths and seepage loss rates based on climate data at Victoria River Downs Station in the Victoria catchment Dam details are in Table 5-16. Annual cost assumes a 7% discount rate. Time periods of 4, 6 and 9 months refer to length of time water is stored or required for irrigation. 5.4.6 Natural water bodies Wetland systems and waterholes that persist throughout the dry season are natural water bodies characteristic of large parts of the northerly draining catchments of northern Australia. Many property homesteads in northern Australia use natural waterholes for stock and domestic purposes. However, the quantities of water required for stock and domestic supply are orders of magnitude less than those required for irrigated cropping, and it is partly for this reason that naturally occurring persistent water bodies in northern Australia are not used to source water for irrigation. For example, a moderately sized (5 ha) rectangular water body of mean depth 3.5 m may contain about 175 ML of water. Based on the data presented in Table 5-8 and assuming minimal leakage (i.e. 1 mm/day), approximately 72%, 58% and 39% of the volume would be available if a crop were to be irrigated until August, October and January, respectively. Assuming a crop or fodder with a 6-month growing season requires 5 ML/ha of water before losses, and assuming an overall efficiency of 80% (i.e. the waterhole is adjacent to land suitable for irrigation, 95% conveyance efficiency and 85% field application efficiency), a 175 ML waterhole could potentially be used to irrigate about 20 ha of land for half a year if all the water was able to be used for this purpose. A large natural water body of 20 ha and mean depth of 3.5 m could potentially be used to irrigate about 80 ha of land if all the water was able to be used for this purpose. Although the areas of land that could be watered using natural water bodies are likely to be small, the costs associated with storing water are minimal. Consequently, where these waterholes occur at sufficient size and adjacent to land suitable for irrigated agriculture, they can be a very cost- effective source of water. It would appear that where natural water bodies of sufficient size and suitable land for irrigation coincide, natural water bodies may be effective in staging a development (Section 6.3), where several hectares could potentially be developed, enabling lessons learned and mistakes made on a small-scale area before more significant capital investments are undertaken (noting that staging and learning are best to occur over multiple scales). In a few instances it may be possible to enhance the storage potential of natural features in the landscape such as horseshoe lagoons or cut-off meanders adjacent to a river. The main limitation to the use of wetlands and persistent waterholes for the consumptive use of water is that they have considerable ecological significance (e.g. Kingsford, 2000; Waltham et al., 2013), and in many cases there is a limited quantity of water contained within the water bodies. In particular, water bodies that persist throughout the dry season are considered key ecological refugia (Waltham et al., 2013). For a water body situated in a sandy river, a waterhole is likely to be connected to water within the bedsands of the river. Hence, during and following pumping water within the bedsands of a river, the bedsands may in part replenish the waterhole and vice versa. While water within the bedsands of the river may in part replenish a depleted waterhole, in these circumstances it also means that pumping from a waterhole will have a wider environmental impact than just on the waterhole from which water is being pumped. 5.5 Water distribution systems – conveyance of water from storage to crop 5.5.1 Introduction In all irrigation systems, water needs to be conveyed from the water source through artificial and/or natural water distribution systems before ultimately being used on-field for irrigation. This section discusses water losses during conveyance and application of water to a crop, and the associated costs. 5.5.2 Conveyance and application efficiencies Some water diverted for irrigation is lost during conveyance to the field before it can be used by a crop. These losses need to be taken into account when planning irrigation systems and developing likely irrigated areas. The amount of water lost during conveyance depends on the: •river conveyance efficiency, from the water storage to the re-regulating structure or point ofextraction •channel distribution efficiency, from the river offtake to the farm gate •on-farm distribution efficiency, in storing (using balancing storages) and conveying water fromthe farm gate to the field •field application efficiency, in delivering water from the edge of the field and applying it to thecrop. The overall or system efficiency is the product of these four components. Little research on irrigation systems has been undertaken in the Southern Gulf catchments. The time frame of the Assessment did not permit on-ground research into irrigation systems. Consequently, a brief discussion on the components listed above is provided based on relevant literature from elsewhere in Australia and overseas. Table 5-18 summarises the broad range of efficiencies associated with these components. The total conveyance and application efficiency of the delivery of water from the water storage to the crop (i.e. the overall or system efficiency) depends on the product of the four components in Table 5-18. For example, if an irrigation development has a river conveyance efficiency of 80%, a channel distribution efficiency of 90%, an on-farm distribution efficiency of 90% and a field application efficiency of 85%, the overall efficiency is 55% (80% × 90% × 90% × 85%). This means only 55% of all water released from the dam can be used by the crop. Table 5-18 Summary of conveyance and application efficiencies For more information on this table please contact CSIRO on enquiries@csiro.au †River conveyance efficiency varies with a range of factors (including distance) and may be lower than the range quoted here. Under such circumstances, it is unlikely that irrigation would proceed. It is also possible for efficiency to be 100% in gaining rivers. Achieving higher efficiencies requires a re-regulating structure (Section 5.4.3). River conveyance efficiency The conveyance efficiency of rivers is difficult to measure and even more difficult to predict. Although there are many methods for estimating groundwater discharge to surface water, there are few suitable methods for estimating the loss of surface water to groundwater. In the absence of existing studies for northern Australia, conveyance efficiencies as nominated in water resource plans and resource operation plans for four irrigation water supply schemes in Queensland were examined collectively. The results are summarised in Table 5-19. The conveyance efficiencies in Table 5-19 are from the water storage to the farm gate and are nominated efficiencies based on experience delivering water in these supply schemes. These data can be used to estimate conveyance efficiency of similar rivers elsewhere. Table 5-19 Water distribution and operational efficiency as nominated in water resource plans for four irrigation water supply schemes in Queensland For more information on this table please contact CSIRO on enquiries@csiro.au †Ignores differences in efficiency between high- and medium-priority users and variations across the scheme zone areas. ‡Channel conveyance efficiency only. Channel distribution efficiency Across Australia, the mean water conveyance efficiency from the river to the farm gate has been estimated to be 71% (Marsden Jacob Associates, 2003). For heavier-textured soils and well- designed irrigation distribution systems, conveyance efficiencies are likely to be higher. In the absence of larger scheme-scale irrigation systems in the Southern Gulf catchments, it is useful to look at the conveyance efficiency of existing irrigation developments to estimate the conveyance efficiency of irrigation developments in the study area. Australian conveyance efficiencies are generally higher than those found in similarly sized overseas irrigation schemes (Bos and Nugteren, 1990; Cotton Catchment Communities CRC, 2011). The most extensive review of conveyance efficiency in Australia was undertaken by the Australian National Commission on Irrigation and Drainage, which tabulated system efficiencies across irrigation developments in Australia (ANCID, 2001). Conveyance losses were reported as the difference between the volume of water supplied to irrigation customers and the water delivered to the irrigation system. For example, if 10,000 ML of water was diverted to an irrigation district and 8,000 ML was delivered to irrigators, then the conveyance efficiency was 80% and the conveyance losses were 20%. Figure 5-35 shows reported conveyance losses across irrigation areas of Australia between 1999 and 2000, along with the supply method used for conveying irrigation water and associated irrigation deliveries. There is a wide spread of conveyance losses both between years and across the various irrigation schemes. Factors identified by Marsden Jacob Associates (2003) that affect the variation include delivery infrastructure, soil types, distance that water is conveyed, type of agriculture, operating practices, infrastructure age, maintenance standards, operating systems, in- line storage, type of metering used, and third-party impacts such as recreational, amenity and environmental demands. Differences across irrigation seasons are due to variations in water availability, operational methods, climate and customer demands. For more information on this figure please contact CSIRO on enquiries@csiro.au 0% 15% 30% 45% 60% 0100,000200,000300,000400,000500,000600,000700,000800,000Losses 1999 to 2000 (percent) Irrigation deliveries 1999 to 2000 (ML) NSWQldSATasVicWA Figure 5-35 Reported conveyance losses from irrigation systems across Australia The shape of the marker indicates the supply method for the irrigation scheme: square (▪) indicates natural carrier, circle (•) indicates pipe and diamond (♦) indicates channel. The colour of the marker indicates the location of the irrigation system (by state), as shown in the legend. Source: ANCID (2001) Based on these industry data, Marsden Jacob Associates (2003) concluded that, on average, 29% of water diverted into irrigation schemes is lost in conveyance to the farm gate. However, some of this ‘perceived’ conveyance loss may be due to meter underestimation (about 5% of water delivered to provider (Marsden Jacob Associates, 2003)). Other losses were from leakage, seepage, evaporation, outfalls, unrecorded usage and system filling. On-farm distribution efficiency On-farm losses are losses that occur between the farm gate and delivery to the field. These losses usually take the form of evaporation and seepage from on-farm storages and delivery systems. Even in irrigation developments where water is delivered to the farm gate via a channel, many farms have small on-farm storages (i.e. less than 250 ML for a 500 ha farm). These on-farm storages enable the farmer to have a reliable supply of irrigation water with a higher flow rate, and also enable recycling of tailwater. Several studies have been undertaken in Australia for on- farm distribution losses. Meyer (2005) estimated an on-farm distribution efficiency of 78% in the Murray and Murrumbidgee regions, while Pratt Water (2004) estimated on-farm efficiencies to be 94% and 88% in the Coleambally Irrigation and Murrumbidgee Irrigation areas, respectively. For nine farms in these two irrigation areas, however, Akbar et al. (2000) measured channel seepage to be less than 5%. Field application efficiency After water is delivered to a field, it needs to be applied to the crop using an irrigation system. The application efficiency of irrigation systems typically varies between 60% and 90%, with more expensive systems usually resulting in higher efficiency. Three types of irrigation system can potentially be applied in the Southern Gulf catchments: surface irrigation, spray irrigation and micro irrigation (Figure 5-36). Irrigation systems applied in the Southern Gulf catchments need to be tailored to the soil, climate and crops that may be grown in the catchments and matched to the availability of water for irrigation. This is considered in the land suitability assessment figures presented in Section 4.2. System design will also need to consider investment risk in irrigation systems as well as likely returns, degree of automation, labour availability and O&M costs (e.g. the cost of energy). Irrigation systems have a trade-off between efficiency and cost. Table 5-20 summarises the different types of irrigation system, including their application efficiency, indicative cost and limitations. Across Australia the ratio of areas irrigated using surface, spray and micro irrigation is 83:10:7, respectively. Irrigation systems that allow water to be applied with greater control, such as micro irrigation, cost more (Table 5-20) and as a result are typically used for irrigating higher- value crops such as perennial horticulture and vegetables. For example, although only 7% of Australia’s irrigated area uses micro irrigation, it generates about 40% of the total value of produce grown using irrigation (Meyer, 2005). Further details on the three types of irrigation system follow Table 5-20. (a) (b) (c) Figure 5-36 Efficiency of different types of irrigation system (a) For bankless channel surface irrigation systems, application efficiencies range from 60% to 85%. (b) For spray irrigation systems, application efficiencies range from 75% to 90%. (c) For pressurised micro irrigation systems on polymer-covered beds, application efficiencies range from 80% to 90%. Photo: CSIRO Table 5-20 Application efficiencies for surface, spray and micro irrigation systems Application efficiency is the efficiency with which water can be delivered from the edge of the field to the crop. Costs indexed to 2023. For more information on this table please contact CSIRO on enquiries@csiro.au Adapted from Hoffman et al. (2007), Raine and Bakker (1996) and Wood et al. (2007). †Sources: DEEDI (2011a, 2011b, 2011c). Surface irrigation systems Surface irrigation encompasses basin, border strip and furrow irrigation, as well as variations such as bankless channel systems. In surface irrigation, water is applied directly to the soil surface, with check structures (banks or furrows) used to direct water across a field. Control of applied water is dictated by the soil properties, soil uniformity and the design characteristics of the surface system. Generally, fields are prepared by laser levelling to increase the uniformity of applied water and allow ease of management of water and adequate surface drainage from the field. The uniformity and efficiency of surface systems are highly dependent on the system design and soil properties, timing of the irrigation water and the skill of the individual irrigator in operating the system. Mismanagement can severely degrade system performance and lead to systems that operate at poor efficiencies. Surface irrigation has the benefit that it can generally be adapted to almost any crop and usually has a lower capital cost compared with alternative systems. Surface irrigation systems perform better when soils are of uniform texture as infiltration characteristics of the soil play an important part in the efficiency of these systems. Therefore, surface irrigation systems should be designed into homogenous soil management units and layouts (run lengths, basin sizes) tailored to match soil characteristics and water supply volumes. High application efficiencies are possible with surface irrigation systems, provided soil characteristic limitations, system layout, water flow volumes and high levels of management are applied. On ideal soil types and with systems capable of high flow rates, efficiencies can be greater than 85%. On poorly designed and managed systems on soil types with high variability, efficiencies can be less than 60%. Generally, the major cost in setting up a surface irrigation system is land grading and levelling, with costs directly associated with the volume of soil that must be moved. Typical earth-moving volumes are in the order of 800 m3/ha but can exceed 2500 m3/ha. Volumes greater than 1500 m3/ha are generally considered excessive due to costs (Hoffman et al., 2007). Surface irrigation systems are the dominant form of irrigation type used throughout the world. Their potential suitability in the Southern Gulf catchments would be due to their generally lower set-up costs and adaptability to a wide range of irrigated cropping activities. They are particularly suited to the heavier-textured soils on the alluvial plains adjacent to the Gregory and Leichhardt rivers and major tributaries, which reduce set-up or establishment costs of these systems. With surface irrigation, little or no energy is required to distribute water throughout the field, and this ‘gravity-fed’ approach reduces energy requirements of these systems. Surface irrigation systems generally have lower applied irrigation water efficiency than spray or micro systems when compared across an industry and offer less control of applied water; however, well-designed and well-managed systems can approach efficiencies of alternative irrigation systems in ideal conditions. Spray irrigation systems In the context of the Southern Gulf catchments, spray irrigation refers specifically to lateral move and centre pivot irrigation systems. Centre pivot systems consist of a single sprinkler, laterally supported by a series of towers. The towers are self-propelled and rotate around a central pivot point, forming an irrigation circle. Time taken for the pivot to complete a full circle can range from as little as half a day to several days depending on crop water demands and application rate of the system. Lateral or linear move systems are similar to centre pivot systems in construction, but rather than move around a pivot point the entire line moves down a field perpendicular to the lateral direction. Water is supplied by a lateral channel running the length of the field. Lateral lengths are generally in the range of 800 to 1000 m. Their advantage over surface irrigation systems is they can be utilised on rolling topography and generally require less land forming. Both centre pivot and lateral move irrigation systems have been extensively used for irrigating a range of annual broadacre crops and are capable of irrigating most field crops. They are generally not suitable for tree crops or vine crops, or for saline irrigation water applications in arid environments, which can cause foliage damage. Centre pivot and lateral move systems usually have higher capital costs but are capable of very high efficiencies of water application. Generally, application efficiencies for these systems range from 75% to 90% (Table 5-20). They are used extensively for broadacre irrigated cropping situations in high evaporative environments in northern NSW and South West Queensland. These irrigation developments have high irrigation crop water demand requirements, which are similar to those found in the Southern Gulf catchments. A key factor in the suitable use of spray systems is sourcing the energy needed to operate these systems, which are usually powered by electricity or diesel depending on available costs and infrastructure. Where available, electricity is considerably cheaper than diesel for powering spray systems. For pressurised systems such as spray or micro irrigation systems, water can be more easily controlled, and potential benefits of the system through fertigation (application of crop nutrients through the irrigation system (i.e. liquid fertiliser)) are also available to the irrigator. Micro irrigation systems For high-value crops, such as horticultural crops where yield and quality parameters dictate profitability, micro irrigation systems should be considered suitable across the range of soil types and climate conditions in the Southern Gulf catchments. Micro irrigation systems use thin-walled polyethylene pipe to apply water to the root zone of plants through small emitters spaced along the drip tube. These systems are capable of precisely applying water to the plant root zone, thereby maintaining a high level of irrigation control and applied irrigation water efficiency. Historically, micro irrigation systems have been extensively used in tree, vine and row crops, with limited applications in complete-cover crops such as grains and pastures due to the expense of these systems. Micro irrigation is suitable for most soil types and can be practised on steep slopes. Micro irrigation systems are generally of two varieties: above-ground and below-ground (where the drip tape is buried beneath the soil surface). Below- ground micro irrigation systems offer advantages in reducing evaporative losses and improving trafficability. However, below-ground systems are more expensive and require higher levels of expertise to manage. Properly designed and operated micro irrigation systems are capable of very high application efficiencies, with field efficiencies of 80% to 90% (Table 5-20). In some situations, micro irrigation systems offer water and labour savings and improved crop quality (i.e. more marketable fruit through better water control). Management of micro irrigation systems, however, is critical. To achieve these benefits requires a much greater level of expertise than for other traditional systems such as surface irrigation systems, which generally have higher margins of error associated with irrigation decisions. Micro irrigation systems also have high energy requirements, with most systems operating at pressure ranges from 135 to 400 kPa, with diesel or electric pumps most often used. 5.6 References Akbar S, Beecher HG, Cullis B and Dunn B (2000) Using of EM surveys to identify seepage sites in on-farm channel and drains. Proceedings of the Irrigation Australia 2000 Conference, Australia. ANCID (2001) Australian irrigation water provider benchmarking report for 1999/2000. An ANCID initiative funded by Land and Water Australia and Department of Agriculture, Fisheries and Forestry−Australia, Australia. AWA (2018) Desalinisation factsheet. Australia Water Association. Viewed 10 April 2018, http://www.awa.asn.au/AWA_MBRR/Publications/Fact_Sheets/Desalination_Fact_Sheet.aspx. Barnes GT (2008) The potential for monolayers to reduce the evaporation of water from large water storages. Agricultural Water Management 95, 339–353. Benjamin J (2018) Farm-scale dam design and costs. 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. Bos MG and Nugteren J (1990) On irrigation efficiencies, ILRI publication 19. International Institute for Land Reclamation and Improvement, Netherlands. Chanson H (1998) Extreme reservoir sedimentation in Australia: a review. International Journal of Sediment Research 13(3), 55–63. Cotton Catchment Communities CRC (2011) Storage seepage and evaporation. Final summary of results. Cotton storages project 2011. Cotton Catchment Communities CRC. DEEDI (2011a) More profit per drop – irrigated farming systems. Bankless channels – Bullamon Plains. Queensland Department of Employment, Economic Development and Innovation. Viewed 1 September 2013, Hyperlink to: More profit per drop – irrigated farming systems . DEEDI (2011b) More profit per drop – irrigated farming systems. Centre pivot/lateral move irrigation. Queensland Department of Employment, Economic Development and Innovation. Viewed 1 September 2013, More Profit Per Drop website . DEEDI (2011c) More profit per drop – irrigated farming systems. Drip irrigation. Queensland Department of Employment, Economic Development and Innovation. Viewed 12 November 2013, More Profit Per Drop website . Dillon P, Pavelic P, Page D, Beringen H and Ward J (2009) Managed aquifer recharge: an introduction. Waterlines Report Series no. 13. National Water Commission, Canberra. Doherty J (1999) Gravity dams. In: Cole B (ed) A history of dam technology in Australia 1850 – 1999. Institution of Engineers, Australia and Australian National Committee on Large Dams. Drinkwater KF and Frank KT (1994) Effects of river regulation and diversion on marine fish and invertebrates. Aquatic Conservation Freshwater and Marine Ecosystems 4, 135–151. Evans WR, Evans RS and Holland GF (2013) Groundwater governance: a global framework for country action. Thematic paper 2: conjunctive use and management of groundwater and surface water within existing irrigation commands: the need for a new focus on an old paradigm. GEF ID 3726. Fell R, MacGregor P, Stapledon D and Bell G (2005) Geotechnical engineering of dams. Balkema Publishers. Gallant JC, Dowling TI, Read AM, Wilson N, Tickle P and Inskeep C (2011) 1 second SRTM derived digital elevation models user guide. Geoscience Australia. Viewed 19 March 2018, Geoscience Australia website . Gibbs M, Hughes J, Yang A, Wang B, Marvanek S and Petheram C (2024) River model scenario analysis for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Gillanders BM and Kingsford MJ (2002) Impact of changes in flow of freshwater on estuarine and open coastal habitats and the associated organisms. Oceanography and Marine Biology: an annual review 40, 233–309. Hoffman GJ, Evans RG, Jensen ME, Martin DL and Elliott RL (2007) Design and operation of farm irrigation systems. 2nd edition. American Society of Agricultural and Biological Engineers, St. Joseph, Michigan. IAA (2007) Guidelines for ring tank storages. In: Barrett H (ed) Irrigation Association of Australia Ltd. Jackson S, Stoeckl N and Straton A (2008) The changing value of Australian tropical rivers. Geographical Research 46, 275-290 Kingsford RT (2000) Ecological impacts of dams, water diversions and river management on floodplain wetlands in Australia. Austral Ecology 25, 109–127. Knapton A, Taylor AR and Crosbie RS (2024) Groundwater flow modelling of the Cambrian Limestone Aquifer in the Undilla Sub-basin, Georgina Basin using a finite element groundwater flow model. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Knapton A, Taylor AR, Petheram C and Crosbie RS (2023) An investigation into the effects of climate change and groundwater development scenarios on the water resources of the Roper catchment using two finite element groundwater flow models. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia., https://publications.csiro.au/publications/publication/PIcsiro:EP2023-5642. Lennon L, Evans R, George R, Dean F and Parsons S (2014) The role of managed aquifer recharge in developing northern Australia. Ozwater’14, Proceedings of Australia’s International Water Conference and Exhibition, Australian Water Association, Australia. Lewis B (2002) Farm dams: planning, construction and maintenance. Landlinks Press, Collingwood, Victoria. Lyons P, Barber M, Braedon P and Pert P (2024) Indigenous water values, rights, interests and development goals in the Southern Gulf Catchments: A literature review and implications for future research. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia Marsden Jacob Associates (2003) Improving water-use efficiency in irrigation conveyance systems: a study of investment strategies. Marsden Jacob Associates, Land and Water Australia, Australia. Meyer WS (2005) The irrigation industry in the Murray and Murrumbidgee basins. Technical report no. 03/05. CSIRO Cooperative Research Centre for Irrigation Futures, Australia. Nace RL (1972) Water problems and developments of the past. Journal of the American Water Resources Association 8(1), 101–109. NRMMC-EPHC-NHMRC (2009) Australian guidelines for water recycling: managed aquifer recharge. National water quality management strategy document No. 24. Natural Resource Management Ministerial Council, Environmental Protection and Heritage Council, National Health and Medical Research Council. Petheram C, Rogers L, Eades G, Marvanek S, Gallant J, Read A, Sherman B, Yang A, Waltham N, McIntyre-Tamwoy S, Burrows D, Kim S, Podger S, Tomkins K, Poulton P, Holz L, Bird M, Atkinson F, Gallant S and Kehoe M (2013) Assessment of surface water storage options 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. Petheram C, Rogers L, Read A, Gallant J, Moon A, Yang A, Gonzalez D, Seo L, Marvanek S, Hughes J, Ponce Reyes R, Wilson P, Wang B, Ticehurst C and Barber M (2017) Assessment of surface water storage options in the Fitzroy, Darwin and Mitchell catchments. 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. Pratt Water (2004) The business of saving water: report of the Murrumbidgee Valley Water Efficiency Feasibility Project. Pratt Water, Campbellfield, Victoria. QWIS (1969) Progress report on preliminary investigations of dam sites. Gunpowder Creek and Gregory River. Queensland Irrigation and Water Supply Commission. November 1969. QWIS (1974) Gunpowder Creek 108.7 km Dam Project. Report on site access. Queensland Irrigation and Water Supply Commission. Job No. R1778. November, 1974. QWRC (1984) Farm water supplies design manual. Volume I farm storages. 2nd edition. In: Horton AJ and Jobling GA (eds) Farm Water Supplies Section, Irrigation Branch, Queensland Water Resources Commission. Raiber M, Taylor AR, Dupuy M, Priestley S, Crosbie RS, Knapton A, Hodgson G and Barry K (2024) Characterising groundwater resources of the Gilbert River Formation, Camooweal Dolostone and Thorntonia Limestone in the Southern Gulf catchments, Queensland and Northern Territory. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Raine SR and Bakker DM (1996) Increased furrow irrigation efficiency through better design and management of cane fields. Proceedings Australian Society of Sugar Cane Technologists 1996, 119–124. Raymond OL (2018) Australian Geological Provinces 2018.01 edition. Geoscience Australia, Canberra. Australia G, (https://ecat.ga.gov.au/geonetwork/srv/eng/catalog.search#/metadata/116823). Robins JB, Halliday IA, Staunton-Smith J, Mayer DG and Sellin MJ (2005) Freshwater-flow requirements of estuarine fisheries in tropical Australia: a review of the state of knowledge and application of a suggested approach. Marine and Freshwater Research 56, 343–360. Ross A and Hasnain S (2018) Factors affecting the cost of managed aquifer recharge (MAR) schemes. Sustainable Water Resources Management. DOI: 10.1007/s40899-017-0210-8. Scarborough VL and Gallopin G (1991) A water storage adaptation in the Maya Lowlands. Science 251(4), 658–662. DOI:10.1126/science.251.4994.658. Schnitter NJ (1994) A history of dams: the useful pyramids. AA Balkema, Rotterdam, Brookfield. Merrin L, Stratford D, Kenyon R, Pritchard J, Linke S, Ponce Reyes R, Buckworth R, Castellazzi P, Costin B, Deng R, Gannon R, Gao S, Gilbey S, Lachish S, McGinness H and Waltham N (2024) Ecological assets of the Southern Gulf catchments to inform water resource assessments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Thomas M, Philip S, Zund P, Stockmann U, Hill J, Gregory L, Watson I and Thomas E (2024) Soils and land suitability for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Tomkins K (2013) Estimated sediment infilling rates for dams in northern Australia based on a review of previous literature. 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. Vanderzalm JL, Page DW, Gonzalez D, Barry KE, Dillon PJ, Taylor AR, Dawes WR, Cui T and Knapton A (2018) Assessment of managed aquifer recharge (MAR) opportunities in the Fitzroy, Darwin and Mitchell catchments. 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. 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, Canberra, Australia. Viewed 7 February 2017, Hyperlink to: Waterhole ecology in the Flinders and Gilbert catchments . Wasson RJ (1994) Annual and decadal variation of sediment yield in Australia, and some global comparisons. Variability in stream erosion and sediment transport, Proceedings of the Canberra Symposium. IAHS Publ. no. 224, 269–279. Wood M, Wang Q and Bethune M (2007) An economic analysis of conversion from border-check to centre-pivot irrigation on dairy farms in the Murray dairy region. Irrigation Science 26(1), 9–20. Yang A, Petheram C, Marvanek S, Baynes F, Rogers L, Ponce Reyes R, Zund P, Seo L, Hughes J, Gibbs M, Wilson PR, Philip S and Barber M (2024) Assessment of surface water storage options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO Australia. Part IV Economics of development and accompanying risks Chapters 6 and 7 describe economic opportunities for water development in the Southern Gulf catchments, and the accompanying constraints and risks: • economic opportunities and constraints (Chapter 6) • a range of risks to development (Chapter 7). Agricultural production in the Southern Gulf catchments is dominated by extensive grazing of beef cattle, valued at $240 million in 2020–21 and covers about 77% of the study area. Photo: CSIRO – Nathan Dyer 6 Overview of economic opportunities and constraints in the Southern Gulf catchments Authors: Chris Stokes, Shokhrukh Jalilov, Diane Jarvis Chapter 6 examines the types of opportunities for the development of irrigated agriculture in the catchment of the Southern Gulf riversFigure 6-1). 1, that is Settlement Creek, Gregory—Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groups, that are most likely to be financially viable. The chapter considers the costs of building the required infrastructure (both within the scheme and beyond), the financial viability of various types of schemes (including lessons learned from past large dam developments in Australia) and the regional economic impacts (the direct and flow-on effects for businesses across the catchment) ( 1 Only those islands greater than 1000 ha are mapped. Figure 6-1 Schematic diagram of key components affecting the commercial viability of a potential greenfield irrigation development "\\fs1-cbr.nexus.csiro.au\{lw-rowra}\work\10_Reporting\1_All\9_Graphics_artist\3_Vic and SoG\C Bruce Vic CR Chp6_8_2024.jpg" The chapter focuses on the costs and benefits that are the subject of normal market transactions, but it does not provide a full economic analysis. Commercial factors are likely to be among the most important criteria in deciding between potential development opportunities. Options clearly identifiable at the pre-feasibility stage as not being commercially viable could be deprioritised. More-detailed and project-specific agronomic, ecological, social, cultural and regulatory assessments could then focus on those opportunities identified as showing the most commercial promise. The non-market impacts and risks associated with any financially viable development opportunities, discussed in Chapter 7, must also be considered. 6.1 Summary 6.1.1 Key findings Scheme-scale financial viability New investment in irrigation development in the Southern Gulf catchments would depend on finding viable combinations of low-cost water sources, low-cost farming development opportunities, and high-productivity farms; finding opportunities for reducing cropping costs and attracting price premiums for produce; and managing a wide range of risks. Financial analyses have indicated that large dams in the Southern Gulf catchments are unlikely to be viable if public investors target full cost recovery at a 7% internal rate of return (IRR) and do not provide assistance, which would make water from the most cost-effective dam sites too expensive for irrigators. However, large dams could be marginally viable if public investors accepted a 3% IRR. On-farm water sources provide better prospects than large dams: where sufficiently cheap water development opportunities can be found, they could support viable broadacre farms and horticulture with low development costs. Horticulture with high development costs (e.g. fruit orchards) in the Southern Gulf catchments would be more challenging unless farm financial performance could be boosted by (i) finding niche opportunities for premium produce prices, (ii) making savings in production and marketing costs, and/or (iii) obtaining high yields. Farm performance can be affected by a number of risks, including water reliability, climate variability, price fluctuations, and the need to adapt farming practices to new locations. Setbacks that occur soon after an irrigation scheme has been established have the largest effect on scheme viability. There is a strong incentive for using well-proven crops and technologies when starting any new irrigation development, and for being thoroughly prepared for those agronomic risks of establishing new farmland that can be anticipated. Risks that cannot be avoided must be managed, mitigated where possible, and accounted for when determining the realistic returns that may be expected from a scheme and the capital buffers that would be required. Cost–benefit analysis of large public dams A review of recent large public dams built in Australia has highlighted some areas where cost– benefit analyses (CBAs) for water infrastructure projects that could be improved upon, particularly the need for more-realistic forecasting of the demand for water. This chapter provides information for benchmarking a number of the processes commonly used in such CBAs, including demand forecasting, which can be used to check when proposals for new dams are being unrealistically optimistic (or pessimistic). Regional economic impacts Any new irrigated agriculture development and its supporting infrastructure will have knock-on benefits to the regional economy beyond direct economic growth from the new farms and construction. The initial construction phase of a new irrigation development in the Southern Gulf catchments could provide an additional approximately $1.1 million of indirect regional benefits, over and above the direct benefits, for each million dollars spent on construction within the local region. The ongoing production phase of a new irrigation development could provide an additional (approximately) $0.46 to $1.82 million of indirect regional benefits for each million dollars of direct benefits from the increased agricultural activity (gross revenue), depending on the type of agricultural industry. The indirect regional benefits would be reduced if some of the extra expenditure generated by a new development was leaked to outside the catchment. Each $100 million increase in agricultural activity could create approximately 100 to 852 jobs. 6.2 Introduction Large new infrastructure projects in Australia are expected to be increasingly accountable and transparent. This trend extends to the planning and building of new water infrastructure, and the way water resources are managed and priced (e.g. Infrastructure Australia, 2021a, 2021b; NWGA, 2022, 2023), and includes greater scrutiny of the costs and benefits of potential large new public dams. The difficulty in accurately estimating costs and the chance of incurring unanticipated expenses during construction, and of not meeting the projected water demands and achieving revenue trajectories when completed, put the viability of developments at risk if they are not thoroughly planned and assessed. For example, in a global review of dam-based megaprojects, Ansar et al. (2014) found that the forecast costs were systematically biased downwards, with three-quarters of projects running over budget and the mean of the actual costs being almost double the initial estimates. This is typical for most types of large infrastructure projects, not just dams (see review in Section 6.4.1). Ultimately, economic factors are likely to be among the most important criteria in deciding the scale and types of potential development opportunities in the Southern Gulf catchments. An assessment of 13 agricultural developments in northern Australia found that, while the natural environments were challenging for agriculture, the most important factors determining the viability of developments were management, planning and finances (Ash et al., 2014). At the pre- feasibility stage, options that can be clearly identified as not being financially viable could be deprioritised. The expensive, more-detailed and project-specific agronomic, ecological, social, cultural and regulatory assessments could then focus on the more promising opportunities. This chapter aims to assist future planning and evaluation of investments in new irrigated agriculture developments by highlighting the types of projects that are more likely to be viable, and quantifying the costs, benefits and risks involved. It provides a generic information resource that is broadly applicable to a variety of irrigated agriculture development opportunities but does not examine any specific options in detail. The results are presented in a way that allows readers to identify the costs, risks, and farm productivity values specific to the project opportunities in which they are interested, to evaluate their likely financial viability. The information also provides a set of benchmarks for establishing realistic assumptions and the thresholds of financial performance required for water and farm developments, individually and in combination, to be financially viable. This chapter builds on earlier material in Chapter 4 (assessing the viability of new irrigated agriculture opportunities in the Southern Gulf catchments at the enterprise level) and in Chapter 5 (assessing the opportunities for developing water sources to support those farms). Section Rather than analysing the cost–benefit of specific irrigation scheme proposals, this chapter presents generic tables for evaluating multiple alternative development configurations, providing the threshold farm gross margins and water costs and pricing that would be required in order to cover infrastructure costs. These tables serve as tools that allow users to answer their own questions about agricultural land and water development. Examples of the questions that can be asked, and which tables provide the answers, are given in Table 6-1. Table 6-1 Types of questions that users can answer using the tools in this chapter For each question, the relevant table number is given, together with an example answer for a specific development scenario. More questions can be answered with each tool by swapping around the factors that are known and the factor being estimated. (All initial estimates assume farm performance is 100% in all years, i.e. before accounting for risks. See Table 6-3 for the supporting generalised assumptions.) QUESTION (WITH EXAMPLE ANSWER) RELEVANT TABLE 1) How much can various types of farms afford to pay per ML of water they use? Table 6-4 A broadacre farm with a gross margin (GM) of $4000/ha and water consumption of 8 ML/ha could afford to pay $135/ML while achieving a 10% internal rate of return (IRR). 2) How much would the operator of a large off-farm dam have to charge for water? Table 6-6 If off-farm water infrastructure had a capital cost of $5000 for each ML/y supply capacity (yield) at the dam wall, the (public) water supplier would have to charge $537 for each ML to cover its costs (at a 7% target IRR). 3) For an on-farm dam with a known development cost, what is the equivalent $/ML price of water? Table 6-8 If a farm dam had a capital development cost of $1500 for each ML/y supply capacity (yield), water could be purchasable at a cost of $190 for each ML (given a 10% target IRR). 4) (a) What farm GM would be required in order to fully cover the costs of an off-farm dam? (b) What proportion of the costs of off-farm water infrastructure could farms cover? Table 6-5 If off-farm infrastructure had a capital cost of $50,000/ha to build, broadacre farms would need to generate a GM of $5701/ha in order to fully cover the water supply costs (while meeting a target 7% IRR for the water supplier (public investor) and a 10% IRR for the irrigator (private investor)). With the same target IRRs, a broadacre farm with a GM of $4000/ha could contribute the equivalent of $20,000 to $30,000 per ha towards the capital costs of building the same $50,000/ha dam (~50% of the full costs of building and operating that infrastructure). 5) What GM would be required in order to cover the costs of developing a new farm, including a dam or bores? Table 6-7 A horticultural farm with low overheads ($1500/ha) that cost $40,000/ha to develop (e.g. $30,000/ha to establish the farm and $10,000/ha to build the on-farm water supply for irrigating it) would require a GM of $6702/ha to attain a 10% IRR. QUESTION (WITH EXAMPLE ANSWER) RELEVANT TABLE 6) How would risks associated with water reliability affect the farm GMs above? Table 6-9 If an on-farm dam could fully irrigate the farm in 70% of years and could irrigate 50% of the farm in the remaining years, all farm GMs in the answers above would need to be multiplied by 1.18 (i.e. would be 18% higher), and the price irrigators could afford to pay for water would need to be divided by 1.18. For example, in Q4, the GM required in order to cover the costs of the farm development would increase from $5701/ha to $6727/ha after accounting for the risks of water reliability. 7) How would the risks associated with ‘learning’ (initial farm underperformance) affect estimates? Table 6-11 If a farm with a 10% target IRR achieved a GM that was 50% of its full potential in the first year, and gradually improved to achieve its full potential over 10 years, then the GMs above would need to be multiplied by a factor of 1.26 (i.e. would be 26% higher). For example, in Q6, the required farm GM would increase to $8476/ha after accounting for the risks of both water reliability and learning (a combined 49% higher than the value before accounting for risks). 6.3 Balancing scheme-scale costs and benefits Designing a new irrigation development in the Southern Gulf catchments would require balancing three key determinants of irrigation scheme financial performance to find combinations that might collectively constitute a viable investment. The determinants are: 1.farm financial performance (relative to development costs and water use) (Chapter 4) 2.capital cost of development, for both water resources and farms (Chapter 5 and Section 6.3.1) 3.risks (and the associated required level of investment return) (Section 6.3.5). The determinants considered have been limited to those with greater certainty and/or lower sensitivity, so that the results can be applied to a wide range of potential developments. A key finding of the irrigation scheme financial analyses is that no single factor within the above list is likely to be able to provide a silver bullet for meeting the substantial challenge of designing a commercially viable new irrigation scheme. Balancing the benefits to meet costs in order to identify viable investments would likely require contributions from each of the above factors and careful selection to piece together a workable combination. This section provides background information on the analysis approach used, to help readers understand how these factors influence irrigation scheme financial performance. 6.3.1 Approach and terminology Scheme financial evaluations use a discounted cashflow framework to evaluate the commercial viability of irrigation developments. The framework, detailed in the companion technical report on agricultural viability and socio-economics (Webster et al., 2024), is intended to provide a purely financial evaluation of the conditions required to produce an acceptable return from an investor’s perspective. It is not a full economic evaluation of the costs and benefits to other industries, nor does it consider ‘unpriced’ impacts that are not the subject of normal market transactions, or the equity of how costs and benefits are distributed. For the discussion that follows, the costs and benefits of an irrigation scheme were taken to include all those from the development of the land and water resources to the point of sale for farm produce. This section explains the terminology and standard assumptions used. A ‘discounted cashflow analysis’ considers the lifetime of costs and benefits following capital investment in a new project. Costs and benefits that occur at various times are expressed in constant real dollars (December 2023 Australian dollars), with a discount rate being applied to streams of costs and benefits. The ‘discount rate’ is the percentage by which future costs and benefits are discounted each year (compounded) to convert them to their equivalent present value. For an entire project, the ‘net present value’ (NPV) can be calculated by subtracting the present value of the stream of all costs from the present value of the stream of all benefits. The ‘benefit- to-cost ratio’ (BCR) of a project is the present value of all the benefits of a project divided by the present value of all the costs involved in achieving those benefits. To be commercially viable (at the nominated discount rate), a project would require an NPV that is greater than zero (in which case the BCR would be greater than one). The IRR is the discount rate at which the NPV is zero (and the BCR is 1). For a project to be considered commercially viable, it needs to meet its target IRR, and the NPV has to be greater than zero at a discount rate appropriate to the risk profile of the development and alternative investment opportunities available to investors. A target IRR of 7% is typically used when evaluating large public investments (with the sensitivity analysis set at 3% and 10%) (Infrastructure Australia, 2021b). Private agricultural developers usually target an IRR of 10% or more (to compensate for the investment risks involved). A back-calculation approach is used in the tables below to present the threshold GMs and water prices that would be required in order for investors to achieve specified target IRRs (the NPV would be equivalent to zero at these discount rates). The ‘project evaluation periods’ used in this chapter matched the ‘life spans’ of the main infrastructure assets: 100 years for large off-farm dams and 40 years for on-farm developments. To simplify the tracking of asset replacements, four categories of life spans were used: 15 and 40 years for farms and 25 and 100 years for off-farm infrastructure. It was assumed that the shorter-life-span assets would be replaced at the end of their life, and that their costs would have been accounted for in full by the actual year of their replacement. At the end of the evaluation period, a ‘residual value’ was calculated to account for any shorter-life-span assets that have not reached the end of their working life. Residual values were calculated as the proportional asset life remaining multiplied by the original asset price. The ‘capital costs’ of infrastructure were assumed to be the costs at completion (accounted for in full in the year of delivery), such that the assets commenced operations the following year. In some cases, the costs of developing the farmland and setting up the buildings and equipment were considered separately from the costs of the water source, so that various water source options could be compared on a like-for-like basis. Where an off-farm water source was used, the separate investor in that water source would receive payments for water at a price that the irrigator could afford to pay. The main ‘costs for operating’ a large dam and the associated water distribution infrastructure are (i)fixed costs for administering and maintaining the infrastructure, expressed here as percentageof the original capital cost, and (ii) variable costs associated with pumping water into distributionchannels. At the farm scale, fixed overhead costs are incurred each year, whether or not a crop is planted in a particular field that year. ‘Fixed costs’ are dominated by the fixed component of labour costs, but also include maintenance, insurance, professional services, and registrations. An additional allowance is made for annual operation and maintenance (O&M), budgeted at 1% of the original capital value of all assets (with an additional variable component in maintenance costs when machinery is used for cropping operations). A ‘farm annual gross margin’ (GM) is the difference between the gross income from crop sales and the variable costs of growing a crop each year. ‘Net farm revenue’ is calculated by subtracting the fixed overhead costs from the GM. ‘Variable costs’ vary in proportion to the area of land planted, the amount of crop harvested and/or the amount of water and other inputs applied. Farm GMs can vary substantially within and between locations, as described in Chapter 4. The GMs presented here are the values obtained before subtracting the variable costs of supplying water to farms; these water supply costs are, instead, accounted for in the capital costs of developing water resources. (The equivalent unit costs of supplying each ML of water are presented separately below.) The CBA analyses first considered the case of irrigation schemes built around public investment in a large off-farm dam in the Southern Gulf catchments and then considered the case developments using ‘on-farm’ dams and bores. Cost and benefit streams, totalled across the scheme, were tracked as separate components, allowing for both on-farm and off-farm sources of new water development. For farms, these streams were: (i) the capital costs of land development, farm buildings and equipment (including replacement costs and residual values), (ii) the fixed overhead costs, applied to the full area of developed farmland, and (iii) the total farm GM (across all farms in the scheme), applied to the mean proportion of land in production each year. If an ‘on-farm water source’ was being considered, then those costs were added to the farm costs. Farm developers were treated as private investors who would seek a commercial return. When an ‘off-farm water source’ (large dam >25 GL/year) was evaluated, its investor was treated as a separate public investor to whom payment was made by farmers for water supplied (which served as an additional stream of costs for farmers, and a stream of benefits for the water supplier, at their respective target IRRs). For the public off-farm developer, the streams of costs were: (i) the capital costs of developing the water and associated enabling infrastructure (including replacement costs and residual values), and (ii) the costs of maintaining and operating those assets. Threshold gross margins and water pricing to achieve target internal rate of return New irrigation schemes in the Southern Gulf catchments would be costly to develop, so many technically feasible options are unlikely to be profitable at the returns and over the time periods expected by many investors. The results presented below suggest it would be difficult for any farming options to fully cover the costs of a large off-farm dam development. However, there are greater prospects of viable developments using on-farm sources of water for broadacre and cost- efficient horticulture. The costs of developing water and land resources for a new irrigation development vary widely depending on a range of case-specific factors that are dealt with in other parts of this Assessment. These factors include the type and nature of the water source, type of water storage, geology, topography, soil characteristics, water distribution system, type of irrigation system, type of crop to be grown, local climate, land preparation requirements, and level to which infrastructure is engineered. The financial analyses, therefore, have taken a generic approach in exploring the consequences for the development costs of this variation, and other key factors that determine whether or not an irrigation scheme would be viable (e.g. farm performance and the level of returns sought by investors). The analyses used the discounted cashflow framework described above to back- calculate and fit the water prices and farm GMs that would be required for respective public (off- farm) and private investors (irrigators) to achieve their target IRRs. The results are summarised in tables showing the thresholds that must be met for a particular combination of water development and farm development options to meet the investor’s target returns. The tables allow viable pairings to be identified based on either threshold costs of water or required farm GMs. Financial viability for these threshold values was defined and calculated as investors achieving their target IRR (or, equivalently, that the investment would have an NPV of zero and a BCR of one at the target discount rate). Assumptions The first analyses considered the case of irrigation schemes built around public investment in a large off-farm dam in the Southern Gulf catchments. The analyses then considered the case of developments using on-farm dams and bores. To keep the results as relevant as possible to a wide range of different development options and configurations, the analyses here do not assume the scale at which a water development would be undertaken. Instead, all costs are expressed per hectare of irrigated farmland and per ML per year of water supply capacity, facilitating comparisons between scenarios (which can differ substantially in size). To illustrate how this slightly abstract generic approach can be applied to specific development projects, three worked examples show the indicative off-farm infrastructure costs that would be involved in development of the most cost-effective dam sites in close proximity to soils suitable for irrigated agriculture identified in the Southern Gulf catchments (Table 6-2). Table 6-2 Indicative capital costs for developing three irrigation schemes based on the most cost-effective dam sites identified in the Southern Gulf catchments FSL = full supply level. For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au To enhance like-for-like comparisons across the various development scenarios, a set of standard assumptions have been made about the breakdown of development costs (by life span) and associated ongoing operating costs (Table 6-3). Three indicative types of farming enterprise represent different levels of capital investment, associated with the intensity of production and the extent to which farming operations are performed on-farm or outsourced (Table 6-3). The capital costs and fixed costs are higher for horticulture than for broadacre farming, but the more expensive irrigation systems used (such as drippers) apply water more precisely and efficiently to crops. The indicative ‘Broadacre’ farm could, for example, represent hay or cotton farming using furrow irrigation on heavier clay soils. The indicative capital-intensive ‘Horticulture-H’ farm could, for example, represent high-value fruit-tree orchards with high standard on-farm packing and cold room facilities, and include accommodation for seasonal workers travelling to very remote Southern Gulf catchments farms. The indicative less-capital-intensive ‘Horticulture-L’ farm option could, for example, represent a row crop, such as melons, with packing directly to bins and use of off-farm accommodation for seasonal workers (which reduces the upfront capital cost of establishing the farm, but increases the ongoing costs for outsourced services, which reduces farm GMs). Table 6-3 Assumed indicative capital and operating costs for new off- and on-farm irrigation infrastructure Three types of farming enterprise represent a range of increasing intensity, value and cost of production. Indicative base capital costs for establishing new farms (excluding water costs) allow on- and off-farm water sources to be added and compared on an equal basis. Annual operation and maintenance (O&M) costs are expressed as a percentage of the capital costs of assets. The Horticulture-H farm, with higher development costs, includes on-farm packing facilities, cold storage, and accommodation for seasonal workers. The Horticulture-L farm, with lower development costs, does not include these assets and would have to outsource these services if required (reducing the farm gross margin). IRR = internal rate of return. SCHEME COMPONENT ITEM VALUE UNIT O&M COST (% capital cost/y) Off-farm infrastructure development capital and operating costs (large dam and enabling infrastructure) Capital costs Total capital costs (split by life span below) Indicative >50,000 (analysed range: 20,000 to 150,000) $/ha Longer life span infrastructure (100 y) 85 % 0.4 Shorter-life-span infrastructure (40 y) 15 % 1.6 Operating costs O&M (by life span categories) % capital cost $/ha/y Off-farm water source pumping costs ~2 (additional) $/ML/m Target IRR Base (with sensitivity range) 7 % Farm development capital and operating costs Broadacre Horticulture-L (low capital) Horticulture-H (high capital) Capital costs Base (excluding water source) 9,000 25,000 70,000 $/ha Water source (on- or off-farm) Indicative >4,000 (analysed range: 3,000 to 15,000) $/ha Longer-life-span infrastructure (40 y) 50 50 50 % 1.0 SCHEME COMPONENT ITEM VALUE UNIT O&M COST (% capital cost/y) Shorter-life-span infrastructure (15 y) 50 50 50 % 1.0 Operating costs O&M (by life-span categories) % capital cost $/ha/y Farm water source pumping costs ~2 (additional) $/ML/m Fixed costs 600 1,500 6,500 $/ha/y Water use Crop water use (before losses) 6 6 6 ML/ha/y On-farm water use efficiency 70 90 90 % Gross margin Indicative gross margin 4,000 7,000 11,000 $/ha/y Target IRR Base (with sensitivity range) 10 10 10 % For consistency, all costs of delivering water to the farm at the level of the soil surface are treated as the costs of the water source (so the costs of the various water source options can be compared on a like-for-like basis). Subsequent farm pumping costs of distributing and applying the supplied water to crops are treated as part of the variable costs of growing crops and are already accounted for in the crop GMs presented in Chapter 4. The pumping costs for the water supplier are highly situation-specific for the various water sources. In particular, these pumping costs are affected by the elevation of the water source relative to the point of distribution to the farm: for example, the height water needs to be pumped from a weir to a distribution channel, or from a farm dam to a field; or the dynamic head required to lift bore water to the field surface. For this reason, water source pumping costs have not been included in the summary tables of water pricing, but should be added separately as required at a cost of approximately $2 per ML per m dynamic head. This is mainly a consideration for groundwater bores, but also applies when water needs to be lifted from rivers or irrigation channels. For more information on water infrastructure costs, see Chapter 5 (and the companion technical reports referenced there). For more information on crop GMs, see Chapter 4 (and the companion technical reports referenced there). The analyses presented below consider (i) the case of irrigation schemes built around a large dam and its associated supporting off-farm infrastructure (Section 6.3.3); (ii) the case of self-contained, modular farm developments with their own on-farm source of water (Section 6.3.4). For both cases, the water price that irrigators can afford to pay provides a useful common point of reference for identifying the most affordable water sources for various types of farm developments (Section 6.3.2). The initial analyses assumed that all farmland was in full production and performed at 100% of its potential (and assumed 100% reliable water supplies) from the start of the development. Section 6.3.5 provides a set of adjustment factors that quantify the risks of various sources of underperformance that can be anticipated. 6.3.2 Price irrigators can afford to pay for a new water source Table 6-4 shows the price that the three different types of farms could afford to pay for water, while meeting a target 10% IRR, for three different levels of farm water use and productivity. For prices to be sustained at this level throughout the life of the water source, the associated farm GM (in the first column of Table 6-4) would also need to be maintained over this period. The table is, therefore, most useful when assessing the long-term price that can be sustained to pay off long- lived water infrastructure (rather than temporary spikes in farm GMs during runs of favourable years). The lowest GM in the first column of Table 6-4 for each farm type is the value below which the farm would not be viable, even if water was free. This does not necessarily mean that such GMs could readily be achieved in practice: for the capital-intensive Horticulture-H farm, in particular, it would be challenging in the Southern Gulf catchments to reach the $17,000 per ha per year GM to cover the farm’s other costs, even before considering the costs of water. Table 6-4 Price irrigators can afford to pay for water, based on the type of farm, the farm water use and the farm annual gross margin (GM), while meeting a target 10% internal rate of return (IRR) Analyses assume water volumes are measured on delivery to the farm gate or surface: pumping costs involved in getting water to the farmland surface would be an additional cost of supplying the water (indicatively $2 per ML per m dynamic head), while pumping costs in distributing and applying the water to the crop are considered part of the variable costs included in the GM. Indicative GMs that the three types of farms could attain in the Southern Gulf catchments are $4000 and $7000 per ha per year for Broadacre and Horticulture-L farms, respectively (blue-shaded rows), and $11,000 per ha per year for Horticulture-H (Table 6-3, Chapter 4). Note that the Horticulture-H farm cannot pay anything for water until it achieves a GM above $17,000 per ha per year. GROSS MARGIN PRICE IRRIGATORS CAN AFFORD TO PAY ($/ha/y) ($/ML at farm gate/surface) Farm water use (ML/ha, including on-farm distribution and application losses) 4 5 6 7 8 9 10 12 Broadacre ($9,000/ha development costs, $600/ha/y fixed costs, 70% on-farm efficiency) 2,000 25 20 17 14 12 11 10 8 2,500 86 69 57 49 43 38 34 29 3,000 147 118 98 84 74 65 59 49 3,500 209 167 139 119 104 93 83 70 4,000 270 216 180 154 135 120 108 90 5,000 392 314 262 224 196 174 157 131 Horticulture-L ($25,000/ha development costs, $1,500/ha/y fixed costs, 90% on-farm efficiency) 5,000 39 31 26 22 19 17 16 13 6,000 241 193 161 138 121 107 97 80 7,000 444 355 296 254 222 197 178 148 8,000 646 517 431 369 323 287 259 215 10,000 1051 841 701 601 526 467 421 350 12,000 1456 1165 971 832 728 647 583 485 Horticulture-H ($70,000/ha development costs, $6,500/ha/y fixed costs, 90% on-farm efficiency) 17,000 203 162 135 116 101 90 81 68 20,000 810 648 540 463 405 360 324 270 25,000 1823 1458 1215 1042 911 810 729 608 30,000 2835 2268 1890 1620 1418 1260 1134 945 40,000 4860 3888 3240 2777 2430 2160 1944 1620 These water prices are likely to be most useful for public investors in large dams, because the sequencing of development creates asymmetric risks between the water supplier and the irrigators. Irrespective of the planned water pricing for a dam project, once the dam is built, irrigators have the choice of whether to develop new farms; they are unlikely to act to their own detriment by making an investment if they cannot do so at a water price that will allow them to obtain a commercial rate of return. These water prices, together with estimates of likely attainable farm GMs (available in Chapter 4), provide a useful benchmark for checking assumptions about any potential public dam developments in the Southern Gulf catchments. For on-farm water sources, these water prices can assist in identifying water development options that cropping operations could reasonably be expected to afford. The tables in the next sections allow comparisons of irrigation development options by converting the capital costs of developing on- and off-farm water sources to volumetric costs ($/ML supplied). All water prices are based on volumes supplied to the farm gate or surface (after losses in transit) per metered ML supplied. 6.3.3 Financial targets required in order to cover full costs of large, off-farm dams The first generic assessment considered the case of public investment in a large dam in the Southern Gulf catchments and whether the costs of that development could be covered by water payments from irrigators (priced at their capacity to pay). The public costs of development include the cost of the dam and water distribution, and of any other supporting infrastructure required. Costs are standardised per unit of farmland developed, noting that for the same outlay, a smaller area could be developed for a crop with a higher water use (i.e. the water development costs per hectare would be higher). Target farm gross margins for off-farm public water infrastructure Table 6-5shows what farm annual GMs would be required for various costs of water infrastructure development at the public investors’ target IRR. As expected, higher farm GMs are required in order to cover higher capital costs and/or to attain a higher target IRR. The tables in this section can be used to assess whether water development opportunities and farming opportunities in the Southern Gulf catchments are likely to combine in financially viable ways. Indicative farm GMs that could be achieved in the Southern Gulf catchments are approximately $4000, $7000 and $11,000 per ha per year for Broadacre, less-capital-intensive Horticulture-L (including penalising GMs if outsourcing occurs) and capital-intensive Horticulture-H, respectively (Table 6-3). A dam and supporting infrastructure would likely require at least $50,000/ha of capital investment (Table 6-2). None of the three farming types is likely to be viable at these farm GMs and water development costs (at a 7% target IRR for the public investor). However, Broadacre and Horticulture-L farming might be marginally viable at a 3% target IRR for the public investor. Broadacre and lower-cost Horticulture-L could both achieve a target 10% IRR for the farm investments while contributing $20,000 to $30,000 per ha (40%–60%) towards the cost of a dam that cost $50,000/ha to build (including enabling infrastructure and ongoing O&M costs). That is a higher proportion of costs than irrigators have historically contributed towards irrigation schemes in some other parts of Australia (typically approximately a quarter of capital costs (Vanderbyl, 2021)), and would involve a decision by the Australian Government and the NT and Queensland governments based on their expectations, priorities, and investment criteria. 390 | Water resource assessment for the Southern Gulf catchments Table 6-5 Farm gross margins (GMs) required in order to cover the costs of off-farm water infrastructure (at the supplier’s target internal rate of return (IRR)) Assumes 100% farm performance on all farmland in all years, once construction is complete. Costs of supplying water to farms are consistently treated as costs of water source development (and not part of the farm GM calculation). Risk adjustment multipliers are provided in Section 6.3.5. Blue-shaded cells indicate the capital costs that could be afforded by farms with GMs of $4000 (Broadacre), $7000 (Horticulture-L) and $11,000 (Horticulture-H) per ha per year. Blue-shaded column headers indicate the most cost-effective dam development options in the Southern Gulf catchments (Table 6-2). TARGET IRR FARM GROSS MARGIN REQUIRED IN ORDER TO PAY FOR OFF-FARM WATER INFRASTRUCTURE (%) ($/ha/y) Total capital costs of off-farm water infrastructure ($/ha) 20,000 30,000 40,000 50,000 70,000 100,000 125,000 150,000 Broadacre ($9,000/ha development costs, $600/ha/y fixed costs, 70% on-farm efficiency) 3 2,604 3,016 3,428 3,840 4,664 5,900 6,930 7,960 5 2,977 3,569 4,160 4,751 5,933 7,707 9,185 10,663 7 3,359 4,139 4,920 5,701 7,263 9,605 11,558 13,510 10 3,941 5,013 6,085 7,157 9,301 12,516 15,196 17,876 12 4,333 5,601 6,869 8,137 10,673 14,478 17,648 20,818 Horticulture-L ($25,000/ha development costs, $1,500/ha/y fixed costs, 90% on-farm efficiency) 3 5,584 5,996 6,408 6,820 7,645 8,881 9,911 10,941 5 5,985 6,576 7,167 7,759 8,941 10,715 12,193 13,671 7 6,370 7,150 7,931 8,712 10,274 12,616 14,569 16,521 10 6,952 8,024 9,096 10,168 12,312 15,528 18,208 20,887 12 7,345 8,613 9,881 11,149 13,685 17,489 20,659 23,829 Horticulture-H ($70,000/ha development costs, $6,500/ha/y fixed costs, 90% on-farm efficiency) 3 16,618 17,068 17,518 17,967 18,867 20,217 21,342 22,467 5 17,164 17,789 18,413 19,038 20,288 22,162 23,724 25,286 7 17,610 18,416 19,222 20,027 21,638 24,055 26,070 28,084 10 18,215 19,301 20,387 21,472 23,644 26,901 29,615 32,330 12 18,607 19,884 21,161 22,438 24,992 28,823 32,015 35,207 Target water pricing for off-farm public water infrastructure Table 6-6 shows the price that a public investor in off-farm water infrastructure would have to charge to fully cover the costs of development of off-farm water infrastructure, expressed per unit of supply capacity at the dam wall. Pricing assumes that the full supply of water (i.e. reservoir yield) would be used and paid for every year over the entire lifetime of the dam, after accounting for water losses between the dam and the farm. It can be challenging for farms to sustain the high levels of revenue over such long periods (100 years) to justify the costs of building expensive dams. For these base analyses, the water supply is assumed to be 100% reliable; risk adjustment multipliers to account for reliability of supply are provided in Section 6.3.5. For example, in the Southern Gulf catchments some of the most cost-effective dam opportunities would cost approximately $5000 per ML per year of supply capacity at the dam wall, including the cost of the required supporting off-farm water infrastructure (Table 6-2). This would require farms to pay $537 per ML extracted to fully cover the costs of the public investment (at the base 7% target IRR for public investments, Table 6-6). Comparing this figure with what irrigators can afford to pay (Table 6-4) shows that it is unlikely any farming options could cover the costs of a dam in the Southern Gulf catchments at the GMs farms are likely to be able to achieve (Table 6-3, Chapter 4). When a scheme is not viable (BCR < 1), the water cost and pricing tables can be used to generate approximate estimations of the BCR and the likely proportion of public development costs that farms would be able to cover. For example, a Broadacre farm that uses 8 ML/ha (measured at delivery to the farm) and has a GM of $4000 per ha per year could afford to pay $135/ML extracted (Table 6-4), which would cover 25% ($135/$537) of the $537/ML price (Table 6-6) required to cover the full costs of the public development. The BCR would, therefore, be 0.25 (the ratio of the amount the net farm benefits can cover to the full costs of the scheme). As for the example in Table 6-5, the proportion of the capital costs of infrastructure projects they would realistically expect to recover from users would be a decision for the public investor. Table 6-6 Water pricing required in order to cover costs of off-farm irrigation scheme development (dam, water distribution, and supporting infrastructure) at the investors target internal rate of return (IRR) Assumes the conveyance efficiency from dam to farm is 70% and that supply is 100% reliable. Risk adjustment multipliers for water supply reliability are provided in Table 6-9. Pumping costs between the dam and the farm would need to be added (e.g. ~$30/ML extra to lift water ~15 m from the weir pool to distribution channels). ‘$ CapEx per ML/y at dam’ is the capital expenditure on developing the dam and supporting off-farm infrastructure per ML per year of the dam’s supply capacity measured at the dam wall. Blue-shaded cells indicate $/ML cost of water. Blue- shaded column headers are indicative of the most cost-effective large dam options available in the Southern Gulf catchments (Table 6-2). TARGET IRR WATER PRICE THAT WOULD NEED TO BE CHARGED IN ORDER TO COVER OFF-FARM INFRASTRUCTURE COSTS (%) ($/ML charged at farm gate) Capital costs of off-farm infrastructure ($ CapEx per ML/y at dam) 3,000 4,000 5,000 6,000 8,000 10,000 12,000 14,000 16,000 3 162 215 269 323 431 538 646 754 861 5 239 319 399 479 638 798 958 1117 1277 7 322 429 537 644 859 1073 1288 1502 1717 10 448 598 747 897 1196 1495 1794 2093 2392 6.3.4 Financial targets required in order to cover costs of on-farm dams and bores The second generic assessments considered the case of on-farm sources of water. Indicative costs for on-farm water sources, including supporting on-farm distribution infrastructure, vary between $4000 and $15,000 per ha of farmland. Costs depend on the type of water source, how favourable the local conditions are for its development, and the irrigation requirement of the farming system. Since the farm and water source would be developed by a single investor, the first analyses considered the combined cost of all farm development together (without separating out the water component). 392 | Water resource assessment for the Southern Gulf catchments Target farm gross margins required in order to cover full costs of greenfield farm development with water source Table 6-7 shows the farm GMs that would be required in order to cover various costs of farm development at the investor’s target IRR. Note that private on-farm water sources are typically engineered to a lower standard than public water infrastructure and have lower upfront capital costs, higher recurrent costs (higher O&M and asset replacement rates) and lower reliability. Based on the indicative farm GMs provided earlier (Table 6-3) and a 10% target IRR, a Broadacre farm with a $4000 per ha per year GM could cover total on-farm development capital costs of approximately $20,000/ha. A lower-capital-cost Horticulture-L farm with a GM of $7000 per ha per year could afford approximately $40,000/ha of initial capital costs, and a capital-intensive Horticulture-H farm with a GM of $11,000 per ha per year could pay approximately $30,000/ha for farm development (Table 6-7). This indicates that on-farm water sources may have better prospects of being viable than large public dams in the Southern Gulf catchments, particularly for broadacre farms and horticulture with lower development costs, if good sites can be identified for developing sufficient on-farm water resources at a low-enough cost. Table 6-7 Farm gross margins (GMs) required in order to achieve target internal rates of return (IRR), given various capital costs of farm development (including an on-farm water source) Assumes 100% farm performance on all farmland in all years, once construction is complete. Risk adjustment multipliers are provided in Section 6.3.5. Blue-shaded cells indicate the capital costs that could be afforded by farms with GMs of $4000 (Broadacre), $7000 (Horticulture-L) and $11,000 (Horticulture-H) per ha per year. TARGET IRR FARM GROSS MARGIN REQUIRED IN ORDER TO ACHIEVE THE FARMER'S TARGET IRR (%) ($/ha/y) Total capital costs of farm development, including water source ($ CapEx/ha) 10,000 15,000 20,000 30,000 40,000 50,000 70,000 100,000 Broadacre ($600/ha/y fixed costs, 70% on-farm efficiency) 5 1,516 1,957 2,398 3,279 4,160 5,042 6,804 9,449 7 1,669 2,181 2,694 3,718 4,742 5,767 7,815 10,888 10 1,923 2,554 3,185 4,447 5,709 6,972 9,496 13,282 12 2,105 2,821 3,537 4,968 6,400 7,832 10,696 14,991 15 2,389 3,238 4,087 5,785 7,483 9,181 12,578 17,672 20 2,882 3,963 5,044 7,206 9,368 11,530 15,854 22,340 Horticulture-L ($1,500/ha/y fixed costs, 90% on-farm efficiency) 5 2,469 2,909 3,350 4,231 5,113 5,994 7,757 10,401 7 2,637 3,149 3,661 4,685 5,710 6,734 8,783 11,856 10 2,915 3,546 4,177 5,439 6,702 7,964 10,488 14,274 12 3,114 3,830 4,546 5,978 7,409 8,841 11,705 16,001 15 3,424 4,273 5,122 6,820 8,519 10,217 13,613 18,708 20 3,962 5,043 6,124 8,286 10,448 12,610 16,934 23,420 TARGET IRR FARM GROSS MARGIN REQUIRED IN ORDER TO ACHIEVE THE FARMER'S TARGET IRR (%) ($/ha/y) Total capital costs of farm development, including water source ($ CapEx/ha) 10,000 15,000 20,000 30,000 40,000 50,000 70,000 100,000 5 7,760 8,201 8,642 9,523 10,404 11,286 13,048 15,692 7 8,012 8,524 9,036 10,060 11,085 12,109 14,158 17,231 10 8,427 9,058 9,689 10,951 12,213 13,475 15,999 19,785 12 8,720 9,436 10,152 11,584 13,016 14,448 17,312 21,607 15 9,177 10,026 10,875 12,573 14,271 15,970 19,366 24,461 20 9,963 11,044 12,125 14,287 16,449 18,611 22,935 29,421 Volumetric water cost equivalent for on-farm water source Table 6-8 converts the capital cost of developing an on-farm water source (per ML of annual supply capacity) into an equivalent cost for each individual ML of water supplied by the water source. The table can be used to estimate how much a farm could spend on developing the required water resources by comparing the costs per ML with what farms can afford to pay for water (Table 6-4). For example, a Broadacre farm with a GM of $4000 per ha per year, an annual farm water use of 8 ML/ha, and a target 10% IRR could afford to pay $135/ML for its water supply (Table 6-4), which would allow capital costs of approximately $1000 for each ML/year supply capacity for developing an on-farm supply (Table 6-8). Approximate indicative costs for developing on-farm water sources range from $500/ML to $2000/ML (based on the range of per hectare costs above). This alternative approach confirms that there are likely to be viable farming opportunities using on-farm water development in the Southern Gulf catchments. Table 6-8 Equivalent costs of water per ML for on-farm water sources with various capital costs of development, at the internal rate of return (IRR) targeted by the investor Assumes the water supply is 100% reliable. Risk adjustment multipliers for water supply reliability are provided in Table 6-9. Pumping costs to the field surface would be extra (e.g. ~$2 per ML per m dynamic head for bore pumping). Blue-shaded cells indicate $/ML cost of water. TARGET IRR WATER VOLUMETRIC COST-EQUIVALENT UNIT FOR VARIOUS CAPITAL COSTS OF WATER SOURCE (%) ($/ML) Capital costs for on-farm water infrastructure ($ CapEx per ML/y at farmland surface) 300 400 500 700 1000 1250 1500 1750 2000 3 22 29 37 51 74 92 110 129 147 5 26 35 44 61 87 109 131 153 175 7 31 41 51 72 102 128 154 179 205 10 38 51 63 89 127 159 190 222 254 12 43 58 72 101 144 180 216 252 288 15 51 68 85 120 171 213 256 299 342 20 65 87 109 152 217 271 326 380 434 6.3.5 Risks associated with variability in farm performance This section assessed the impacts of two types of risks on scheme financial performance: those that reduce farm performance through the early establishment and learning years, and those occurring periodically throughout the life of the development. The effect of these risks is to reduce the expected revenue and expected GMs. Setbacks that occur soon after a scheme is established were found to have the largest effect on scheme viability, particularly at higher target IRRs. There is a strong incentive to start any new irrigation development with well-established crops and technologies, and to be thoroughly prepared for the those agronomic risks of establishing new farmland that can be anticipated. The analyses showed that delaying full development for longer periods than the learning time had only a slight negative effect on IRRs, whereas proceeding to full development before learning was complete had a much larger impact. This implies that it is prudent to err on the side of delaying full development (particularly given that, in practice, it is only possible to know when full performance has been achieved in retrospect). An added benefit of staging is the limiting of losses when small- scale testing proves initial assumptions of benefits to be overoptimistic and indicates that full- scale development could never be profitable (even after attempts to overcome unanticipated challenges). For an investment to be viable, farm GMs must be sustained at high levels over long periods. Thus, variability in farm performance poses risks that must be considered and managed. GMs can vary between years because of either short-term initial underperformance or periodic shocks. Initial underperformance is likely to be associated with learning, as farming practices need to be adapted to local conditions and to overcome initial challenges in order to reach their long-term potential. In addition, unavoidable periodic risks are associated with water reliability, climate variability, flooding, outbreaks of pests and diseases, periodic technical or equipment failures, and fluctuations in commodity prices and market access. Unreliability of water supply, is less easy to avoid than other periodic risks. Risks that cannot be avoided must be managed, mitigated where possible, and accounted for in determining the realistic returns that can be expected from an irrigation development. Such accounting would include having adequate capital buffers for survival through challenging periods. Another perceived risk for investors is the potential of future policy changes and delays in regulatory approvals. Reducing this, or any other sources of risk, in the Southern Gulf catchments would help make marginal investment opportunities more attractive. The results of the analyses of both periodic and learning risks are shown below. Right to farm and other sovereignty risks, especially with regard to access to water, may become key factors in future years, based on experience from elsewhere, but these are not the subject of the risk discussion presented here. Throughout this section, farm performance in a given year is quantified as the proportion of the long-term mean GM a farm attains: 100% performance is when this level is reached, and 0% equates to a performance where revenues only balance variable costs (GM = zero). Risks from periodic underperformance The analyses considered periodic risks generically, without assuming any of the particular causes listed above. To quantify their effects on scheme financial performance, the periodic risks were characterised by three components: • reliability – the proportion of ‘good’ years, in which ‘full’ 100% farm performance was achieved, with the remainder of years being termed ‘failed’ years, in which some negative impact was experienced • severity – the farm performance in a ‘failed’ year, in which some type of setback occurred • timing – in ‘early’ timing (in relation to a 10-year cycle), the ‘failed’ years came early in each 10- year cycle (e.g. 80% reliability meant that ‘failed’ years occurred in the first 2 years of the scheme and in the first 2 years of each 10-year cycle after that). In ‘late’ timing, the ‘failed’ years came at the end of each 10-year cycle. In ‘random’ timing, each year was allocated the long- term mean farm performance of ‘good’ and ‘failed’ years (frequency weighted). Table 6-9 summarises the effects of a range of reliabilities and severities for periodic risks on scheme viability. Periodic risks had a consistent proportional effect on target GMs, irrespective of development options or costs, so the results were simplified as a set of risk adjustment multipliers. The multipliers can, therefore, be applied to the target farm GMs in Section 6.3.2 (the GMs required in order to cover the capital costs of development at the investor’s target IRRs at 100% farm performance) to account for the effects of various risks. These same adjustment factors could be applied to the water prices that irrigators can afford to pay (Table 6-4), but would be used as divisors to reduce the price that irrigators could pay for water. Table 6-9 Risk adjustment factors for target farm gross margins (GMs), accounting for the effects of the reliability and severity (level of farm performance in ‘failed’ years) of the periodic risk of water reliability Results are not affected by discount rates. ‘Good’ years = 100% farm performance; ‘failed’ years = <100% performance. ‘Failed year performance’ is the mean farm GM for years in which some type of setback is experienced, relative to the mean GM when the farm is running at ‘full’ performance. FAILED YEAR PERFORMANCE (%) RISK ADJUSTMENT MULTIPLIER FOR TARGET FARM GROSS MARGINS (VS BASE 100% RELIABILITY TABLES) (unitless ratio) Reliability (proportion of ‘good’ years) 1.00 0.90 0.85 0.80 0.70 0.60 0.50 0.40 0.30 0.20 85 1.00 1.02 1.02 1.03 1.05 1.06 1.08 1.10 1.12 1.14 75 1.00 1.03 1.04 1.05 1.08 1.11 1.14 1.18 1.21 1.25 50 1.00 1.05 1.08 1.11 1.18 1.25 1.33 1.43 1.54 1.67 25 1.00 1.08 1.13 1.18 1.29 1.43 1.60 1.82 2.11 2.50 0 1.00 1.11 1.18 1.25 1.43 1.67 2.00 2.50 3.33 5.00 As expected, the greater the frequency and severity of ‘failed’ years, the greater the impact on the scheme viability and the greater the increase in farm GMs required in order to offset these impacts. As an example, the reliability of water supply is one of the more important sources of unavoidable variability in the productivity of irrigated farms. Water reliability (proportion of ‘good’ years, in which the full supply of water is available) is shown as ‘reliability’ in Table 6-9, and the mean percentage of water available in a ‘failed’ year (in which less than the full supply of water is available) is shown as the ‘failed year performance’ in For crops for which the quality of the produce is more important than the quantity, such as horticulture, the approach of reducing the planted land area in proportion to the available water in ‘failed’ years would be reasonable. For perennial horticulture or tree crops, it may be difficult to reduce (or increase) areas on an annual basis. Farmers of these crops would, therefore, tend to opt for systems with a high degree of reliability of water supply (e.g. 95%). For many broadacre crops, deficit irrigation could partially mitigate impacts on farm performance in years with reduced water availability, as could carryover effects from inputs (such as fertiliser) in a ‘failed’ year that reduce input costs the following year (see Section 4.3.4). Table 6-10 shows how the timing of periodic impacts affects scheme viability, providing risk adjustment factors for a range of reliabilities for an impact that had 50% severity with late timing, early timing and random (long-term frequency, weighted mean performance) timing. These results indicate that any negative disturbances that reduce farm performance will have a larger effect if they occur soon after the scheme is established, and that this effect is greater at higher target IRRs. For example, at a 7% target IRR and 70% reliability with ‘late’ timing (in which setbacks occur in the last 3 of every 10 years), the GM multiplier is 1.13, meaning the annual farm GM would need to be 13% higher than if farm performance were 100% reliable. In contrast, for the same settings with ‘early’ timing, the GM multiplier is 1.23, meaning the farm GM would need to be 23% higher than if farm performance were 100% reliable. The impacts of early setbacks are more severe than the impacts of late setbacks. Table 6-10 Risk adjustment factors for target farm gross margins (GMs) accounting for the effects of reliability and the timing of periodic risks Assumes 50% farm performance during ‘failed’ years, in which 50% farm performance means 50% of the GM at ‘full’ potential production. IRR = internal rate of return. TARGET IRR (%) TIMING OF FAILED YEARS RISK ADJUSTMENT MULTIPLIER FOR TARGET FARM GROSS MARGINS (VS BASE 100% RELIABILITY TABLES) (unitless ratio) Reliability (proportion of ‘good’ years) 1.00 0.90 0.80 0.70 0.60 0.50 0.40 0.30 0.20 3 Late 1.00 1.05 1.10 1.16 1.22 1.30 1.39 1.50 1.63 Random – no bias 1.00 1.05 1.11 1.18 1.25 1.33 1.43 1.54 1.67 Early 1.00 1.06 1.13 1.20 1.28 1.37 1.47 1.58 1.70 7 Late 1.00 1.04 1.08 1.13 1.19 1.26 1.35 1.46 1.59 Random – no bias 1.00 1.05 1.11 1.18 1.25 1.33 1.43 1.54 1.67 Early 1.00 1.07 1.15 1.23 1.32 1.41 1.51 1.62 1.74 10 Late 1.00 1.03 1.07 1.12 1.17 1.24 1.32 1.42 1.56 Random – no bias 1.00 1.05 1.11 1.18 1.25 1.33 1.43 1.54 1.67 Early 1.00 1.08 1.16 1.25 1.35 1.45 1.55 1.66 1.77 Risks from initial ‘learning’ period Another form of risk arises from the initial challenges of establishing new agricultural industries in the Southern Gulf catchments; it includes setbacks from delays, such as gaining regulatory approvals, and adapting farming practices to conditions in the Southern Gulf catchments. Some of these risks are avoidable, provided investors and farmers learn from past experiences of development in northern Australia (e.g. Ash et al., 2014), avoid previous mistakes and select farming options that are already well proven in analogous northern Australian locations. However, even well-prepared developers are likely to face initial challenges in adapting to the unique circumstances of a new location. Newly developed farmland can take some time to reach its productive potential, when soil nutrient pools have been established, soil limitations ameliorated, suckers and weeds controlled, and pest and weed management systems established. ‘Learning’ (used here to broadly represent all aspects of overcoming initial sources of farm underperformance) was assessed in terms of two simplified generic characteristics: •initial level of performance – the proportion of the long-term mean GM that the farm achievesin its first year •time to learn – the number of years taken to reach the long-term mean farm performance. Performance was represented as increasing linearly over the learning period from the starting level to the long-term mean performance level (100%). The effect of learning on scheme financial viability was considered for a range of initial levels of farm performance and learning times. As described above, learning had consistent proportional effects on target GMs, so the results were simplified as a set of risk adjustment factors (Table 6-11). As expected, the impacts on scheme viability are greater the lower the starting level of farm performance and the longer it takes to reach the long-term performance level. Since these impacts, by their nature, are weighted to the early years of a new development, they have more impact at higher target IRRs. To minimise the risks of learning impacts, there is a strong incentive to start any new irrigation development with well-established crops and technologies, and to be thoroughly prepared for those agronomic risks of establishing new farmland that can be anticipated. Higher-risk options (e.g. novel crops, equipment or practices that are not currently in profitable commercial use in analogous environments) could be tested and refined on a small scale until locally proven. Table 6-11 Risk adjustment factors for target farm gross margins (GMs), accounting for the effects of learning risks Learning risks were expressed as the level of initial farm underperformance and the time taken to reach full performance levels. Initial farm performance is the initial GM as a percentage of the GM at ‘full’ performance. IRR = internal rate of return. TARGET IRR (%) INITIAL FARM PERFORMANCE (%) RISK ADJUSTMENT MULTIPLIER FOR TARGET FARM GROSS MARGINS (VS BASE 100% RELIABILITY TABLES) (unitless ratio) Learning time (years to 100% performance) 2 4 6 8 10 15 3 85 1.01 1.02 1.03 1.03 1.04 1.05 75 1.02 1.03 1.04 1.05 1.07 1.10 50 1.04 1.06 1.09 1.12 1.14 1.21 25 1.06 1.10 1.14 1.19 1.23 1.35 TARGET IRR INITIAL FARM PERFORMANCE (%) RISK ADJUSTMENT MULTIPLIER FOR TARGET FARM GROSS MARGINS (VS BASE 100% RELIABILITY TABLES) (unitless ratio) 7 85 1.02 1.03 1.04 1.05 1.05 1.07 75 1.03 1.05 1.06 1.08 1.09 1.13 50 1.06 1.10 1.13 1.17 1.21 1.29 25 1.09 1.15 1.22 1.28 1.35 1.51 0 1.12 1.21 1.31 1.41 1.52 1.83 10 85 1.02 1.03 1.05 1.06 1.07 1.09 75 1.04 1.06 1.08 1.10 1.11 1.15 50 1.08 1.12 1.17 1.21 1.26 1.35 25 1.12 1.20 1.28 1.36 1.44 1.65 0 1.16 1.28 1.41 1.55 1.69 2.10 As indicated in the examples above, the influence of each risk individually can be quite modest. However, the combined influence of all foreseeable risks must be accounted for in planning, and the cumulative effect of these risks can be substantial. For example, the last question in Table 6-1 shows that the combined effect of just two risks requires farm GMs to be approximately 50% higher than they would be without the risks. See Stokes and Jarvis (2021) for the effects of a common suite of risks on the financial performance of a Bradfield-style irrigation scheme. 6.4 Cost–benefit considerations for water infrastructure viability 6.4.1 Lessons from recent Australian dams CBA is widely used to help decision makers evaluate the net benefits likely to arise from implementing a proposed project, particularly for investments in large-scale public infrastructure. Despite this wide usage of CBAs, there are a few examples for which the estimated costs and benefits used to justify the project have been revisited at a later date. Such ex-post evaluations allow the outcomes of completed projects to improve planning, management and risk mitigation in future projects (Infrastructure Australia, 2021a). The few examples in which water infrastructure CBAs have been evaluated have focused on exploring the accuracy of the forecast capital costs. An international study of large water infrastructure projects showed that actual construction costs exceeded contracted costs by a mean of 96% (Ansar et al., 2014). Similarly, an Australian-focused study found mean cost overruns of 120% (Petheram and McMahon, 2019). There is evidence of a systematic tendency across a range of large infrastructure projects for proponents to substantially under estimate development costs (Ansar et al., 2014; Flyvbjerg et al., 2002; Odeck and Skjeseth, 1995; Wachs, 1990; Western Australian Auditor General, 2016). Ex-post evaluations of project benefits are even scarcer. One international study found that large dam developments frequently underperformed, whereby ‘irrigation services have typically fallen short of physical targets, did not recover their costs and have been less profitable in economic terms than expected’ (World Commission on Dams, 2000). In particular, this study highlighted inaccurate and overestimated forecasting of future irrigation demand for water from dam developments. Review of recent Australian dams The Roper River Water Resource Assessment technical report on agricultural viability and socio- economics (Stokes et al., 2023) conducted a systematic review of the five most recently built dams in Australia (Figure 6-2, Table 6-12) to address the gap in the ex-post evaluations. The goal was to assess how well Australian dam projects have achieved their anticipated benefits and to make the learnings available for future planning. These lessons provide a context for interpreting CBAs from project proponents and independent analysts, and the financial analyses provided in the previous section. The key lessons from that review are summarised below, and the full details are reported in Webster et al. (2024). Locations of five dams used in costing review map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\2_Victoria\1_GIS\1_Map_docs\Se-V-503_Map_Australia_and_river_basins_new dams_V1.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 6-2 Locations of the five dams used in this review The dams are numbered in blue as 1: New Harvey Dam, 2: Paradise Dam, 3: Meander Dam, 4: Wyaralong Dam and 5: Enlarged Cotter Dam. Table 6-12 Summary characteristics of the five dams used in this review The documents reviewed for each dam are cited in the companion technical report on agricultural viability and socio- economics (Webster et al., 2024). CBA = cost–benefit analysis. NEW HARVEY DAM PARADISE DAM MEANDER DAM WYARALONG DAM ENLARGED COTTER DAM State/territory WA Qld Tas Qld ACT Date completed 2002 2005 2008 2011 2012 Capacity (GL) 59 300 43 103 78 New dam or redevelopment of existing dam? Replaces Harvey weir (built 1916, extended 1931), capacity of ~10 GL New New New Replaces original Cotter Dam (built 1915, extended 1951), capacity of ~4 GL Primary use(s) proposed for water from dam Irrigated agriculture Irrigated agriculture, water supply Irrigated agriculture, environmental flows, hydro- electric power Water supply to South East Queensland Water supply for Canberra Type of key project documents used for this review Proposed water allocation plans (no CBA available) CBA and economic impact assessment CBA Environmental Impact Statement (EIS) (no CBA available) EIS (which included CBA information, but the actual CBA report was unavailable) Summary of key issues identified This review highlighted a number of issues with the historical use of CBAs for recently built dams in Australia and suggested how CBAs could be more rigorously addressed (Table 6-13). These issues arise because of the complexity of the forecasts and estimates required in planning large infrastructure projects and because of pressures on proponents that can introduce systematic biases. However, this report acknowledges that flaws with the use of CBAs in large public infrastructure investment decisions are not unique to regional Australia or to water infrastructure – they are systemic and occur in relation to many different types of infrastructure globally. Under such circumstances, it would be inequitable to apply more rigour to CBAs only for some select investments, geographic regions, and infrastructure classes before the same standards are routinely applied in all cases. And there is no incentive for individual proponents to apply more rigour to CBAs if their proposals would suffer from unfavourable comparisons with alternative or competing investments with exaggerated cost-to-benefit ratios (CBRs). In the short term, the main value of the information provided here is to enable more critical interpretation and evaluation of CBAs so that more-informed decisions can be made about the likely viability (and relative ranking) of projects in practice. In particular, it highlights several aspects of CBAs regarding which the claims of proponents warrant critical scrutiny. The longer term value of this analysis is that it has identified many issues similar to those raised in past review cycles of Infrastructure Australia’s CBA best-practice guidelines and in the recommendations that are being progressively added to those guidelines to improve how large public investments are evaluated (Infrastructure Australia, 2021a, 2021b). Table 6-13 Summary of key issues and potential improvements arising from a review of recent dam developments KEY ISSUE POTENTIAL IMPROVEMENTS 1 There is a lack of clear documentary evidence regarding the actual outcome of dam developments compared with the assumptions made in ex-ante proposals, Environmental Impact Statements (EISs) and cost– benefit analyses (CBAs). Ex-post evaluations or post- completion reviews have either not been prepared or not been made publicly available. Conducting ex-post evaluations of developments and making these publicly available (as recommended by 2021 guidance from Infrastructure Australia (Infrastructure Australia, 2021a, 2021b) and in the 2022 National Water Grid Investment Framework (NWGA 2022)) would enable lessons learned to be shared and benefit future developments. 2 Predicted increases in water demand from specific developments generally do not appear to arise at the scale and/or within the time frame forecast. While the reasons for this are varied and context-dependent, there does appear to be a systematic bias towards overestimation of the magnitude and rate at which new benefits would flow. Recognising the tendency towards a systematic bias of overstating benefits and understating costs, CBAs in project proposals could be improved by: (i) further efforts to present unbiased financial analysis (e.g. independent review) and ensuring appropriate sensitivity analysis is included in all proposals, (ii) developing broadly applicable and realistically achievable benchmarks for evaluating proponents’ assumptions and financial performance claims, (iii) using past experiences and lessons learned from previous projects with a similar context to inform the analysis presented in the proposals (building on Issue 1 above), and (iv) presenting a like-for-like comparison of cost-to-benefit ratios (CBRs) for the proposed case vs standard alternatives (such as water buybacks or a smaller dam, possibly better matched to realistic future demand). 3 The systematic bias towards optimism in proposals is exacerbated by mismatches between forecast demand and the full supporting infrastructure required to enable this demand to be realised, resulting in additional capital investment (pipelines, treatment plants, etc.) being required that was not costed in the original proposal. The same improvements as for Issue 2 in recognising and addressing inherent bias apply here. 4 Developments are justified based on a complex mix of multiple market and non-market benefits, many of which are hard to monetise and capture in a single net present value (NPV) figure. CBAs could be improved by presenting clear information on the full portfolio of benefits (and costs and disbenefits) anticipated to arise from a project. While the quantitative part of the CBA would analyse the easily monetised costs and benefits (with metrics such as CBR and NPV), benefits that are hard to monetise could also be formally presented in whatever form is most appropriate to the magnitude and nature of the particular benefit. This presentation would enable the relative importance of each element of the mix to be weighed and given appropriate consideration, rather than attention being focused on a single NPV figure, which may have omitted key elements of the project. 5 Improved water security and reliability of supply is often the most important benefit offered by dam developments, while also being the hardest to monetise. Dams provide a form of insurance against the risk that water may not be available when needed in the future. Assessing the value of this insurance requires consideration of the cost of lack of water supply when needed and the likelihood that this could occur. CBAs could be improved by providing clear information on exactly how the development will serve to improve water security, the likelihood that such insurance will be required (i.e. an estimate of the risk), and the estimated social and economic impacts if the insurance was not there when required. Such information could be presented alongside, and given equal prominence with, other information regarding the proposal, including the estimated NPV. This is preferable to attempting to ‘force’ the benefit into an NPV calculation that is ill-equipped to deal with such a benefit. 6.4.2 Demand trajectories for water use For irrigated agriculture to expand in the Southern Gulf catchments, additional water will be required. Forecasting that growth in demand is essential, both for planning new water infrastructure and for evaluating individual water infrastructure proposals. This will ensure assumed demand trajectories for water, and the associated value that can be generated from irrigated agriculture to justify the costs of that infrastructure, are reasonable. Australian Bureau of Statistics data series on historical agricultural production and water use were analysed to derive trends and relationships for benchmarking realistic growth trajectories in Queensland ( (a) Australia (b) Queensland Trend in gross value of ag production, Aust \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\3_Economic\SoGWRA-Charts_Economic.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au 010,00020,00030,0001981-901991-002001-102011-21GVAP ($M) DecadeCrops (horticulture)Crop (other)Livestock Trend in gross value of ag production, Qld \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\3_Economic\SoGWRA-Charts_Economic.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au 020004000600080001981-901991-002001-102011-21GVAP ($M) DecadeCrops (horticulture)Crop (other)Livestock Figure 6-3 Trends in gross value of agricultural production (GVAP) in (a) Australia and (b) Queensland over 40 years (1981–2021) Data points are decade averages of annual values. The ‘Crop (other)’ category is predominantly broadacre farming. Source: (ABS, 2022) Horticultural produce is typically perishable and expensive to store and transport, and must meet stringent phytosanitary (plant health) standards for export, so most Australian horticultural produce (~70%) is sold domestically for consumption shortly after harvest. Growth in horticultural industries is, therefore, constrained by growth in demand from local consumers. The current rate of growth in the value of Australian horticulture is $2.7 billion per decade (step changes in gross value of agricultural production (GVAP) from 1981–90 to 2011–21 are shown in Figure 6-3). Any new irrigated development would compete for some share of that growth, providing a benchmark guide for the scale of new horticulture that could realistically be included in any new irrigation scheme. It also provides a benchmark for the trajectory at which high-value horticulture (and the associated demand for high-priority water) could grow towards the ultimate scheme potential. In addition, the scale of new horticultural expansion for any single crop is limited by seasonal gaps in supply, so the horticulture in any single location is typically a mix of products that fill the niche market gaps that the location can supply (usually dictated by climate, but sometimes a result of other factors, such as backloading opportunities; see Chapter 4), rather than being a monoculture of the most valuable crop alone. Data on how the value of irrigated agriculture has increased with increasing irrigation water availability over time provide an indicative benchmark of how much gross value such a mix of new agricultural activities could generate for each new GL of irrigation water that becomes available (Figure 6-4). Based on the trendlines in Figure 6-4, each extra new GL of water use could produce: • between $2.1 and $3.7 million of gross value from mixed fruit industries • between $5.6 and $10.3 million of gross value from mixed vegetable industries • between $2.5 and $5.0 million of gross value from mixed horticulture (combined), or • between $0.8 and $1.7 million of gross value from a typical mix of agriculture overall. Growth trends in the value of broadacre crops are stronger than those for horticulture (Figure 6- 3); they are a combination of increases in both product volumes and the value per unit product. Unlike horticultural crops, bulk broadacre commodities are stored and traded on large global markets (with multiple competing international buyers), which could easily absorb the scale of increases in production that would be possible from the Southern Gulf catchments. However, supply chains, rather than markets, pose a challenge for new broadacre production. Despite northern Australia being geographically closer than southern Australia to many key markets, the supply chains for northern Australian produce are longer, because most agricultural exports leave through southern ports. For example, the Port of Townsville currently does not handle bulk food- grade containers (for either import or export). The challenge is to develop transport and handling capacity for exports and balance that with compatible imports to avoid the added cost of dead freighting empty containers (CRCNA, 2020). (a) Fruits (c) Fruits and vegetables combined (b) Vegetables (d) Total agriculture Trend in gross value of irrigated ag with water applied, fruit \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\3_Economic\SoGWRA-Charts_Economic.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Trend in gross value of irrigated ag with water applied, fruit and vegetable \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\3_Economic\SoGWRA-Charts_Economic.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Trend in gross value of irrigated ag with water applied, vegetable \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\3_Economic\SoGWRA-Charts_Economic.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Trend in gross value of irrigated ag with water applied, total agriculture \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\4_Data\3_Economic\SoGWRA-Charts_Economic.xlsx For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 6-4 National trends for increasing gross value of irrigated agricultural production (GVIAP) as available water supplies have increased for (a) fruits, (b) vegetables, (c) fruits and vegetables combined, and (d) total agriculture Source: (ABS, 2022) 6.4.3 Costs of enabling infrastructure A range of infrastructure would be required to support the development of a new irrigation scheme in the Southern Gulf catchments, both within the scheme itself and beyond. Any infrastructure that is not included in the initial water development contract but is required to enable the new water resources to be used effectively (and to achieve their anticipated benefits) will require construction after the contracted project is complete, often at public expense. The types of infrastructure addressed here are those that would not typically be included in a formal CBA or be built by the water infrastructure developer or farmers. Within the context of a large irrigation development, such enabling infrastructure can be considered ‘hard’ or ‘soft’, which can be broadly defined as follows: • Hard infrastructure refers to the physical assets necessary for a development to function. It can include water storage, roads, irrigation supply channels, energy, and processing infrastructure, such as sugar mills, cotton gins, abattoirs and feedlots. • Soft infrastructure refers to the specialised services required for maintaining the economic, health, cultural and social standards of a population. These are indirect costs of a development and are usually less obvious than hard infrastructure costs. They can include expenses that continue after the construction of a development has been completed. Soft infrastructure can include: – physical assets, such as community infrastructure (e.g. schools, hospitals, housing) – non-physical assets, such as institutions, supporting rules and regulations, compensation packages, and law enforcement and emergency services. New processing infrastructure and community infrastructure are particularly pertinent to large, remote, greenfield developments, and these costs to other providers of infrastructure can be substantial, even after a new irrigation scheme has been developed. For example, a review of the Ord-East Kimberley Development Plan (for expansion of the Ord irrigation system by ~15,000 ha) found additional costs of $114 million to the WA Government beyond the planned $220 million state investment in infrastructure already provided to directly support the expansion (Western Australian Auditor General, 2016). This section provides an indication of the additional public and private infrastructure required to support a new irrigation development (once the main water infrastructure and farms are built) and the costs of the additional investments required. The intention is to highlight potentially overlooked costs of infrastructure that is required to realise the benefits of development and population growth in a region, rather than to diminish the potential benefits. Costs of hard infrastructure Establishing new irrigated agriculture in the Southern Gulf catchments would involve the initial costs of land development, water infrastructure (which could include distribution and re- regulation or balancing of storages), and farm set-up (for equipment and facilities on each new farm). It may also involve costs associated with constructing processing facilities, extending electricity networks, and upgrading road transport. The costs of water storage and conveyance are provided in Chapter 5. Indicative costs for processing facilities are provided in Table 6-14, and indicative costs for roads and electricity infrastructure are provided in Table 6-15. Indicative costs for transporting goods to key markets are listed in Table 6-16. All table data are summaries of information provided in the companion technical report on agricultural viability and socio-economics (Webster et al., 2024). Table 6-14 Indicative costs of agricultural processing facilities ITEM CAPITAL COST OPERATING COST COMMENT Meatworks $35 million $340/head Operational capacity 100,000 head/y Cotton gin $32 million $1.1 million/y plus $24 to $35 per bale Operational capacity of 1,500 bales/day Operating costs depend on the scale of the gin, and the source of energy Sugar mill $409 million $34 million/y Operational capacity of 1000 t cane/h, 6-month crushing season Basic mill producing sugar only (no electricity or ethanol) Table 6-15 Indicative costs of road and electricity infrastructure ITEM CAPITAL COST COMMENT Roads Seal dirt road $0.31 to $2.4 million per km Upgrade and widen dirt road to sealed road New bridges and floodway $27.4 million Costs of bridges and floodways vary widely Electricity New generation capacity may also be required Transmission lines $0.34 to $1.57 million per km High-voltage lines deliver bulk flow of electricity from generators over long distances Distribution lines $0.22 to $0.49 million per km Lower-voltage lines distribute power from substations over shorter distances to end users Substation $1.3 to $12.2 million Transformers and switchgear connect transmission and distribution networks Table 6-16 Indicative road transport costs between the Southern Gulf catchments and key markets and ports The top section of the table gives trip costs from Mount Isa to key destinations. The bottom section gives distance- based costs of getting goods from within the catchment to Mount Isa (on unsealed roads) and approximate distance- based costs of getting goods from Mount Isa on sealed roads to other destinations (not specifically listed). DESTINATION TRANSPORT COST Unrefrigerated Refrigerated Cattle Transport costs from Mount Isa ($/t) Adelaide 404 480 364 Brisbane 295 337 266 Cairns 181 235 163 Darwin 242 293 218 Fremantle 748 793 673 Karumba 87 103 78 Melbourne 423 467 380 Port Hedland 505 612 455 Sydney 398 426 358 Townsville 143 161 129 Wyndham 285 344 257 Transport costs by distance ($ per t per km) Properties to Mount Isa 0.28 0.32 0.25 Mount Isa to key markets/ports 0.16 0.18 0.14 Costs of soft infrastructure The availability of community services and facilities would play an important role in attracting people to (or deterring them from) living in a new development in the Southern Gulf catchments. If local populations increase as a result of new irrigated developments, then the demand for public services would increase, and provision of those services would need to be anticipated and planned for. Indicative costs for constructing a variety of facilities that may be required for supporting population growth are listed in Table 6-17. Each 1000 people in Australia require 2.3 (in ‘Major cities’) to 4.0 (in ‘Remote and Very remote areas’) hospital beds, served by 16 full-time equivalent (FTE) hospital staff, and $3.5 million/year funding to maintain current mean national levels of hospital service (AIHW, 2023). Health care services in remote locations generally focus on providing primary care and some secondary care. More specialised tertiary services tend to be concentrated in referral hospitals, which are generally located in big cities but also serve the surrounding area. Primary schools tend to be smaller and more widespread than secondary schools, which are larger and more centralised. Table 6-17 Indicative costs of community facilities Costs are quoted for Darwin as a reference capital city for northern Australia. Costs in remote parts of northern Australia, including the Southern Gulf catchments, are estimated to be approximately 30% to 60% higher than those quoted for Darwin. School costs were estimated separately based on a number of locations across northern Australia. See the companion technical report on agricultural viability and socio-economics (Webster et al., 2024) for details. ITEM CAPITAL COST COMMENT Hospital $0.2 to $0.5 million per bed Higher-end costs include a major operating theatre and a larger hospital area per bed School $27,000 to $35,000 per student Secondary schools tend to be larger and more centralised than primary schools House (each) $585,000 to $850,000 Single- or double-storey house, 325 m2 Unit (each) $230,000 to $395,000 Residential unit (townhouse), 90 to 120 m2 Offices $2400 to $3450 per m2 1 to 3 storeys, outside central businesses district The demand for community services is growing, both from population increases in Australia and rising community expectations. New infrastructure would be built to service that demand, irrespective of any development in the Southern Gulf catchments. However, if new irrigation projects encourage people to live in the Southern Gulf catchments, this could then shift the locations at which some services would be delivered and the associated infrastructure built. The costs of delivering services and building infrastructure are generally higher in very remote locations like the Southern Gulf catchments. The net cost of any new infrastructure built to support development in the Southern Gulf catchments is the difference in the cost of shifting some infrastructure to this very remote location (rather than the full cost of the facilities (Table 6- 17) which would otherwise have been built elsewhere). 6.5 Regional-scale economic impact of irrigated development New irrigated development in the Southern Gulf catchments could provide economic benefits to the region in terms of both increased economic activity and jobs. The size of the total economic benefit experienced would depend on the scale of the development, the type of agriculture that was established, and how much spending from the increased economic activities occurred within the region. Regional economic impacts are an important consideration for evaluating potential new water development projects. It was estimated that each million dollars spent on construction within the Southern Gulf catchments would generate an additional $1.06 to $1.09 million of indirect benefits ($2.06 to $2.18 million total regional benefits, including the direct benefit of each million dollars spent on construction). It was estimated that each million dollars of direct benefit from new agricultural activity would generate an additional $0.46 to $1.82 million in regional economic activity (depending on the particular agricultural industry). The full, catchment-wide impact of the economic stimulus provided by an irrigated agriculture or aquaculture development project extends far beyond the impact on those businesses and workers directly involved in either the short term (construction phase) or the longer term (operational phase). Businesses directly benefiting from the project would need to increase their purchases of the raw materials and intermediate products used by their growing outputs. Should any of these purchases be made within the surrounding region, this would provide a stimulus to those businesses from which they purchase, contributing to further economic growth within the region. Furthermore, household incomes would increase as a result of the employment of local residents as a consequence of the direct and/or production-induced business stimuli. As a proportion of this additional household income would be spent in the region, economic activity within the region would be further stimulated. Accordingly, the larger the initial amount of money spent within the region, and the larger the proportion of that money re-spent locally, the greater the overall benefits that would accrue to the region. The size of the impact on the local regional economy can be quantified by regional economic multipliers (derived from I–O tables that summarise expenditure flows between industry sectors and households within the region): a larger multiplier indicates larger regional benefits. These multipliers can be used to estimate the value of increased regional economic activity likely to flow from a stimulus to particular industries, focusing on construction in the short term and various types of agriculture in the longer term. It is also possible to estimate the increase in household incomes in the region, and then estimate the approximate number of jobs represented by the increased economic activity, including both those directly related to the increase in agriculture and those generated indirectly within other industries in the region. Not all expenditure generated by a large-scale development will occur within the local region. The greater the leakage (i.e. the amount of direct and indirect expenditure occurring outside the region), the smaller the resulting economic benefit enjoyed by the region. Conversely, the greater the retention of the initial expenditure and subsequent indirect expenditure within the region, the greater the economic benefit and the number of jobs created within the local region. However, a booming local economy can also bring with it a number of issues that can place upward pressure on prices (including materials, houses and wages) in the region, negating some of the positive impacts of the development. If some of the unemployed or underemployed people within the Southern Gulf catchments could be engaged as workers during the construction or operational phases of the development, this could reduce pressure on local wages and reduce the leakage resulting from the use of fly-in fly-out (FIFO) or drive-in drive-out (DIDO) workers, retaining more of the benefit from the project within the local region. However, the current low unemployment rate within the Southern Gulf catchments (Chapter 3) suggests there may be difficulties in sourcing local workers from within the region. The overall regional benefit created by a particular development depends on both the one-off benefits from the construction phase and the ongoing annual benefits from the operational phase. The benefits from the operational phase may take a number of years to reach the expected level, as new and existing agricultural enterprises learn and adapt to making full use of the new opportunities presented by the development. It is important to note that the results presented here are based on illustrative scenarios incorporating broad assumptions, are derived from an I–O model developed for an I–O region that is much larger than the Southern Gulf catchments study area, and are subject to the limitations of the method. 6.5.1 Estimating the size of regional economic benefits To develop regional multipliers for the Southern Gulf catchments, it was necessary to use the available information and models for the Queensland section of the Southern Gulf catchments I–O region. The regional I–O table developed by the Office of the Government Statistician of the Queensland Government provided coverage for the north-west region of Queensland (Office of the Government Statistician, Queensland Government, 2004 (Figure 6-5)). These results can be considered broadly applicable to the NT component of the Assessment area, noting the caveats for the I–O model below. For more details, see the companion technical report on agricultural viability and socio-economics (Webster et al., 2024). Data are presented to show how the economic circumstances of the Southern Gulf catchments compare with those of the I–O region (Table 6-18). The I–O model covers a much wider geographic scale than the Southern Gulf catchments (307,082 km2 for north-west Queensland, compared with 108,097 km2 for the Southern Gulf catchments). However, the regionally important city of Mount Isa falls within both regions. Both geographic scale and degree of urbanisation can affect the relative complexity of the economic structures in each region. There are wide variations in the size of the multipliers for the various industries within the Queensland and north-west Queensland I–O regions. Industries with larger local regional multipliers would be expected to benefit more from development within the I–O region. For example, agricultural industries generated smaller multipliers than mining and construction for both I–O models. However, a simple comparison of I–O multipliers can be misleading when considering the different benefits from regional investment, because some impacts provide a short-term, one-off benefit (e.g. the construction phase of a new irrigation development) while others provide a sustained stream of benefits over the longer term (e.g. the production phase of a new irrigation scheme). A rigorous comparison between specific regional investment options would require NPVs of the full cost and benefit streams to be calculated. Extent of regional input models map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\5_Agriculture_economics\4_S_Gulf\1_GIS\1_Map_docs\Se-S-506_Map_Australia_and_economic_regions_v1.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 6-5 Queensland’s north-west region used in the input–output (I–O) analyses relative to the Southern Gulf catchments Assessment area Table 6-18 Key 2021 data comparing the Southern Gulf catchments with the related I–O analysis region SOUTHERN GULF CATCHMENTS† NORTH-WEST QUEENSLAND I–O REGION‡ Land area (km2) 108,097 307,081.5 Population 22,493 30,377 Percentage male 51.74 52.11 Percentage Indigenous 27.28 27.03 Median age 31 32 Median household income $109,429 $104,312 Contribution of agriculture, forestry and fishing to employment in the region 2.67% 8.56% Major industries of employment – top three industries in region (by % of employment 2021) Largest employer in region Mining Mining 2nd largest employer in region Health care and social assistance Health care and social assistance 3rd largest employer in region Education and training Public administration and safety Gross value of total agriculture in region§ $243 million $945 million † Statistics for the Southern Gulf catchments region (ABS, 2021) have been estimated using the weighted mean of ABS 2021 census data obtained by SA2 statistical region, with weighting based on the proportion of relevant ABS SA2 statistical regions falling within the catchments region. ‡ ABS 2021 census data (ABS, 2021). § ABS value of agricultural commodities produced 2020–21 by region, report VACPDCASGS202021 (ABS, 2022). I–O = input–output. 6.5.2 Indirect benefits during the construction phase of a development Initially, building new infrastructure (on-farm and off-farm development, including construction of related supporting infrastructure, such as roads, schools and hospitals) comes at a cost. But the additional expenditure within a region (which puts additional cash into people’s and businesses’ pockets) would increase regional economic activity. This creates a fairly short-term economic benefit to the region during the construction phase, provided that at least some of the expenditure occurs within the region and is not all lost from the region due to leakage. The regional impacts of the construction phase of potential developments were estimated using a scenario approach for the scales of development. The analyses modelled regional impacts for five different indicative sizes of developments in the Southern Gulf catchments. Total capital costs, including costs of labour and materials required by the project, ranged from $250 million to $4 billion. The smallest scale of development in Table 6-19, with a capital cost of $250 million, broadly represents approximately 20 new farm developments with their own on-farm water sources enabling approximately 10,000 ha of irrigation for horticulture and broadacre farming (based on costing information from the companion technical report on agricultural viability and socio-economics (Webster et al., 2024)). The second-smallest scenario, with a $500 million capital cost, could represent a similar development to the first but with 20,000 ha of new irrigated farmland; this level of investment could also include a new processing facility (such as a cotton gin) required by (and supported by) this scale of agricultural development. Alternatively, the $500 million development could represent a large off-farm water infrastructure development (e.g. see Table 6-2) and related farm establishment costs. The larger scales of development, at $1 billion or $2 billion, shown in Table 6-19, indicate outcomes from combining potential developments in various ways (such as one large off-farm dam and multiple on-farm water sources). They also include investment in indirect supporting infrastructure across the region, such as roads, electricity and community infrastructure (see indicative costs in Section 6.4.3). Table 6-19 Regional economic impact estimated by I–O analysis for the total construction phase of an irrigated agricultural development based on estimated Type ll multipliers determined from the north-west Queensland I–O models Estimates represent an upper bound, because some assumptions of I–O analysis are violated in the case of such a large public investment in a region where existing irrigated agricultural activity is so low. Leakage to other regions and other countries is accounted for by reducing the proportion of expenditure (and benefits) within the I–O region. I–O = input–output. DEVELOPMENT CAPITAL COST ($ billion) TOTAL REGIONAL ECONOMIC ACTIVITY WITHIN I–O REGION AS A RESULT OF THE CAPITAL COST OF THE DEVELOPMENT ($ billion) Southern Gulf catchments based on north-west Queensland I–O model Proportion of total scheme-scale capital cost made locally within the I–O region 65% 50% 35% 0.250 0.26 0.20 0.14 0.500 0.53 0.41 0.28 1.000 1.06 0.81 0.57 2.000 2.11 1.63 1.14 4.000 4.23 3.25 2.28 The proportion of expenditure during the construction phase that would be spent within the region depends on the types of costs, including labour, materials and equipment. It is likely that wages would be paid to workers sourced both from within the region and from elsewhere. The likely proportion of labour costs for each source of workers would depend on the availability of appropriately skilled labour within the region. For example, a highly populated region (more than 100,000 people) with a high unemployment rate (more than 10%) and a skilled labour force is likely to be able to supply a large proportion of the workers required from within the region. However, a sparsely populated region like the Southern Gulf catchments is more likely to need to attract many workers from outside the region, either on a FIFO or DIDO basis or by encouraging migration to the region. Similarly, some regions may be better able to supply a large proportion of the required materials and equipment from within the region, whereas construction projects in other locations may not be able to source what they need locally and instead need to import a significant proportion into the region from elsewhere. The low representation of the required supplying industries in the Southern Gulf catchments means that most construction supplies would be likely to be sourced from other parts of Australia (and internationally). A review of five large dam projects across the country showed that the proportions of local construction expenditure sourced within a region (as opposed to being imported, with no impact on the local regional economy) varied significantly. Thus, the analyses considered three levels for the proportion spent locally: 65% (i.e. low leakage), and 50%, and 35% spent locally (i.e. high leakage). However, note that leakage might be higher (i.e. <35% spent locally) for a very remote region like the Southern Gulf catchments. In cases of high leakage, the knock-on benefits would instead occur in the regions supplying the goods and services (a large proportion of which are likely to be elsewhere in the state). Table 6-19 shows estimates of the regional economic benefit for the construction phase of a new development for five scales of scheme capital cost ($0.25 billion to $4 billion) and the three levels of leakage described above. Clearly, the proportion of scheme construction costs spent within the region has a significant impact on the size of the regional economic benefit experienced. If a large proportion of the initial expenditure in the Southern Gulf catchments leaked outside, the benefit of the initial construction investment would be less concentrated in the local Southern Gulf catchments’ economy and would spread to those locations supplying the goods and services. 6.5.3 Indirect benefits during the operational phase of a development Regional impacts of irrigation development are presented for scenarios using four indicative scales of increase in GVAP ($25, $50, $100 and $200 million per year, indicative of potential outcomes). At the low end ($25 million/year), this could represent 10,000 ha of new plantation timber, while the high end ($200 million/year) could represent 10,000 ha of mixed broadacre cropping and horticulture (based on farm financial estimates for the various crops presented in Chapter 4), with other crop options falling in between. Estimated regional impacts are shown as the total increased economic activity (Table 6-20) and the associated estimated increases in incomes and employment (Table 6-21) for each category of agricultural activity (beef cattle, agriculture excluding beef cattle, and aquaculture, forestry and fishing. As can be seen from the economic impacts (Table 6-20), an irrigation scheme that increases the output of the beef cattle industry could have a larger impact on total regional economic activity than a scheme that promotes agriculture excluding beef cattle, while the smallest regional economic benefit would derive from a development focused on aquaculture, forestry and fishing. Table 6-20 Estimated regional economic impact per year in the Southern Gulf catchments resulting from four scales of direct increase in agricultural output (rows) for the different categories of agricultural activity (columns) using the I–O model for north-west Queensland Increases in agricultural output are assumed to be net of the annualised value of contributions towards the construction costs. Estimates are based on Type ll multipliers determined from the I–O model for each year of agricultural production. Estimates represent an upper bound, because some assumptions of I–O analysis are violated in the case of such a large public investment in a region where existing agricultural activity is so low. Leakage to other regions and other countries is accounted for by reducing the proportion of expenditure (and benefits) within the region. DIRECT INCREASE IN AGRICULTURAL OUTPUT PER YEAR NET OF CONTRIBUTION TO CONSTRUCTION COSTS ($ million) TOTAL ANNUAL VALUE OF INCREASED ECONOMIC ACTIVITY IN I–O REGION – DIRECT, PRODUCTION-INDUCED AND CONSUMPTION-INDUCED ($ million) Southern Gulf catchments based on north-west Queensland I–O model Type of agricultural development Beef cattle Agriculture excluding beef cattle Aquaculture, forestry and fishing 25 46 38 33 50 91 76 66 100 182 152 132 200 364 303 265 The results for employment (Table 6-21) are closely related to those for impacts on regional economic activity, but the two measures do reveal some differences. Additional FTE jobs arising in the region may require additional community infrastructure (e.g. schools, health services) if workers move to fill these jobs from other parts of the country, resulting in population growth. However, additional infrastructure would not be necessary should these additional jobs be filled by currently unemployed or underemployed local people. Estimates of the expected increases in incomes were divided between Indigenous and non-Indigenous households, using methods outlined by Jarvis et al. (2018), with most increases expected to flow to non-Indigenous households (Table 6-21). For example, if new irrigation development in the Southern Gulf catchments directly enabled an extra $100 million of cropping output per year, the region could benefit from an extra $152 million of economic activity recurring annually (Table 6-20) and generate approximately 294 new FTE ongoing jobs, depending on the type of agriculture (Table 6-21). Table 6-21 Estimated impact on annual household incomes and full-time equivalent (FTE) jobs within the Southern Gulf catchments resulting from four scales of direct increase in agricultural output (rows) for the various categories of agricultural activity (columns) Increases in agricultural output are assumed to be net of the annualised value of contributions towards the construction costs. Estimates are based on Type ll multipliers determined from two independent I–O models for each year of agricultural production. Estimates represent an upper bound, because some assumptions of I–O analysis are violated in the case of such a large public investment in a region where existing agricultural activity is so low. Leakage to other regions and other countries is accounted for by reducing the proportion of expenditure (and benefits) within the region. DIRECT INCREASE IN AGRICULTURAL OUTPUT PER YEAR NET OF ANY CONTRIBUTION TO CONSTRUCTION COSTS ($ million) TOTAL ANNUAL VALUE OF INCREASED ECONOMIC ACTIVITY IN I–O REGION – DIRECT, PRODUCTION-INDUCED AND CONSUMPTION-INDUCED ($ million or FTE) Southern Gulf catchments based on north-west Queensland I–O model Type of agricultural development Beef cattle Agriculture excluding beef cattle Aquaculture, forestry and fishing Additional incomes expected to flow to Indigenous households from development ($ million) 25 0.2 0.1 0.1 50 0.4 0.1 0.1 100 0.8 0.3 0.3 200 1.6 0.6 0.6 Additional incomes expected to flow to non-Indigenous households from development ($ million) 25 3.5 5.0 2.9 50 7.0 10.0 5.9 100 14.0 20.1 11.8 200 28.0 40.2 23.5 Additional jobs estimated to be created (FTE) 25 53 73 43 50 107 147 87 100 213 294 174 200 427 588 347 6.6 References ABS (2021) Water account, Australia, 2019–20 financial year. Australian Bureau of Statistics, Canberra. Viewed 19 December 2022, Hyperlink to: WAter account, Australia, 2019-20 financial year . ABS (2022) Value of agricultural commodities produced, Australia 2021–22. Australian Bureau of Statistics, Canberra. Viewed 19 December 2022, Hyperlink to: Agricultural commodities, Australia . AIHW (2023) Australia’s hospitals at a glance: web report. Australian Institute of Health and Welfare, Canberra. Viewed 1 March 2023, Hyperlink to: Australia's hospitals at a glance: web report . Ansar A, Flyvbjerg B, Budzier A and Lunn D (2014) Should we build more large dams? The actual costs of hydropower megaproject development. Energy Policy 69, 43–56. Hyperlink to: Should we build more large dams? The actual costs of hydropower megaproject development . Ash A, Gleeson T, Cui H, Hall M, Heyhoe E, Higgins A, Hopwood G, MacLeod N, Paini D, Pant H, Poulton P, Prestwidge D, Webster T and Wilson P (2014) Northern Australia: food and fibre supply chains study project report. CSIRO and ABARES, Australia. Viewed 13 September 2024. https://www.csiro.au/en/research/natural-environment/land/food-and-fibre CRCNA (2020). Northern Australian broadacre cropping situational analysis. ST Strategic Services and Pivotal Point Strategic Directions (Issue July). Cooperative Research Centre for Developing Northern Australia, Townsville. Viewed 13 September 2024. https://crcna.com.au/wp-content/uploads/2024/05/NA-Cropping-situational-August.pdf Devlin K (2024) Conceptual arrangements and costings of hypothetical irrigation developments in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Flyvbjerg B, Holm MS and Buhl S (2002) Underestimating costs in public works projects: error or lie? Journal of the American Planning Association 68(3), 279–295. Hyperlink to: Underestimating costs in public works projects: error or lie? Infrastructure Australia (2021a) Post completion review. Stage 4 of the Assessment Framework. Infrastructure Australia, Canberra. Viewed 1 March 2023, Hyperlink to: Post completion review. Stage 4 of the Assessment Framework . Infrastructure Australia (2021b) Guide to economic appraisal. Technical guide of the Assessment Framework. Infrastructure Australia, Canberra. Viewed 1 March 2023, Hyperlink to: Guide to economic appraisal . Jarvis D, Stoeckl N, Hill R and Pert P (2018) Indigenous land and sea management programs: can they promote regional development and help ‘close the (income) gap’? Australian Journal of Social Issues 53(3), 283–303. Hyperlink to: Indigenous land and sea management programs: can they promote regional development and help ‘close the (income) gap’? NWGA (National Water Grid Authority) (2022) National Water Grid investment framework. Australian Government Department of Climate Change, Energy, the Environment and Water, Canberra. Viewed 1 March 2023, Hyperlink to: National Water Grid Investment Framework . NWGA (National Water Grid Authority) (2023) Project administration manual: National Water Grid Fund. Australian Government Department of Climate Change, Energy, the Environment and Water, Canberra. Viewed 1 March 2023, https://www.nationalwatergrid.gov.au/framework. Odeck J and Skjeseth T (1995) Assessing Norwegian toll roads. Transportation Quarterly 49(2), 89– 98. Viewed 13 September 2024. Hyperlink to: Assessing Norwegian toll roads Office of the Government Statistician Queensland Government (2004) Queensland regional input– output tables 1996–97 34 industries. Queensland Government, Brisbane. Petheram C, and McMahon TA (2019) Dams, dam costs and damnable cost overruns. Journal of Hydrology X 3, 100026. Hyperlink to: Dams, dam costs and damnable cost overruns . Stokes C and Jarvis D (2021) Economic assessment. In: Petheram C, Read A, Hughes J, Marvanek S, Stokes C, Kim S, Philip S, Peake A, Podger G, Devlin K, Hayward J, Bartley R, Vanderbyl T, Wilson P, Pena Arancibia J, Stratford D, Watson I, Austin J, Yang A, Barber M, Ibrahimi T, Rogers L, Kuhnert P, Wang B, Potter N, Baynes F, Ng S, Cousins A, Jarvis D and Chilcott C. An assessment of contemporary variations of the Bradfield Scheme. A technical report to the National Water Grid Authority from the Bradfield Scheme Assessment. CSIRO, Canberra. Viewed 13 September 2024. https://publications.csiro.au/publications/publication/PIcsiro:EP2021-2556 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 and Poulton P (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. Viewed 13 September 2024. https://publications.csiro.au/publications/publication/PIcsiro:EP2024-1414 Vanderbyl T (2021) Southern Gulf: Queensland water plans and settings. A report from the CSIRO Southern Gulf Water Resource Assessment to the Government of Australia. CSIRO, Australia. Wachs M (1990) Ethics and advocacy in forecasting for public policy. Business and Professional Ethics Journal 9(1), 141–157. Hyperlink to: Ethics and advocacy in forecasting for public policy Webster A, Jarvis D, Jalilov S, Philip S, Oliver Y, Watson I, Rhebergen T, Bruce C, Prestwidge D, McFallan S, Curnock M and Stokes C (2024) Financial and socio-economic viability of irrigated agricultural development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Western Australian Auditor General (2016) Ord-East Kimberley Development. Report 20: September 2016. Office of the Auditor General Western Australia, Perth. Viewed 15 December 2022, Hyperlink to: Ord-East Kimberley Development . World Commission on Dams (2000) Dams and development. A new framework for decision making. Retrieved from Earthscan Publications Ltd, UK. Viewed 20 December 2022, Hyperlink to: Dams and development. A new framework for decision making . Yang A, Petheram C, Marvanek S, Baynes F, Rogers L, Ponce Reyes R, Zund P, Seo L, Hughes J, Gibbs M, Wilson PR, Philip S and Barber M (2024) Assessment of surface water storage options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO Australia 7 Ecological, biosecurity, off-site and irrigation- induced salinity risks Rob Kenyon, Rocio Ponce Reyes, Danial Stratford, Linda Merrin, Lynn Seo, Matt Gibbs, Simon Linke, Justin Hughes, Heather McGuinness, Ian Watson, Cuan Petheram, Peter Zund, John Virtue, Katie Motson Chapter 7 discusses a range of potential risks to be considered before establishing a greenfield agriculture or aquaculture development. These include ecological implications of altered flow regimes, biosecurity considerations, irrigation drainage and aquaculture discharge water, and irrigation-induced salinity. The key components and concepts of Chapter 7 are shown in Figure 7-1. Figure 7-1 Schematic diagram of the components where key risks can manifest when considering the establishment of a greenfield irrigation or aquaculture development For more information on this figure please contact CSIRO on enquiries@csiro.au Numbers in blue refer to sections in this report. 7.1 Summary This chapter provides information on the ecological, biosecurity, off-site and downstream impacts and irrigation-induced salinity risks to the catchments of the Southern Gulf rivers – that is, Settlement Creek, Gregory–Nicholson River and Leichhardt River, the Morning Inlet catchments and the Wellesley island groups1 – from greenfield agriculture or aquaculture development. It is principally concerned with the risks from these developments to the broader environment but also considers biosecurity risks to the enterprises themselves. 1 Only those islands greater than 1000 ha are mapped The ecological impacts of vegetation clearing associated with irrigated agriculture are not explicitly examined in the Assessment as it is considered of secondary concern to potential impacts on water dependent ecological assets. This is because irrigated agriculture occupies a very small proportion of the landscape (typically less than 0.5%) but can result in a disproportionately high degree of regulation of river flow. Consequently, the Assessment placed greatest effort in understanding the potential ecological impacts of changes in streamflow on aquatic dependent ecosystems. 7.1.1 Key findings Ecological implications of altered flow regimes The freshwater, terrestrial and near-shore marine zones of the Southern Gulf catchments contain important and diverse species, habitats, industries and ecosystem functions supported by the patterns, volumes and quality of river flows. Although irrigated agriculture only occupies a small percentage of the landscape, changes in the flow regime can have profound effects on flow- dependent flora and fauna, their habitats and ecosystem service provision. These effects may extend considerable distances onto the floodplain and downstream, including into the marine environment. Hypothetical future scenarios for water harvesting, instream dams and groundwater development produced a range of water volumes and patterns of flow with a variety of impacts on ecology. The findings are summarised below: • The level of impact resulting from water resource development was highly dependent on the type of development, the extraction volume and the mitigation measures implemented. For most assets, water extraction had negligible to minor impacts on key aspects of river flows. • Large instream dams had a greater mean impact on surface-flow-dependent ecology averaged across the Southern Gulf catchments than did water harvesting. Large instream dams resulted in significantly larger local impacts on flow ecology in reaches below the dam wall than did water harvesting. A dam on the perennial Gregory River had major impacts on downstream flows that were important to several species or groups, while a dam on Gunpowder Creek, a major tributary of the Leichhardt River, had negligible to minor impacts on important flows for most assets. • Mitigation measures can reduce the negative impacts of flow modification on catchment ecology and key biota by maintaining aspects of flow that have been identified as critical to ecosystem service provision. Mitigation strategies, such as transparent flows (environmental flows that pass the dam wall) from dams that mimic historical flow patterns, guaranteed annual end-of-system flow volume prior to water extraction, and river flow pump-initiation thresholds below which water extraction cannot occur have been identified as effective in reducing risks. Water harvesting outcomes were sensitive to the volume of extraction. Mitigation measures also changed ecology outcomes across the Southern Gulf catchments, with mitigation reducing the effect on flows from moderate to minor, or minor to negligible. Harvesting 50 to 300 GL of water resulted in negligible to minor changes to most asset means across the catchments with impacts often accumulating downstream past multiple extraction points. Threadfin, prawn species, mud crabs and mullet were among the ecological assets most affected by flow change for water harvesting (moderate change). For low water harvest extraction volumes, suitable levels of end-of- system flow requirements, commence-to-pump thresholds and pump rates improved mean outcomes across ecological assets and negligible change at catchment scales. Mitigation strategies demonstrate the importance of protecting minimum flows and annual ‘first flows’ for many of the ecological assets and that deployment of mitigation has considerable potential to reduce impacts on water-dependent ecosystems. For instream dams, the location of the dam in the catchment matters as there is potential for extreme risks of local impacts. Improved outcomes were associated with maintaining attributes of the natural flow regime via the provision of transparent flows. A single dam on either Gunpowder Creek or the Gregory River resulted in negligible to minor (respectively) mean change to assets flows at the catchment scale, but local impacts were often with considerably higher and reached extreme levels. Mangroves, mullet, cryptic wader birds, prawns and mud crabs were among the ecological assets most affected by instream dams. Providing transparent flows (environmental low-flows) improved flow regimes for ecological service provision at both local and catchment scales: mean outcomes for mangroves could be improved from major to moderate mean change, and outcomes for catfish, cryptic wader birds and inchannel waterhole assets from moderate to minor mean change across the catchment. Beyond flow, other impacts and considerations are also important. At catchment scales, the direct impacts of irrigation on the terrestrial environment are typically small. However, indirect impacts, such as weeds, pests and landscape fragmentation, particularly to riparian zones, and changes in fire regimes may be considerable. Changes in water quality may also affect ecology but are not considered in the quantitative analysis. The combined changes of a potentially drier future climate and hypothetical water resource development produced greater impacts than did each factor on its own. Biosecurity considerations Biosecurity is the prevention and management of pests, weeds and diseases, both terrestrial and aquatic, to limit their economic, environmental, social and cultural impacts. Economic impacts include reduced crop yield and product quality, interference with farming operations, loss of market access and costs of implementing control measures. Environmental impacts include loss of biodiversity and changes to ecosystem processes such as fire regimes. Social and cultural impacts of pests, weeds and diseases include diminished value of areas for recreational or traditional uses. Despite its relative isolation, there are many human-mediated and natural pathways by which pests, weeds and diseases can spread to and within the Southern Gulf catchments. New pests, weeds and diseases may spread from adjacent regions, other parts of Australia or even neighbouring countries. Biosecurity is a shared responsibility that requires governments, industries and the community to each take steps to limit the introduction and spread, detect and respond to incursions, and manage the impacts of key biological threats. A further complication for the cross-border catchment is being aware of biosecurity requirements for both the NT and Queensland, including the latter’s general biosecurity obligation. A range of current and potential pests, weeds and diseases could affect irrigated cropping in the Southern Gulf catchment. These include fall armyworm (Spodoptera frugiperda), which consumes C4 grass crops; cucumber green mottle mosaic virus, which infects a wide range of cucurbit crops; incursion risks from overseas such as citrus canker (Xanthomonas citri subsp. citri) and exotic fruit flies; and parthenium weed (Parthenium hysterophorus), which is a crop competitor, seed contaminant and allergen. Farm biosecurity planning to identify, prevent, detect and manage key pest, weed and disease threats is fundamental to a successful enterprise. Such planning includes following government and industry best practice regarding movement of plants, plant products and machinery, control of declared species, pesticide use, farm stewardship and market access requirements. Preventive biosecurity practices are crucial in aquaculture facilities as diseases can be difficult to eliminate. There are many diseases of production concern, whether overseas, having entered Australia (e.g. white spot syndrome virus of crustaceans) or naturally occurring in Australian ecosystems. Aquaculture biosecurity planning needs to consider hygiene actions needed for key pathways of disease entry, early detection and diagnosis, quarantining and treatment. Invasive species, whether pest, weed or disease, are commonly characterised as occurring across multiple land uses in a landscape. Their impacts will vary between land uses, but their coordinated control requires action across all tenures. Various high-impact weeds listed as declared (NT) or restricted (Queensland) are present in, or threaten to invade, the Southern Gulf catchments, including aquatic plants, grasses, shrubs and trees. There are also pest vertebrates (e.g. large feral herbivores, exotic fish), pest invertebrates (e.g. exotic ants) and plant pathogens (e.g. Phytophthora) that can cause disease. NT and Queensland government legal requirements to prevent or control invasive pests, weeds and diseases need to be followed. Regional and local irrigation and industry infrastructure development, including road networks, should include prevention and management of invasive species in their environmental planning processes. Choice of crops and aquaculture species should also consider the invasive risk and any management required to prevent their spread into the environment. Off-site and downstream impacts Agriculture can affect the water quality of downstream freshwater, estuarine and marine ecosystems. The principal pollutants from agriculture are nitrogen, phosphorus, total suspended solids, herbicides and pesticides. Most of the science in northern Australia concerned with the downstream impacts of agricultural development has been undertaken in the eastern-flowing rivers that flow into the Great Barrier Reef lagoon. Comparatively little research on the topic has been done in the rest of northern Australia. Degraded water quality can cause a loss of aquatic habitat, biodiversity and ecosystem services. Increased nitrogen and phosphorus can cause plankton blooms and weed infestation, increase hypoxia (low oxygen levels) and result in fish deaths. Pesticides, used to increase agricultural productivity, can harm downstream aquatic ecosystems, flora and fauna. Fertiliser nutrients and pesticides can enter surface water bodies and groundwater via infiltration, leaching, and runoff from rainfall events and irrigation. Losses via runoff or deep drainage are the main pathways by which agricultural pollutants enter water bodies. Management of irrigation or agricultural drainage waters is a key consideration when evaluating and developing new irrigation systems, and it should be given careful consideration during the planning and design process. Seasonal hydrology, particularly ‘first-flush’ events following irrigation or significant rainfall, plays a critical role in determining water quality. Studies have shown that pesticide concentrations in runoff are highest following initial irrigation events but decrease in subsequent events. Similarly, nitrogen concentrations in runoff are often higher following early-season rainfall, when crops have not yet fully absorbed available nitrogen, leading to increased transport in runoff. Minimising drainage water by using best-practice irrigation design and management should be a priority in any new irrigation development in northern Australia. While elevated contaminants and water quality parameters can harm the environment and human health, there are several processes by which aquatic ecosystems can partially process contaminants and regulate water quality. Denitrification is a natural process that can remove and reduce nitrogen concentrations within a water body. Phosphorus, however, does not have a microbial reduction process equivalent to denitrification. Instead, if it is not temporarily taken up by plants, phosphorus can be adsorbed onto the surface of inorganic and organic particles and stored in the soil, or deposited in the sediments of water bodies, such as wetlands. Aquaculture can be affected by poor water quality and can also contribute to poor water quality unless aquaculture operations are well managed. Aquaculture species are particularly vulnerable to some of the insecticides and other chemicals used in agricultural, horticultural and mining sectors, and in industrial and domestic settings. Aquaculture management is designed to discharge water that contains low amounts of nutrients and other contaminants. The aim is for discharge waters to have similar physiochemical parameters to the source water. Because aquaculture management in northern Australia has largely been developed to ensure that the waters of the Great Barrier Reef lagoon do not receive excessive contaminants, there is considerable experience in operating aquaculture enterprises in northern Australia under world’s best practice. Irrigation-induced salinity Naturally occurring areas of salinity, or ‘primary salinity’, occur in the landscape, and their ecosystems are adapted to the saline conditions. Any change to landscape hydrology, including clearing and irrigation, can mobilise salts, resulting in environmental degradation in the form of ‘secondary salinity’. This occurs where rising groundwater mobilises salts in the soil and substrate materials, moving the salts into the plant root zone and/or discharging salts on lower slopes, in drainage depressions or in nearby streams. Soil knowledge and best-practice management of irrigation timing and application rates can reduce the risk of irrigation-induced salinity. Note that the material in this chapter provides general information regarding soils suitable for irrigation development. The risk of secondary salinisation at a specific location in the Southern Gulf catchments can only be properly assessed by undertaking detailed field investigations at a local scale. Existing salinity is not prominent in the Assessment area apart from the salt plains along the coast, which are not considered for irrigation development. However, the cracking clay soils on the Armraynald Plain, particularly the black soils along the Gregory River backplain, have subsoils that are high in salt and susceptible to irrigation-induced secondary salinity. These cracking clay soils can be successfully irrigated if they can be managed to prevent waterlogging and the mobilisation of salts in the profile. The clay soils (soil generic group (SGG) 9, see Section 2.3) on the Barkly Tableland have low subsoil salt levels. Where they are underlain by porous limestone and dolomite, a build-up of salts due to irrigation is not expected. The sandy, loamy and sand or loam over friable brown, yellow and grey clay soils on the Doomadgee Plain also have negligible salts within the soil profile. However, due to other risk factors, care would need to be exercised when clearing the silver box, bloodwood and broad-leaf paperbark savanna landscapes for rainfed or irrigated cropping. Groundwater aquifers contained by underlying ferricrete, the likelihood of soils having variable depths, and the very gently undulating plain make it difficult to manage irrigation water discharge on lower slopes and in drainage depressions, causing salts to accumulate in these areas in the long term. In places where these soils are shallow, it would be necessary to monitor the depth of watertables and manage irrigation rates accordingly. In addition, over-irrigation is likely to have off-site impacts in the long term, as the lateral flow of water can ‘wick’ from the lower slopes in these landscapes to form scalds. From these scalds, salts can potentially be mobilised towards nearby streams. 7.2 Introduction Water and irrigation development can result in complex and, in some cases, unpredictable changes to the surrounding environment and communities. For instance, before the construction of the Burdekin Falls Dam, the Burdekin Project Committee (1977) and Burdekin Project Ecological Study (Fleming et al., 1981) concluded that the dam would improve water quality and clarity in the lower river and that para grass (Brachiaria mutica), an invasive weed from Africa that was then present at relatively low levels, could become a useful ecological element as a result of increased water delivery to the floodplain. However, the Burdekin Falls Dam has remained persistently turbid since construction in 1987, greatly altering the water quality and ecological processes of the river below the dam and the many streams and wetlands into which that water is pumped on the floodplain (Burrows and Butler, 2007). Para grass and, more recently, hymenachne (Hymenachne amplexicaulis), an ecologically similar plant from South America, have become serious weeds of the floodplain wetlands, rendering innumerable wetlands unviable as habitat for most aquatic biota that formerly occurred there (Tait and Perna, 2001; Perna, 2003, 2004). Thus, there are limitations to the advice that can be provided in the absence of specific development proposals, so this section provides general advice on the considerations or externalities most strongly affected by water resource and irrigation developments. It is not possible to discuss every potential change that could occur. In particular the ecological impacts of vegetation clearing associated with irrigated agriculture are not explicitly examined as it is considered of secondary concern to potential impacts on water dependent ecological assets. This is because irrigated agriculture occupies a very small proportion of the landscape (typically less than 0.5%) but can result in a disproportionately high degree of regulation of river flow. Consequently, the Assessment placed greatest effort in understanding the potential ecological impacts of changes in streamflow on aquatic dependent ecosystems. It is noted, however, that areas of high agricultural potential may also be highly valued with respect to biodiversity conservation (Kutt et al., 2009). For these and other reasons the northern jurisdictions have formal processes in place for the approval (or not) of clearing native vegetation. Clearing approvals are only provided by the jurisdictions where they consider the ecological impact to be minimal given the extent and protection of vegetation type in the region (e.g. Queensland Government, 2023a). The remainder of the chapter is structured as follows: • Section 7.3 Ecological implications of altered flow regimes examines how river regulation affects inland and freshwater assets in the Southern Gulf catchments and marine assets in the near- shore marine environment. It also examines how the impacts can be mitigated. • Section 7.4 Biosecurity considerations discusses the risks presented to an irrigation development by disease, pests and weeds, and the risks that new agriculture or aquaculture enterprise in the Southern Gulf catchments may present to the wider industry and broader catchment. • Section 7.5 Off-site and downstream impacts considers how agriculture can affect the water quality of downstream freshwater, estuarine and marine ecosystems. • Section 7.6 Irrigation-induced salinity briefly discusses the risk of irrigation-induced salinity to irrigation development and to the downstream environment in the Southern Gulf catchments. Other externalities associated with water resource and irrigation development discussed elsewhere in this report include the direct impacts of the development of a large dam and reservoir on: • Indigenous cultural heritage (Section 3.4) • water quality (Section 7.5) • the movement of aquatic species (Section 5.4) • terrestrial ecosystems within the reservoir inundation area (Section 5.4). These externalities are rarely factored into the true costs of water resource or irrigation development. Even in parts of southern Australia where data are more abundant, it is very difficult to express these costs in monetary terms, as perceived changes are strongly driven by values, which can vary considerably within and between communities and fluctuate over time. Therefore, the material in this chapter is presented as a stand-alone analysis to help inform conversations and decisions between communities and government. Note that this chapter primarily focuses on key risks resulting from irrigated agriculture and aquaculture, although the section on biosecurity considers both risks to the enterprise and risks emanating from the enterprise into the broader environment. Other risks to irrigated agriculture and aquaculture are discussed elsewhere in this report, including risks associated with: • flooding (Section 2.5) • sediment infill of large dams (Section 5.4) • reliability of water supply (sections 5.4 and 6.3) • timing of runs of failed years on the profitability of an enterprise (Section 6.3). Material within this chapter is largely based on the companion technical reports on ecology (Merrin et al., 2024; Ponce Reyes et al., 2024) but also draws upon findings presented in the Northern Australia Water Resource Assessment technical reports on agricultural viability (Ash et al., 2018) and aquaculture viability (Irvin et al., 2018). Further information can be found in those reports. 7.3 Ecological implications of altered flow regimes 7.3.1 Water resource development and flow ecology The ecology of a river is intricately linked to its flow regime, and its species are broadly adapted to the prevailing conditions under which they occur. Changes in freshwater flows can affect the persistence or ephemerality of rivers, the volumes of river flows, and patterns of floodplain inundation and discharges. These changes directly affect species, habitats and ecosystem functions. Freshwater-flow-dependent flora, fauna and habitats are defined here as those sensitive to changes in flow and sustained by either surface water or groundwater flows or a combination of both. In rivers and floodplains, activities like water capture, storage, release, conveyance and extraction can significantly alter the environmental ecohydrology on which rivers function. Water regulation is frequently considered one of the biggest threats to aquatic ecosystems worldwide (Bunn and Arthington, 2002; Poff et al., 2007). Water resource development can cause changes in flow during both wet and dry periods, including the magnitude, timing, duration and frequency of flows (Jardine et al., 2015; McMahon and Finlayson, 2003). These changes can affect flora, fauna and habitats, and effects often extend far downstream, reaching near-shore coastal and marine areas as well as floodplains (Burford et al., 2011; Nielsen et al., 2020; Plagányi et al., 2024; Pollino et al., 2018). Water resource development can also result in changes to water quality, which is discussed in Section 7.5. The environmental risks associated with water resource development are particularly complex in northern Australia. This is in part because of the diversity of species and habitats distributed across and within the catchments and near-shore zones. In addition, wet-season precipitation is critical to sustain the ecology of Southern Gulf landscapes throughout the approximately 8-month annual dry season (Section 2.4). From the ecosystem scale, such as the coastal mangrove/estuary/salt flat complex, to the microhabitat scale, such as riverine pond refugia, communities and species depend on wet-season rainfall and runoff to invigorate the catchment- to-coast ecosystem following 9 months of negligible rainfall, high evaporation and no freshwater runoff or inflows (Duke et al., 2019; McJannet et al., 2014). Across northern Australia, monsoon rains relieve the extended period of dry conditions in the landscape (Petheram et al., 2008; Petheram et al., 2012). As a result, water resource development can divert water critical for ecosystem services, leading to a wide range of direct and indirect environmental impacts. Impacts on flow-dependent species and habitats can include flow regime change, loss of habitat, loss of function (such as connectivity), changes to water quality and the establishment of pest species. Instream dams create large bodies of standing water that inundate terrestrial habitat and result in the loss of the original stream and riverine conditions (Nilsson and Berggren, 2000; Schmutz and Sendzimir, 2018). Dams can capture flood pulses and reduce the volume and extent of water that transports important nutrients into estuaries and coastal waters via flood plumes (Burford et al., 2016; Burford and Faggotter, 2021; Tockner et al., 2010). Even minor instream barriers, such as road causeways, can disrupt migration and movement pathways, causing fragmentation of populations and loss of essential habitat for species that need passage along the river (Crook et al., 2015; Pelicice et al., 2015). Increased human activity associated with water resource development, such as irrigation, can introduce additional pressures, including biosecurity risks from invasive species that may spread into new or modified habitats or may be at increased risk of establishment (Pyšek et al., 2020). This section provides an analysis of the risks associated with flow regime change in the Southern Gulf catchments to freshwater, estuarine and near-shore marine ecology and terrestrial systems dependent upon river flows. The impacts of habitat loss within hypothetical dam impoundments and of loss of connectivity due to the development of new instream barriers and impact associated with land use change due to the creation of an impoundment are discussed in the companion technical report (Yang et al., 2024). Existing and other potentially threatening processes for ecological assets, including their possible synergistic impacts, are discussed qualitatively in the companion ecological descriptions report (Merrin et al., 2024). For more details of the ecological asset analysis and details of analysis for all assets, see Ponce Reyes et al. (2024). 7.3.2 Ecology of the Southern Gulf catchments The Southern Gulf catchments span 108,200 km² across the NT and Queensland, encompassing several significant protected areas (Merrin et al., 2024). These areas feature national parks like Boodjamulla and Finucane Island national parks, and 13 wetlands listed in the Directory of Important Wetlands in Australia. Among these wetlands is the Southern Gulf Aggregation (the largest continuous estuarine wetland in Australia), where the Gregory River crosses the Nicholson River Basin and Leichhardt River Basin (the largest perennial river in semi-arid and arid Queensland), and the Wentworth Aggregation (Department of Agriculture‚ Water and the Environment, 2021). The ecology of the Southern Gulf catchments is shaped by its wet-dry tropical climate, characterised by an extended dry season and a wet season during which the most rainfall occurs. In the dry season, river flows decrease and many of the catchment’s streams recede into isolated waterholes. However, in parts of the catchment, water persists during the dry season, supported by groundwater-fed perennial rivers like the Gregory and O’Shannassy rivers and Lawn Hill Creek, along with other water sources such as creeks, permanent lakes and inchannel waterholes. These water sources provide essential refuge habitats for a diverse array of aquatic species, particularly in this semi-arid environment (McJannet et al., 2023; Waltham et al., 2013). A significant landscape within the Southern Gulf catchments is the low-gradient, tidally influenced coastline characterised by some of the most extensive supratidal salt flats and chenier plain formations of the Australian coastline (Short, 2020). The salt flats are fronted by a beach-barrier shoreline (west of Mornington Island, 87% of the coastline is sand beach) and mangrove-fringed mudflats (east of Mornington Island, 13.5% of the coastline is sand beach). The salt flat complex is most extensive east of Mornington Island and can extend 30 to 50 km inland. It is dissected by several major rivers and multiple mangrove-lined creeks, creating a distinctive supra-littoral and littoral estuarine and marine environment that provides critical habitat to many species. The coastal ecosystem is characterised by an extremely low gradient, both in terrestrial habitats landward of the salt flat complex and in intertidal and subtidal habitats to seaward, contributing to the southern gulf of Carpentaria coastline being the most low-energy region of the open Australian coastline (Short, 2020). Within the multiple estuaries, mangrove forests, intertidal and supratidal sea-flats of the coastal zone of the Southern Gulf catchments, primary production from microalgae within the estuarine water column and subtidal, intertidal and supratidal sediments rivals productivity from other sources, such as the macrophytes (Burford et al., 2016; Burford et al., 2012). Annual flood flows deposit thousands of tonnes of nitrogen and phosphorus within flood plumes into the Gulf of Carpentaria, particularly in years of high floods (Burford et al., 2012; Burford and Faggotter, 2021). In addition, annual wetting of the salt flats due to rainfall and overbank flows invigorate extensive senescent algal crusts that cover the hundreds of square kilometres of salt flats, contributing an additional 13% to the primary productivity of estuarine/salt flat complex of the southern Gulf of Carpentaria (Burford et al., 2016). Primary productivity contributes to populations of mollusc, crustacean and annelid meiofauna and macrobenthos within estuarine and coastal sediments of the Southern Gulf (Lowe et al., 2022; Venarsky et al., 2022), forming the basis of the coastal food web. The volume and persistence of wet- and dry-season river flows have strong influences on the abundance and species diversity of infauna within estuarine and shallow coastal sandflats (Duggan et al., 2014; Lowe et al., 2022). This coastal matrix ecosystem is critical habitat for many bird, reptile, fish and crustacean species. Wader birds that migrate between the northern and southern hemispheres use the littoral habitats of Southern Gulf catchments as a crucial stopover on their biannual flyway (Garnett, 1987; Tait, 2005). Many of the bird species are threatened, endangered or critically endangered. Key fish species such as barramundi, threadfin, barred-grunter, sawfish, mullet and catfish, as well as crustaceans such as cherubin (in the brackish ecotone), mud crabs and multiple species of penaeid prawns, are abundant within estuarine and coastal habitats (Leahy and Robins, 2021; Robins et al., 2020; Staples and Vance, 1986). Freshwater and saltwater crocodiles are common within the catchments (Read et al., 2004). During the wet season, significant flooding occurs in the Southern Gulf catchments, inundating floodplains and connecting wetlands to the river channels, which enhances primary and secondary productivity. This process is particularly notable in the lower parts of the catchment, including the floodplain wetlands and the extensive intertidal salt flats along the mainland coastline, particularly south of Bentinck and Sweers islands. These nutrient-laden flood discharges into the marine waters of the south-western Gulf of Carpentaria support high levels of marine productivity, which in turn sustains important fisheries and ecological processes. The Southern Gulf catchments have high biodiversity. They support at least 170 species of fish, 150 species of waterbirds, 30 species of aquatic and semi-aquatic reptiles, 60 species of amphibians and 100 macroinvertebrate families (van Dam et al., 2008b). Notable species include the freshwater or largetooth sawfish (Pristis pristis; listed as Vulnerable under the Commonwealth Environment Protection and Biodiversity Conservation Act 1999 (EPBC Act)) and the Gulf snapping turtle (Elseya lavarackorum; Endangered). The catchments also serve as crucial stopover habitats for migratory shorebird species listed under the EPBC Act, including the eastern curlew (Numenius madagascariensis; Critically Endangered) and the Australian painted snipe (Rostratula australis; Endangered). The ecology of the Southern Gulf catchments is further detailed in the companion report on ecological assets (Merrin et al., 2024). 7.3.3 Scenarios of hypothetical water resource development and future climate The ecology analysis used modelled hydrology from a river model for the Gregory– Nicholson and Leichhardt catchments (see companion technical reports on river model calibration and simulation (Gibbs et al., 2024a,b)) to explore the possible impacts of water resource development in the Southern Gulf catchments using a range of hypothetical scenarios. The scenarios were configured to explore how different types and scales of water resource development, such as instream infrastructure (i.e. large dams) and water harvesting (i.e. pumping river water into offstream farm- scale storages) affect water-dependent ecosystems. The impact of a hypothetical development on water-dependent ecological assets is inferred and reported here in terms of a catchment- weighted, percentage change in key ‘flow dependencies’ (at key stages of the life cycle) relative to a baseline, averaged over the Assessment time period (i.e. 1 September 1890 to 31 August 2022). The catchment-weighted value is calculated by spatially weighting percentage change in key flow dependency calculations using the modelled likely habitat of each asset across the study area. This is referred to as the spatially weighted mean impact on key flow dependencies. The ecological analysis used Scenario AE as a reference (Section 1.4.3), recognising that current conditions likely reflect a stabilised state following past development in the catchments (Gibbs et al., 2024b). Changes in flow dependencies do not necessarily correlate with changes in asset condition, as this depends on the relative importance of river flow compared to other factors such as rainfall, soil water, groundwater and local runoff. Section 1.2.2 discusses the plausibility of development pathways and should be consulted when evaluating the likelihood of a hypothetical development scenario occurring. Broad scenario definitions used in the Assessment are described in Section 1.4.3 and summarised in Table 7-1. The scenarios were chosen to cover a range of potential ecological outcomes for the selected assets. The location of the river modelling nodes referred to in this section are shown in Figure 7-2. Further details of the river system model simulations are provided in the companion technical report on river model simulation (Gibbs et al., 2024b). Key terms used in Section 7.3 Water harvesting – an operation where water is pumped or diverted from a river into an offstream storage. Offstream storages – usually fully enclosed, circular or rectangular earthfill embankment structures situated close to major watercourses or rivers to minimise the cost of pumping. Large engineered instream dams – usually constructed from earth, rock or concrete materials as a barrier across a river to store water in the reservoir created and intercept a drainage line (Yang et al., 2024). Annual diversion commencement flow requirement (DCFR) – the cumulative volume passing the most downstream node in catchments with water harvest (on the Leichhardt River node 9130071, Albert River node 9129040 and Nicholson River node 9121090) from the start of the water year that is required before water harvest pumping can commence. Pump-start threshold – a daily flow threshold above which pumping or diversion of water can commence. This is usually implemented as a strategy to minimise the ecological impact of water harvesting. Pump capacity – the capacity of the pumps expressed as the number of days it would take to pump the entire node irrigation target. Reach irrigation volumetric target – the maximum volume of water extracted in a river reach over a water year. Note that the end use need not be irrigation; users could also be involved in aquaculture, mining, urban or industrial activities. System irrigation volumetric target – the maximum volume of water extracted across the entire study area over a water year. Note that the end use need not be irrigation; users could also be involved in aquaculture, mining, urban or industrial activities. Transparent flow – a strategy to mitigate the ecological impacts of large instream dams by allowing all reservoir inflows below a flow threshold to pass ‘through’ the dam. Note that each potential water resource development pathway results in different changes to flow regimes. This is due to differences in rainfall and upstream catchment sizes, inflows, the attenuation of flow through the river system (including accumulating inflows with river confluences), and also the many ways each water resource development could unfold and be implemented and managed. These scenarios were not analysed because they were considered likely or recommended by CSIRO; rather, they were selected to explore some of the interactions between location and the types and scale of development, to provide insights into how different types and scales of water resource development and mitigation measures may influence ecology outcomes across the catchment. Some of the hypothetical scenarios listed in Table 7-1 provide the minimum level of dedicated environmental provisions and have been optimised for water yield reliability, without considering policy settings or additional restrictions that may help mitigate the impacts on water-dependent ecosystems. These scenarios are useful for considering impacts across various development strategies in the absence of mitigation strategies or policy settings (or could be representative of regulatory non-compliance). Furthermore, management and regulatory requirements in a real- world setting would likely provide a range of safeguards for environmental outcomes, possibly establishing a combination of transparent flows (river flows that are managed to pass a regulating structure to support ecology), end-of-system requirements, extraction limits and/or minimum flow thresholds. Each of these safeguards, if implemented, would likely improve the environmental outcomes. Furthermore, many of the scenarios explored, while technically feasible, exceed the level of development that would reasonably occur (see Section 1.2.2). These scenarios were included as a stress test of the system and can be useful for benchmarking or contrasting various levels of change. The development scenarios are hypothetical and are for the purpose of exploring a range of options and issues in the Southern Gulf catchments. In the event of any future development occurring, further work would need to be undertaken to assess environmental impacts associated with the specific development across a broad range of environmental considerations. River system model nodes, flow-ecology dependencies, map \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\4_S_Gulf\1_GIS\1_Map_docs\Ec-S-535_Potential dams and Nodes_v9.mxd For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 7-2 Locations of the river system modelling nodes at which flow–ecology dependencies were assessed (numbered) and the locations of hypothetical water resource developments in the Southern Gulf catchments The flow ecology of the environmental assets was assessed in subcatchments downstream of the river system nodes. The locations of assets across the catchment are documented in Merrin et al. (2024). Marine assets used a combined end-of-system node (9100000), which combined flows from the Nicholson (9121090 and 9129040) and Leichhardt (9130071) rivers. Table 7-1 Water resource development and climate scenarios explored in this ecology analysis Gibbs et al. (2024a) and Gibbs et al. (2024b) describe the river system modelling and additional scenario details. SCENARIO DESCRIPTION ASSUMES FULL USE OF EXISTING LICENCES TRANSPARENT FLOW THRESHOLD (% OF MEAN INFLOW) TARGET EXTRACTION VOLUME (GL) ANNUAL DIVERSION COMMENCEMENT FLOW REQUIREMENT (GL) PUMP-START THRESHOLD (ML/d) PUMP CAPACITY (d) Scenario A Historical climate and no hypothetical development AN No development – natural conditions No (no use) na 0 na na na AE Current (2023) levels of development No (current) (29.9 GL) No 0 0 variable variable A Full use of existing entitlements Yes (113.5 GL) No 0 0 variable variable Scenario B Historical climate and hypothetical future development B-DGPC B-DGR Single hypothetical dams‡ Yes (113.5 GL) No na‡ na na na B-D2 Two hypothetical dams together (B-DGPC and B- DGR) Yes (113.5 GL) No na‡ na na na B-DGPCT B-DGRT Single hypothetical dams with transparent flows Yes (113.5 GL) Q = 20 na‡ na na na B-D2T Two hypothetical dams together with transparent flows (B- DGPCT and B-DGRT) Yes (113.5 GL) Q = 20 na‡ na na na B-WT150P600R30E0 Water harvesting varying target extraction volume (T), annual diversion commencement flow requirement (E), pump- start threshold (P), and/or pump capacities (R) Yes (113.5 GL) na T = 50, 150, 300 E = 0, 150… P = 200, 600… R = 10, 20… Scenario C Future climate and current level of development CEdry Dry GCM†† projection (see Section 2.4.5) No (current) (29.9 GL) No 0 0 variable variable Scenario D Future climate and hypothetical future development D-DGPC Single hypothetical dams‡ with Scenario Cdry Yes (113.5 GL) No na‡ na na na DESCRIPTION SCENARIO ASSUMES FULL USE OF EXISTING LICENCES TRANSPARENT FLOW THRESHOLD (% OF MEAN INFLOW) TARGET EXTRACTION VOLUME (GL) ANNUAL DIVERSION COMMENCEMENT FLOW REQUIREMENT (GL) PUMP-START THRESHOLD (ML/d) PUMP CAPACITY (d) D-D2 Two hypothetical dams (same as B-D2) with Scenario Cdry Yes (113.5 GL) No na‡ na na na D-D2T Two hypothetical dams (same as B-D2) with Scenario Cdry with transparent flows Yes (113.5 GL) Q = 20 na‡ na na na D- WT150P600R30E0 Water harvesting with Scenario Cdry Yes (113.5 GL) na T = 50, 150, 300 E = 0, 150, 250… P = 200, 600… R = 30… ‡No target volume for hypothetical dam scenarios; instead a target extraction volume that could be met with 85% reliability was identified. ††GCM = general circulation model. na = not applicable. 7.3.4 Ecology outcomes and implications The ecology activity used an asset-based approach to analysis, building on the work of Pollino et al. (2018) and Stratford et al. (2024). For the Southern Gulf catchments, 21 ecological assets (Table 7-2) were selected for analysis across 79 nodes, including the end-of-system node for near-shoremarine assets (Figure 7-2). Both the ecology asset descriptions technical report (Merrin et al., 2024) and the ecology asset analysis technical report (Ponce Reyes et al., 2024) should beconsulted in conjunction with the material provided here. The selected ecological assets encompass freshwater, marine and terrestrial habitats that depend on river flows to varying extents, and they were modelled with regards to changes to surface water. Assets were included if they were distinctive, representative, describable and significant within the catchment. The assets’ flow ecologies and locations were described in Merrin et al. (2024), which also provides species and habitat distribution maps, including species distribution models developed for many of the species. Each asset had different needs and linkages to the flow regime, and these assets occurred across different parts of the catchment or the near-shore marine zone. Understanding the flow ecology of assets and their locations across the catchment was important for identifying the potential risks of changes in catchment hydrology, as not all types of changes will affect assets equally. Table 7-2 Ecological assets used in the Southern Gulf catchments Water Resource Assessment and the different ecology groups used in this analysis Twenty-one assets are modelled in the ecology analysis; assets may be assigned to more than one group. Description of the ecological assets and their distribution is provided in Merrin et al. (2024). Assets marked with an asterisk are presented in this report. Analysis and interpretation for all assets is provided in Ponce Reyes et al. (2024). ASSET GROUP ASSET SYSTEMS Fish, sharks and rays Barramundi (Lates calcarifer) Freshwater and marine Bullshark Freshwater and marine Catfish (order Siluriformes) Freshwater Grunters (family Terapontidae) Freshwater Mullet (family Mugilidae) Freshwater and marine Sawfishes (Pristis and Anoxypristis spp.)* Freshwater and marine Threadfin (Polydactylus macrochir) Marine Waterbirds Colonial and semi-colonial nesting wading waterbirds Freshwater Cryptic wading waterbirds Freshwater Shorebirds Freshwater and marine Swimming, grazing and diving waterbirds* Freshwater Prawns, turtles and other species Banana prawns (Penaeus merguiensis) Marine Endeavour prawns (Metapenaeus endeavouri and M. ensis) Marine Tiger prawns (Penaeus semisulcatus and P.esculentus) Freshwater turtles (family Chelidae) Freshwater Mud crabs (Scylla serrata) Marine Flow-dependent habitats Floodplain wetlands* Freshwater Inchannel waterholes Freshwater Mangroves Marine Saltpans and salt flats Marine Seagrass Marine Surface-water-dependent vegetation Freshwater and terrestrial The flow dependencies (hydrometrics) modelling calculated for each asset an index of flow regime change resulting from the different scenarios using asset-specific hydrometrics using metrics based upon Kennard et al. (2010). Hydrometrics are statistical measures of the long-term flow regime and can include aspects of flow magnitude, duration, timing, frequency and rate of change (Kennard et al., 2010). Merrin et al. (2024) details each asset’s ecology and relationship to flow, including: • habitat dependencies (e.g. floodplain inundation, refuge, recharging of groundwater) • life cycle processes (e.g. flow to trigger spawning) • migration and movement pathways (e.g. high flows to enable migration into floodplain wetlands and along the river) • flow to support productivity and food resources (e.g. nutrient plumes into coastal areas). Hydrometrics were calculated for each scenario and used to quantify relative change in important parts of the flow regime as percentile change relative to the distribution of annual values of Scenario AE, calculated over the Assessment period (i.e. 1 September 1890 to 31 August 2022). The index of change is calculated as: Percentile change=x − scenario medianscenario median × 100 (1) Where x is the median of metric i, for the hypothetical scenario, and all values are for individual nodes. The assets’ important metrics are combined by averaging, with each metric being weighted, considering the knowledge base to support it and its significance to the asset’s ecology. The percentile change is weighted downstream of nodes by the habitat value of each reach in which the asset occurs based upon results of species distribution models, and the change in flow dependency is calculated for each node. The species distribution models were developed using a combination of Random Forests, Generalised Linear Models (GLMs), and Maxent algorithms (see Stratford et al., 2024a). These models were applied to a 2.5 km buffer surrounding the rivers within the Southern Gulf catchments to quantify habitat suitability. The change in the flow dependencies was weighted by habitat suitability for each asset between the river system model nodes of each river reach. As such, river reaches with important asset habitat quality or values are weighted higher than marginal habitat. Aggregation of these weighted flow dependency values is undertaken to calculate the catchment means of asset–flow dependencies from the individual node values (see Ponce Reyes et al., 2024 for more details). Hydrometrics have been broadly used in ecohydrology assessments in national and international contexts for a range of purposes, including water allocation planning, and in ecohydrology research and literature (Leigh and Sheldon, 2008; Marsh et al., 2012; Olden and Poff, 2003). For this analysis, the flow dependencies modelling considered reach and catchment-wide changes in each asset’s important flow dependencies across the subcatchments in which the assets occur, including the near-shore marine zone. The impact of a hypothetical development on water- dependent ecological assets is inferred and reported here in terms of a habitat-weighted percentile change in asset-specific important flow dependencies. For interpretation of results, larger values represent greater change in the parts of the flow regime that are important for the asset, and qualitative descriptors are provided in Table 7-3. As the values are percentile change from the distribution in Scenario AE, the asset’s flow dependency values can be referenced against the historical variability. For example, a value of 25 for a metric represents a change from the median (50th percentile of the historical distribution) to the 25th percentile. Using mean annual flow as an example metric, the value of 25 would represent the scenario median now being similar to the driest 25% of years for this metric. Table 7-3 Reporting qualitative values for the flow dependencies modelling as rank percentile change of the hydrometrics Values consider the change in mean hydrometric value against the natural distribution observed in the modelled baseline series of 132 years. For more information including metric and habitat weighting see Ponce Reyes et al. (2024). For more information see Ponce Reyes et al. (2024). PERCENTILE VALUE RATING IMPLICATION >0-2 Negligible The mean for the asset’s metrics under the scenario has negligible change as considered against the modelled historical conditions and lies well within the normal conditions experienced at the model node. The asset’s hydrometrics are within two percentile of the historical Scenario AE mean 2-5 Minor The change is minor with the mean for the asset’s metrics for the scenario between two and five percentile of Scenario AE and the historical distribution of the hydrometrics 5-15 Moderate The change is moderate with the mean for the asset’s metrics under the scenario between five and 15 percentile of Scenario AE and the historical distribution of the hydrometrics 15-30 Major The change is major with the mean for the asset’s metrics for the scenario between 15 and 30 percentile of Scenario AE and the historical distribution of the hydrometrics >30 Extreme The change is extreme with the mean for the asset’s metrics under the scenario being very different from the modelled historical conditions, with and metrics occurring well outside typical conditions at the modelled node. The scenario mean is more than 30 percentile from the historical Scenario AE mean In addition to comparing against the historical variability of Scenario AE, benchmarking the level of change in asset flow dependencies is achieved by comparing to asset results modelled at the end- of-system of the Ord River, taking into account the modelled changes in flow associated with the development of the Ord River Irrigation Scheme (with and without Top Dam) near the end-of- system. In addition, three natural periods of low-flow conditions are assessed and plotted alongside the hypothetical development and projected climate scenario values. For the Southern Gulf catchments, these were the periods with the lowest 30-year flow, lowest 50-year flow and lowest 70-year flow across the historical climate (Scenario AE). Additional context is provided by calculating the change in asset flow dependency that has been modelled to occur in the Gregory and Leichhardt catchments since European settlement (i.e. by comparing key flow dependency metrics under Scenario AN and Scenario AE). These are benchmarks, so flow conditions and outcomes of change would not necessarily be equivalent to these if development were to occur, but they provide a useful comparison of the potential level of change under the scenarios. Note that this ecology analysis is broad in scale, and the results include significant uncertainty. This uncertainty is due to a range of factors, including, but not limited to, incomplete knowledge, variability within and between catchments, and limitations associated with modelling processes and data. Furthermore, thresholds, temporal processes, interactions, synergistic effects and feedback responses in the ecology of the system may not be adequately captured in the modelling process. There is also uncertainty associated with the projected future climates, such as rainfall patterns and any additional synergistic and cumulative threatening processes that may emerge and interact across scales of space and time, including producing potentially novel outcomes. The region that the Southern Gulf catchments occur within is vast and diverse, and the knowledge base of species occurrences is limited. More broadly, the understanding of freshwater ecology in northern Australia is still developing. Provided below is a sample of outcomes for three representative assets for the Southern Gulf catchments: sawfishes; swimming, diving and grazing waterbirds; and floodplain wetlands. For more details and for results on other assets see Ponce Reyes et al. (2024). Sawfishes Four species of sawfish inhabit the Gulf of Carpentaria, primarily in inshore marine habitats and estuaries. Tropical Australian waters are one of the last strongholds for sawfishes (Phillips et al., 2011). The two largest species, largetooth or freshwater sawfish (Pristis pristis) and green sawfish (P. zijsron) are listed as Critically Endangered on the IUCN Red List of Threatened Species and Vulnerable under the Commonwealth EPBC Act. The dwarf sawfish (P. clavata) is listed as Critically Endangered (IUCN) and Vulnerable (EPBC Act), while the narrow sawfish (Anoxypristis cuspidata) is listed as Critically Endangered (IUCN) and not listed under the EPBC Act. Freshwater sawfish juveniles inhabit riverine reaches before moving to coastal and marine environments as adults. Juvenile dwarf sawfish have been recorded in upper estuaries and lower rivers, while adults are found offshore. Green sawfish are common in estuaries and occasionally in riverine habitats. In the Southern Gulf catchments, sawfishes occupy estuarine, freshwater and offshore habitats. Estuarine and riverine connectivity is critical for the survival of freshwater sawfish, which pup in estuarine and inshore waters (Dulvy et al., 2016; Morgan et al., 2017). Sawfishes are vulnerable to human activities and also used as a food source (Naughton et al., 1986). In Australia, only Indigenous Australians are allowed to capture sawfishes. Freshwater sawfish, in particular, are affected by variability in the flow regime despite sustained riverine and estuarine connectivity during the wet season. Strong upstream recruitment of juveniles to riverine habitats only occurs during the highest flood flows (Lear et al., 2019). The higher the volume of flood flows, the greater the sustained body condition of sawfish during the subsequent dry season (Lear et al., 2021). The key threats to sawfishes are associated with the loss of high-level flood flows to support upstream recruitment and with any reduction in low-level dry-season flows that would reduce instream connectivity or create barriers among deep-water pools and reduce their persistence or water quality during the dry season. The analysis considers change in flow regime and related habitat changes but does not consider the loss of potential habitat associated with the creation of a dam impoundment or instream structures (see also Yang et al. (2024) for dam impoundments). Flow dependencies analysis Sawfishes were assessed across a total of 3948 km of river reaches in the Southern Gulf catchments and the marine region, with flows from 63 model nodes (Ponce Reyes et al., 2024). The locations for modelling sawfishes were selected based on the species distribution models of freshwater or largetooth sawfish (Pristis pristis) (see Merrin et al. (2024)). Hypothetical water resource development in the Southern Gulf catchments led to varying changes in key flow metrics important for sawfish (see details and results for all the scenarios in Ponce Reyes et al. (2024)). Hypothetical dam scenarios showed a range of change in key flow metrics from negligible (0.6 under Scenario B-DGPCT) to moderate (5.2; Scenario B-D2). Change in key flow metrics under water harvesting scenarios was negligible, with values ranging from 0.5 (Scenario B-WT50P600R30E250) to 1.4 (Scenario B-WT150P200R30E0). Under Scenario Cdry, there was a moderate change (6.5) in flow metrics. The variation in flow regime changes across dam, water harvesting, and climate scenarios is due to differences in spatial patterns of flow and the distribution of important habitat for sawfish (Figure 7-4). Water harvesting and changes in important flows for sawfish The change in important flows for sawfish under water harvesting varied depending on extraction targets, pump-start thresholds, pump rates and annual diversion commencement flow requirements (Ponce Reyes et al., 2024). Under a low extraction target of 50 GL (Scenario B-WT50P600R30E0), the change in flow was negligible (0.5), and it increased to 0.9 with an extraction target of 300 GL (Scenario B-WT300P600R30E0) (Figure 7-5). Raising the pump-start threshold from 200 ML/day (Scenario B-WT150P200R30E0) to 600 ML/day (Scenario B-WT150P600R30E0) reduced asset dependent flow changes, with values ranging from 1.4 to 0.7. Protecting key parts of the flow regime, such as maintaining low flows and limiting extraction volumes, is crucial for sawfish ecology. Flow modifications, particularly reduced high flows and shorter peak durations, can disrupt species reliant on floodplain inundation and wetland connectivity. Additionally, water impoundment or upstream extraction may reduce the depth and persistence of critical riverine pools during the dry season. Figure 7-3 Riparian vegetation along the Leichhardt River – riparian zones are often more fertile and productive than surrounding terrestrial vegetation Photo: CSIRO – Nathan Dyer For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 7-4 Spatial heatmap of habitat-weighted changes in flow for sawfish, considering the assets important locations across the catchment Scenarios are: (a) B-WT150P600R30E0, (b) B-WT150P600R30E150, (c) B-WT300P600R30E0, (d) B-D2, (e) CEdry and (f) D-WT150P600R30E0. River shading indicates the level of flow change of sawfish important metrics weighted by the habitat value of each reach. \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\0_Working\Lynn_ecology\3_Plots\3_SG\Plot1\Plot1C\v6\Sawfish_noBoxPlot_by_MeanChange_orderByLocation_v6b.png Figure 7-5 Habitat-weighted change in sawfish flow dependencies by scenario across model nodes Colour intensity represents the level of change occurring in important flow metrics as percentile change from the historical conditions weighted by the importance of each reach for sawfishes. Equivalent colour intensity (i.e. corresponding to the asset flow dependency change value) for the Ord River below Lake Kununurra shown bottom right. Scenarios are ordered on the left axis by the magnitude of change with the mean across nodes shown on the right axis. Horizontal grey bars and number correspond to the mean change across all model node locations. Only the 30 highest impact nodes are shown (x-axis). Results under Scenario A corresponding to changes in asset flow dependency for the lowest 30-year (A lowest 30), 50-year (A lowest 50) and 70-year (A lowest 70) time periods provide a reference for the modelled changes under different hypothetical development and projected future climate scenarios. AN-AE corresponds to the change in asset flow dependency in the Nicholson and Leichhardt catchments that has already occurred since European settlement. Dams and changes in important flows for sawfish Under the hypothetical dam scenarios, Scenario B-DGPC (without transparent flows) resulted in a negligible change in important flow metrics (0.7) when averaged across the 63 sawfish assessment nodes. Implementing transparent flows (Scenario B-DGPCT) reduced this to 0.6. Scenario B-DGR resulted in a minor flow change (4.9), which was reduced to negligible (1.6) with transparent flows (Scenario B-DGRT). Scenario B-D2, involving two dams, showed a moderate change (5.2), which was reduced to 3.1 with transparent flows (Scenario B-D2T). As expected, Scenario B-D2 resulted in a larger mean change in flow dependency than either of the individual dam scenarios due to the combined effects of two dams and the larger portion of the catchment affected. Transparent flows consistently mitigated impacts, improving environmental outcomes for sawfish. The greatest habitat-weighted changes occurred at node 9121012, with an extreme change (48.2). Nodes directly downstream of the dams saw major (16.3) and extreme (45.4) changes, which were reduced to moderate levels (11.7 and 12.9) with transparent flows. Dams placed higher in the catchment reduced potential impacts, but sawfish remain vulnerable to a combination of factors, including changes in flow and connectivity loss (see Ponce Reyes et al. (2024)). Climate change and water resource development for important flows for sawfish Scenario Cdry resulted in a moderate change (6.5) in important flow metrics for sawfish across the 63 sawfish assessment nodes, which is higher than the changes seen in scenarios B-DGPC and B- WT150P600R30E0 (both negligible; 0.7). However, note that local changes in flows under some water resource development scenarios can be considerably higher than the catchment means. Combined impacts of climate change and water resource development resulted in moderate changes of 7.1 (Scenario D-DGPC) and 7 (Scenario D-WT150P600R30E0), indicating a significant impact on sawfish flow dependencies via flow reduction. These changes, especially reductions in wet- season flood flows and late-dry-season low flows, could reduce neonate recruitment upstream and disrupt connectivity to wetlands. Dams further impede access to juvenile habitats (see Yang et al. (2024)), while large flood flows are critical for the recruitment, growth and survival of largetooth sawfish within riverine freshwater habitats (Lear et al., 2019; Lear et al., 2021). Flow modifications, particularly reductions in high flows and shortened peak water levels, can negatively affect species relying on floodplain inundation and wetland connectivity (Close et al., 2014; Hunt et al., 2012; Jellyman et al., 2016; Morgan et al., 2016; Novak et al., 2017). Among Australian tropical rivers at similar latitudes, water resource development and a drying climate have been modelled to have significant negative impacts on sawfish populations (Plagányi et al., 2024). Measures to protect important parts of the flow regime can support ecology; for example, reducing the extraction target puts limits on the volume of water extracted in any water year, and increasing the pump-start threshold protects the low flows that are important for sawfish ecology. Flow modifications, particularly the reduction of high flows and shortened duration of the peak water levels (upper 25% of flows), can affect species such as the sawfish that rely on floodplain inundation and wetland connectivity. Furthermore, the maintenance of depth and persistence of important riverine pools during the dry season may be reduced by water impoundment or upstream extraction. Swimming, diving and grazing waterbirds The swimming, diving and grazing waterbirds group comprises species with a relatively high level of dependence on semi-open, open and deeper water environments, who commonly swim when foraging (including diving, filtering, dabbling, grazing) or when taking refuge. In northern Australia, this group comprises 49 species from 11 families, including ducks, geese, swans, grebes, pelicans, darters, cormorants, shags, swamphens, gulls, terns, noddies and jacanas (see species list in Merrin et al. (2024)). Reduced extent, depth and duration of inundation of waterhole and other deep-water environments are likely to reduce habitat availability and food availability for swimming, diving and grazing waterbirds. Reduced high-level flows increases competition, and predation also increases the risk of disease and parasite spread. Conversely, species in this group that nest at water level or just above, such as magpie geese (Anseranas semipalmata), are particularly at risk of nests drowning when water depths increase unexpectedly. The analysis focuses on flow regime changes and excludes habitat changes associated with a dam creation (see also Yang et al. (2024)). Flow dependencies analysis Swimming, diving and grazing waterbirds were assessed across a total of 3948 km of river reaches in the Southern Gulf catchments using flows from 62 model nodes The locations for modelling these birds were selected based on species distribution models of the magpie goose (Anseranas semipalmata) (see Merrin et al. (2024)). Hypothetical water resource development in the Southern Gulf catchments led to varying levels of change in key flow metrics important for these waterbirds. Across all 62 analysis nodes, dam scenarios showed changes ranging from negligible (0.5; Scenario B-DGPCT) to minor (3.5; Scenario B-D2). In contrast, water harvesting scenarios resulted in negligible changes 0.4 to 1.1 for scenarios B-WT50P600R30E150 and B-WT150P200R30E0, respectively. Scenario Cdry resulted in a moderate change(5.1) in important flow metrics. The impact associated with these scenarios varied due to spatial differences in flow regime changes and the distribution of important habitat for swimming, diving and grazing waterbirds (Figure 7-6). Water harvesting and changes in important flows for swimming, diving and grazing waterbirds The hypothetical water harvesting scenarios for swimming, diving and grazing waterbirds resulted in a negligible mean change ranging from 0.4 to 1.1 for scenarios B-WT50P600R30E150 and B- WT150P200R30E0, respectively. Changes in important flows under water harvesting scenarios varied depending on extraction targets, pump-start thresholds, pump rates and annual diversion commencement flow requirements (Figure 7-6). With a low extraction target of 50 GL (Scenario B-WT50P600R30E0), the mean weighted change in flows remained negligible at 0.4. This increased slightly to 0.7 with a higher extraction target of 300 GL (Scenario B-WT300P600R30E0). The pump-start threshold is important for protecting low flows by preventing pumping when the river is below this threshold. With a 150 GL extraction target, increasing the pump-start threshold from 200 ML/day (Scenario B-WT150P200R30E0) to 600 ML/day (Scenario B-WT150P600R30E0) reduced the mean change from 1.1 to 0.6 (negligible) (Figure 7-6). Measures such as limiting extraction targets and increasing the pump-start threshold can help protect the low flows that are important for swimming, diving and grazing waterbirds ecology. 440 | Water resource assessment for the Southern Gulf catchments \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\0_Working\Lynn_ecology\3_Plots\3_SG\Plot1\Plot1C\v6\Swimmers..divers..and.grazers_noBoxPlot_by_MeanChange_orderByLocation_v6b.png Figure 7-6 Habitat-weighted change in swimming, diving and grazing waterbirds flow dependencies by scenario across model nodes Colour intensity represents the level of change occurring in important flow metrics as percentile change from the historical conditions weighted by the importance of each reach for swimming, diving and grazing waterbirds. Equivalent colour intensity (i.e. corresponding to the asset flow dependency change value) for the Ord River below Lake Kununurra shown bottom right. Scenarios are ordered on the left axis by the magnitude of change with the mean across nodes shown on the right axis. Horizontal grey bars and number correspond to the mean change across all model node locations. Only the 30 highest impact nodes are shown (x-axis). Results under Scenario A corresponding to changes in asset flow dependency for the lowest 30-year (A lowest 30), 50-year (A lowest 50) and 70-year (A lowest 70) time periods provide a reference for the modelled changes under different hypothetical development and projected future climate scenarios. AN-AE corresponds to the change in asset flow dependency in the Nicholson and Leichhardt catchments that has already occurred since European settlement. Dams and changes in important flows for swimming, diving and grazing waterbirds Under the dam scenarios, Scenario B-DGPC resulted in a negligible change (0.6) in important flow metrics when averaged across the 62 swimming, diving and grazing waterbirds assessment nodes. With the implementation of transparent flows in Scenario B-DGPCT, the change was further reduced to 0.5 (negligible). Scenario B-DGR, in contrast, resulted in a minor change (3.2), which was reduced to negligible (0.8) with the transparent flows under Scenario B-DGRT. Scenario B-D2, which includes both the B-DGPC and B-DGR dams, resulted in a larger mean flow change across the catchment compared to either single dam scenario. The mean change across the catchment without transparent flows was minor (3.5) and this was reduced to negligible (1.7; Scenario B-D2T) with transparent flows. The greater change under Scenario B-D2 compared to the single dam scenarios is due to the combined effects of the two dams and their larger impact on the catchment. Scenario B-D2T (with transparent flow) resulted in a smaller change than scenarios without transparent flows, demonstrating the importance of providing flows to support environmental outcomes for swimming, diving and grazing waterbirds (Figure 7-6). Under dam scenarios, node 9121015 showed the greatest habitat-weighted changes in important flows for swimming, diving and grazing waterbirds (Figure 7-6) with the change recorded as extreme (51.5). Nodes directly downstream of the dams under scenarios B-DGPC and B-DGR resulted in moderate (8.7) and major (19.5) changes in important flows, respectively. These changes were reduced to moderate (5.1) and minor (3.7), respectively, when modelled with transparent flows. This pattern reflects the combined effect of flow changes directly downstream of the dams and the benefits of providing flows to support environmental outcomes. Climate change and water resource development for important flows for swimming, diving and grazing waterbirds Scenario Cdry resulted in a moderate change (5.1) in important flow metrics based on the mean across the 62 swimming, diving and grazing waterbirds assessment nodes (Figure 7-6). This indicates that the dry climate scenario led to a larger mean change than did Scenario B-DGPC and B-WT150P600R30E0, both of which showed negligible changes (0.6). However, note that local changes in flows can be considerably higher than the catchment means under some water resource development scenarios. When considering the impacts of climate change combined with water resource development, scenarios D-DGPC and D-WT150P600R30E0 both resulted in moderate changes (5.4), which is higher than the impacts of Scenario Cdry or either of scenarios B-D2 and B-WT160P200R30 alone. Species in the swimming, grazing and diving waterbirds group are sensitive to changes in the depth, extent and duration of perennial semi-open and open deeper water environments such as wetlands and waterholes (Marchant and Higgins, 1990; McGinness, 2016). They can also be sensitive to changes in the type, density or extent of the fringing aquatic or semi-aquatic vegetation in and around these habitats. These changes can occur when water is extracted directly from these habitats or when the time between connecting flows or rainfall events that fill these habitats is extended (Kingsford and Norman, 2002). Climate change as explored through the Scenario Cdry and extremes are likely to interact with changes induced by water resource development, including inundation of freshwater habitats by seawater, and inundation of nests by extreme flood events or seawater intrusion (Nye et al., 2007; Poiani, 2006; Traill et al., 2009a; Traill et al., 2009b). The reduced extent, depth and duration of inundation of waterholes and other 442 | Water resource assessment for the Southern Gulf catchments deep-water environments is likely to reduce their habitat and food availability, increasing competition and predation and also increasing risk of disease and parasite spread. Conversely, species in this group that nest at water level or just above, such as magpie geese, are particularly at risk of nests drowning when water depths increase unexpectedly (Douglas et al., 2005; Poiani, 2006; Traill et al., 2010; Traill et al., 2009a; Traill et al., 2009b). Floodplain wetlands For this analysis, floodplain wetlands are defined as freshwater lakes, ponds, swamps and floodplains with water that can be permanent, seasonal or intermittent, and can be natural or artificial. The Southern Gulf catchments contain 13 nationally significant wetlands listed in the Directory of Important Wetlands in Australia, although none are listed under the Ramsar Convention (see Merrin et al. (2024)). Wetlands provide permanent, temporary or refugia habitat for a range of species, are important for driving both primary and secondary productivity, and provide a range of additional ecosystem services (Junk et al., 1989; Mitsch et al., 2015; Nielsen et al., 2015; van Dam et al., 2008a; Ward and Stanford, 1995). Floodplain wetlands are influenced by the timing, duration, extent and magnitude of floodplain inundation, which significantly affect their ecological values, including species diversity, productivity and habitat structure (Close et al., 2015; Tockner et al., 2010). In the southern Gulf of Carpentaria, during high-level flood flows, floodplain wetlands connect with coastal salt flats as a ‘shallow lake’ continuum. Freshwater fauna move downstream and brackish-water-tolerant estuarine species move from the river channels onto the inundated, productive salt flats to forage and reproduce, taking advantage of the food web that depends on the floodwater-stimulated algal crusts that invigorate from their dry-season senescence (Burford et al., 2016; Burford et al., 2010). The key threats to floodplain wetlands are changes in flood regimes, specifically the timing, duration, extent and magnitude of floodplain inundation and the ensuing effect on the habitat structure, size and permanence of the wetlands and thus on species diversity and community productivity. Flow dependencies analysis Floodplain wetlands were assessed across a total of 2027 km of river reaches in the Southern Gulf catchments with contributing flows from 34 model nodes (see Ponce Reyes et al. (2024)). Key river reaches for floodplain wetlands in the Assessment catchments were modelled downstream of nodes 9139000, 9130140 and 9130050, which were selected based upon wetland and floodplain mapping (see Merrin et al. (2024)). Hypothetical water resource development in the Southern Gulf catchments resulted in varying changes to key flow metrics affecting floodplain wetlands. The mean weighted change in flow of the hypothetical dam scenarios across all 34 floodplain wetlands analysis nodes ranged from negligible (0.3; Scenario B-DGPCT) to minor (3.6; Scenario B-D2T). In contrast, water harvesting scenarios ranged from negligible (0.6; Scenario B-WT50P600R30E0) to minor (2.3; Scenario B-WT300P600R30E0). Scenario Cdry resulted in a moderate change in important flow metrics (12.9) for floodplain wetlands. The level of flow regime change associated with dam construction, water harvesting and climate scenarios varied due to the differing spatial patterns of flow regime change and the distribution of important habitat for floodplain wetlands (Figure 7-7). \\FS1-CBR.nexus.csiro.au\{lw-rowra}\work\7_Ecology\0_Working\Lynn_ecology\3_Plots\3_SG\Plot1\Plot1C\v6\Floodplain.wetlands_noBoxPlot_by_MeanChange_orderByLocation_v6b.png Figure 7-7 Habitat-weighted change in floodplain wetlands flow dependencies by scenario across model nodes Colour intensity represents the level of change occurring in important flow metrics as percentile change from the historical conditions weighted by the importance of each reach for floodplain wetlands. Equivalent colour intensity (i.e. corresponding to the asset flow dependency change value) for the Ord River below Lake Kununurra shown bottom right. Scenarios are ordered on the left axis by the magnitude of change with the mean across nodes shown on the right axis. Horizontal grey bars and number correspond to the mean change across all model node locations. Only the 30 highest impact nodes are shown (x-axis). Results under Scenario A corresponding to changes in asset flow dependency for the lowest 30-year, 50-year and 70-year time periods provide a reference for the modelled changes under different hypothetical development and projected future climate scenarios. AN-AE corresponds to the change in asset flow dependency in the Nicholson and Leichhardt catchments that has already occurred since European settlement. Water harvesting and changes in important flows for floodplain wetlands The hypothetical water harvesting scenarios resulted in a mean change across floodplain wetland assessment nodes from negligible (0.6; Scenario B-WT50P600R30E0) to minor (2.3; Scenario B-WT300P600R30E0). The changes varied depending on extraction targets, pump-start thresholds, pump rates, and annual diversion commencement flow requirements (Figure 7-7). For a lowextraction target of 50 GL (Scenario B-WT50P600R30E0), the flow change was negligible (0.6), and thisincreased to minor (2.3) with a higher extraction target of 300 GL (Scenario B-WT300P600R30E0). Thepump-start threshold is important for protecting low flows by preventing pumping when the riveris below this threshold. With a 150 GL extraction target, raising the pump-start threshold from200 ML/day (Scenario B-WT150P200R30E0) to 600 ML/day for (Scenario B-WT150P600R30E0) slightlyreduced the flow change from 1.6 to 1.5 (Figure 7-7). Measures to protect important parts of theflow regime can support ecology; for example, reducing the extraction target puts limits on thevolume of water extracted in any water year, and increasing the pump-start threshold protectsthe low flows. Dams and changes in important flows for floodplain wetlands Under the dam scenarios, when averaged across the 34 floodplain wetlands assessment nodes, Scenario B-DGPC (without transparent flows) resulted in a negligible change (0.3) in flow metrics for floodplain wetlands. This change remained the same with transparent flows (Scenario B-DGPCT). Scenario B-DGR resulted in a minor change (3.2), which slightly increased to 3.3 with transparent flows under Scenario B-DGRT. In Scenario B-D2, which includes both dams, a minor (3.4) mean change occurred, and this increased to 3.6 with transparent flows. Scenario B-D2 resulted in a larger mean flow change across the catchment than either of the single dam scenarios. This increase was due to the combined effects on flows downstream and the larger portion of the catchment affected. Habitat-weighted flow changes were most significant at node 9121050 and directly downstream (Figure 7-7), with a major change (26.5; Scenario B-DGR). This was reduced to 22.5 (major) with transparent flows (Scenario B-DGRT). This pattern reflects the combined effect of flow changes directly downstream of the dams and the benefits of providing flows to support environmental outcomes. Dams can have a significant impact on floodplain wetlands, as they capture runoff from rainfall events that would otherwise spill onto floodplains during larger events and facilitate the connection of the wetlands to the main river channel. The reduction in flood magnitude due to dams can change the connectivity between the river channel and the floodplain wetlands, significantly affecting the size of the inundated area. A loss of connectivity between the river channel and the floodplain wetland may also occur. This disconnection can alter the frequency and duration of wetland inundation, potentially leading to changes in the structure, function and biodiversity of these wetland habitats (Poff and Zimmerman, 2010; Richter et al., 1996). Climate change and water resource development for important flows for floodplain wetlands Scenario Cdry resulted in a moderate change (12.9) in important flow metrics for floodplain wetlands based on the mean across the 34 floodplain wetland assessment nodes (Figure 7-7). This indicates that the dry climate scenario led to a larger mean change across all catchment nodes than scenarios B-DGPC (negligible; 0.3) and B-WT150P600R30E0 (negligible; 1.5). However, it is important to note that local changes in flows under some water resource development scenarios can be considerably higher than the catchment means. Considering the combined impacts on flow associated with climate change and water resource development, scenarios D-DGPC and D-WT150P600R30E0 resulted in moderate (13.2 and 13.9, respectively) changes when weighted across all floodplain wetland assessment nodes. This shows that the changes of scenarios D-DGPC or DdryWT150P600R30E0 were higher than those of Scenario Cdry or either of scenarios B-D2 and B-WT160P200R30 alone. 7.3.5 Management of impacts on ecology The magnitude and spatial extent of ecological impacts arising from water resource development are highly dependent on the type and location of development, the extraction volume and how the type of changes in the flow regime affect different aspects of flow ecology. Mitigation measures seek to protect key parts of the flow regime and can be important for sustaining ecology under water resource development. This section explores the effectiveness of different mitigation measures, including providing transparent flows for dams, different rules for water harvesting and different overall targets for water extraction. Instream dam development Two potential locations for instream dams (Gregory River and Gunpowder Creek) were selected for modelling and analysis (Yang et al., 2024) and simulated following the hydrology modelling approach outlined in Gibbs et al. (2024b). Their locations are shown in Figure 7-2. The objective of this analysis is to test the effect of different dam locations and configurations on changes to streamflow to understand the effect on downstream ecology. Dams are modelled individually (i.e. scenarios B-DGR and B-DGPC), as well as both together (Scenario B-D2) to better understand cumulative impacts and to have variants with and without the mitigation measure of providing transparent flows (see Ponce Reyes et al. (2024) for definitions). Instream dams create a range of impacts on streamflow associated with the capture and extraction of water, affecting the timing and magnitude of downstream flows. The change of downstream flow associated with instream dams is explored here across broad asset groups, and results are shown as the mean of asset values. Impacts associated with loss of connectivity due to the dam wall and loss of habitat associated with the dam inundation extent are discussed in Yang et al. (2024). The dam scenarios and the resulting ecology flow dependencies are discussed in more detail for each asset in Ponce Reyes et al. (2024). Assessment of the individual dams found varying levels of effect on ecology flow dependencies (Table 7-4). Scenario B-DGPC resulted in no asset group having changes in important flow dependencies greater than negligible averaged across their assessment nodes, although local impacts on flows were often higher. However, Scenario B-DGR resulted in major change for the ‘other species’ group including the turtles and prawn species. The dams vary in size, inflows and capture volumes, and the location of the dam in the catchment also influences outcomes, particularly in the freshwater reaches. Impacts on flow directly downstream of modelled dams can often be high and may cause extreme changes in ecological flow dependencies. Areas further downstream have contributions from unimpacted tributaries, and for the marine region, from flows from other catchments that help support ecological outcomes of flow regimes. Dams further up the catchment may, however, affect a larger proportion of streams and river reaches in terms of flow regime change but may have lower impacts in terms of connectivity. The Southern Gulf catchments, in particular, are complex and braided, and they have many tributaries that may be unimpacted within the freshwater sections of the catchment. Effects on important flows are not 446 | Water resource assessment for the Southern Gulf catchments equivalent across assets, and large local impacts may lead to changes in ecology across other parts of the catchment due to the connected nature of ecological systems. The cumulative change in flow dependencies associated with two dams (Scenario B-D2) are greater than those of individual dams (Table 7-4). However, the largest contribution of change to Scenario B-D2 originates from B-DGR. Cumulative change in flows on ecology may be associated with acombination of a larger portion of the catchments being affected by changes in flows and residualflows being lower due to the overall greater level of water use and losses from dams (Table 7-4). Table 7-4 Scenarios of different hypothetical instream dam locations showing mean changes of ecology flows for groups of assets across each asset’s respective catchment assessment nodes Higher values represent greater change in flows important to the assets of each group. Values are asset means across their respective catchment assessment nodes (see Appendix A in Ponce Reyes et al. (2024)). Some assets are considered in multiple groups, where the mean across the nodes is used. Asset means include values from all nodes that the asset is assessed in, including in reaches that may not be affected by flow regime change. SCENARIO DESCRIPTION ALL ASSET MEAN FISH WATERBIRDS OTHER SPECIES HABITATS FRESHWATER ASSETS MARINE ASSETS B-DGPC Gunpowder Creek dam 0.8 0.4 0.8 0.1 0.5 0.8 0.2 B-DGPCT Gunpowder Creek dam with transparent flows 0.7 0.3 0.7 0.1 0.5 0.7 0.2 B-DGR Gregory River dam 4.1 8.9 4.0 18.2 6.4 4.4 12.7 B-DGRT Gregory River dam with transparent flows 1.5 1.8 1.7 3.8 2.8 1.7 3 B-D2 Both B-DGPC and B- DGR 4.5 9.1 4.4 18.3 6.7 4.7 12.8 B-D2T Both B-DGPC and B- DGR with transparent flows 2.7 5.5 2.7 12.7 4.1 2.9 8.4 Measures to mitigate the flow-related impacts of large instream dams, such as transparent flows (inflows let to pass the dam wall for environmental purposes), resulted in lower change to ecological flow dependencies broadly across all assets compared to instream dams without these measures (Table 7-4). Particularly strong benefits are shown in the reduced change to important flow dependencies resulting from transparent flows for members of the fish and waterbird groups (see Ponce Reyes et al. (2024) for groupings and more detail). Instream dams capture inflows and change downstream flow regimes. Transparent flows are a type of environmental flow provided as releases from dams that mimic or maintain some aspects of natural flows. Staged offtakes to maintain natural water temperatures are required. However, providing transparent flows from a dam it lowers the volume of water in storage and thereby increases the capture of early flood events in the following year. This might also result in the lowering of flood peaks, resulting in smaller inundation events during periods of floods affecting assets such as surface-water- dependent vegetation (see Ponce Reyes et al. (2024)). Modelling transparent flows uses inflow thresholds on dams and was designed primarily to preserve lower flows during periods of natural inflow. Inflow thresholds used in the transparent flows analysis were similar to the commence-to- pump thresholds used in water harvest scenarios, facilitating comparison. Transparent flow scenarios are provided across both individual dams and under Scenario B-D2 (Gibbs et al., 2024b). Water harvesting For water harvesting scenarios, several measures can mitigate the impacts of flow-related changes from extraction. These include limiting system targets to reduce extraction across the catchment, implementing a pump-start threshold to restrict pumping during low river flows, setting an annual diversion commencement flow requirement to allow a volume of water to pass through the system before pumping, and limiting the pump rate for extraction (see Ponce Reyes et al. (2024) and Gibbs et al. (2024b) for more details). These measures improve environmental outcomes compared to scenarios without them. Reducing system targets decreases the changes in flows across asset groups, while larger extraction volumes lead to moderate increases in flow dependencies across the catchment’s ecological assets. Groups like turtles, prawns and other species and the marine assets (Table 7-2), experience greater changes at higher system targets of 400 to 500 GL/year (Figure 7-9) (see Ponce Reyes et al. (2024) for details on individual assets). Providing minimum flow thresholds or annual diversion commencement requirements can help mitigate these changes. An annual diversion commencement flow requirement of 250 GL improves ecological flows across asset groups with smaller requirements proportionally reducing flow changes (see Ponce Reyes et al. (2024)). The largest benefits for smaller irrigation targets are often seen with an initial 100 GL requirement, as this delays the start of pumping, retaining early wet-season flows and shortening the water harvest period (Gibbs et al., 2024b). Increasing the pump-start threshold to 1000 ML/day significantly reduces changes in flow dependencies across several asset groups, particularly for fish, sharks, rays and freshwater- dependent habitats (Table 7-2) (see Ponce Reyes et al. (2024) for details on individual assets). Compared to lower thresholds of 200 or 400 ML/day, the improvements are particularly notable beyond 600 ML/day, where the impacts on freshwater-dependent habitats and marine environments are reduced (Figure 7-9) (see Ponce Reyes et al. (2024) for more details). Higher pump-start thresholds appear most effective when system targets are limited to about 200 GL/year and when annual diversion commencement requirements are low or absent, as substantial flows may have already passed through the system before the pump threshold is activated (see Ponce Reyes et al. (2024) for more details). This indicates that optimising the pump- start threshold is crucial for reducing the strain on ecosystems, particularly during periods of low flow. Varying levels of end-of-system (EOS) flow requirements also affect different ecological assets in relation to system targets. For fish, sharks and rays, increasing the EOS requirement from zero to 250 GL/day results in minimal changes in flow dependencies, particularly at lower system targets (~100 GL/year). However, other groups, such as turtles, prawns and other species, experience significant impacts at lower EOS requirements (below 50 GL/day). As the EOS requirement increases, these impacts are mitigated, especially at higher system targets. Flow-dependent habitats and marine environments are also sensitive to lower EOS requirements, with notable reductions in ecological change when the EOS requirement reaches 150 to 250 GL/day. Overall, impacts are more evenly distributed across EOS requirements with the least change occurring when EOS requirements are higher and system targets are kept below 200 GL/year. This suggests that increasing EOS flow requirements provides substantial ecological benefits, particularly for groups sensitive to flow changes, such as turtles, prawns and flow-dependent habitats (see Ponce Reyes et al. (2024) for more details). Finally, limiting pump capacity helps reduce changes in flow dependencies during water harvesting. Slower pump rates and minimum thresholds restrict water extraction during the wet season. Additionally, limiting pump capacity at higher extraction volumes further reduces the total volume extracted and lessens ecological impacts (Gibbs et al., 2024b). Figure 7-8 Lake Moondarra near Mount Isa is used for urban water supply and is a popular water and recreational reserve Photo: CSIRO – Nathan Dyer For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 7-9 Mean change associated with each asset’s important metrics across water harvesting increments of system target and pump-start threshold with no annual diversion commencement flow requirement and pump rate of 30 days Colour intensity represents the mean level of change in important flow dependencies with the scenario given the habitat importance of each node for each asset. 7.4 Biosecurity considerations 7.4.1 Introduction Biosecurity is the prevention and management of pests, weeds and diseases, both terrestrial and aquatic, to limit the risk of detrimental economic, environmental, social and/or cultural impacts. ‘Pests’ is a broad term encompassing pest insects, other invertebrates (e.g. nematodes, mites, molluscs) and vertebrates (e.g. mammals, birds, fish). Weeds broadly include invasive plants and algae. Diseases are caused by pathogens or parasites such as bacteria, fungi and viruses. Any development of the water resources within the catchments of the Southern Gulf for plant industries or aquaculture must take account of biosecurity risks that may threaten production or markets. Development in the study area may also pose broader biosecurity risks to other industries, the environment or communities and these risks must be prevented and/or managed. The catchments of the Southern Gulf extend across Queensland and the NT, so the following sections consider biosecurity policies and regulations for both jurisdictions. Biosecurity practices to protect the Southern Gulf catchments occur at a range of scales. At the national level, the Australian Government imposes quarantine measures to regulate the biosecurity risks associated with entry of goods, materials, plants, animals and people into Australia. The Queensland and NT governments also have biosecurity legislation to limit the entry of new pests, weeds and diseases into their jurisdictions, and to require the control of certain species that are already established. There can also be requirements at the regional level, such as participating in weed management programs (NT Government, 2021; SGNRM and NWQROC, 2022). At the local scale, individual properties ideally follow routine biosecurity protocols, and work with other similar enterprises in implementing industry-wide biosecurity measures. While the Southern Gulf catchments are relatively isolated compared with other regions of Australia, they still have physical connections to the neighbouring regions in Queensland and the NT, across northern Australia more broadly, with the rest of the country and with neighbouring countries such as Indonesia. Examples of such connections are the sharing of specialist cropping machinery between agricultural regions, transport of crop products, tourist visits into remote areas, international trade and tourism, mining exploration, shifting cattle between pastoral properties, army training exercises and movements between Indigenous communities. These connections can be pathways for entry of new pests, weeds or diseases. This section introduces the impacts, spread and management of pests, weeds and diseases of irrigated cropping and aquaculture, as well as invasive species that pose a risk to the Southern Gulf catchments. Given the focus on water-intensive primary industries, biosecurity for terrestrial livestock industries is not included. Impacts of pests, weeds and diseases In primary industries, pests, weeds and diseases can cause economic losses by reducing crop yield and product quality, interfering with farm operations and, loss of market access plus the costs of control measures. The national economic impact of established weeds and vertebrate pests on Australian agriculture has been estimated at over $5.3 billion/year (Hafi et al., 2023). Insect pests are also a substantial economic burden nationally (Bradshaw et al., 2021). The environmental impacts of pests, weeds and diseases, collectively termed ‘invasive species’, include loss of native plants and animals (from competition, predation and infection), degradation of habitats and disruption of ecosystem processes (e.g. changed fire or moisture regimes). Invasive species are the greatest threat to Australia’s threatened flora and fauna (Ward et al., 2021). For example, myrtle rust (Austropuccinia psidii) has potential to cause the extinction of some rare, native myrtaceous shrubs and trees (Makinson et al., 2020). Social impacts of pests, weeds and diseases include loss of public amenity and access to outdoor areas, damage to infrastructure and public safety risks. Cultural impacts include a loss of traditional foods, impaired access for hunting and damage to cultural sites. For example, Gamba grass (Andropogon gayanus) is an African grass originally introduced for pasture in the NT and is now a Weed of National Significance (WoNS). WoNS are nationally agreed weed priorities that have been a focus for prevention and improved management (Hennecke, 2012; CISS, 2021). Gamba grass forms tall, dense stands that burn intensely, posing significant risks to public safety, community and primary industries infrastructure, Indigenous heritage sites, native ecosystems and grazing lands (Setterfield et al., 2013). Pathways of movement Pests, weeds and diseases spread via movement of adults and juveniles (e.g. vertebrate pests), with movement of their hosts (e.g. infected aquaculture broodstock or nursery stock for planting, harvested produce infested with insect larvae) and by movement of propagules (e.g. fungal spores, insect eggs, weed seeds, viral particles). Such movements provide many pathways by which pests, weeds and diseases could be introduced to the Southern Gulf catchments, potentially causing new outbreaks. Just as importantly, there is also the potential for pests, weeds and diseases from the Southern Gulf catchments to spread to other areas in Queensland and the NT, and elsewhere in Australia. Human-mediated spread Human activities are the key means of long-distance and local movement. Pests and propagules, including those within transported soil, can ‘hitchhike’ on or in vehicles, construction and farm machinery, shipping containers and other equipment brought into a region. The ease of movement on vehicles and machinery means that the road network (including access roads to camping areas, railways, pipelines and powerlines) can be a frequent source of new infestations. Propagules may contaminate livestock, seed or nursery stock for establishing crops, hay, road base and landscaping supplies (including turf and ornamental plants). Weed infestations can also arise from invasive garden, crop and pasture plants. Aquatic pests and diseases may become established due to deliberate species release into the environment for fishing, inadvertent transport on fishing equipment or vessels, or dumping of aquarium contents. Incursions of new pests, weeds and diseases from overseas are most likely to occur through contamination of imported goods or containers, or be carried by people (e.g. propagules on shoes or clothing, smuggling of seed or fruit). Natural spread Natural dispersal via wind, water and wild animals usually occurs over short distances. Extreme weather events such as floods and cyclones can disperse pests, weeds and diseases over long distances in addition to causing major environmental disturbances that increase the likelihood of invasive species becoming established. Ebner et al. (2020) flagged the floodplain lowlands of the Gulf of Carpentaria as a high risk for pest fish movement due to a natural and relatively regular connectivity of adjacent catchments during periods of torrential rain. Furthermore, irrigation infrastructure such as dams, pipelines and channels may facilitate distant spread via water movements of some aquatic pests, weeds and diseases, within and across catchments. Some animal pests, such as locusts and fall armyworm (Spodoptera frugiperda), naturally migrate long distances. Northern Australia is close to the southern coasts of Indonesia, Timor-Leste and Papua New Guinea (PNG). These neighbouring countries have a range of serious plant pests and diseases that are not present in Australia, including exotic fruit flies and citrus canker (Xanthomonas citri subsp. citri). The likelihood of their arrival by long-distance wind dispersal is uncertain, particularly with regards to novel atmospheric conditions and extreme weather events occurring under climate change. However, their economic consequences in Australia would be severe were they to establish in Australia. Thus, ongoing biosecurity vigilance in northern Australia through government, industry and community surveillance is vital (DAFF, 2024a; PHA, 2021). 7.4.2 Pest, weed and disease threats to the Southern Gulf catchments The Southern Gulf catchments principally face biosecurity risks from pests, weeds and diseases already present in the catchment and those that occur in neighbouring regions of northern Australia. However, pests, weeds and diseases could also come from other parts of Australia with similar climates and/or production systems, or from overseas. Examples of pests, weeds and diseases that pose a risk to the Southern Gulf catchments are highlighted in the following sections. Whether any one of these would have a significant impact at the property level depends on the local environment, land use and agricultural or aquatic enterprise. However, there is a legal requirement to prevent and manage any pest, weed or disease that is formally ‘declared’ under jurisdictional biosecurity legislation, regardless of its local impact. Furthermore, in Queensland there is a general biosecurity obligation (GBO) under the Biosecurity Act 2014 (Qld) that all persons must take all reasonable and practical steps under their control to not pose a biosecurity risk to others, in all locations (e.g. workplace, home, visiting). This applies regardless of whether a particular pest, weed or disease has been declared as prohibited or restricted matter in Queensland. Plant industries The priority pests and diseases for cropping in the Southern Gulf catchments depend on what is grown. Table 7-5 gives examples of high-impact pest and disease threats to particular crops, and their current distribution and legal status under the Queensland Biosecurity Act and the NT Plant Health Act 2008 (DITT, 2023; Queensland Government, 2022). Information on pests and diseases of plant industries is given on the Queensland and NT governments’ websites (Queensland Government, 2022; NT Government, 2024a). Plant Health Australia is a centralised resource on exotic (i.e. overseas) biosecurity risks to Australia’s plant industries. Research and development corporations, including the Grains Research and Development Corporation, the Cotton Research and Development Corporation, AgriFutures Australia and Hort Innovation also provide extension publications on identifying and managing biosecurity threats. Many pests and diseases have a high host specificity to a particular crop, but there are also generalists that can use many crops as hosts. Local native species can also pose risks of impacts. For example, naturally occurring pathogens of various native wild rices may infect cultivated rice (Chapman et al., 2020) or native animals may graze on crops. Irrigation brings the potential for year-round cropping, which can provide a ‘green bridge’ in the dry season to enable pests or diseases, including native insects and diseases, to persist and increase locally, and to potentially spread to other areas. A significant new generalist pest of cropping is fall armyworm, which has become widely established across northern Australia since a national incursion was detected in 2020. It is likely to be present year round in the Southern Gulf catchments, with a lower incidence in the dry season (PHA, 2020). Fall armyworm caterpillars favour C4 grass crops (e.g. maize, sorghum, rice) and pastures but may also feed on broadleaved crops such as soybean, melon, green bean and cotton. Young crops are most at risk of severe damage and can require immediate insecticide treatment if invaded at levels above the damage threshold. Cucumber green mottle mosaic virus (CGMMV) infects a wide range of cucurbit crops, including various melons, cucumber, pumpkin and squash, and can also be hosted by a range of broadleaved crop weeds. It causes plant stunting and fruit discolouration, malformation and rotting. CGMMV is present on a few farms in Queensland but is more widespread in the NT. It has also been found interstate. Its presence on-farm can make access to interstate markets more difficult as many jurisdictions have imposed quarantine requirements. Infected plants cannot be treated, so preventive farm biosecurity measures are vital (NT Government, 2024a; Queensland Government, 2019). Types of weed threats differ between plant industries according to production methods. For example, annual grain and cotton crops tend to have annual weeds (grasses and herbs) and herbaceous perennials which persist and spread vegetatively via underground rhizomes. Perennial horticulture disturbs the soil less, so typically has more perennial grasses and perennial broadleaved weeds. The highest priority weeds in cropping tend to be those that are most difficult to control, such as herbicide-resistant biotypes or species that are otherwise tolerant to routinely used herbicides. For example, some annual grasses that invade cotton crops have developed resistance to certain herbicides, including barnyard grass (Echinochloa spp.) and feathertop Rhodes grass (Chloris virgata) (CRDC, 2023). Various native vertebrates may consume grain and horticultural crops that are becoming established and damage tree crops. These vertebrate pests include birds (waterfowl, cockatoos), macropods (kangaroos, wallabies) and rodents. Large flocks of magpie geese (Anseranas semipalmata) can be particularly destructive, trampling, grazing and uprooting plants and consuming fruit (Clancy, 2020). Table 7-5 Examples of significant pest and disease threats to plant industries in the Southern Gulf catchments Links to further information are current as of March 2024 BIOSECURITY THREAT AND LEGAL STATUS (NT/QLD) For NT: Plant Health Act 2008 – D = declared; nd = not declared For Queensland: Biosecurity Act 2014 – P = prohibited matter; R = restricted matter; FNP = far northern pest; GBO = general biosecurity obligation CURRENT STATUS CROPS AT RISK INVERTEBRATE PESTS Asian citrus psyllid Diaphorina citri nd/P, GBO Incursion risk from overseas (including Indonesia and PNG) Citrus Cluster caterpillar Spodoptera litura nd/GBO Widespread in northern Australia Cotton, pulses, brassicas Fall armyworm Spodoptera frugiperda nd/GBO Widely established across northern Australia following first detection in 2020 Grasses (cereal and fodder), cotton, soybean, melon, green beans Fruit flies, various species including: Mediterranean fruit fly Ceratitis capitata D/P, GBO Melon fruit fly Bactrocera cucurbitae D/P, GBO Oriental fruit fly Bactrocera dorsalis D/P, GBO New Guinea fruit fly Bactrocera trivialis D/P, GBO Queensland fruit fly Bactrocera tryoni D/GBO Mediterranean fruit fly established in WA. Queensland fruit fly endemic in the NT and Queensland. Melon, oriental, New Guinea and other exotic fruit fly incursion risks from overseas (incl. Indonesia, PNG) and the Torres Strait Fruit and fruiting vegetable crops Bollworms Helicoverpa spp., Pectinophora spp. nd/GBO Widespread in northern Australia Cotton, pulses, brassica, sunflower, forage sorghum, grain sorghum, maize Guava root-knot nematode Meloidogyne enterolobii D/GBO Recent detections in the NT (Darwin region) and Queensland Cucurbits, solanaceous crops, sweet potato, cotton, guava, ginger Leaf miners: American serpentine leaf miner Liriomyza trifolii nd/GBO Serpentine leaf miner Liriomyza huidobrensis nd/GBO Vegetable leaf miner Liriomyza sativae D/FNP, GBO Serpentine and American serpentine leaf miners are recent incursions now present in various locations across Australia. Vegetable leaf miner is present and under control in the far northern biosecurity zone, Queensland Vegetables, cotton Mango pulp weevil Sternochetus frigidus D/GBO Incursion risk from overseas (including Indonesia) Mango Mango shoot looper Perixera illepidaria nd/GBO Recent incursion in Queensland and the NT Mango, lychee Melon thrips Thrips palmi D/GBO Limited presence in the NT north of Alligator River township and in Queensland Vegetables Spur throated locust Austracris guttulosa nd/GBO Native to northern Australia Grasses (cereal and fodder), sunflowers, soybeans, cotton DISEASES Alternaria leaf blight Alternaria alternata nd/GBO Present in northern Australia Cotton Banana freckle Phyllosticta cavendishii D/P, GBO Under eradication in the NT Banana Brown spot Cochliobolus miyabeanus nd/GBO Endemic on wild rices in northern Australia Rice Citrus canker Xanthomonas citri subsp. citri D/P, GBO Eradicated from the NT and Queensland. Incursion risk from overseas (incl. Indonesia, Timor-Leste and PNG) Citrus Cucumber green mottle mosaic virus (CGMMV) nd/R, GBO Present in certain areas in the NT, Queensland and other states Cucurbits Fusarium wilt Fusarium oxysporum f. sp. vasinfectum D/GBO Not present in the NT Cotton Huanglongbing Candidatus Liberibacter asiaticus D/GBO Incursion risk from overseas (including Indonesia, Timor-Leste and PNG) Citrus Panama disease tropical race 4 (Panama TR4) Fusarium oxysporum f. sp. cubense D/GBO Established throughout the NT. Limited distribution in Queensland and mandatory code of practice for its management and control Banana Rice blast Pyricularia oryzae nd/GBO Endemic on wild rices in northern Australia Rice Aquaculture A wide range of diseases and parasites are of concern to Australian aquaculture (DAWE, 2020a), including those not known to be in Australia, those now established (i.e. endemic) in Australia and those native to Australian ecosystems. Barramundi farmers need to consider preventing and managing the biosecurity risks of a range of endemic parasites and viral, bacterial and fungal pathogens that naturally occur in northern Australia (Irvin et al., 2018). In addition, national quarantine measures are vital to prevent exotic disease risks for barramundi from entering Australia (Landos et al., 2019). Prawn aquaculture in northern Australia is most at risk from white spot syndrome virus (WSSV), for which there have been national incursion responses at prawn farms and hatcheries in south- east Queensland and northern NSW. However, there are also many other exotic crustacean diseases (DAWE, 2020a). Endemic viruses (and endemic genotypes of viruses also found overseas) that occur naturally in Australian waters can also trigger mortalities or reduce productivity (Irvin et al., 2018). Invasive species Invasive species, whether pest, weed or disease, are commonly characterised as occurring across multiple land uses in a landscape. Their impacts will vary between land uses, but their coordinated control requires action across all tenures. Weeds Table 7-6 lists regional weed priorities in the Southern Gulf catchments and their legal status under the NT Weed Management Act 2001 and the Queensland Biosecurity Act 2014 (NT Government, 2021; SGNRM and NWQROC, 2022). Of those listed, parthenium weed (Parthenium hysterophorus), Bathurst burr (Xanthium spinosum) and Noogoora burr (X. occidentale) are direct competitors and potential contaminants in dryland and irrigated crops. Parthenium weed also poses a health risk to animals and people as a severe allergen. Many species in Table 7-6 are WoNS. Aquatic weeds can hamper the efficient function of irrigation infrastructure and also cause severe ecological impacts through dense infestations in waterways and wetlands. More-constant water flows from within-stream reservoirs can also change riparian conditions from seasonally ephemeral to perennial, predisposing native vegetation to invasion by weeds that thrive in moist environments. Terrestrial vertebrate pests Various large feral herbivores are present in the Southern Gulf catchments, including buffalo (Bubalus bubalis), horses (Equus caballus), donkeys (E. asinus) and pigs (Sus scrofa). They can directly affect agricultural production through grazing impacts, severe soil erosion and damaged infrastructure such as fencing and irrigation channels. Feral animal damage to habitats is a key disturbance mechanism that facilitates weed invasion, particularly in riparian and wetland areas. Feral pigs in particular are a major threat to irrigated cropping. Their daily water requirement means that they concentrate during the dry season around watercourses and man-made water supplies (Bengsen et al., 2014). Cane toads (Rhinella marina) are already established in the Southern Gulf catchments (Kearney et al., 2008), but would likely become more abundant around irrigation developments where they could access year-round moisture (Bengsen et al., 2014). Aquatic pests Freshwater aquatic pests such as non-native fish, molluscs and crustaceans can affect biodiversity and ecosystem functioning. While these pests may not directly affect irrigated cropping, the associated infrastructure required (e.g. dams, channels, drains) brings increased risk of deliberate release by people for recreational fishing or the disposal of aquarium contents. This infrastructure can also provide enhanced habitat and pathways for the persistence and dispersal of aquatic pests and weeds in the catchment (Ebner et al., 2020). Table 7-6 Regional weed priorities and their management actions in Southern Gulf catchments WEED AND LEGAL STATUS WoNS = Weed of National Significance For NT: Weed Management Act 2001 – D = declared; † = has a statutory management plan under the Act For Queensland: Biosecurity Act 2014 – P = prohibited matter; R = restricted matter; GBO = only the general biosecurity obligation applies REGIONAL ACTION HABITATS AT RISK NT QLD AQUATIC (e.g. river, wetland, dam) WETTER AREAS (e.g. riparian, floodplain, drain) DRIER AREAS (e.g. grassland, woodland) AQUATIC/SEMI-AQUATIC HERB Cabomba Cabomba caroliniana WoNS D/R P P ✓ ✓ Limnocharis Limnocharis flava D/R P P ✓ ✓ Sagittaria Sagittaria platyphylla WoNS D/R P P ✓ ✓ Salvinia Salvinia molesta WoNS D/R P C ‡ ✓ ✓ Water mimosa Neptunia plena D/R P P ✓ ✓ Water hyacinth Pontederia crassipes WoNS D/R P P ✓ ✓ GRASS Aleman grass Echinochloa polystachya -/GBO P P ✓ ✓ Gamba grass Andropogon gayanus WoNS D†/R E P ✓ ✓ Giant rat’s tail grass Sporobolus spp. -/R P P ✓ ✓ Grader grass Themeda quadrivalvis D†/GBO C C ‡ ✓ ✓ Hymenachne Hymenachne amplexicaulis WoNS D/R P P ✓ ✓ Thatch grass Hyparrhenia rufa D/GBO P n ✓ ✓ BROADLEAVED HERB Bathurst burr Xanthium spinosum D/GBO n C ‡ ✓ ✓ Devils claw Martynia annua D/GBO P n ✓ Noogoora burr Xanthium occidentale D/GBO n M ✓ ✓ Parthenium weed Parthenium hysterophorus WoNS D/R P C ‡ ✓ ✓ CLIMBER/VINE Rubber vine Cryptostegia grandiflora WoNS D/R P C ✓ ✓ Ornamental rubber vine Cryptostegia madagascariensis D/R P P ✓ ✓ TREE/SHRUB WEED AND LEGAL STATUS WoNS = Weed of National Significance For NT: Weed Management Act 2001 – D = declared; † = has a statutory management plan under the Act For Queensland: Biosecurity Act 2014 – P = prohibited matter; R = restricted matter; GBO = only the general biosecurity obligation applies REGIONAL ACTION HABITATS AT RISK NT QLD AQUATIC (e.g. river, wetland, dam) WETTER AREAS (e.g. riparian, floodplain, drain) DRIER AREAS (e.g. grassland, woodland) Athel pine Tamarix aphylla WoNS D†/R E C ‡ ✓ Barleria Barleria prionitis D/GBO n C ‡ ✓ ✓ Bellyache bush Jatropha gossypiifolia WoNS D†/R C C ✓ ✓ Calotrope Calotropis procera, C. gigantea D/GBO M M ✓ Chinee apple Ziziphus mauritiana D†/R C C ‡ ✓ ✓ Coffee bush Leucaena leucocephala -/GBO M P ✓ Lantana Lantana camara WoNS D/R P P Madras thorn Pithecellobium dulce -/R n E Mesquite Prosopis spp. WoNS D†/PR E C ‡ ✓ ✓ Mimosa Mimosa pigra WoNS D†/R E P ✓ Neem Azadirachta indica D†/GBO C C ✓ ✓ Parkinsonia Parkinsonia aculeata WoNS D/R M M ✓ ✓ Pond apple Annona glabra WoNS D/R P P ✓ ✓ Prickly acacia Vachellia nilotica WoNS D/R E C ‡ ✓ ✓ Siam weed Chromolaena odorata D/R P P ✓ ✓ Sicklepod Senna obtusifolia, S. hirsuta, S. tora D/R n P ✓ ✓ Yellow bells Tecoma stans -/R n C ‡ ✓ ✓ Yellow oleander Cascabela thevetia -/R M C ‡ ✓ ✓ CACTI/SUCCULENTS Bunny ears Opuntia microdasys WoNS D/R P E ✓ Coral cactus Cylindropuntia fulgida WoNS D/R P C Engelmann’s prickly pear Opuntia engelmannii WoNS D/P P E Harrisia cactus Harrisia martinii, H. tortuosa, H. pomanensis D/R P C ‡ Mother of millions Bryophyllum delagoense -/R P C‡ Other opuntioid cacti Opuntia spp., Cylindropuntia spp., Austrocylindropuntia spp. WoNS D/PR P P For regional actions: P = alert weed for prevention and early intervention E = eradication target (few infestations known) C = strategic control target (control and containment of core infestations, eradication of outlier populations, prevention elsewhere) M = widely established; regional management focused on protecting assets at risk n = not a regional priority in that jurisdiction ‡ Restricted to part of catchment only (SGNRM and NWQROC, 2022); otherwise a P Sources: NT Government (2021); SGNRM and NWQROC (2022); pers. comm. NT and Queensland government officers Table 7-7 lists examples of high-risk pest fish for the Southern Gulf catchments. Certain species are formally declared as ‘noxious’ under the NT Fisheries Regulations 1992, or as prohibited or restricted matter in Queensland under the Biosecurity Act. Those not declared noxious in the NT are still covered by a general precautionary provision that excludes import into the NT and possession of non-native fish that are not on the Australian Government’s list of permitted live freshwater ornamental fish (DAFF, 2023), or otherwise not listed as a permitted import in Schedule 7 of the Regulations. Those not categorised as prohibited or restricted matter in Queensland are still subject to the general biosecurity obligation. Table 7-7 High-risk freshwater pest fish threats to the Southern Gulf catchments PEST FISH AND LEGAL STATUS (NT/QLD) LEGAL STATUS IF ANY (NT/QLD) CURRENT DISTRIBUTION Alligator gar Atractosteus spatula N/R Not known to be in the wild in Australia Black pacu Piaractus brachypomus E/R Not known to be in the wild in Australia. Risk of incursion from PNG Carp Cyprinus carpio N/R N/R Not known to be in the wild in the NT Cichlids, including tilapia: Giant cichlid Boulengerochromis microlepis Jaguar cichlid Parachromis managuensis Pearl cichlid Geophagus brasiliensis Mozambique tilapia Oreochromis mossambicus Nile tilapia Oreochromis niloticus Spotted tilapia Pelmatolapia mariae Texas cichlid Herichthys cyanoguttatus N/R E/GBO E/GBO N/R N/P N/R E/GBO Jaguar, pearl and Texas cichlid and Mozambique and spotted tilapia present in the wild in Queensland Mozambique tilapia and pearl cichlid also present in the wild in WA. Giant cichlid not known to be in the wild in Australia Nile tilapia an incursion risk from northern Torres Strait Climbing perch Anabas testudineus N/R Risk of incursion from northern Torres Strait Gambusia (mosquito fish) Gambusia holbrooki N/R Not known to be established in the NT (eradicated in Darwin and Alice Springs). Recorded in the wild across Queensland and parts of WA Guppy Poecilia reticulata GBO Recorded in Darwin and Nhulunbuy in the NT. Likely to be present elsewhere Marbled lungfish Protopterus aethiopicus N/R Not known to be in the wild in Australia Oriental weatherloach Misgurnus anguillicaudatus N/R Not known to be in the wild in the NT Oscar Astronotus ocellatus GBO Not known to be in the wild in the NT. Present in the wild in Queensland Platy Xiphophorus maculatus GBO Present in the wild in Darwin and Nhulunbuy in the NT, and in eastern Queensland Siamese fighting fish Betta splendens GBO Established in the Adelaide River catchment in the NT. Not known to be in the wild elsewhere in Australia Spotted gar Lepisosteus oculatus N/R Not known to be in the wild in Australia Swordtail Xiphophorus hellerii GBO Present in the wild in Darwin and Nhulunbuy in the NT, and in eastern Queensland For NT: Fisheries Regulations 1992 – N = noxious, E = excluded for import and possession For Queensland: Biosecurity Act 2014 – P = prohibited matter, R = restricted matter, GBO = only the general biosecurity obligation applies Source: NT Government, 2024a.; Queensland Government, 2023b Terrestrial invertebrates Terrestrial invertebrates can also be high-impact invasive species. For example, certain exotic ants can form ‘super colonies’ from which they outcompete native ants, consume native invertebrates, small invertebrates and seeds, and affect people through stinging and infesting buildings. Some ant species ‘farm’ sap-sucking scale insects that are pests of horticultural crops and native plants. Non-native ants can be introduced in pot plants, soil or among other materials. Yellow crazy ant (Anoplolepis gracilipes) is established in parts of eastern Queensland and has also been detected in Darwin and Arnhem Land. Browsing ant (Lepisiota frauenfeldi) has been the subject of a national eradication program, including in Darwin and Kakadu in the NT and at the Port of Brisbane. Other national eradication programs continue for red imported fire ant (RIFA; Solenopsis invicta) in south-east Queensland and electric ant (Wasmannia auropunctata) in far north Queensland (Outbreak, 2024; Environment and Invasives Committee, 2019). The national eradication program for RIFA has cost $596 million (Outbreak, 2023b). Diseases Examples of diseases that affect multiple species of native, ornamental and crop plants are myrtle rust and phytophthora (Phytophthora spp.). They can cause the death of plants, including established shrubs and trees. 7.4.3 Preventing, responding to and managing biosecurity threats Biosecurity can be categorised into three broad approaches: •Prevention - taking measures to stop movement along pathways of spread, whether thatbe at the international or state border, to and within a catchment, or between and withinproperties •Incursion response - undertaking surveillance to detect new pests, weeds or diseases andattempting eradication upon detection, where feasible and cost-beneficial to do so •Ongoing management – managing a pest, weed or disease that is firmly established in anarea (i.e. is not feasible to eradicate), with control measures regularly applied to containfurther spread and/or mitigate impacts. The invasion curve (Figure 7-10) is commonly used as a visual representation of biosecurity actions taken at various stages of pest invasion. It applies at any scale from national down to an individual property. Prevention and eradication generally cost far less than the ongoing management which is needed for widely established species (i.e. containment and impact mitigation), although improved management tools may substantially reduce long-term costs. For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 7-10 The invasion curve with biosecurity actions taken at various stages Source: J. Virtue Shared responsibility at all levels Effective biosecurity requires a collaborative approach between government, industry and community, from the organisational to the individual level. Such ‘shared responsibility’ includes taking action to limit the risk of entry and spread of new pests, weeds and diseases, routinely looking for biosecurity threats and reporting if and when detected, and collaborating in coordinated control programs across land tenures. Everyone has a duty of care (whether legal or moral) to not pose a biosecurity risk to others, including to not harbour invasive species that may threaten economic, environmental, social or cultural impacts on neighbouring land uses. The biosecurity systems of Queensland and the NT (NT Government, 2016; Queensland Gov, 2024a) are nested in the national biosecurity system of Australia’s border quarantine and states’ domestic quarantine and control program arrangements (DAFF, 2022a). Broadly defined, the national biosecurity system consists of the combined Australian, state and territory governments’ biosecurity legislative frameworks that seek to prevent pests, weeds and diseases entering, establishing, spreading and having an impact in Australia. It involves cooperation and collaboration between jurisdictions, and working with and supporting industry and community to involve multiple organisations across Australia as biosecurity partners. Various national agreements, plans and governance arrangements drive this shared responsibility ethos. The following sections describe prevention, incursion response and ongoing management activities for plant industries, aquaculture and invasive species in the Southern Gulf catchments, within local, state and national contexts. Biosecurity in plant industries Farm biosecurity planning In practice, most plant industry biosecurity activities – whether prevention, preparedness, surveillance, elimination, containment or ongoing management – occur at the property level. This level is where the relationship between expenditure on crop protection and maintaining profit- driven productivity and market access is most direct. Developing and implementing a farm biosecurity plan is an effective means to prevent the introduction and establishment of new pests, weeds and diseases, and to limit the spread and impacts of those which are already established. Standard guidance is available on developing such plans (AHA and PHA, n.d.), which cover hygiene practices for pathways of introduction (e.g. certified seed, machinery and vehicle washdowns, restricted movement of visitors), routine surveillance and quick responses to any on-farm identified biosecurity risks. Associated with implementing these are signage (e.g. Figure 7-11), staff training, mapping, visitor management, record keeping, reporting and annual activity planning. A farm biosecurity plan is informed by the key biosecurity threats to the crops being grown and broader invasive species risks. A plan should cover both incursion risks and those pests, weeds and diseases already present. It also needs to align with government regulatory requirements and industry standards. For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 7-11 Farm biosecurity signage available through www.farmbiosecurity.com.au Regulatory prevention Government regulation and policy for plant biosecurity in the Southern Gulf catchments is primarily governed by the Queensland Biosecurity Act or the NT Plant Health Act, depending on where a property is located. It is important to be aware of any differences in legal requirements between the jurisdictions in moving equipment and plant materials across the borders. In Queensland, plant pests, weeds and diseases may be listed in the Biosecurity Act as ‘prohibited’ (not established in the state) or ‘restricted’ (established in parts of the state and subject to strict controls). There are also declared ‘far northern pests’ that pose a risk of entry into Queensland from PNG, via the Torres Strait and Cape York Peninsula. Both declared and non-declared pests, weeds and diseases are also subject to the Act’s general biosecurity obligation. In Queensland, everyone has a general biosecurity obligation and must take all reasonable and practical steps to prevent or minimise biosecurity risk and to minimise the likelihood of causing an adverse biosecurity event and any harmful effects that risk could have, including not doing anything that could make matters worse. The Northern Territory plant health manual provides a consolidated list of all declared plant pests and diseases in that jurisdiction (DITT, 2023), specifying those which must be reported if detected in the NT. Each jurisdiction has entry conditions for all commercial and non-commercial movement of plants and plant products and other potential vectors of plant pests and diseases (DITT, 2023; Queensland Government, 2023c). For example, the NT restricts the entry of maize and soybean seed due to disease risks and the entry of nursery stock due to risk of introducing scale insects and sucking insects. Soil attached to used farm machinery, containers and earth-moving machinery may carry pests or diseases such as nematodes, snails, Phytophthora or Fusarium. Hence these items are legally required to be clean of soil, and a permit may be required for their entry into NT. In relation to used machinery for cotton production, the NT seeks to retain its ‘area freedom’ status for cotton fusarium wilt (Fusarium oxysporum f. sp. vasinfectum), which is established in cotton-growing areas in Queensland and NSW (Le et al., 2020). Examples of Queensland movement restrictions include a hygiene requirement for equipment from interstate that has been used for planting, producing or harvesting a cucurbit crop, to minimise the transmission of CGMMV. There are also restrictions on moving soil or other carriers of RIFA or electric ant (e.g. mulch, compost) out of their respective biosecurity zones within Queensland (Queensland Government, 2023d). To access interstate markets, produce must meet the respective quarantine specifications and protocols, so as not to inadvertently introduce pests or diseases declared in those jurisdictions. This typically requires an inspection and the issue of a certificate verifying that conditions have been met or that the property is in an area known to be free of a specific pest of concern (DITT, 2023). For example, SA has movement restrictions (as of February 2024) on the entry of melons and other hosts of melon thrips (Thrips palmi) from jurisdictions where it is known to occur, including the NT and Queensland (PIRSA, 2024). Current information on moving plant goods interstate is compiled on the Australian Interstate Quarantine website (Subcommittee on Domestic Quarantine and Market Access, 2024). Exports to overseas markets similarly must meet Australian standards and any additional entry requirements from the importing countries for the product (DAFF, 2024b). This includes certification and supporting documentation relating to area freedom and/or treatments applied for specific pests, weeds and/or diseases. Depending on the country, there also may be maximum residue limits, or even nil tolerances, for specific pesticides. Exports are regulated by the Australian Government through the Commonwealth Export Control Act 2020 and associated rules for particular produce and products. Incursion response Most plant industries have national biosecurity, surveillance and/or preparedness plans for high- risk exotic pests and diseases that pose national incursion risks (PHA, 2024a). Their entry into Australia is prevented by the Australian Government’s pre-border and border quarantine requirements under the Commonwealth Biosecurity Act 2015. The Australian Government’s Northern Australia Quarantine Strategy is an ongoing surveillance program that seeks to detect incursions from countries to Australia’s north (DAFF, 2024a). Plant Health Australia (PHA) is the custodian of the Emergency Plant Pest Response Deed (EPPRD; Anon., 2024), which specifies how governments and affected industries undertake collaborative national eradication responses, including cost sharing and decision making. PLANTPLAN provides accompanying national guidelines for managing responses to emergency plant pest incidents at national, state or territory, and local levels (PHA, 2022). For example, banana freckle (Phyllosticta cavendishii) is currently the subject of an EPPRD national eradication program in the NT (Outbreak, 2023a). Ongoing management Best management practice guides for control of established pests, weeds and diseases are available through the research and development corporations, other industry organisations and state primary industries departments, with some specific to cropping in northern Australia (e.g. DPIF, 2014; NT Farmers, 2022). These extension materials focus on integrated management approaches that combine a range of control practices (e.g. chemical, physical and biological control methods). The cotton industry has a broader online best management practice assurance system (myBMP, 2024), which includes modules on integrated pest management and pesticide management. Additionally, the Grains Farm Biosecurity Program is an initiative to improve the management of, and preparedness for, biosecurity risks in the grains industry at the farm and industry levels (PHA, 2024b). Pesticides must be approved for use by the Australian Pesticides and Veterinary Medicines Authority (APVMA) and applied in a manner that aligns with state or territory requirements on the use of such chemicals. This includes minimising spray drift, following label requirements for work health and safety, and ensuring appropriate applicator skills and licences. There are maximum permissible levels for certain pesticides in specified agricultural produce, achieved by following pesticide label (or APVMA permit) requirements regarding approved crops, rates and frequency of application, and withholding periods (NT Government, 2024c; Queensland Government, 2023e). A key consideration for ongoing management on-farm is ensuring chemical control tools are used tactically to limit the risk of developing insecticide, herbicide and fungicide resistance (Grains Research and Development Corporation, 2024; CropLife Australia, 2021). For example, growers cultivating Bollgard® 3 and Roundup Ready Flex® cotton must follow on-farm stewardship packages (Bayer, 2023). Growers whose crop production is affected by native animals may require NT Government permits before taking any lethal control measures. On-property storage of harvested grain needs consideration of physical and chemical means to prevent beetle and weevil pests (Grain Storage Extension Project, 2024). Biosecurity in aquaculture Plan for prevention Prevention in aquaculture starts with enterprise-level biosecurity planning. This is vital in protecting aquaculture facilities from disease and parasite risks, which can be difficult to eliminate, let alone manage, once established. Planning guides have been developed for various industries, including barramundi (Landos et al., 2019) and oyster hatcheries. Preventing entry of pathogens into facilities is vital. Growers need to understand the various disease risks and where they could come from. Wild-captured broodstock poses a very high risk of introducing endemic diseases; stock known to be free from specific diseases should be sourced (Cobcroft et al., 2020). Diseases may enter a facility via contaminated equipment, workers handling diseased fish, water harbouring pathogens, or wild animals such as birds entering ponds (Irvin et al., 2018). Untreated source water is a key pathway for disease entry, with pathogen risks coming from wild stocks or, potentially, from a nearby, upstream aquaculture facility (Irvin et al., 2018). Pathogen monitoring should be ongoing, and emergency response plans should be developed to isolate any detected disease occurrence and implement thorough disinfestation procedures. Commercial aquaculture in the NT is regulated through the NT Fisheries Act 1998 and the Fisheries Regulations 1992. The Regulations prohibit movement or sale of diseased fish and require reporting of any legally notifiable diseases detected in aquaculture facilities. Movement of all stock is under a permit system, and health assessments are conducted to manage the risk of disease movement through movement of aquaculture stock. Similar legal provisions apply in Queensland under the Biosecurity Act and Fisheries Act 1994. Under the Queensland Biosecurity Act’s general biosecurity obligation, aquaculture enterprises are legally obligated to take reasonable and practical measures to minimise the risk and spread of disease. Aquaculture farms are required to have a disease management plan as a standard condition of development approval under the state’s Planning Act 2016. Approval from Fisheries Queensland is required to move live aquatic animals within the state or interstate, and such movements must comply with a specified health protocol (Queensland Government, 2024b). Actual or suspected disease incidents must be reported. Incursion response AQUAPLAN is the national aquatic animal health strategic plan; it aims to improve border, enterprise and regional biosecurity measures, and build surveillance, diagnostic capacity and emergency preparedness (DAFF, 2022b). There is also national policy guidance on minimising the movement of disease when translocating live aquatic animals for aquaculture and other purposes (DAWE, 2020b). The NT Government can declare a control area in the event of an actual or likely notifiable disease outbreak in an aquaculture facility, providing for limits on further fish movement, halting the release of aquaculture water, and/or requiring mandatory treatment or destruction measures for fish and contaminated equipment. Similar biosecurity emergency restrictions and prohibitions can be applied under the Queensland Biosecurity Act. Having aquaculture biosecurity plans is not just about protecting the enterprise. There is also a duty of care to protect nearby wild fisheries which may be exposed to disease from discharge waters, escapee infected animals or fish movement via predatory birds. Prompt isolation of affected ponds and preventing water flow from these to the surrounding environment are vital in the event of a disease outbreak. The escape of white spot syndrome virus from prawn farms in south-east Queensland and northern NSW led to restrictions on commercial and recreational fishing of crustaceans in adjacent catchments, with substantial local economic impacts. Ongoing management Treatment options are limited for aquatic diseases, particularly viral pathogens. Veterinary medicines, such as antibiotics for bacterial disease in barramundi, are available. However, their use can require veterinary permission in order to manage risks of antimicrobial resistance, both in the aquaculture facility itself and the broader food chain. Fungal and external protozoan pathogens may be able to be suppressed using altered salinity bathing. Most fundamental is the need for a high-quality rearing environment, with optimal water conditions and feed supply, to reduce the risk of stress-induced disease outbreaks (Irvin et al., 2018). Biosecurity for invasive species Irrigation development planning Regional and local irrigation and industry infrastructure development, including road networks, should include prevention and management of invasive species in their environmental planning processes. This includes meeting legal obligations under the various Acts mentioned above, a stocktake of present distribution of declared species, and risk mitigation to limit pathways of introduction of new invasive species during construction and ongoing maintenance. Ongoing monitoring should be implemented for both terrestrial and aquatic pests (vertebrate and invertebrate) and weeds. Weeds The Southern Gulf catchments fall within scope of the Katherine regional weeds strategy 2021– 2026 (NT Government, 2021) and the North West Queensland regional biosecurity plan 2022–2027 (SGNRM and NWQROC, 2022). These define priority declared weed control programs within the study area as coordinated by the NT Government or by local governments in Queensland. Under the Queensland Biosecurity Act, all landholders are responsible for taking all reasonable and practical steps to minimise the risks associated with invasive plants under their control, as per the general biosecurity obligation. Prohibited plants are not known to be established in Queensland and must be reported. Restricted plants are high-risk weeds that are established in parts of Queensland, and legal requirements are in place to prevent their further spread, including reporting their presence and/or prohibiting their sale, movement and/or cultivation. Under the NT Weed Management Act, every landholder is legally obliged to take all reasonable measures to prevent land from being infested with a declared weed and to prevent a declared weed from spreading. There are also prohibitions on buying, selling, cultivating, moving or propagating any declared weed, and a legal requirement to notify a declared weed’s presence if it is new to a property. Certain declared plants, such as Gamba grass, neem (Azadirachta indica) and bellyache bush (Jatropha gossypiifolia), also have statutory management plans (Table 7-6), which require specific management actions to be implemented by all landholders. The NT Government website and other Australian websites (e.g. www.weeds.org.au) provide best management practice information on how to control declared weeds and other invasive plants, including registered herbicides and biological control agents. In particular, much information is available on management on WoNS (CISS, 2021). In selecting new crops and pastures for planting in the Southern Gulf catchments, landholders should consider the crops’ weed risks to the surrounding environment. Many northern Australia pasture grasses can be invasive and cause significant biodiversity and cultural impacts in the landscape (DSEWPC, 2012). An example method for considering weed risks is the WA Government’s environmental weed risk assessment process for plant introductions to pastoral lease land (Moore et al., 2022). Cotton is not considered to pose a significant environmental weed threat in northern Australia (Office of the Gene Technology Regulator, 2024). It has been sporadically recorded across northern Australia on roadsides, near cropping fields, in irrigation drains and adjacent to natural watercourses (Atlas of Living Australia, 2024). However, modern varieties’ ability to establish and reproduce is constrained by dense lint around seeds impeding germination, seed predation, seasonal drought, competition from established plants, herbivory and fire (Eastick and Hearnden, 2006; Rogers et al., 2007). Nonetheless, it is recommended that transport of harvested cotton is covered to reduce the likelihood of spread outside cultivation (Addison et al., 2007). Terrestrial vertebrate pests Large feral herbivores are controlled through mustering, trapping, baiting and/or aerial or ground shooting programs, depending on the approved humane control methods for particular species (CISS, 2024; NT Government, 2024d; Queensland Government, 2018). For long-term suppression, programs need to be conducted over multiple years at a subregional scale across all infested properties, taking account of animal movements and sub-populations. Ongoing control is then needed to maintain low densities. 7.5 Off-site and downstream impacts 7.5.1 Introduction Northern Australian river systems are distinctive as they have highly variable flow regimes, unique species composition, low human population densities and, in some cases, naturally high turbidity (Brodie and Mitchell, 2005). Primary influences on groundwater and surface water quality include increased sediment loads associated with land clearing, grazing, agriculture and late dry-season fires, and nutrient pollution from agricultural and pastoral land use (Dixon et al., 2011). These can affect the water quality of not just groundwater and rivers, lakes and wetlands but also estuarine and marine ecosystems. The principal pollutants from agriculture are nitrogen, phosphorus, total suspended solids, herbicides and pesticides (Lewis et al., 2009; Kroon et al., 2016; Davis et al., 2017). Water losses via runoff or deep drainage are the main pathways by which agricultural pollutants enter water bodies. The type and quantity of pollutants lost from an agricultural system and ultimately the quality of the receiving surface water and groundwater is significantly influenced by a wide range of factors, including environment factors such as climate, hydrology, soils, hydro geochemistry and topography as well as land use and management factors such as crops, cropping system, method of application of irrigation water, tailwater management, quality of source water, location and proximity to drainage lines and conservation and irrigation practices. Due to the high dependency of the location, design, implementation and operation of an irrigation development on water quality predicting water quality impacts associated with irrigated agriculture is very difficult. Rather the influence of these environment and management factors on water quality are discussed in more detail in the companion technical report on water quality (Motson et al., 2024). Most of the science in northern Australia concerned with the downstream impacts of agricultural development has been undertaken in the eastern-flowing rivers that flow into the Great Barrier Reef lagoon. Comparatively little research on the topic has been done in the rest of northern Australia and there is need for caution in transposing findings from north-eastern Australia, which is different in terms of climate, geomorphology and patterns of settlement to those parts of northern Australia west of the Great Dividing Range. Nonetheless experience from north-eastern Australia has been that the development of agriculture has been associated with declining water quality (Lewis et al., 2009; Mitchell et al., 2009; De’ath et al., 2012; Waterhouse et al., 2012; Thorburn et al., 2013; Kroon et al., 2016). Pollutant loads in north-eastern Australian rivers (typically those in which agriculture dominates as a land use) are estimated to have increased considerably since European settlement in the 1850s for nitrogen (2 to 9 times baseline levels), phosphorus (3 to 9 times), suspended sediment (3 to 6 times) and pesticides (~17,000 kg) (Kroon et al., 2016). 7.5.2 Managing irrigation drainage Degraded water quality can cause a loss of aquatic habitat, biodiversity, and ecosystem services. Increased nitrogen and phosphorus can cause plankton blooms and weed infestation, increase hypoxia (low oxygen levels) and result in fish deaths. Pesticides, used to increase agricultural productivity, can harm downstream aquatic ecosystems, flora and fauna. As with fertiliser nutrients, pesticides can enter surface water bodies and groundwater via infiltration, leaching, and runoff from rainfall events and irrigation. These chemicals can be toxic to non-target species, such as aquatic life and humans, affecting nervous systems, immune systems, photosynthesis and growth (Cantin et al., 2007; Kaur et al., 2019; Naccarato et al., 2023). They can be carcinogenic (Mohanty and Jena, 2019) and cause multiple sub-lethal effects that can disrupt the ecological balance of aquatic systems and degrade aquatic communities (Giglio and Vommaro, 2022; Miller et al., 2020; Wang et al., 2022). Other water quality variables that can have a significant effect on the health of aquatic species, communities and ecosystems include salinity, pH, and suspended sediments. Increased salinity, indicated by increased EC and TDS, can interfere with osmoregulatory processes, harming those species not adapted to saline conditions (Hart et al., 2003). Variations in the pH of a water body can negatively affect an organism’s biochemical processes, leading to altered behaviour, functioning, growth, and even survival (US EPA, 2024). In aquatic ecosystems, elevated loads of suspended sediment can smother habitats and benthic invertebrates, affect the feeding and respiratory systems of aquatic species, and reduce light penetration, affecting photosynthetic activity (Chapman et al., 2017). Table 7-8 provides a summary of how changes in key water quality variables affect aquatic ecology and human health. Table 7-8 Water quality variables reviewed – their impacts on the environment, aquatic ecology and human health WATER QUALITY VARIABLE THREATS TO AQUATIC ECOLOGY AND HUMAN HEALTH REFERENCE Nutrients Nitrogen Forms of nitrogen in freshwater systems include: nitrate (NO3), nitrite (NO2), ammonia (NH3) and ammonium (NH4). In excessive quantities, contributes to eutrophication and algal blooms, which can deplete oxygen and create hypoxic/anoxic conditions harmful to aquatic life. Health threat to humans, particularly infants, and mammals Carpenter et al. (1998) Phosphorus High concentrations may lead to eutrophication and algal blooms, which can deplete oxygen and create hypoxic/anoxic conditions Mainstone and Parr (2002) WATER QUALITY VARIABLE THREATS TO AQUATIC ECOLOGY AND HUMAN HEALTH REFERENCE Salinity Can affect osmoregulatory processes of aquatic species, harming aquatic life not adapted to saline water. Significant increases in salinity may compromise the integrity of freshwater ecosystems Hart et al. (2003) Other Total Suspended Solids Can smother habitats, reduce light penetration (through increasing turbidity), and affect the feeding and respiratory systems of aquatic organisms Chapman et al. (2017) pH Variations can negatively affect aquatic life stages, affecting their biochemical processes. Preferred pH range of 6.4–8.4 for aquatic species US EPA(2024) harmful to aquatic life Dissolved Organic Carbon A proxy for dissolved organic matter, affecting water clarity, temperature, biogeochemical processes, food webs and ecosystem productivity. Dissolved Organic Carbon may exacerbate eutrophication and hypoxia in aquatic ecosystems, and cause problems in drinking water treatment processes Palviainen et al. (2022) Pesticide groups Arylurea Includes pesticides such as Diuron® and tebuthiuron. May inhibit photosynthesis in plants and aquatic species. These pesticides are less soluble in water and better absorbed by the soil Cantin et al. (2007), Fojut et al. (2012) Carbamates Broad-spectrum pesticides that affect nerve impulse transmission and are highly toxic to vertebrate species. Suspected carcinogens and mutagens. Relatively low persistence; not easily adsorbed to soil particles Kaur et al. (2019), Rad et al. (2022) Chloroacetanilides Affects cell division, disrupting aquatic plant growth; also toxic to aquatic insects. Persistent. Low binding affinity to soil particles but highly water soluble; therefore, it has a high capacity for leaching into the groundwater and ending up in surface water. Carcinogens with moderate to high chronic toxicity ANZG (2020), Mohanty and Jena (2019) Dinitroanilines Broad-spectrum herbicides with low water solubility; considered non- mobile in soil. Affect seed germination and root growth in plants. Variable, species-specific toxicity ranging from slightly to highly acute. Hazardous to animals and humans in sub-lethal concentrations. Known bioaccumulation in and acute toxicity to aquatic organisms Giglio and Vommaro (2022) Neonicotinoids Highly toxic to invertebrates, particularly aquatic insects. Sub-lethal toxicity in fish. High solubility. High chronic risk to global freshwater ecosystems. Suspected to be carcinogenic Wang et al. (2022) Organochlorines Persistent organic pollutants that can bioaccumulate in fatty tissues. These pesticides are toxic to humans and other animals, and they are highly toxic to most aquatic life DCCEEW (2021) Organophosphates Broad-spectrum pesticides that control a wide range of pests via multiple functions. Organophosphate insecticides are toxic to both vertebrates and invertebrates, disrupting nerve impulse transmission Kaur et al. (2019) Phenylpyrazole These pesticides disrupt nerve impulse transmission. Toxic to aquatic organisms and birds. Phenylpyrazole pesticides, such as Fipronil, have been found to degrade stream communities. Moderate water solubility and hydrophobicity. Slightly mobile in soils. Moderate persistence Gao et al. (2020), Miller et al. (2020) Triazine Inhibits photosynthesis in plants, potentially leading to reduced plant growth and blocks food intake by insect pests. Short to moderate persistence depending on soil pH. Adverse and sub-lethal effects on terrestrial and aquatic non-target organisms, affecting growth and the nervous and immune systems Naccarato et al. (2023) 7.5.3 Impacts of irrigated agriculture on water quality Fertiliser applications in irrigated agriculture can significantly affect nutrient levels in drainage waters, leading to increased concentrations of total phosphorous and total nitrogen in surface waters during the irrigation season (Barbieri et al., 2021; Mosley and Fleming, 2010). The type of cropping system employed also plays a crucial role in determining groundwater nutrient concentrations. For example, variations in cropping practices, such as mulch-till versus ridge-till systems, can result in substantial differences in nitrate levels, underscoring the importance of adopting best management practices for protecting groundwater quality (Albus and Knighton, 1998). Surface water quality is similarly affected by nutrient inputs, with concentrations of total phosphorous often decreasing as streamflow increases, suggesting a dilution effect (Skhiri and Dechmi, 2012). However, this relationship can be inconsistent, as dilution effects may not be evident when only storm event streamflow is considered. Instead, total phosphorous concentrations are influenced by a combination of factors, including rainfall duration and intensity, as well as irrigation and fertiliser application practices. The interplay of these factors highlights the complex interactions between rainfall, irrigation, and nutrient management in determining both surface water and groundwater quality outcomes. Controlled pesticide use is crucial for managing its impact on surface water quality. When pesticide application rates are managed and irrigation schedules are aligned with crop growth stages, their concentrations are typically low. Pesticide-specific application practices also influence runoff concentrations: pesticides that are applied to, and therefore intercepted by, the crop canopy have significantly lower surface water concentrations relative to those applied to bare soil (Moulden et al., 2006). Seasonal hydrology, particularly ‘first-flush’ events following irrigation or significant rainfall, plays a critical role in determining water quality (Davis et al., 2013; Yeates, 2016). Studies have shown that pesticide concentrations in runoff are highest following initial irrigation events but decrease in subsequent events (Davis et al., 2013). Despite this dilution, pesticide concentrations in receiving waters can still exceed recommended levels. Similarly, nitrogen concentrations in runoff are often higher following early-season rainfall, when crops have not yet fully absorbed available nitrogen, leading to increased transport in runoff (Yeates, 2016). These findings underscore the importance of implementing sustainable irrigation management practices and highlight the need for continuous monitoring and adaptive management to mitigate the impacts of agricultural activities on water quality. Ensuring effective management is vital for protecting water resources and maintaining the ecological integrity of aquatic ecosystems and communities amid agricultural intensification. The potential for irrigated agriculture to cause secondary salinisation is discussed in Section 7.6. Managing irrigation drainage Surface drainage water is water that runs off irrigation developments as a result of over-irrigation or rainfall. This is mostly an issue where water is applied using surface irrigation methods (e.g. furrow, flood) rather than spray or micro-irrigation methods (e.g. drip, micro-spray). This excess water can potentially affect the surrounding environment by modifying flow regimes and changing water quality. Hence, management of irrigation or agricultural drainage waters is a key consideration when evaluating and developing new irrigation systems and should be given careful consideration during the planning and design processes. Regulatory constraints on the disposal of agricultural drainage water from irrigated lands are being made more stringent as this disposal can potentially have significant off-site environmental effects (Tanji and Kielen, 2002). Hence, minimising drainage water through the use of best-practice irrigation design and management should be a priority in any new irrigation development in northern Australia. This involves integrating sound irrigation systems, drainage networks and disposal options so as to minimise off- site impacts. Surface drainage networks must be designed to cope with the runoff associated with irrigation, and also the runoff induced by rainfall events on irrigated lands. Drainage must be adequate to remove excess water from irrigated fields in a timely manner and hence reduce waterlogging and potential salinisation, which can seriously limit crop yields. In best-practice design, surface drainage water is generally reused through a surface drainage recycling system where runoff tailwater is returned to an on-farm storage or used to irrigate subsequent fields within an irrigation cycle. The quality of drainage water depends on a range of factors including water management and method of application, soil properties, method and timing of fertiliser and pesticide application, hydrogeology, climate and drainage system (Tanji and Kielen, 2002). These factors need to be taken into consideration when implementing drainage system water recycling and also when disposing of drainage water to natural environments. A major concern with tailwater drainage is the agricultural pollutants derived from pesticides and fertilisers that are generally associated with intensive cropping and are found in the tailwater from irrigated fields. Crop chemicals can enter surface drainage water if poor water application practices or significant rainfall events occur after pesticide or fertiliser application (Tanji and Kielen, 2002). Thus, tailwater runoff may contain phosphate, organic nitrogen and pesticides that have the potential to adversely affect flora and fauna and ecosystem health, on land and in waterways, estuaries or marine environments. Tailwater runoff may also contain elevated levels of salts, particularly if the runoff has been generated on saline surface soils. Training irrigators in responsible application of both water and agrochemicals is therefore an essential component of sustainable management of irrigation. As tailwater runoff is either discharged from the catchment or captured and recycled, it can result in a build-up of agricultural pollutants that may ultimately require disposal from the irrigation fields. In externally draining basins, the highly seasonal nature of flows in northern Australia does offer opportunities to dispose of poor-quality tailwater during high-flow events. However, downstream consequences are possible, and no scientific evidence is available to recommend such disposal as good practice. Hence, consideration should be given to providing an adequate understanding of the downstream consequences of disposing of drainage effluent, and options must be provided for managing disposal that minimise impacts on natural systems. 7.5.4 Natural processing of water contaminants While elevated contaminants and water quality parameters can harm the environment and human health, there are several processes by which aquatic ecosystems can partially process contaminants and regulate water quality. Denitrification, for example, is the process of anaerobic microbial respiration which, in the presence of carbon, reduces nitrogen to nitrous oxide and dinitrogen gas (Martens, 2005). Therefore, denitrification is a naturally occurring process that can remove and reduce nitrogen concentrations within a water body. Pesticides can also be naturally removed from water via chemical oxidation, microbial degradation, or ultraviolet photolysis, although some chemically stable pesticides are highly persistent, and their microbial degradation is slow (Hassaan and El Nemr, 2020). Phosphorus, however, does not have a microbial reduction process equivalent to denitrification. Instead, if it is not temporarily taken up by plants, phosphorus can be adsorbed onto the surface of inorganic and organic particles and stored in the soil, or deposited in the sediments of water bodies, such as wetlands (Finlayson, 2006). This phosphorus can be remobilised into solution and re-adsorbed, resulting in ‘legacy’ phosphorus that can affect water quality for many years (Records et al., 2016). 7.5.5 Water quality considerations relevant to aquaculture Aquaculture can be impacted by poor water quality and can also contribute to poor water quality unless aquaculture operations are well managed. A summary is provided below, however, for more information see Northern Australia Water Resource Assessment report on aquaculture viability (Irvin et al., 2018). Chemical contaminant risk to aquaculture and the environment Hundreds of different chemicals, including oils, metals, pharmaceuticals, fertilisers and pesticides (i.e. insecticides, herbicides, fungicides), are used in different agricultural, horticultural and mining sectors, and in industrial and domestic settings, throughout Australia. Releasing these chemical contaminants beyond the area of target application can contaminate soils, sediments and waters in nearby environments. In aquatic environments, including aquaculture environments, fertilisers have the potential to cause non-point source pollution. Eutrophication is caused by nutrients that trigger excessive growth of plant and algal species, which then form hypoxic ‘dead zones’ and potentially elevated levels of toxic un-ionised ammonia (Kremser and Schnug, 2002). This can have a significant impact on the health and growth of animals in aquaculture operations, as well as in the broader environment. Of concern to aquaculture in northern Australia are the risks posed to crustaceans (e.g. prawns and crabs) by some of the insecticides in current use. These are classified based on their specific chemical properties and modes of action. The different classes of insecticides have broad and overlapping applications across different settings. The toxicity of organophosphate insecticides is not specific to target insects, raising concerns about the impacts on non-target organisms such as crustaceans and fish. Despite concerns about human health impacts and potential carcinogenic risks, organophosphates are still one of the most broadly used types of insecticide globally, and they are still used in Australia for domestic pest control (Weston and Lydy, 2014; Zhao and Chen, 2016). Pyrethroid insecticides have low toxicity to birds and mammals, but higher toxicity to fish and arthropods. Phenylpyrazole insecticides also pose risks to non-target crustaceans (Stevens et al., 2011). Neonicotinoid insecticides are being used in increasing amounts because they are very effective at eliminating insect pests, yet they pose low risks to mammals and fish (Sánchez-Bayo and Hyne, 2014). Monitoring data from the Great Barrier Reef catchments indicate that the concentration of neonicotinoid insecticides in marine water samples is rapidly increasing with widespread use. One significant concern for aquaculture is the risk that different insecticides, when exposed to non-target organisms, may interact to cause additive or greater-than-additive toxicity. Aquaculture discharge water and off-site impacts Discharge water is effluent from land-based aquaculture production (Irvin et al., 2018). It is water that has been used (culture water) and is no longer required in a production system. In most operations (particularly marine), bioremediation is used to ensure that water discharged off-farm into the environment contains low amounts of nutrients and other contaminants. The aim is for discharge waters to have similar physiochemical parameters to the source water. Discharge water from freshwater aquaculture can be easily managed and provides a water resource suitable for general or agriculture-specific irrigation. Discharge water from marine aquaculture is comparatively difficult to manage and has limited reuse applications. The key difference in discharge management is that marine (salty) water must be discharged at the source, whereas the location for freshwater discharge is less restrictive and potential applications (e.g. irrigation) are numerous. Specific water discharge guidelines vary with aquaculture species and jurisdiction. For example, Queensland water discharge policy minimum standards for prawn farming include standards for physiochemical indicators (e.g. oxygen and pH) and nutrients (e.g. nitrogen, phosphorus and suspended solids) and total volume (EHP, 2013). The volume of water required to be discharged or possibly diverted to a secondary application (e.g. agriculture) is equivalent to the total pond water use for the season minus total evaporative losses and the volume of recycled water used during production. A large multidisciplinary study of intensive Australian prawn farming, which assessed the impact of effluent on downstream environments (CSIRO, 2013), found that Australian farms operate under world’s best practice for the management of discharge water. The study found that discharge water had no adverse ecological impact on the receiving environment and that nutrients could not be detected 2 km downstream from the discharge point. While Australian prawn farms are reported as being among the most environmentally sustainable in the world (CSIRO, 2013), the location of the industry adjacent to the World Heritage–listed Great Barrier Reef and related strict policy on discharge has been a major constraint to the industry’s expansion. Strict discharge regulation, which requires zero net addition of nutrients in waters adjacent to the Great Barrier Reef, has all but halted expansion in the last decade. An example of the regulatory complexity in this study area is the 14-year period taken to obtain approval to develop a site in the Burdekin shire in north Queensland (APFA, 2016). In a report to the Queensland Government (Department of Agriculture and Fisheries Queensland, 2013), it was suggested that less-populated areas in northern Australia, which have less conflict for the marine resource, may have potential as areas for aquaculture development. The complex regulatory environment in Queensland was a factor in the decision by Project Sea Dragon to investigate greenfield development in WA and NT as an alternative location for what would be Australia’s largest prawn farm (Seafarms, 2016). Today, most farms (particularly marine) use bioremediation ponds to ensure that water discharged off-farm into the environment contains low amounts of nutrients and other contaminants. The prawn farming industry in Queensland has adopted a code of practice to ensure that discharge waters do not result in irreversible or long-term impacts on the receiving environment (Donovan, 2011). 7.6 Irrigation-induced salinity 7.6.1 Introduction Salts occur naturally in all soils and landscapes. The amount of salt present depends on local climate, salt store in the geology, landscape position, soil type, the depth to the watertable and the quality of the groundwater below the watertable. Naturally occurring areas of salinity, or ‘primary salinity’, occur in the landscape, with ecosystems adapted to the saline conditions. Any change to landscape hydrology, including clearing and irrigation, can mobilise salts, resulting in environmental degradation in the form of ‘secondary salinity’. Secondary salinity manifests itself in two main forms: irrigation-induced salinity and land clearing induced salinity. In the case of irrigation-induced salinity, an increase in drainage below the root zone following applications of irrigation water can raise watertables and bring salts to the soil surface. Excessive drainage of irrigation water below the root zone tends to more likely occur in coarser-textured soils, (Petheram et al., 2002). In Australia, over-irrigation (or overwatering) due to poor irrigation practices, together with leakage of water from irrigation distribution networks and drainage channels, has caused the watertable level to rise in many irrigated areas. Significant parts of all major irrigation areas have watertables that are now close to the land surface (in the vicinity of 2 to 3 m from the land surface) (Christen and Ayars, 2001). Salts can concentrate in the root zone over time as a result of evaporation, which can be a compounding factor. In addition, the process by which the salts accumulate in the root zone is accelerated if the groundwater has high salt concentrations. In the Assessment area, some clay soils do have high levels of salt within 1 to 2 m of the surface, depending on their location. Clay soils that have slowly or very slowly permeable subsoils (SGG 9) can accumulate salts in the subsoil over time, even in areas receiving high rainfall during the wet season. Caution should be applied in the irrigation of poorly drained clay soils (SGG 9). The cracking clay soils on the Armraynald Plain, particularly the black soils along the Gregory River backplain, have subsoils that are high in salt and susceptible to irrigation-induced secondary salinity. No evidence of surface salt was found by this survey or past surveys (Christian et al., 1954; Perry et al., 1964) of the Southern Gulf catchments, apart from on or near the marine plains (Karumba Plains) (see Figure 7-12). However, development of irrigation in the Assessment area is in its infancy. It should be noted that this section on irrigation-induced salinity provides general information regarding soils suitable for irrigation development. The risk of secondary salinisation at a specific location in the Southern Gulf catchments can only be properly assessed by undertaking detailed field investigations at a local scale. 7.6.2 Potential sources of salt Salt stores in soils The salts in the soil and landscape in the Southern Gulf catchments are derived from rainfall, wind, the weathering of primary minerals, and former marine sediments in the substrate. The amount of salt in the landscape (‘salt store’) depends on the origins of the salts, the degree of geological weathering, the climate (particularly the rainfall and the prevailing wind direction), position in the landscape, landscape permeability (soils and rock), and watertable dynamics. The Southern Gulf catchments have extensive natural surface salinity, on the extratidal salt plains along the coast of the Gulf of Carpentaria (Karumba Plain; see Section 2.2.2 and Figure 7-12). These plains extend up to 35 km inland and are not considered suitable for irrigation development. For more information on this figure, please contact CSIRO on enquiries@csiro.au Figure 7-12 An extratidal flat on the Karumba Plain near Burketown, with white salt crystals either side of a shallow natural drainage line. Salt-tolerant samphire and mangrove species can be seen in the background. Photo: CSIRO The grey clay soils (SGG 9) north of the Nardoo – Burketown Road on the Armraynald Plain have a high risk of salinity if irrigated, due to their proximity to the salty tidal creeks of the Karumba Plain. Drainage from irrigation could cause the watertable to rise, intercepting some of the tidal creeks and causing a spread of salt beyond the creeks. The risk is much less south of the Nardoo– Burketown Road. Figure 7-13 Example of salt-affected areas on the northern part of the Armraynald Plain Photo: Esri, Maxar, Earthstar Geographics, and the GIS User Community The black clay soils (SGG 9) on the backplain of the Gregory River contain salts at depths of more than 1.5 m. The salts are likely to be derived from the weathering of the underlying Pleistocene marine and terrestrial sediments and from cyclic (rain and wind-blown) sources. If these soils are irrigated, there is a risk that the watertable will rise and carry this salt up to within the plant root zone. The grey clay soils (SGG 9) around Gregory and the brown clay soils on the east of the Leichhardt River on the Armraynald Plain have much less salt in the subsoil, and thus the risk of irrigation- induced salinity is lower. The clay soils (SGG 9) on the Barkly Tableland have low subsoil salt levels. Where they are underlain by porous limestone and dolomite, a build-up of salts due to irrigation is not expected. The sandy soil (SGG 6), loamy soil (SGG 4) and sand or loam over friable brown, yellow and grey clay soils (SGG 1.2) on the Doomadgee Plain also have negligible salts within the soil profile. However, due to other risk factors, care would need to be exercised when clearing the silver box, bloodwood and broad-leaf paperbark savanna landscapes for rainfed or irrigated cropping. Groundwater aquifers contained by underlying ferricrete, the likelihood of soils having variable depths, and the very gently undulating plain make it difficult to manage irrigation water discharge on lower slopes and in drainage depressions, causing salts to accumulate in these areas over time. In places where these soils are shallow, it would be necessary to monitor the depths of rising watertables and manage irrigation rates accordingly. In addition, excessive irrigation rates are likely to have off-site impacts in the long term, as the lateral flow of water can ‘wick’ from the lower slopes in these landscapes to form scalds. From these scalds, salts can potentially be mobilised towards nearby streams. Irrigation water as a potential source of salt In many irrigation developments around the world, poor-quality irrigation water is the source of the salt in secondary salinisation. In the Southern Gulf catchments, however, the river water is relatively fresh, and the aquifers with potential for groundwater resource development are also relatively fresh (see Section 2.5.2), so are unlikely to be a source of salt. However, there is a risk that in some cases certain levels of localised groundwater extraction from aquifers could result in the entrainment of poorer quality water from the surrounding aquifers or aquitards over time, thereby reducing the overall quality of the groundwater applied to crops. The potential for this occurring would require a local, site-specific investigation. 7.6.3 Rise in watertable level and changes in groundwater discharge due to irrigation development The extent to which the watertable rises close to the surface depends on: •the initial depth to the watertable •the amount of groundwater recharge (from root zone drainage) •the size of the irrigation area (which dictates the total volume of water added to the landscape) •the lateral distance to rivers (which act as a drainage boundary, thus reducing the height of thegroundwater mound under irrigation) •aquifer parameters, including saturated hydraulic conductivity, aquifer thickness and specificwater yield. Generic modelling results evaluating the risk of watertable rise are documented in the Flinders and Gilbert Agricultural Resource Assessment technical report on surface water – groundwater connectivity (Jolly et al., 2013). Site-specific assessments need to be undertaken to determine the extent of the secondary salinity risk. 7.7 References Addison SJ, Farrell T, Roberts GN and Rogers DJ (2007) Roadside surveys support predictions of negligible naturalisation potential for cotton (Gossypium hirsutum) in north‐east Australia. Weed Research, 47(3), 192–201. DOI:10.1111/j.1365-3180.2007.00561.x. AHA (Animal Health Australia) and PHA (Plant Health Australia) (n.d.) Farm biosecurity action planner. AHA and PHA, Canberra. Viewed 26 February 2024, https://www.farmbiosecurity.com.au/toolkit/planner/. Albus WL and Knighton RE (1998) Water quality in a sand plain after conversion from dryland to irrigation: tillage and cropping systems compared. Soil and Tillage Research 48(3), 195–206. Anon. (2024) Government and Plant Industry Cost Sharing Deed in respect of Emergency Plant Pest Responses. Viewed 26 February 2024, https://www.planthealthaustralia.com.au/biosecurity/emergency-plant-pest-response- deed/. ANZG (2020) Toxicant default guideline values for aquatic ecosystem protection: Metolachlor in freshwater. Australian and New Zealand Guidelines for Fresh and Marine Water Quality. Viewed 30 September 2024, https://www.waterquality.gov.au/sites/default/files/documents/metolachlor_fresh_dgv- technical-brief_0.pdf. APFA (2016) Productivity Commission inquiry into the regulation of Australian marine fisheries and aquaculture sectors. Viewed 27 September 2024, Productivity Commission Inquiry (2016) into the Regulation of Australian Marine Fisheries and Aquaculture Sectors. Submission from Australian Prawn Farmers Association (APFA) Ash A, Bristow M, Laing A, MacLeod N, Niscioli A, Paini D, Palmer J, Prestwidge D, Stokes C, Watson I, Webster T and Yeates S (2018) Agricultural viability: Darwin catchments. 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. Atlas of Living Australia (2024) Atlas of Living Australia: occurrence records. https://support.ala.org.au/support/home Barbieri MV, Peris A, Postigo C, Moya-Garcés A, Monllor-Alcaraz LS, Rambla-Alegre M, Eljarrat E and López de Alda M (2021) Evaluation of the occurrence and fate of pesticides in a typical Mediterranean delta ecosystem (Ebro River Delta) and risk assessment for aquatic organisms. Environmental Pollution 274, 115813. DOI: 10.1016/j.envpol.2020.115813. Bayer (2023) Growing cotton in Northern Australia: 2023/24 grower guide. Viewed 26 February 2024, https://www.crop.bayer.com.au/products/biotechnology-traits/bollgard-3-with- roundup-ready-flex-cotton#tab-4. Bengsen AJ, Gentle MN, Mitchell JL, Pearson HE and Saunders GR (2014) Impacts and management of wild pigs Sus scrofa in Australia. Mammal Review 44(2),135–147. DOI:10.1111/mam.12011. Bradshaw CJA, Hoskins AJ, Haubrock PJ, Cuthbert RN, Diagne C, Leroy B, Andrews L, Page B, Cassey P, Sheppard AW and Courchamp F (2021) Detailed assessment of the reported economic costs of invasive species in Australia. NeoBiota 67, 511–550. DOI:10.3897/neobiota.67.58834. 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. Bunn SE and Arthington AH (2002) Basic principles and ecological consequences of altered flow regimes for aquatic biodiversity. Environmental Management 30(4), 492-507. Burdekin Project Committee (1977) Resources and potential of the Burdekin River Basin, Queensland. Main report and 5 volumes of appendices. Australian Government Publishing Service, Canberra. Burford M, Valdez D, Curwen G, Faggotter S, Ward D and Brien KO (2016) Inundation of saline supratidal mudflats provides an important source of carbon and nutrients in an aquatic system. Marine Ecology Progress Series 545, 21-33. Burford M, Webster I, Revill A, Kenyon R, Whittle M and Curwen G (2012) Controls on phytoplankton productivity in a wet–dry tropical estuary. Estuarine, Coastal and Shelf Science 113, 141-151. Burford MA and Faggotter SJ (2021) Comparing the importance of freshwater flows driving primary production in three tropical estuaries. Marine Pollution Bulletin 169, 112565. Burford MA, Kenyon RA, Whittle M and Curwen G (2010) Flow and fisheries: River flow impacts on estuarine prawns in the Gulf of Carpentaria. FRDC Project 2007/003 Final Report. Griffith University and CSIRO Marine and Atmospheric Research, Brisbane. Burford MA, Revill AT, Palmer DW, Clementson L, Robson BJ and Webster IT (2011) River regulation alters drivers of primary productivity along a tropical river-estuary system. Marine and Freshwater Research 62(2), 141-151. DOI: 10.1071/MF10224. Burrows DW and Butler B (2007) Determining end-point goals and effective strategies for rehabilitation of coastal wetlands: examples from the Burdekin River, north Queensland. In: Wilson AL, Dehaan RL, Watts RJ, Page KJ, Bowmer KH and Curtis A (eds) Proceedings of the 5th Australian Stream Management Conference. Charles Sturt University, Thurgoona, New South Wales, 49–54. Cantin NE, Negri AP and Willis BL (2007) Photoinhibition from chronic herbicide exposure reduces reproductive output of reef-building corals. Marine Ecology Progress Series 344, 81–93. Carpenter SR, Caraco NF, Correll DL, Howarth RW, Sharpley AN and Smith VH (1998) Nonpoint pollution of surface waters with phosphorus and nitrogen. Ecological Applications 8(3), 559– 568. DOI: 10.2307/2641247. Chapman B, Henry R, Wurm P, Bellairs S, Crayn D, Smyth H, Furtado A, Sivapalan S, Ford R and Matchett T (2020) A situational analysis for developing a rice industry in Northern Australia. Project A.1.1718120, CRC for Developing Northern Australia (CRCNA). Viewed 14 February 2024, https://crcna.com.au/projects/northern-australia-rice-industry-situational-analysis- study/ Chapman PM, Hayward A and Faithful J (2017) Total suspended solids effects on freshwater lake biota other than fish. Bulletin of Environmental Contamination and Toxicology 99(4), 423– 427. DOI: 10.1007/s00128-017-2154-y. Christen EW and Ayars J (2001) Subsurface drainage system design and management in irrigated agriculture: best management practices for reducing drainage volume and salt load. Technical report 38/01. CSIRO Land and Water, Griffith NSW. Hyperlink to 'Subsurface drainage system design and management in irrigated agriculture: best management practices for reducing drainage volume and salt load' . Viewed 5 August 2024. Christian CS, Noakes LC, Perry RA, Slatyer RO, Stewart GA and Traves DM (1954) Survey of the Barkly Region, Northern Territory and Queensland, 1947–48. Land Research Series No. 3. CSIRO, Melbourne. CISS (Centre for Invasive Species Solutions) (2024) Management toolkits. pestSMART website, CISS. Viewed 27 February 2024, https://pestsmart.org.au/pest-animals/. CISS (Centre for Invasive Species Solutions) (2021) Profiles for Weeds of National Significance. Weeds Australia website, CISS. Viewed 13 March 2024, https://weeds.org.au/weeds- profiles/. Clancy TF (2020) Wildlife management program for the magpie goose (Anseranas semipalmata) in the Northern Territory of Australia 2020–2030. Northern Territory Department of Environment and Natural Resources. Viewed 20 February 2024, https://depws.nt.gov.au/programs-and-strategies/magpie-goose-management. 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. Close P, Dobbs R, Tunbridge D, Speldewinde P, Warfe D, Toussaint S and Davies P (2014) Customary and recreational fishing pressure: large-bodied fish assemblages in a tropical, intermittent Australian river. Marine and Freshwater Research 65(4), 466-474. Cobcroft J, Bell R, Diedrich A, Jerry D and Fitzgerald J (2020) Northern Australia aquaculture industry situational analysis: project A.1. 1718119. CRC for Developing Northern Australia (CRCNA). Viewed 6 February 2024, https://crcna.com.au/projects/northern-australia- aquaculture-industry-situational-analysis-study/ CRDC (Cotton Research and Development Corporation) (2023) 2023–24 Cotton pest management guide. CRDC. Viewed 19 February 2024, https://www.crdc.com.au/publications/cotton-pest- management-guide. 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. CropLife Australia (2021) Resistance management strategies: insecticides, fungicides, herbicides 2021–2022. CropLife Australia. Viewed 20 March 2024, https://www.croplife.org.au/resources/programs/resistance-management-strategies/. CSIRO (2013) The environmental management of prawn farming in Queensland – worlds best practice. Viewed 27 September 2024, https://era.daf.qld.gov.au/id/eprint/2064/2/1._Prawn_farming_environmental_research_1995-2002-sec.pdf DAFF (Department of Agriculture, Fisheries and Forestry) (2022a) National biosecurity strategy. DAFF, Canberra. Viewed 26 February 2024, https://www.biosecurity.gov.au/about/national- biosecurity-committee/nbs. DAFF (Department of Agriculture, Fisheries and Forestry) (2022b) AQUAPLAN 2022–2027: Australia’s national strategic plan for aquatic animal health. DAFF, Canberra. Viewed 26 February 2024, https://www.agriculture.gov.au/agriculture-land/animal/aquatic/aquaplan. DAFF (Department of Agriculture, Fisheries and Forestry) (2023) Permitted live freshwater ornamental fish suitable for import. DAFF, Canberra. Viewed 11 March 2024, https://www.agriculture.gov.au/biosecurity-trade/policy/legislation/list-of-permitted-live- freshwater-and-marine-ornamental-fish#daff-page-main. DAFF (Department of Agriculture, Fisheries and Forestry) (2024a) Northern Australia quarantine strategy (NAQS). DAFF website. Viewed 11 March 2024, https://www.agriculture.gov.au/biosecurity-trade/policy/australia/naqs. DAFF (Department of Agriculture, Fisheries and Forestry) (2024b) Exporting from Australia. DAFF website. Viewed 20 March 2024, https://www.agriculture.gov.au/biosecurity- trade/export/from-australia. Davis A, Thorburn P, Lewis S, Bainbridge Z, Attard S, Milla R and Brodie J (2013) Environmental impacts of irrigated sugarcane production: herbicide run-off dynamics from farms and associated drainage systems. Agriculture, Ecosystems & Environment 180, 123–135. Davis AM, Pearson RG, Brodie JE and Butler B (2017) Review and conceptual models of agricultural impacts and water quality in waterways of the Great Barrier Reef catchment area. Marine and Freshwater Research 68, 1–19. DOI: 10.1071/MF15301. DAWE (Department of Agriculture, Water and the Environment) (2020a) Aquatic animal diseases significant to Australia: identification field guide. 5th edn. DAWE, Canberra. Viewed 10 March 2024, https://www.agriculture.gov.au/agriculture-land/animal/aquatic/guidelines- and-resources/aquatic_animal_diseases_significant_to_australia_identification_field_guide. DAWE (Department of Agriculture, Water and the Environment) (2020b) National policy guidelines for the translocation of live aquatic animals. DAWE, Canberra. Viewed 26 February 2024, https://www.agriculture.gov.au/agriculture-land/animal/aquatic/biosecurity-plan- guidelines#national-policy-guidelines-for-the-translocation-of-live-aquatic-animals. DCCEEW (2021) Organochlorine pesticides (OCPs) – trade or common use names. Australian Government. Viewed 30 September 2024, https://www.dcceew.gov.au/environment/protection/publications/ocp-trade-names. De’ath G, Fabricius KE, Sweatman H and Puotinen M (2012) The 27-year decline of coral cover on the Great Barrier Reef and its causes. Proceedings of the National Academy of Sciences of the United States of America 109, 17995–17999. DOI: 10.1073/pnas.1208909109. Department of Agriculture and Fisheries Queensland (2013) About biosecurity. Department of Agriculture and Fisheries. Queensland. Department of Agriculture‚ Water and the Environment (2021) Directory of Important Wetlands in Australia. Available online: . DITT (Department of Industry, Tourism and Trade) (2023) Northern Territory plant health manual. Plant Biosecurity Branch, DITT, NT. Viewed 14 February 2024, https://nt.gov.au/industry/agriculture/food-crops-plants-and-quarantine/plants-and- quarantine/introduction. Dixon I, Dobbs R, Townsend S, Close P, Ligtermoet E, Dostine P, Duncan R, Kennard M and Tunbridge D (2011) Trial of the Framework for the Assessment of River and Wetland Health (FARWH) in the wet–dry tropics for the Daly and Fitzroy Rivers, Tropical Rivers and Coastal Knowledge (TRaCK) research consortium. Charles Darwin University, Darwin. Donovan D (2011) Environmental code of practice for Australian prawn farmers. Viewed 2 June 2023, Hyperlink to: Environmental Code of Practice for Australian Prawn Farmers . 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. DPIF (Department of Primary Industry and Fisheries) (2014) Field guide to pests, beneficials, diseases and disorders of vegetables in northern Australia. DPIF, Darwin. Viewed 20 March 2024, https://industry.nt.gov.au/__data/assets/pdf_file/0009/227772/vegfieldguide-sml.pdf. DSEWPC (Department of Sustainability, Environment, Water, Population and Communities) (2012) Threat abatement plan to reduce the impacts on northern Australia’s biodiversity by the five listed grasses. DSEWPC, Canberra. Viewed 27 February 2024, https://www.dcceew.gov.au/environment/biodiversity/threatened/publications/threat- abatement-plan-reduce-impacts-northern-australias-biodiversity-five-listed-grasses . Duggan M, Connolly R, Whittle M, Curwen G and Burford M (2014) Effects of freshwater flow extremes on intertidal biota of a wet-dry tropical estuary. Marine Ecology Progress Series 502, 11-23. Duke NC, Field C, Mackenzie JR, Meynecke J-O and Wood AL (2019) Rainfall and its possible hysteresis effect on the proportional cover of tropical tidal-wetland mangroves and saltmarsh–saltpans. Marine and Freshwater Research 70(8), 1047-1055. DOI: https://doi.org/10.1071/MF18321. Dulvy NK, Davidson LN, Kyne PM, Simpfendorfer CA, Harrison LR, Carlson JK and Fordham SV (2016) Ghosts of the coast: global extinction risk and conservation of sawfishes. Aquatic Conservation: Marine and Freshwater Ecosystems 26(1), 134-153. Eastick RJ and Hearnden MN (2006) Potential for weediness of Bt cotton in northern Australia. Weed Science, 54(6), 1142–1151. DOI:10.1614/WS-06-077R.1. 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), James Cook University, Cairns. Viewed 20 March 2024, https://era.daf.qld.gov.au/id/eprint/8276/. EHP (2013) Operational policy marine prawn aquaculture. Department of Environment and Heritage Protection, Queensland Government. Viewed 2 June 2023, Hyperlink to: Operation policy marine prawn aquaculture Environment and Invasives Committee (2019) National invasive ant biosecurity plan 2018–2028. Australian Government, Canberra. Viewed 23 February 2024, https://www.dcceew.gov.au/environment/biodiversity/threatened/publications/tap/invasive-ants. Finlayson CM, Lowry J, Bellio MG, Nou S, Pidgeon R, Walden D, Humphrey C and Fox G (2006) Biodiversity of the wetlands of the Kakadu Region, northern Australia. Aquatic Sciences 68(3), 374–399. Fleming PM, Gunn RH, Reece AM, and McAlpine JR (eds) (1981) Burdekin project ecological study. CSIRO and Department of National Development and Energy, Canberra. Fojut TL, Palumbo AJ and Tjeerdema RS (2012) Aquatic life water quality criteria derived via the UC Davis method: III. Diuron. In: Tjeerdema RS (ed) Aquatic life water quality criteria for selected pesticides. Reviews of Environmental Contamination and Toxicology, vol. 216. Springer, Boston, USA, 105–141. Viewed 30 September 2024, https://doi.org/10.1007/978- 1-4614-2260-0_3. Gao J, Wang F, Jiang W, Miao J, Wang P, Zhou Z and Liu D (2020) A full evaluation of chiral phenylpyrazole pesticide flufiprole and the metabolites to non-target organism in paddy field. Environmental Pollution, 264, 114808. DOI: 10.1016/j.envpol.2020.114808. Garnett S (1987) Aerial Surveys of Waders (Aves: Charadriiformes) Along the Coast of North- eastern Australia. Aus. Wildl. Res. 14, 521-528. Gibbs M, Hughes J and Yang A (2024a) River model calibration for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Gibbs M, Hughes J, Yang A, Wang B, Marvanek S and Petheram C (2024b) River model scenario analysis for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Giglio A and Vommaro ML (2022) Dinitroaniline herbicides: a comprehensive review of toxicity and side effects on animal non-target organisms. Environmental Science and Pollution Research 29(51), 76687–76711. DOI: 10.1007/s11356-022-23169-4. Grain Storage Extension Project (2024) Stored grain information hub. Stored Grain Project website. Viewed 29 February 2024, https://storedgrain.com.au/. Grains Research and Development Corporation (2024) WeedSmart website. Viewed 20 March 2024, https://www.weedsmart.org.au/. Hafi A, Arthur T, Medina M, Warnakula C, Addai D and Stenekes N (2023) Cost of established pest animals and weeds to Australian agricultural producers. Australian Bureau of Agricultural and Resource Economics and Sciences (ABARES), Canberra. Viewed 5 February 2024, https://www.agriculture.gov.au/abares/research-topics/biosecurity/biosecurity- economics/cost-of-established-pest-animals-and-weeds-to-australian-agricultural-producers Hart BT, Lake PS, Webb JA and Grace MR (2003) Ecological risk to aquatic systems from salinity increases. Australian Journal of Botany 51(6), 689–702. Hassaan MA and El Nemr A (2020) Pesticides pollution: classifications, human health impact, extraction and treatment techniques. Egyptian Journal of Aquatic Research 46(3), 207–220. DOI: 10.1016/j.ejar.2020.08.007. Hennecke BR (2012) Assessing new Weeds of National Significance candidates. Proceedings of the 18th Australasian Weeds Conference. Weed Society of Victoria Inc., Melbourne. Viewed 20 March 2024, https://caws.org.nz/18th-australasian-weeds-conference-2012/. Hunt R, Jardine T, Hamilton S and Bunn S (2012) Temporal and spatial variation in ecosystem metabolism and food web carbon transfer in a wet-dry tropical river. Freshwater Biology 57(3), 435-450. DOI: 10.1111/j.1365-2427.2011.02708.x. Irvin S, Coman G, Musson D, Doshi A and Stokes C (2018) Aquaculture viability. 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. Viewed 14 February 2024, https://publications.csiro.au/publications/publication/PIcsiro:EP181756. Irvin S, Coman G, Musson D, Doshi A, Stokes C (2018). Aquaculture viability. 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. Australia: CSIRO; 2018. csiro:EP181756. https://doi.org/10.25919/5b86ed759d698 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. Jellyman D, Gehrke PC and Harris JH (2016) Australasian freshwater fisheries. In: Craig JF (ed.), Freshwater Fisheries Ecology. John Wiley & Sons, Ltd., Oxford. Jolly I, Taylor A, Rassam D, Knight J, Davies P and Harrington G (2013) Surface water — groundwater connectivity. . 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. 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. Kaur R, Mavi GK, Raghav S and Khan I (2019) Pesticides classification and its impact on environment. International Journal of Current Microbiology and Applied Science 8(3), 1889– 1897. Kearney M, Phillips BL, Tracy CR, Christian KA, Betts G and Porter WP (2008) Modelling species distributions without using species distributions: the cane toad in Australia under current and future climates. Ecography 31(4), 423–434. DOI:10.1111/j.0906-7590.2008.05457.x. Kennard MJ, Mackay SJ, Pusey BJ, Olden JD and Marsh N (2010) Quantifying uncertainty in estimation of hydrologic metrics for ecohydrological studies. River Res. Applic. 26, 137-156. Kingsford R and Norman F (2002) Australian waterbirds—products of the continent’s ecology. Emu 102(1), 47-69. Kremser U and Schnug E (2002) Impact of fertilizers on aquatic ecosystems and protection of water bodies from mineral nutrients. Landbauforsch Völkenrode 52(2), 81–90. Kroon FJ, Thorburn P, Schaffelke B and Whitten S (2016) Towards protecting the Great Barrier Reef from land-based pollution. Global Change Biology 22, 1985–2002. DOI: 10.1111/gcb.13262. Kutt A, Felderhof L, VanDerWal J, Stone P and Perkins G (2009) Chapter 4. Terrestrial ecosystems of northern Australia. Northern Australia Land and Water Science Review, October 2009. Landos M, Calogeras C, Ruscoe J-A, Hayward S and Ruscoe J (2019) National biosecurity plan guidelines for Australian barramundi Farms. Department of Agriculture, Canberra. Viewed 20 February 2024, https://www.agriculture.gov.au/agriculture- land/animal/aquatic/biosecurity-plan-guidelines#barramundi-farms. Le DP, Tran TT, Gregson A and Jackson R (2020) TEF1 sequence-based diversity of Fusarium species recovered from collar rot diseased cotton seedlings in New South Wales, Australia. Australasian Plant Pathology 49(3), 277–284. Hyperlink to: TEF1 sequence-based diversity of Fusarium species recovered from collar rot diseased cotton seedlings in New South Wales, Australia . Leahy SM and Robins JB (2021) River flows affect the growth of a tropical finfish in the wet-dry rivers of northern Australia, with implications for water resource development. Hydrobiologia 848(18), 4311-4333. 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), 17071. Lear KO, Morgan DL, Whitty JM, Beatty SJ and Gleiss AC (2021) Wet season flood magnitude drives resilience to dry season drought of a euryhaline elasmobranch in a dry-land river. Science of The Total Environment 750, 142234. Leigh C and Sheldon F (2008) Hydrological changes and ecological impacts associated with water resource development in large floodplain rivers in the Australian tropics. River Res. Applic. 24(2008), 1251-1270. Lewis SE, Brodie JE, Bainbridge ZT, Rohde KW, Davis AM, Masters BL, Maughan M, Devlin MJ, Mueller JF and Schaffelke B (2009) Herbicides: a new threat to the Great Barrier Reef. Environmental Pollution 157, 2470–2484. DOI: 10.1016/j.envpol.2009.03.006. Lowe V, Frid CLJ, Venarsky M and Burford MA (2022) Responses of a macrobenthic community to seasonal freshwater flow in a wet-dry tropical estuary. Estuarine, Coastal and Shelf Science 265, 107736. DOI: https://doi.org/10.1016/j.ecss.2021.107736. Mainstone CP and Parr W (2002) Phosphorus in rivers — ecology and management. Science of the Total Environment 282–283, 25–47. DOI: 10.1016/S0048-9697(01)00937-8. Makinson RO, Pegg GS and Carnegie AJ (2020) Myrtle rust in Australia – a national action plan. Australian Plant Biosecurity Science Foundation, Canberra. Viewed 11 March 2024, https://www.apbsf.org.au/myrtle-rust/. Marchant S and Higgins P (1990) Handbook of Australian, New Zealand and Antarctic Birds. Volume 1, Ratites to ducks. Oxford University Press, Melbourne. Marsh N, Sheldon F, Wettin P, Taylor C and Barma D (2012) Guidance on ecological responses and hydrological modelling for low-flow water planning, March 2012. The National Water Commission, Canberra. Martens DA (2005) Denitrification. In: Hillel D and Hatfield JL (eds) Encyclopedia of soils in the environment (Vol. 3). Elsevier, 378–382. DOI: 10.1016/B0-12-348530-4/00138-7. McGinness HM (2016) Waterbird responses to flooding, stressors and threats. A report prepared for the Murray-Darling Freshwater Research Centre as part of the Environmental Water Knowledge and Research Project. CSIRO, Canberra, 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. McJannet D, Yang A and Seo L (2023) Climate data characterisation for hydrological and agricultural scenario modelling across the Victoria, Roper and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. McMahon TA and Finlayson BL (2003) Droughts and anti-droughts: the low flow hydrology of Australian rivers. Freshwater Biology 48(2003), 1147–1160. Merrin L, Stratford D, Kenyon R, Pritchard J, Linke S, Ponce Reyes R, Buckworth R, Castellazzi P, Costin B, Deng R, Gannon R, Gao S, Gilbey S, Lachish S, McGinness H and Waltham N (2024) Ecological assets of the Southern Gulf catchments to inform water resource assessments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Miller JL, Schmidt TS, Van Metre PC, Mahler BJ, Sandstrom MW, Nowell LH, Carlisle DM and Moran PW (2020) Common insecticide disrupts aquatic communities: A mesocosm-to-field ecological risk assessment of fipronil and its degradates in U.S. streams. Science Advances 6(43), eabc1299. DOI: 10.1126/sciadv.abc1299. Mitchell A, Reghenzani J, Faithful J, Furnas M and Brodie J (2009) Relationships between land use and nutrient concentrations in streams draining a wet-tropics catchment in northern Australia. Marine and Freshwater Research 60, 1097–1108. DOI: 10.1071/MF08330. Mitsch WJ, Bernal B and Hernandez ME (2015) Ecosystem services of wetlands. Taylor & Francis. Mohanty SS and Jena HM (2019) A systemic assessment of the environmental impacts and remediation strategies for chloroacetanilide herbicides. Journal of Water Process Engineering 31, 100860. DOI: 10.1016/j.jwpe.2019.100860. Moore GA, Munday C and Barua P (2022) Environmental weed risk assessment protocol for growing non-indigenous plants in the Western Australian rangelands. Department of Primary Industries and Regional Development, Perth. Viewed 20 March 2024, https://library.dpird.wa.gov.au/bulletins/279/. Morgan D, Whitty J, Allen M, Ebner B, Keleher J, Gleiss A and Beatty S (2016) Wheatstone Environmental Offsets-Barriers to sawfish migrations. A report for Chevron Australia and the Western Austalian Marine Science Institution. Freshwater Fish Group & Fish Health Unit, Centre for Fish & Fisheries Research, Murdoch University. Morgan DL, Somaweera R, Gleiss AC, Beatty SJ and Whitty JM (2017) An upstream migration fought with danger: freshwater sawfish fending off sharks and crocodiles. Ecology 98(5), 1465-1467. Mosley LM and Fleming N (2010) Pollutant loads returned to the Lower Murray River from flood- irrigated agriculture. Water, Air, and Soil Pollution 211(1–4), 475–487. DOI: 10.1007/s11270- 009-0316-1. Motson K, Mishra A and Waltham N (2024) A review of water quality studies relevant to northern Australia. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Moulden JH, Yeates SJ, Strickland GR and Plunkett GM (2006) Developing an environmentally responsible irrigation system for cotton in the Ord River Irrigation Area. Proceedings of ANCID 2006, 16–19 October, Darwin, NT. The Australian National Committee on Irrigation and Drainage, Sydney. myBMP (2024) myBMP website. Viewed 26 February 2024, https://www.mybmp.com.au/. Naccarato A, Vommaro ML, Amico D, Sprovieri F, Pirrone N, Tagarelli A and Giglio A (2023) Triazine herbicide and NPK fertilizer exposure: accumulation of heavy metals and rare earth elements, effects on cuticle melanization, and immunocompetence in the model species Tenebrio molitor. Toxics 11(6), 499. 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. Nielsen D, Merrin L, Pollino C, Karim F, Stratford D and O’Sullivan J (2020) Climate change and dam development: Effects on wetland connectivity and ecological habitat in tropical wetlands. Ecohydrology 2020, 13pgs. DOI: 10.1002/eco.2228. 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. Nilsson C and Berggren K (2000) Alterations of riparian ecosystems caused by river regulation. BioScience 50(9), 783-792. Novak P, Bayliss P, Crook DA, Garcia EA, Pusey BJ and MM D (2017) Do upstream migrating, juvenile amphidromous shrimps, provide a marine subsidy to river ecosystems? Freshwater Biology 62(5), 880-893. DOI: 10.1111/fwb.12907. NT Farmers (2022) Northern Australia broadacre cropping manual. Northern Territory Farmers Association. Viewed 26 February 2024, https://ntfarmers.org.au/wp- content/uploads/2023/01/NA-Broadacre-Cropping-Manual-Final-Digital.pdf. NT Government (2016) Northern Territory biosecurity strategy 2016–2026. NT Government. Viewed 30 January 2024, https://industry.nt.gov.au/industries/primary- industry/agriculture/agricultural-developments/nt-biosecurity-strategy-2016-2026. NT Government (2021) Katherine Regional Weeds Strategy 2021–2026. NT Government Department of Environment, Parks and Water Security. Viewed 2 January 2024, Hyperlink to: Katherine Regional Weeds Strategy 2021-2026 . NT Government (2024a) Plant diseases and pests. NT.GOV.AU website. Viewed 27 February 2024, https://nt.gov.au/industry/agriculture/food-crops-plants-and-quarantine/plant-diseases- and-pests/pests-and-diseases. NT Government (2024b) Cucumber green mottle mosaic virus (CGMMV). NT.GOV.AU website. Viewed 19 February 2024, https://nt.gov.au/industry/agriculture/food-crops-plants-and- quarantine/cucumber-green-mottle-mosaic-virus/cucumber-green-mottle-mosaic-virus. NT Government (2024c) Using chemicals responsibly. NT.GOV.AU website. Viewed 26 February 2024, https://nt.gov.au/industry/agriculture/farm-management/using-chemicals- responsibly/managing-chemical-residues. NT Government (2024d) Feral animals. NT.GOV.AU website. Viewed 27 February 2024, https://nt.gov.au/environment/animals/feral-animals. Nye ER, Dickman CR and Kingsford RT (2007) A wild goose chase-temporal and spatial variation in the distribution of the Magpie Goose (Anseranas semipalmata) in Australia. Emu 107(1), 28- 37. DOI: 10.1071/MU05012. Office of the Gene Technology Regulator (2024) The Biology of Gossypium hirsutum L. and Gossypium barbadense L. (cotton). Viewed 2 September 2024, https://www.ogtr.gov.au/resources/collections/biology-documents Olden JD and Poff L (2003) Redundancy and the choice of hydrologic indices for characterizing streamflow regimes. River Res. Applic. 19(2003), 101-121. Outbreak (2023a) Banana freckle (Phyllosticta cavendishii). Outbreak Animal and Plant Pests and Diseases website. Viewed 20 March 2024, https://www.outbreak.gov.au/current- outbreaks/banana-freckle. Outbreak (2023b) Red imported fire ant (Solenopsis invicta), Outbreak Animal and Plant Pests and Diseases website. Viewed 10 March 2024, https://www.outbreak.gov.au/current- outbreaks/red-imported-fire-ant. Outbreak (2024) Respond to, prevent and prepare for animal and plant pest and disease outbreaks. Outbreak Animal and Plant Pests and Diseases website. Viewed 23 February 2024, https://www.outbreak.gov.au/. Palviainen M, Peltomaa E, Laurén A, Kinnunen N, Ojala A, Berninger F, Zhu X and Pumpanen J (2022) Water quality and the biodegradability of dissolved organic carbon in drained boreal peatland under different forest harvesting intensities. Science of the Total Environment 806, 150919. DOI: 10.1016/j.scitotenv.2021.150919. Pelicice FM, Pompeu PS and Agostinho AA (2015) Large reservoirs as ecological barriers to downstream movements of Neotropical migratory fish. Fish and Fisheries 2015(16), 697-715. Perna C (2003) Fish habitat assessment and rehabilitation in the Burdekin Delta distributary streams. Report 03/22. Australian Centre for Tropical Freshwater Research, James Cook University, Townsville. Perna C (2004) Impacts of agriculture and restoration of the habitat values, water quality and fish assemblages of a tropical floodplain. MSc Thesis, James Cook University, Townsville. Perry RA, Sleeman JR, Twidale CR, Prichard CE, Slatyer RO, Lazarides M and Collins FH (1964) General report on lands of the Leichhardt–Gilbert area, Queensland, 1953–54. In Land Research Series No. 11. CSIRO, Melbourne. Petheram C, McMahon TA and Peel MC (2008) Flow characteristics of rivers in northern Australia: Implications for development. Journal of Hydrology 357(1), 93-111. DOI: https://doi.org/10.1016/j.jhydrol.2008.05.008. Petheram C, Rustomji P, Chiew FHS and Vleeshouwer J (2012) Rainfall–runoff modelling in northern Australia: A guide to modelling strategies in the tropics. Journal of Hydrology 462- 463, 28-41. DOI: https://doi.org/10.1016/j.jhydrol.2011.12.046. Petheram C, Walker G, Grayson R, Thierfelder T and Zhang L (2002) Towards a framework for predicting impacts of land-use on recharge: 1. A review of recharge studies in Australia. Australian Journal of Soil Research 40(3), 397–417. DOI: 10.1071/SR00057. PHA (Plant Health Australia) (2020) Fall armyworm continuity plan for the Australian grains industry. Version 1, November 2020. PHA, Canberra. Viewed 2 January 2024, https://www.planthealthaustralia.com.au/fall-armyworm/. PHA (Plant Health Australia) (2021) Exotic pest identification and surveillance guide for tropical horticulture. Version 1.0, February 2021. PHA, Canberra. Viewed 20 March 2024, https://www.planthealthaustralia.com.au/biosecurity/surveillance/exotic-pest- identification-and-surveillance-guide-for-tropical-horticulture/. PHA (Plant Health Australia) (2022) PLANTPLAN: Australian emergency plant pest response plan. PHA, Canberra. Viewed 26 February 2024, https://www.planthealthaustralia.com.au/biosecurity/incursion-management/plantplan/. PHA (Plant Health Australia) (2024a) Industries. PHA website. Viewed 26 February 2024, https://www.planthealthaustralia.com.au/industries/. PHA (Plant Health Australia) (2024b) Protecting grains from diseases, pests and weeds. Grains farm biosecurity program website. Viewed 28 March 2024, https://grainsbiosecurity.com.au/. Phillips NM, Chaplin JA, Morgan DL and Peverell SC (2011) Population genetic structure and genetic diversity of three critically endangered Pristis sawfishes in Australian waters. Marine Biology 158(4), 903-915. DOI: 10.1007/s00227-010-1617-z. PIRSA (South Australian Department of Primary Industries and Regions ) (2024) Plant quarantine standard South Australia. Version 17.4. PIRSA, Adelaide. Viewed 26 February 2024, https://pir.sa.gov.au/biosecurity/plant_health/plant_quarantine_standard_and_updated_conditions. Plagányi É, Kenyon R, Blamey L, Robins J, Burford M, Pillans R, Hutton T, Hughes J, Kim S and Deng RA (2024) Integrated assessment of river development on downstream marine fisheries and ecosystems. Nature Sustainability 7(1), 31-44. Poff NL and Zimmerman JK (2010) Ecological responses to altered flow regimes: a literature review to inform the science and management of environmental flows. Freshwater Biology 55(1), 194-205. Poff NL, Olden JD, Merritt DM and Pepin DM (2007) Homogenization of regional river dynamics by dams and global biodiversity implications. PNAS 104(4), 5732-5737. Poiani A (2006) Effects of floods on distribution and reproduction of aquatic birds. Advances in Ecological Research, Vol 39 39, 63-83. DOI: 10.1016/s0065-2504(06)39004-6. Pollino C, Barber E, Buckworth R, Cadiegues M, Deng R, Ebner B, Kenyon R, Liedloff A, Merrin L, Moeseneder C, Morgan D, Nielsen D, O’Sullivan J, Ponce Reyes R, Robson B, Stratford D, Stewart-Koster B and Turschwell M (2018) 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, Canberra. Ponce Reyes R, Stratford D, Linke S, Merrin L, Kenyon R, Buckworth R, Deng RA, Hughes J, McGinness H, Pritchard J, Seo L and Waltham N (2024) Assessment of the potential ecological outcomes of water resource development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Pyšek P, Hulme PE, Simberloff D, Bacher S, Blackburn TM, Carlton JT, Dawson W, Essl F, Foxcroft LC, Genovesi P, Jeschke JM, Kühn I, Liebhold AM, Mandrak NE, Meyerson LA, Pauchard A, Pergl J, Roy HE, Seebens H, van Kleunen M, Vilà M, Wingfield MJ and Richardson DM (2020) Scientists’ warning on invasive alien species. Biological Reviews 2020(95), 1511–1534. DOI: 10.1111/brv.12627. Queensland Government (2018) Controlling invasive animals. Business Queensland website. Viewed 29 February 2024, https://www.business.qld.gov.au/industries/farms-fishing- forestry/agriculture/biosecurity/animals/invasive/control. Queensland Government (2019) Cucumber green mottle mosaic virus. Business Queensland website. Viewed 27 February 2024, https://www.business.qld.gov.au/industries/farms- fishing-forestry/agriculture/biosecurity/plants/priority-pest-disease/cucumber-green- mottle-mosaic-virus. Queensland Government (2022) Priority plant pests and diseases. Business Queensland. Viewed 27 February 2024, https://www.business.qld.gov.au/industries/farms-fishing- forestry/agriculture/biosecurity/plants/priority-pest-disease. Queensland Government (2023a) Clearing codes. Viewed 4 July 2024, https://www.qld.gov.au/environment/land/management/vegetation/clearing- approvals/codes/choose#:~:text=The%20most%20common%20example%20of,to%20clear% 20in%20these%20areas. Queensland Government (2023b) Your legal obligation for invasive freshwater animals. Queensland Department of Agriculture and Fisheries. Viewed 4 July 2024, https://www.daf.qld.gov.au/__data/assets/pdf_file/0008/1398842/prohibited-restricted- invasive-fish.pdf . Queensland Government (2023d) Restrictions on moving plant material, soil and related equipment. Business Queensland website. Viewed 28 March 2024, https://www.business.qld.gov.au/industries/farms-fishing- forestry/agriculture/biosecurity/plants/moving/restrictions. Queensland Government (2023e) Using chemicals safely and effectively. Business Queensland website. Viewed 29 February 2024, https://www.business.qld.gov.au/industries/farms- fishing-forestry/agriculture/sustainable/chemical/use/safety. Queensland Government (2023c) Biosecurity Manual: Version 21.0. Business Queensland website. Viewed 26 February 2024, https://www.business.qld.gov.au/industries/farms-fishing- forestry/agriculture/biosecurity/plants/moving/restrictions/into-qld. Queensland Government (2024a) Biosecurity. Business Queensland website. Viewed 29 February 2024, https://www.business.qld.gov.au/industries/farms-fishing- forestry/agriculture/biosecurity. Queensland Government (2024b) Disease in aquaculture farms. Business Queensland website. Viewed 28 March 2024, https://www.business.qld.gov.au/industries/farms-fishing- forestry/fisheries/aquaculture/disease. Rad SM, Ray AK and Barghi S (2022) Water pollution and agriculture pesticide. Clean Technologies 4(4), 1088–1102. Read M, Wright B and Enoch C (2004) Crocodiles in Queensland- An Overview. Crocodiles:Proceedings of the 17th Working Meeting of the Crocodile Specialist Group, IUCN- The World Conservation Union. Gland, Switzerland and Cambridge UK. Records RM, Wohl E and Arabi M (2016) Phosphorus in the river corridor. Earth-Science Reviews 158, 65–88. DOI: 10.1016/j.earscirev.2016.04.010. Richter BD, Baumgartner JV, Powell J and Braun DP (1996) A method for assessing hydrologic alteration within ecosystems. Conservation biology 10(4), 1163-1174. 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. Department of Agriculture and Fisheries. Brisbane. Rogers DJ, Reid RE, Rogers JJ and Addison SJ (2007) Prediction of the naturalisation potential and weediness risk of transgenic cotton in Australia. Agriculture, Ecosystems & Environment, 119, 177–189. DOI:10.1016/j.agee.2006.07.007. Sánchez-Bayo F and Hyne RV (2014) Detection and analysis of neonicotinoids in river waters – development of a passive sampler for three commonly used insecticides. Chemosphere 99, 143–151. DOI: 10.1016/j.chemosphere.2013.10.051. Schmutz S and Sendzimir J (eds) (2018) Riverine Ecosystem Management Science for Governing Towards a Sustainable Future. Springer Open. Seafarms (2016) Seafarms group – project sea dragon. Viewed 27 September 2024, https://seafarms.com.au/project-sea-dragon/ Setterfield SA, Rossiter-Rachor NA, Douglas MM, Wainger L, Petty AM, Barrow P, Shepherd IJ and Ferdinands KB (2013) Adding fuel to the fire: the impacts of non-native grass invasion on fire management at a regional scale. PLoS ONE, 8(5), e59144. DOI:10.1371/journal.pone.0059144. SGNRM (Southern Gulf NRM) and NWQROC (North West Queensland Regional Organisation of Councils) (2022) North West Queensland Regional Biosecurity Plan 2022–2027. SGNRM and NWQROC. Viewed 2 January 2024, https://www.nwqroc.com.au/news/regional-biosecurity- plan-a-first-for-north-west-queensland. Short AD (2020) Australian coastal systems: beaches, barriers and sediment compartments. Springer, Switzerland. Skhiri A and Dechmi F (2012) Impact of sprinkler irrigation management on the Del Reguero river (Spain) II: phosphorus mass balance. Agricultural Water Management 103, 130–139. DOI: 10.1016/j.agwat.2011.11.004. Staples D and Vance D (1986) Emigration of juvenile banana prawns Penaeus merguiensis from a mangrove estuary and recruitment to offshore areas in the wet-dry tropics of the Gulf of Carpentaria, Australia. Marine Ecology Progress Series 27(239), 52. Stevens MM, Burdett AS, Mudford EM, Helliwell S and Doran G (2011) The acute toxicity of fipronil to two non-target invertebrates associated with mosquito breeding sites in Australia. Acta Tropica 117(2), 125–130. DOI: 10.1016/j.actatropica.2010.11.002. Stratford D, Merrin L, Linke S, Kenyon R, Ponce Reyes R, Buckworth R, Deng RA, McGinness H, Pritchard J, Seo L and Waltham N (2024) Assessment of the potential ecological outcomes of water resource development in the Roper catchment. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Subcommittee on Domestic Quarantine and Market Access (2024) Moving plant goods interstate. Australian Interstate Quarantine website. Viewed 28 March 2024, https://interstatequarantine.org.au/producers/moving-plant-goods/ Tait J (2005) Identifying existing wetland management practices and needs on southern gulf pastoral properties. Report to WWF – Australia and Southern Gulf Catchments. Tait J and Perna C (2001) Fish habitat management challenges on an intensively developed tropical floodplain: Burdekin River North Queensland. RipRap 19, 14–21. Tanji KK and Kielen NC (2002) Agricultural water management in arid and semi-arid areas. Irrigation and drainage paper 61. Food and Agriculture Organization of the United Nations, Rome. Thorburn PJ, Wilkinson SN and Silburn DM (2013) Water quality in agricultural lands draining to the Great Barrier Reef: a review of causes, management and priorities. Agriculture, Ecosystems and Environment 180, 4–20. DOI: 10.1016/j.agee.2013.07.006. Tockner K, Pusch M, Borchardt D and Lorang MS (2010) Multiple stressors in coupled river– floodplain ecosystems. Freshwater Biology 55, 135-151. Traill LW, Bradshaw CJA, Delean S and Brook BW (2010) Wetland conservation and sustainable use under global change: a tropical Australian case study using magpie geese. Ecography 33(5), 818-825. DOI: 10.1111/j.1600-0587.2009.06205.x. Traill LW, Bradshaw CJA, Field HE and Brook BW (2009a) Climate Change Enhances the Potential Impact of Infectious Disease and Harvest on Tropical Waterfowl. Biotropica 41(4), 414-423. DOI: 10.1111/j.1744-7429.2009.00508.x. Traill LW, Whitehead PJ and Brook BW (2009b) How will climate change affect plant-herbivore interactions? A tropical waterbird case study. Emu 109(2), 126-134. DOI: 10.1071/MU09003. US EPA (2024) Causal Analysis/Diagnosis Decision Information System (CADDIS): pH. United States Environmental Protection Agency. Viewed 30 September 2024, https://www.epa.gov/caddis/ph. van Dam RA, Bartolo R and Bayliss P (2008a) Chapter 2: Identification of ecological assets, pressures and threats. In: Bartolo R, Bayliss P and Van Dam RA (eds), Ecological risk assessment for Australia’s northern tropical rivers. Land and Water Australia, Canberra. van Dam RA, Bartolo R and Bayliss P (2008a) Identification of ecological assets, pressures and threats. Environmental Research Institute of the Supervising Scientist: Darwin, Australia, 14- 161. Venarsky MP, Lowe V, Frid CLJ and Burford MA (2022) Flow regimes among rivers influences benthic biota biodiversity, but not abundance or biomass, in intertidal mudflats and sandflats in wet-dry tropical estuaries. Estuarine, Coastal and Shelf Science 271, 107858. DOI: https://doi.org/10.1016/j.ecss.2022.107858. 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. Wang YYL, Xiong J, Ohore OE, Cai Y-E, Fan H, Sanganyado E, Li P, You J, Liu W and Wang Z (2022) Deriving freshwater guideline values for neonicotinoid insecticides: Implications for water quality guidelines and ecological risk assessment. Science of the Total Environment 828, 154569. DOI: 10.1016/j.scitotenv.2022.154569. 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. Ward M, Carwardine J, Yong CJ, Watson JEM, Silcock J, Taylor GS, Lintermans M, Gillespie GR, Garnett ST, Woinarski J, Tingley R, Fensham RJ, Hoskin CJ, Hines HB, Roberts JD, Kennard MJ, Harvey MS, Chapple DG and Reside AE (2021) A national‐scale dataset for threats impacting Australia’s imperilled flora and fauna. Ecology and Evolution 11(17), 11749–61. DOI:10.1002/ece3.7920. Waterhouse J, Brodie J, Lewis S and Mitchell A (2012) Quantifying the sources of pollutants in the Great Barrier Reef catchments and the relative risk to reef ecosystems. Marine Pollution Bulletin 65, 394–406. DOI: 10.1016/j.marpolbul.2011.09.031. Weston DP and Lydy MJ (2014) Toxicity of the insecticide fipronil and its degradates to benthic macroinvertebrates of urban streams. Environmental Science & Technology 48(2), 1290. DOI: 10.1021/es4045874. Yang A, Petheram C, Marvanek S, Baynes F, Rogers L, Ponce Reyes R, Zund P, Seo L, Hughes J, Gibbs M, Wilson PR, Philip S and Barber M (2024) Assessment of surface water storage options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO Australia. Yeates S (2016) Comparison of delayed release N fertiliser options for cotton on clay soils with urea; yield, N fertiliser uptake and N losses in runoff – Burdekin 2013–2015. Cotton Research and Development Project: CSP1302. Zhao J and Chen B (2016) Species sensitivity distribution for chlorpyrifos to aquatic organisms: Model choice and sample size. Ecotoxicology and Environmental Safety 125, 161–169. DOI: 10.1016/j.ecoenv.2015.11.039. Appendices Lawn Hill Gorge on Lawn Hill Creek is a popular tourist attraction Photo: Manfred Gottschalk – Alamy stock photo Assessment products More information about the Southern Gulf Water Resource Assessment can be found at https://www.csiro.au/southerngulf. The website provides readers with a communications suite including factsheets, multimedia content, FAQs, reports and links to other related sites, particularly about other research in northern Australia. In order to meet the requirements specified in the contracted ‘Timetable for the Services’, the Assessment provided the following key deliverables: • Technical reports present scientific work at a level of detail sufficient for technical and scientific experts to reproduce the work. Each of the activities of the Assessment has at least one corresponding technical report. • The catchment report (this report) synthesises key material from the technical reports, providing well-informed but non-scientific readers with the information required to make decisions about the opportunities, costs and benefits associated with water resource development. • A summary report is provided for a general public audience. • A factsheet provides key findings for a general public audience. This appendix lists all such deliverables, plus those jointly delivered for the concurrent Victoria River Water Resource Assessment. Please cite as they appear. Methods report CSIRO (2021) Proposed methods report for the Southern Gulf catchments. A report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid Authority. CSIRO, Australia. Technical reports Devlin K (2023) Pump stations for flood harvesting or irrigation downstream of a storage dam. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Devlin K (2024) Conceptual arrangements and costings of hypothetical irrigation developments in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Gibbs M, Hughes J and Yang A (2024) River model calibration for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Gibbs M, Hughes J, Yang A, Wang B, Marvanek S and Petheram C (2024) River model scenario analysis for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Hayward J (2024) Potential for farm-scale hybrid renewable energy supply options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Karim F, Kim S, Ticehurst C, Gibbs M, Hughes J, Marvanek S, Yang A, Wang B and Petheram C (2024) Floodplain inundation mapping and modelling for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Lyons P, Barber M, Braedon P and Pert P (2024) Indigenous water values, rights, interests and development goals in the Southern Gulf Catchments: A literature review and implications for future research. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia McJannet D, Yang A and Seo L (2023) Climate data characterisation for hydrological and agricultural scenario modelling across the Victoria, Roper and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Merrin L, Stratford D, Kenyon R, Pritchard J, Linke S, Ponce Reyes R, Buckworth R, Castellazzi P, Costin B, Deng R, Gannon R, Gao S, Gilbey S, Lachish S, McGinness H and Waltham N (2024) Ecological assets of the Southern Gulf catchments to inform water resource assessments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Motson K, Mishra A and Waltham N (2024) A review of water quality studies relevant to northern Australia. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Ponce Reyes R, Stratford D, Linke S, Merrin L, Kenyon R, Buckworth R, Deng RA, Hughes J, McGinness H, Pritchard J, Seo L and Waltham N (2024) Assessment of the potential ecological outcomes of water resource development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Raiber M, Taylor AR, Dupuy M, Priestley S, Barry K, Crosbie RS, Knapton A and Hodgson G (2024) Characterising groundwater resources of the Gilbert River Formation, Camooweal Dolostone and Thorntonia Limestone in the Southern Gulf catchments, Queensland and Northern Territory. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Rider Levett Bucknall (2024) Water infrastructure-related costs for the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO Australia Speed R and Vanderbyl T (2024) Regulatory requirements for land and water development in the Northern Territory and Queensland. A technical report from the CSIRO Victoria and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO, Australia. Thomas M, Philip S, Zund P, Stockmann U, Hill J, Gregory L, Watson I and Thomas E (2024) Soils and land suitability for the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Vanderbyl T (2021) Southern Gulf: Queensland water plans and settings. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Waschka M and Macintosh A (2024) CSIRO Water Resource Assessments: Indigenous rights and interests in Queensland and the Northern Territory. A report from Barraband Consulting to CSIRO to inform the CSIRO Victoria, Roper and Southern Gulf Water Resource Assessments. CSIRO, Australia. Webster A, Jarvis D, Jalilov S, Philip S, Oliver Y, Watson I, Rhebergen T, Bruce C, Prestwidge D, McFallan S, Curnock M and Stokes C (2024) Financial and socio-economic viability of irrigated agricultural development in the Southern Gulf catchments. A technical report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Yang A, Petheram C, Marvanek S, Baynes F, Rogers L, Ponce Reyes R, Zund P, Seo L, Hughes J, Gibbs M, Wilson PR, Philip S and Barber M (2024) Assessment of surface water storage options in the Victoria and Southern Gulf catchments. A technical report from the CSIRO Victoria River and Southern Gulf Water Resource Assessments for the National Water Grid. CSIRO Australia. Catchment report Watson I, Bruce C, Philip S, Petheram C and Chilcott C (eds) (2024) Water resource assessment for the Southern Gulf catchments. A report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Summary report CSIRO (2024) The Southern Gulf Water Resource Assessment. A summary report from the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Factsheet on key findings CSIRO (2024) The Southern Gulf Water Resource Assessment. Key messages of reports to the CSIRO Southern Gulf Water Resource Assessment for the National Water Grid. CSIRO, Australia. Shortened forms SHORT FORM FULL FORM ABS Australian Bureau of Statistics ACARA Australian Curriculum, Assessment and Reporting Authority AE adult equivalent AEP annual exceedance probability AHD Australian Height Datum ALA Aboriginal Land Act 1991 (Qld) ALRA Aboriginal Land Rights (Northern Territory) Act 1976 (Cth) AMTD adopted middle thread distance ANZECC Australian and New Zealand Environment and Conservation Council APSIM Agricultural Production Systems sIMulator ARMCANZ Agriculture and Resource Management Council of Australia and New Zealand ASC Australian Soil Classification AWC available water capacity AWRC Australian Water Resources Council BCR benefit-to-cost ratio BOM Bureau of Meteorology CBA cost–benefit analysis CBR cost-to-benefit ratio CLA Cambrian Limestone Aquifer CLCAC Carpentaria Land Council Aboriginal Corporation CMB chloride mass balance CMIP Coupled Model Intercomparison Project CV coefficient of variation DENR Department of Environment and Natural Resources (NT) DIDO drive-in drive-out DIWA Directory of Important Wetlands in Australia DN Nominal Diameter DNRME Department of Natural Resources, Mines and Energy (Queensland) DOGIT deeds of grant in trust EBITDA earnings before interest, taxes, depreciation and amortisation EGM96 Earth Gravitational Model 1996 SHORT FORM FULL FORM EIS Environmental Impact Statement ENSO El Niño–Southern Oscillation EOS end of system EPBC Act Environment Protection and Biodiversity Conservation Act 1999 (Cth) FIFO fly-in fly-out FSL full supply level FTE full-time equivalent GAB Great Artesian Basin GABORA Great Artesian Basin and Other Regional Aquifers GCM global climate model GCM-PS global climate model – pattern scaled GDE groundwater-dependent ecosystem GGNTAC Gangalidda and Garawa Native Title Aboriginal Corporation GM gross margin GRA Gilbert River Aquifer GRF Gilbert River Formation GVAP gross value of agricultural production GVIAP gross value of irrigated agricultural production HSD health service district IEO Index of Education and Occupation IER Index of Economic Resources ILUA Indigenous Land Use Agreement IPA Indigenous Protected Area IPCC Intergovernmental Panel on Climate Change IRR internal rate of return IRSD Index of Relative Socio-Economic Disadvantage IUCN International Union for Conservation of Nature JORC Joint Ore Reserves Committee LGA local government area MAR managed aquifer recharge MODIS Moderate Resolution Imaging Spectroradiometer NAWRA Northern Australia Water Resource Assessment n.d. not dated NEM National Electricity Market NGMA Nicholson Groundwater Management Area NPF Northern Prawn Fishery NPV net present value SHORT FORM FULL FORM NRM Natural Resource Management NT Northern Territory NWPS North West Power System O&M annual operation and maintenance PAW plant available water PAWC plant available water capacity PBC Prescribed Body Corporate PE potential evaporation PET potential evapotranspiration PHN primary health network QSNTS Queensland South Native Title Services RCC roller compacted concrete RNTBC Registered Native Title Body Corporate RoNA rest of northern Australia SA South Australia SA2 Statistical Area Level 2 SEIFA Socio-Economic Indexes for Areas SGG soil generic group SILO Scientific Information for Land Owners SOI Southern Oscillation Index SSP Shared Socio-economic Pathway SWAN Surface Water Ambient Water Quality Network SWL standing water level TDS total dissolved solids TraNSIT Transport Network Strategic Investment Tool UNESCO United Nations Educational, Scientific and Cultural Organization WA Western Australia Units UNIT DESCRIPTION $ dollars % per cent c cents cm centimetre d day dS decisiemens DS dry season GL gigalitre ha hectare kg kilogram km kilometre km2 square kilometre kPa kilopascal kV kilovolt kW kilowatt L litre m metre m3 cubic metre mASL metres above sea level mBGL metres below ground level mEGM86 metres (Earth Gravitational Model 1986) mEGM96 metres (Earth Gravitational Model 1996) mg milligram ML megalitre mm millimetre MWh megawatt hour s second t metric tonne y year °C degrees Celsius List of figures Figure 1-1 Map of Australia showing Assessment area (Southern Gulf catchments) and other recent or ongoing CSIRO Assessments ........................................................................................... 3 Figure 1-2 Number of large dams constructed in Australia and northern Australia over time ..... 8 Figure 1-3 Schematic of key components and concepts in the establishment of a greenfield irrigation development ................................................................................................................. 10 Figure 1-4 The Southern Gulf catchments .................................................................................... 13 Figure 2-1 Schematic diagram of key natural components and concepts in the establishment of a greenfield irrigation development ............................................................................................. 19 Figure 2-2 Surface geology of the Southern Gulf catchments ...................................................... 23 Figure 2-3 Physiographic units of the Southern Gulf catchments ................................................ 25 Figure 2-4 Major geological provinces of the Southern Gulf catchments .................................... 29 Figure 2-5 The soil generic groups (SGGs) of the Southern Gulf catchments produced by digital soil mapping .................................................................................................................................. 32 Figure 2-6 Cracking clay (brown Vertosol; SGG 9) Mitchell grass (Astrebla spp.) downs with whitewood (Elaeocarpus sp.) and gutta percha (Palaquium spp.) on the Armraynald Plain physiographic unit, east of the Leichhardt River .......................................................................... 36 Figure 2-7 Brown Dermosol (SGG 2) buffel grass (Cenchrus ciliaris) open woodland with silver leaf box (Eucalyptus pruinosa) on the Armraynald Plain physiographic unit in the middle reaches of the Leichhardt River .................................................................................................................. 36 Figure 2-8 Red sandy soil (SGG 6.1) open woodland of Darwin box (Eucalyptus tectifica), bauhinia (Bauhinia spp.) and Cooktown ironwood (Erythrophleum chlorostachys) near Doomadgee on the Armraynald Plain physiographic unit north of the Nicholson River ............. 37 Figure 2-9 Level to gently undulating cracking clay soils of the Armraynald Plain suitable for broadacre irrigation ...................................................................................................................... 38 Figure 2-10 Soil profile of the brown Vertosol (SGG 9) sampled on the Armraynald Plain physiographic unit east of the Leichhardt River ........................................................................... 39 Figure 2-11 (a) Surface soil pH (top 10 cm) of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction .............................................................. 40 Figure 2-12 (a) Soil thickness of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction ................................................................................ 41 Figure 2-13 (a) Soil surface texture of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction ................................................................................ 42 Figure 2-14 (a) Soil permeability of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction ................................................................................ 43 Figure 2-15 (a) Available water capacity in the upper 100 cm of the soil profile (AWC 100) of the Southern Gulf catchments as predicted by digital soil mapping and (b) reliability of the prediction ...................................................................................................................................... 44 Figure 2-16 (a) Surface rockiness in soils of the Southern Gulf catchments represented by presence or absence as predicted by digital soil mapping and (b) reliability of the prediction .. 45 Figure 2-17 Historical rainfall, potential evaporation and rainfall deficit .................................... 47 Figure 2-18 Historical monthly rainfall (left) and time series of annual rainfall (right) in the Southern Gulf catchments at Mount Isa, Doomadgee, Gregory and Burketown ........................ 50 Figure 2-19 Historical monthly potential evaporation (PE) (left) and time series of annual PE (right) in the Southern Gulf catchments at Mount Isa, Doomadgee, Gregory and Burketown ... 51 Figure 2-20 (a) Coefficient of variation (CV) of annual rainfall and (b) the CV of annual rainfall plotted against mean annual rainfall for 99 rainfall stations around Australia ........................... 52 Figure 2-21 Runs of wet and dry years at Mount Isa, Doomadgee, Gregory and Burketown ..... 54 Figure 2-22 Percentage change in rainfall and potential evaporation per degree of global warming for the 32 Scenario C simulations relative to Scenario A values for the Southern Gulf catchments .................................................................................................................................... 56 Figure 2-23 Spatial distribution of mean annual rainfall across the Southern Gulf catchments under scenarios (a) Cwet, (b) Cmid and (c) Cdry .......................................................................... 56 Figure 2-24 (a) Monthly rainfall and (b) potential evaporation for the Southern Gulf catchments under scenarios A and C ................................................................................................................. 57 Figure 2-25 Simplified schematic diagram of terrestrial water balance in the Southern Gulf catchments .................................................................................................................................... 60 Figure 2-26 Simplified regional hydrogeology of the Southern Gulf catchments ........................ 62 Figure 2-27 Two-dimensional conceptual schematic of the interconnected aquifer system and its variability .................................................................................................................................. 64 Figure 2-28 Gregory River at Riversleigh Road ............................................................................. 65 Figure 2-29 Lawn Hill Gorge .......................................................................................................... 66 Figure 2-30 The Gregory River receives groundwater discharge from the Cambrian Limestone Aquifer ........................................................................................................................................... 67 Figure 2-31 Full extent of the Georgina Basin and Carpentaria Sub-basin of the Great Artesian Basin. Inset map shows full extent of Great Artesian Basin ......................................................... 68 Figure 2-32 Groundwater bore yields for major aquifers across the Southern Gulf catchments 69 Figure 2-33 Groundwater salinity for major aquifers in the Southern Gulf catchments ............. 70 Figure 2-34 Annual recharge estimates for the Southern Gulf catchments ................................. 72 Figure 2-35 Summary of recharge statistics to outcropping areas of key hydrogeological units across the Southern Gulf catchments ........................................................................................... 73 Figure 2-36 Spatial distribution of groundwater discharge classes including surface water – groundwater connectivity across the Southern Gulf catchments ................................................ 75 Figure 2-37 Modelled streamflow under natural conditions ....................................................... 76 Figure 2-38 Streamflow observation data availability in the Southern Gulf catchments ............ 78 Figure 2-39 Median annual streamflow (50% exceedance) in the Southern Gulf catchments under Scenario A ........................................................................................................................... 79 Figure 2-40 (a) 20% and (b) 80% exceedance of annual streamflow in the Southern Gulf catchments under Scenario A ....................................................................................................... 80 Figure 2-41 Catchment area and elevation profile from the upper tributaries in each catchment to the mouth along the (a) Gregory–Nicholson River and (b) Leichhardt River ........................... 80 Figure 2-42 Mean annual (a) rainfall and (b) runoff across the Southern Gulf catchments under Scenario A ..................................................................................................................................... 81 Figure 2-43 Annual runoff at (a) 20%, (b) 50% and (c) 80% exceedance across the Southern Gulf catchments under Scenario A ....................................................................................................... 82 Figure 2-44 Total runoff across the Southern Gulf catchments under Scenario A showing (a) time series of annual runoff and (b) monthly runoff averaged across the catchments .............. 82 Figure 2-45 Flood inundation map of the Southern Gulf catchments .......................................... 84 Figure 2-46 Flood inundation in the Southern Gulf catchments for a flood event of 1 in 38 annual exceedance probability (AEP) in March 2023 ................................................................... 85 Figure 2-47 Peak flood discharge and annual exceedance probability (AEP) at (a) gauge 912107A (Nicholson River at Connolly’s Hole), (b) gauge 912105A (Gregory River at Riversleigh) and (c) gauge 913007B (Leichhardt River at Floraville) ............................................................................ 86 Figure 2-48 Minimum observed September streamflow at two stream gauge locations on the Gregory River ................................................................................................................................ 87 Figure 2-49 Minimum monthly flow over 132 years of simulation for the month of October .... 88 Figure 2-50 Instream waterhole evolution in a reach of the Flinders River ................................. 89 Figure 2-51 The Leichhardt River near Kajabbi looking south towards the Isa highlands. In the highly seasonal climate of the Southern Gulf catchments, springs and persistent waterholes provide important ecological refugia during the dry season ....................................................... 89 Figure 2-52 Location of river reaches containing permanent water in the Southern Gulf catchments .................................................................................................................................... 90 Figure 2-53 Accelerated erosion contributes sediment to streamflow........................................ 92 Figure 2-54 Water quality samples for selected constituents on the Gregory River at Riversleigh (A912005A), Gunpowder Creek at Gunpowder (913006A) and Leichhardt River at Floraville Homestead (913007B) .................................................................................................................. 93 Figure 3-1 Schematic diagram of key components of the living and built environment to be considered in establishing a greenfield irrigation development ................................................ 101 Figure 3-2 Conceptual diagram of selected ecological assets of the Southern Gulf catchments ..................................................................................................................................................... 106 Figure 3-3 Estuarine crocodiles inhabit fresh and saltwater environments ............................... 108 Figure 3-4 Location of protected areas and important wetlands within the Southern Gulf catchments Assessment area...................................................................................................... 109 Figure 3-5 Regional ecosystem biodiversity status in the Queensland part of the Southern Gulf catchments .................................................................................................................................. 110 Figure 3-6 Brolgas flying into the sunset at Lake Moondarra ..................................................... 117 Figure 3-7 Land subject to inundation (potential floodplain wetlands) and important wetlands in the Southern Gulf catchments .................................................................................................... 119 Figure 3-8 This saltpan area in northern Australia is typical in being located between mangrove and saltmarsh areas .................................................................................................................... 122 Figure 3-9 Australian bustards are common in grasslands and woodlands across northern Australia ...................................................................................................................................... 123 Figure 3-10 Location of saltpans in the Southern Gulf catchments marine region .................... 124 Figure 3-11 The O’Shannassy River – one of the northern Australian rivers where catfish are found ........................................................................................................................................... 126 Figure 3-12 Modelled potential species distribution for fork-tailed catfish (Neoarius graeffei) in the Southern Gulf catchments .................................................................................................... 127 Figure 3-13 Magpie goose perched on a fallen tree branch ....................................................... 129 Figure 3-14 Distribution of species listed under the Environment Protection and Biodiversity Conservation Act 1999 and by the NT and Queensland governments in the Southern Gulf catchments .................................................................................................................................. 130 Figure 3-15 Boundaries of the Australian Bureau of Statistics Statistical Area Level 2 (SA2) regions used for demographic data in this analysis.................................................................... 135 Figure 3-16 Land use classification for the Southern Gulf catchments ...................................... 139 Figure 3-17 Regions in the Northern Prawn Fishery and the North West Minerals Province ... 143 Figure 3-18 Main commodity mineral occurrences and exploration tenements in the Southern Gulf catchments .......................................................................................................................... 144 Figure 3-19 Local government areas and the Tropical North Queensland tourism region that statistics on tourism visitation are extracted from ..................................................................... 150 Figure 3-20 Road rankings and conditions in the vicinity of the Southern Gulf catchments ..... 153 Figure 3-21 Roads accessible to Type 2 vehicles in the vicinity of the Southern Gulf catchments: minor roads are not classified ..................................................................................................... 154 Figure 3-22 Common configurations of heavy freight vehicles used for transporting agricultural goods in Australia ........................................................................................................................ 155 Figure 3-23 Road conditions and distance to market impact the economics of development in the Southern Gulf catchments. ................................................................................................... 155 Figure 3-24 Mean speed achieved for freight vehicles on roads in the vicinity of the Southern Gulf catchments .......................................................................................................................... 156 Figure 3-25 Many roads are gravel in the Southern Gulf catchments, and often impassable in the wet season ............................................................................................................................ 158 Figure 3-26 Annual amounts of trucking in the Southern Gulf catchments and the locations of pastoral properties and ports ..................................................................................................... 159 Figure 3-27 Electricity generation and transmission network and pipelines in the Southern Gulf catchments .................................................................................................................................. 160 Figure 3-28 Solar photovoltaic capacity factors in the Southern Gulf catchments .................... 162 Figure 3-29 Wind capacity factors in the Southern Gulf catchments ......................................... 163 Figure 3-30 Location, type and volume of annual licensed surface water and groundwater entitlements across the Southern Gulf catchments ................................................................... 165 Figure 3-31 Exclusive Aboriginal lands and pastoral interests make up the majority of the Southern Gulf catchments .......................................................................................................... 174 Figure 3-32 The extent of native title claims and determinations over the Southern Gulf catchments .................................................................................................................................. 175 Figure 3-33 Indigenous protected areas and other protected areas in the Southern Gulf catchments as of April 2020........................................................................................................ 176 Figure 3-34 Indigenous Land Use Agreements (ILUAs) ............................................................... 177 Figure 4-1 Schematic of agriculture and aquaculture enterprises as well as crop and/or forage integration with existing beef enterprises to be considered in the establishment of a greenfield irrigation development ............................................................................................................... 206 Figure 4-2 Area (ha) of the Southern Gulf catchments mapped in each of the land suitability classes for 14 selected land use combinations (crop group × season × irrigation type) ............ 214 Figure 4-3 Agricultural versatility index map for the Southern Gulf catchments....................... 215 Figure 4-4 Climate comparisons of Southern Gulf catchments’ sites with established irrigation areas at Kununurra (WA) and Mareeba (Queensland) ............................................................... 219 Figure 4-5 Annual cropping calendar for irrigated agricultural options in the Southern Gulf catchments .................................................................................................................................. 221 Figure 4-6 Soil wetness indices that indicate when seasonal trafficability constraints are likely to occur on sands, Kandosols (loamy sands) and Vertosols (high clay) with a Gregory climate for two thresholds (a) 80% and (b) 70% of the maximum plant available water capacity .............. 222 Figure 4-7 Influence of planting date on rainfed grain sorghum yield at Gregory for a (a) Kandosol and (b) Vertosol ........................................................................................................... 224 Figure 4-8 Influence of available irrigation water on grain sorghum yields for planting dates of (a) 1 February and (b) 1 August, for a Vertosol with a Gregory climate .................................... 225 Figure 4-9 Fluctuations in seedless watermelon prices at Melbourne wholesale markets from April 2020 to February 2023 ....................................................................................................... 234 Figure 4-10 Modelled land suitability for Crop Group 7 (e.g. sorghum (grain) or maize) using furrow irrigation in the (a) wet season and (b) dry season ........................................................ 246 Figure 4-11 Sorghum (grain) ....................................................................................................... 248 Figure 4-12 Modelled land suitability for mungbean (Crop Group 10) in the dry season using (a) furrow irrigation and (b) spray irrigation .................................................................................... 250 Figure 4-13 Mungbean ................................................................................................................ 250 Figure 4-14 Modelled land suitability for soybean (Crop Group 10) in the dry season using (a) furrow irrigation and (b) spray irrigation .................................................................................... 253 Figure 4-15 Soybean.................................................................................................................... 253 Figure 4-16 Modelled land suitability for peanut (Crop Group 6) using spray irrigation in the (a) wet season and (b) dry season ................................................................................................... 256 Figure 4-17 Peanut ...................................................................................................................... 256 Figure 4-18 Modelled land suitability for cotton (Crop Group 7) using furrow irrigation in the (a) wet season and (b) dry season ................................................................................................... 260 Figure 4-19 Cotton ...................................................................................................................... 260 Figure 4-20 Modelled land suitability for Rhodes grass (Crop Group 14) using (a) spray irrigation and (b) furrow irrigation ............................................................................................................. 263 Figure 4-21 Rhodes grass ............................................................................................................ 264 Figure 4-22 Modelled land suitability for Cavalcade (Crop Group 13) in the wet season using (a) spray irrigation and (b) furrow irrigation .................................................................................... 266 Figure 4-23 Lablab ....................................................................................................................... 268 Figure 4-24 Modelled land suitability for (a) cucurbits (e.g. rockmelon, Crop Group 3) using trickle irrigation in the dry season and (b) root crops such as onion (Crop Group 6) using spray irrigation in the wet season ........................................................................................................ 270 Figure 4-25 Rockmelon ............................................................................................................... 270 Figure 4-26 Modelled land suitability for (a) mango (Crop Group 1) and (b) lime (Crop Group 2), both grown using trickle irrigation.............................................................................................. 273 Figure 4-27 Mango ...................................................................................................................... 273 Figure 4-28 Modelled land suitability for Indian sandalwood (Crop Group 15) grown using (a) trickle or (b) furrow irrigation ..................................................................................................... 276 Figure 4-29 Indian sandalwood and host plants ......................................................................... 276 Figure 4-30 Black tiger prawns .................................................................................................... 280 Figure 4-31 Barramundi .............................................................................................................. 280 Figure 4-32 Schematic of marine aquaculture farm ................................................................... 282 Figure 4-33 Land suitability in the Southern Gulf catchments for marine species aquaculture in (a) lined ponds and (b) earthen ponds ....................................................................................... 286 Figure 4-34 Land suitability in the Southern Gulf catchments for freshwater species aquaculture in (a) lined ponds and (b) earthen ponds ................................................................................... 287 Figure 5-1 Schematic of key engineering and agricultural components to be considered in the establishment of a water resource and greenfield irrigation development .............................. 296 Figure 5-2 Hydrogeological units with potential for future groundwater resource development ..................................................................................................................................................... 306 Figure 5-3 Groundwater dependent ecosystems along Lawn Hill Creek ................................... 308 Figure 5-4 Thickness of the Georgina Basin in the Southern Gulf catchments .......................... 309 Figure 5-5 Hydrogeological cross-section through the Cambrian Limestone Aquifer in the Georgina Basin and south-west of the Southern Gulf catchments ............................................ 310 Figure 5-6 Depth to standing water level (SWL) of the Cambrian Limestone Aquifer (CLA)...... 311 Figure 5-7 Location of the Undilla Sub-basin groundwater flow model in relation to the Southern Gulf catchments and portions of the model that coincide with the Lawn Hill Creek and Gregory subcatchments and Nicholson Groundwater Management Area ................................ 312 Figure 5-8 Depth to the top of the Gilbert River Aquifer (GRA) ................................................. 314 Figure 5-9 South-west to north-east cross-section traversing the Great Artesian Basin in the Southern Gulf catchments. ......................................................................................................... 315 Figure 5-10 Depth to standing water level of the Gilbert River Aquifer (GRA) .......................... 316 Figure 5-11 Uncontrolled artesian flow from the Gilbert River Aquifer at the Burketown groundwater bore ....................................................................................................................... 318 Figure 5-12 Types of managed aquifer recharge ........................................................................ 319 Figure 5-13 Managed aquifer recharge (MAR) opportunities for the Southern Gulf catchments independent of distance from a water source for recharge ...................................................... 321 Figure 5-14 (a) Managed aquifer recharge (MAR) opportunities in the Southern Gulf catchments within 5 km of major rivers and (b) aquifer underlying the MAR opportunity classes .............. 322 Figure 5-15 Potential storage sites in the Southern Gulf catchments based on minimum cost per megalitre storage capacity .......................................................................................................... 326 Figure 5-16 Potential storage sites in the Southern Gulf catchments based on minimum cost per megalitre yield at the dam wall .................................................................................................. 328 Figure 5-17 Julius Dam on the Leichhardt River ......................................................................... 329 Figure 5-18 Southern Gulf catchments hydro-electric power generation opportunity map ..... 330 Figure 5-19 Cumulative yield at 85% annual time reliability versus cumulative cost of water in $/ML and change in the end-of-system (EOS) volume in the Southern Gulf catchments ......... 335 Figure 5-20 Location of listed species, water-dependent assets and aggregated modelled habitat in the vicinity of the potential dam site on Gunpowder Creek dam AMTD 66 km and reservoir extent ........................................................................................................................... 337 Figure 5-21 Potential dam site on Gunpowder Creek AMTD 66 km: cost and yield at the dam wall .............................................................................................................................................. 338 Figure 5-22 Location of listed species, water-dependent assets and aggregated modelled habitat in the vicinity of the potential dam site on the Gregory River AMTD 174 km ............... 340 Figure 5-23 Potential dam site on Gregory River AMTD 174 km: cost and yield at the dam wall ..................................................................................................................................................... 341 Figure 5-24 Schematic cross-section of sheet piling weir ........................................................... 343 Figure 5-25 Rectangular ringtank and 500 ha of cotton in the Flinders catchment (Queensland) ..................................................................................................................................................... 345 Figure 5-26 Suitability of land for large farm-scale ringtanks in the Southern Gulf catchments 346 Figure 5-27 Annual reliability of diverting annual system and reach target volumes for varying pump start thresholds ................................................................................................................. 348 Figure 5-28 Annual reliability of diverting annual system and reach target volumes for varying pump start thresholds with annual diversion commencement flow requirement of 150 GL ... 350 Figure 5-29 Annual reliability of diverting annual system and reach target volumes for varying pump start thresholds assuming pumping capacity of 10 days ................................................. 351 Figure 5-30 Annual reliability of diverting annual system and reach target volumes for varying pump rates assuming a pump start threshold of 600 ML/day ................................................... 352 Figure 5-31 50% annual exceedance (median) streamflow relative to Scenario A in the Southern Gulf catchments for a pump start threshold of 600 ML/day and a pump capacity of 20 days .. 353 Figure 5-32 80% annual exceedance streamflow relative to Scenario A in the Southern Gulf catchments for a pump start threshold of 600 ML/day and a pump capacity of 20 days ......... 354 Figure 5-33 Most economically suitable locations for large farm-scale gully dams in the Southern Gulf catchments .......................................................................................................... 360 Figure 5-34 Suitability of soils for construction of gully dams in the Southern Gulf catchments ..................................................................................................................................................... 361 Figure 5-35 Reported conveyance losses from irrigation systems across Australia .................. 369 Figure 5-36 Efficiency of different types of irrigation system .................................................... 370 Figure 6-1 Schematic diagram of key components affecting the commercial viability of a potential greenfield irrigation development .............................................................................. 379 Figure 6-2 Locations of the five dams used in this review .......................................................... 400 Figure 6-3 Trends in gross value of agricultural production (GVAP) in (a) Australia and (b) Queensland over 40 years (1981–2021) ..................................................................................... 403 Figure 6-4 National trends for increasing gross value of irrigated agricultural production (GVIAP) as available water supplies have increased for (a) fruits, (b) vegetables, (c) fruits and vegetables combined, and (d) total agriculture ............................................................................................ 404 Figure 6-5 Queensland’s north-west region used in the input–output (I–O) analyses relative to the Southern Gulf catchments Assessment area ........................................................................ 410 Figure 7-1 Schematic diagram of the components where key risks can manifest when considering the establishment of a greenfield irrigation or aquaculture development ............ 417 Figure 7-2 Locations of the river system modelling nodes at which flow–ecology dependencies were assessed (numbered) and the locations of hypothetical water resource developments in the Southern Gulf catchments .................................................................................................... 429 Figure 7-3 Riparian vegetation along the Leichhardt River – riparian zones are often more fertile and productive than surrounding terrestrial vegetation ............................................................ 436 Figure 7-4 Spatial heatmap of habitat-weighted changes in flow for sawfish, considering the assets important locations across the catchment ...................................................................... 437 Figure 7-5 Habitat-weighted change in sawfish flow dependencies by scenario across model nodes ........................................................................................................................................... 438 Figure 7-6 Habitat-weighted change in swimming, diving and grazing waterbirds flow dependencies by scenario across model nodes ......................................................................... 441 Figure 7-7 Habitat-weighted change in floodplain wetlands flow dependencies by scenario across model nodes .................................................................................................................... 444 Figure 7-8 Lake Moondarra near Mount Isa is used for urban water supply and is a popular water and recreational reserve................................................................................................... 449 Figure 7-9 Mean change associated with each asset’s important metrics across water harvesting increments of system target and pump-start threshold with no annual diversion commencement flow requirement and pump rate of 30 days .................................................. 450 Figure 7-10 The invasion curve with biosecurity actions taken at various stages ..................... 461 Figure 7-11 Farm biosecurity signage available through www.farmbiosecurity.com.au ........... 462 Figure 7-12 An extratidal flat on the Karumba Plain near Burketown, with white salt crystals either side of a shallow natural drainage line. Salt-tolerant samphire and mangrove species can be seen in the background. ......................................................................................................... 475 Figure 7-13 Example of salt-affected areas on the northern part of the Armraynald Plain ...... 476 List of tables Table 2-1 Soil generic groups (SGGs), descriptions, management considerations and correlations to Australian Soil Classification (ASC) for the Southern Gulf catchments .................................... 33 Table 2-2 Area and proportions covered by each soil generic group (SGG) in the Southern Gulf catchments .................................................................................................................................... 35 Table 2-3 Projected sea-level rise for the coast of the Southern Gulf catchments ...................... 57 Table 2-4 Streamflow metrics at selected gauging stations in the Southern Gulf catchments ... 77 Table 2-5 Summary of water quality data for the open Southern Gulf catchments sites, with values of minimum, median and maximum for each site and each water quality parameter .... 91 Table 3-1 Categories of biodiversity status of the Queensland regional ecosystems ................ 112 Table 3-2 Freshwater, marine and terrestrial ecological assets with freshwater flow dependences in the Southern Gulf catchments ......................................................................... 116 Table 3-3 Nationally important wetlands in the Southern Gulf catchments ............................. 118 Table 3-4 Definition of threatened categories under the Commonwealth Environment Protection and Biodiversity Conservation Act 1999, the NT wildlife classification system, and the Queensland Nature Conservation Act 1992 ................................................................................ 131 Table 3-5 Major demographic indicators for the Southern Gulf catchments ............................ 134 Table 3-6 Socio-Economic Indexes for Areas (SEIFA) scores of relative socio-economic advantage for the Southern Gulf catchments .............................................................................................. 136 Table 3-7 Key employment data for the Southern Gulf catchments .......................................... 137 Table 3-8 Value of agricultural production estimated for the Southern Gulf catchments and the value of agricultural production for Queensland for 2020–21 ................................................... 140 Table 3-9 Indicative numbers for employment in different types of mining operations in the Southern Gulf catchments .......................................................................................................... 145 Table 3-10 Resource and reserve data for several major deposits in and on the margins of the Southern Gulf catchments .......................................................................................................... 146 Table 3-11 Global water consumption in the mining and refining of selected metals .............. 148 Table 3-12 Overview of commodities (excluding livestock) annually transported into and out of the Southern Gulf catchments .................................................................................................... 157 Table 3-13 Unallocated surface water in the Queensland part of the Southern Gulf catchments ..................................................................................................................................................... 166 Table 3-14 Schools servicing the Southern Gulf catchments ..................................................... 168 Table 3-15 Number and percentage of unoccupied dwellings and population for the Southern Gulf catchments .......................................................................................................................... 169 Table 4-1 Land suitability classes based on FAO (1976, 1985) as used in the Assessment ........ 212 Table 4-2 Crop groups and individual land uses evaluated for irrigation (and rainfed) potential ..................................................................................................................................................... 213 Table 4-3 Qualitative land evaluation observations in Southern Gulf catchments areas A to F shown in Figure 4-3 ..................................................................................................................... 216 Table 4-4 Crop options for which performance was evaluated in terms of water use, yields and gross margins .............................................................................................................................. 218 Table 4-5 Soil water content at sowing, and rainfall for the 90-day period following sowing for three sowing dates, based on a Gregory climate on a Vertosol ................................................. 223 Table 4-6 Performance metrics for broadacre cropping options in the Southern Gulf catchments: applied irrigation water, crop yield and gross margin (GM) for four environments ..................................................................................................................................................... 227 Table 4-7 Breakdown of variable costs relative to revenue for broadacre crop options ........... 231 Table 4-8 Sensitivity of cotton crop gross margins ($/ha) to variation in yield, lint prices and distance to gin ............................................................................................................................. 232 Table 4-9 Sensitivity of forage (Rhodes grass) crop gross margins ($/ha) to variation in yield and hay price ...................................................................................................................................... 232 Table 4-10 Performance metrics for horticulture options in the Southern Gulf catchments: annual applied irrigation water, crop yield and gross margin .................................................... 233 Table 4-11 Sensitivity of watermelon crop gross margins ($/ha) to variation in melon prices and freight costs ................................................................................................................................. 235 Table 4-12 Performance metrics for plantation tree crop options in the Southern Gulf catchments: annual applied irrigation water, crop yield and gross margin ............................... 236 Table 4-13 Likely annual irrigated crop planting windows, suitability and viability in the Southern Gulf catchments .......................................................................................................... 239 Table 4-14 Sequential cropping options for Kandosols .............................................................. 240 Table 4-15 Production and financial outcomes from the different irrigated forage and beef production options for a representative property in the Southern Gulf catchments ................ 243 Table 4-16 Summary information relevant to the cultivation of cereals, using sorghum (grain) as an example .................................................................................................................................. 247 Table 4-17 Summary information relevant to the cultivation of pulses, using mungbean as an example ....................................................................................................................................... 251 Table 4-18 Summary information relevant to the cultivation of oilseed crops, using soybean as an example .................................................................................................................................. 254 Table 4-19 Summary information relevant to the cultivation of root crops, using peanut as an example ....................................................................................................................................... 257 Table 4-20 Summary information relevant to the cultivation of cotton .................................... 261 Table 4-21 Rhodes grass production for hay over 1 year of a 6-year cycle ................................ 264 Table 4-22 Cavalcade production over a 1-year cycle ................................................................ 267 Table 4-23 Summary information relevant to row crop horticulture production, with rockmelon as an example ............................................................................................................................. 271 Table 4-24 Summary information relevant to tree crop horticulture production, with mango as an example .................................................................................................................................. 274 Table 4-25 Summary information for Indian sandalwood production ....................................... 277 Table 4-26 Indicative capital and operating costs for a range of generic aquaculture development options .................................................................................................................. 288 Table 4-27 Gross revenue targets required to achieve target internal rates of return (IRR) for aquaculture developments with different combinations of capital costs and operating costs . 290 Table 5-1 Summary of capital costs, yields and costs per ML supply, including operation and maintenance ............................................................................................................................... 300 Table 5-2 Opportunity-level estimates of the potential scale of groundwater resource development opportunities in the Southern Gulf catchments .................................................. 304 Table 5-3 Summary of estimated costs for a 250 ha irrigation development using groundwater ..................................................................................................................................................... 307 Table 5-4 Mean annual groundwater balance for the Cambrian Limestone Aquifer (CLA) in the Undilla Sub-basin of the Georgina Basin for a 109-year climate sequence (1910–2019) and areas of the CLA that coincide with the Lawn Hill Creek and Gregory subcatchments and Nicholson Groundwater Management Area (NGMA) ................................................................................. 313 Table 5-5 Potential dam sites in the Southern Gulf catchments examined as part of the Assessment ................................................................................................................................. 331 Table 5-6 Summary comments for potential dams in the Southern Gulf catchments ............... 332 Table 5-7 Estimated construction cost of 3 m high sheet piling weir......................................... 343 Table 5-8 Effective volume after net evaporation and seepage for ringtanks of three mean water depths, under three seepage rates, near Century Mine in the Southern Gulf catchments ..................................................................................................................................................... 355 Table 5-9 Indicative costs for a 4000 ML ringtank ...................................................................... 357 Table 5-10 Annualised cost for the construction and operation of three ringtank configurations ..................................................................................................................................................... 357 Table 5-11 Equivalent annual cost per megalitre for two different capacity ringtanks under three seepage rates based on a climate station near the Century Zinc Mine in the Southern Gulf catchments .................................................................................................................................. 358 Table 5-12 Actual costs of four gully dams in northern Queensland ......................................... 362 Table 5-13 Cost of three hypothetical large farm-scale gully dams of capacity 4 GL ................. 363 Table 5-14 High-level breakdown of capital costs for three hypothetical large farm-scale gully dams of capacity 4 GL ................................................................................................................. 363 Table 5-15 Effective volumes and cost per megalitre for a 4 GL storage with different mean depths and seepage loss rates based on a climate station at the Century Zinc Mine in the Southern Gulf catchments .......................................................................................................... 364 Table 5-16 Cost of construction and operation of three hypothetical 4 GL gully dams............. 364 Table 5-17 Equivalent annualised cost and effective volume for three hypothetical 4 GL gully dams with various mean depths and seepage loss rates based on climate data at Victoria River Downs Station in the Victoria catchment ................................................................................... 365 Table 5-18 Summary of conveyance and application efficiencies .............................................. 367 Table 5-19 Water distribution and operational efficiency as nominated in water resource plans for four irrigation water supply schemes in Queensland ........................................................... 368 Table 5-20 Application efficiencies for surface, spray and micro irrigation systems ................. 371 Table 6-1 Types of questions that users can answer using the tools in this chapter ................. 382 Table 6-2 Indicative capital costs for developing three irrigation schemes based on the most cost-effective dam sites identified in the Southern Gulf catchments ........................................ 386 Table 6-3 Assumed indicative capital and operating costs for new off- and on-farm irrigation infrastructure .............................................................................................................................. 387 Table 6-4 Price irrigators can afford to pay for water, based on the type of farm, the farm water use and the farm annual gross margin (GM), while meeting a target 10% internal rate of return (IRR) ............................................................................................................................................. 389 Table 6-5 Farm gross margins (GMs) required in order to cover the costs of off-farm water infrastructure (at the supplier’s target internal rate of return (IRR)) ......................................... 391 Table 6-6 Water pricing required in order to cover costs of off-farm irrigation scheme development (dam, water distribution, and supporting infrastructure) at the investors target internal rate of return (IRR) ........................................................................................................ 392 Table 6-7 Farm gross margins (GMs) required in order to achieve target internal rates of return (IRR), given various capital costs of farm development (including an on-farm water source) .. 393 Table 6-8 Equivalent costs of water per ML for on-farm water sources with various capital costs of development, at the internal rate of return (IRR) targeted by the investor .......................... 394 Table 6-9 Risk adjustment factors for target farm gross margins (GMs), accounting for the effects of the reliability and severity (level of farm performance in ‘failed’ years) of the periodic risk of water reliability ................................................................................................................ 396 Table 6-10 Risk adjustment factors for target farm gross margins (GMs) accounting for the effects of reliability and the timing of periodic risks .................................................................. 397 Table 6-11 Risk adjustment factors for target farm gross margins (GMs), accounting for the effects of learning risks ............................................................................................................... 398 Table 6-12 Summary characteristics of the five dams used in this review................................. 401 Table 6-13 Summary of key issues and potential improvements arising from a review of recent dam developments ..................................................................................................................... 402 Table 6-14 Indicative costs of agricultural processing facilities .................................................. 405 Table 6-15 Indicative costs of road and electricity infrastructure .............................................. 406 Table 6-16 Indicative road transport costs between the Southern Gulf catchments and key markets and ports ....................................................................................................................... 406 Table 6-17 Indicative costs of community facilities .................................................................... 407 Table 6-18 Key 2021 data comparing the Southern Gulf catchments with the related I–O analysis region ............................................................................................................................. 410 Table 6-19 Regional economic impact estimated by I–O analysis for the total construction phase of an irrigated agricultural development based on estimated Type ll multipliers determined from the north-west Queensland I–O models ............................................................................ 411 Table 6-20 Estimated regional economic impact per year in the Southern Gulf catchments resulting from four scales of direct increase in agricultural output (rows) for the different categories of agricultural activity (columns) using the I–O model for north-west Queensland 413 Table 6-21 Estimated impact on annual household incomes and full-time equivalent (FTE) jobs within the Southern Gulf catchments resulting from four scales of direct increase in agricultural output (rows) for the various categories of agricultural activity (columns) ............................... 414 Table 7-1 Water resource development and climate scenarios explored in this ecology analysis ..................................................................................................................................................... 430 Table 7-2 Ecological assets used in the Southern Gulf catchments Water Resource Assessment and the different ecology groups used in this analysis............................................................... 432 Table 7-3 Reporting qualitative values for the flow dependencies modelling as rank percentile change of the hydrometrics ........................................................................................................ 434 Table 7-4 Scenarios of different hypothetical instream dam locations showing mean changes of ecology flows for groups of assets across each asset’s respective catchment assessment nodes ..................................................................................................................................................... 447 Table 7-5 Examples of significant pest and disease threats to plant industries in the Southern Gulf catchments .......................................................................................................................... 455 Table 7-6 Regional weed priorities and their management actions in Southern Gulf catchments ..................................................................................................................................................... 457 Table 7-7 High-risk freshwater pest fish threats to the Southern Gulf catchments .................. 459 Table 7-8 Water quality variables reviewed – their impacts on the environment, aquatic ecology and human health ....................................................................................................................... 468