You are on page 1of 23

AUSTRALIA

Geography

Australia, in the Southern Hemisphere, is comprised by the Australian continent mainland, the
Tasmania island and several smaller islands in the Pacific and Indian Oceans. It is the sixth largest
country in the world. Neighbouring countries include Papua New Guinea, Timor-Leste and Indonesia
to the north, New Caledonia (France), Vanuatu and the Solomon Islands to the north-east, and New
Zealand to the south-east.

Australia’s size confers a wide variety of landscapes and climatic zones, with a dry desert at its centre,
mountain ranges to the south-east, south-west and in eastern areas, and subtropical rainforests to the
north-east. It is the ‘flattest’ continent, with the oldest and least fertile soils. Mount Kosciuszko on the
Great Dividing Range is at 2 228 m, the highest mountain of the country. Commonly known as ‘the
outback’, desert or semi-arid land makes up by far the largest portion of country. With only 2.8
inhabitants per square kilometre, it is among the lowest population density in the world, although along
the temperate south-eastern coastline lives a large proportion of the population.

Eastern Australia is marked by the Great Dividing Range that runs parallel to the coast of Queensland,
New South Wales and much of Victoria. Parts of the range consist of low hills and the ‘highlands’ are
typically no more than 600 m in height. The Great Barrier Reef, the largest coral reef in the world,
extending for over 2000 km, lies close to the north-east coast.

In the northern part of the country, the landscapes, with their tropical climate, consist of grassland,
woodland and desert. At the north-west corner of the continent are the sandstone cliffs and gorges of
the Kimberley, and below them, the Pilbara. At the heart of the country are the uplands of central
Australia.

The climate of Australia is significantly influenced by ocean currents. A tropical climate typified by
summer rainfall pertains to most of the northern part of the country. A Mediterranean climate pertains
to the south-west corner of the country. A temperate climate pertains to most of the south-eastern part
of the country, including Tasmania.

Australia, being rich in natural resources, is a major exporter of agricultural products, particularly wool
and wheat, as well as minerals such as iron, gold, bauxite, base metals and diamonds, and energy in the
forms of natural gas, coal and uranium. The mining sector represents 10% of GDP and the ‘mining
related economy’ (exploration, mining equipment) represents 9% of GDP. Although accounting for
only 3% and 5% of GDP, agriculture and natural resources, respectively, contribute considerably to
exports. The service sector of the economy, accounting for ~70% of GDP, includes tourism, financial
and educational services [1].

Geology

General

Australia can be divided into four orogenic provinces mantled by four platform covers (Figure 1).

The West Australian Orogenic Province is the oldest province and contains the Pilbara, Yilgarn and
Rum Jungle Blocks, which are belts of metamorphosed sediments and volcanics in predominantly
gneiss and granite terrains. The platform covers of the West Australian Orogenic Province, consisting
principally of the Hamersley Basin, mantles the Pilbara Block and comprises Lower Proterozoic basic
volcanics and pyroclastics overlain by chemical and clastic sediments.
FIG. 1. Regional geological setting of Australia showing the distribution of selected uranium deposits and
occurrences. For the general uranium deposit and occurrence legend see World Uranium Geology, Exploration,
Resources and Production, IAEA, 2020. A general global geological legend is shown although not all geological
units necessarily occur on this particular map.

The North Australian Orogenic Province contains the Pine Creek Block, Tennant Creek Block,
Granites–Tanami Block, the Nicholson Block and the Halls Creek Belt. These areas generally consist
of low to moderately metamorphosed Lower Proterozoic sediments and volcanics, which, in some
cases, are intruded by granites. The North Australian Platform Cover was deposited in the Middle
Proterozoic and is preserved in the Kimberley Basin, the McArthur Basin and around the Tennant
Creek, Granites–Tanami and Nicholson Blocks. This platform cover is largely pelitic with minor
dolomites and volcanics.

The Central Australian Orogenic Province consists of several moderate to highly metamorphosed
volcano-sedimentary blocks and belts of Lower–Middle Proterozoic age. In the west, there are the
Ophthalmia–Gascoyne, Northampton and Albany–Fraser Blocks; in the centre there are the Musgrave
and Gawler Blocks; and to the east the Mount Isa Belt and the Willyama, Mount Painter and Wonaminta
Blocks. The Central Australian Platform Cover was deposited over a wide area of western and central
Australia from the Middle Proterozoic through the Palaeozoic.
In the west, there are the Carnarvon, Bangemall and Canning Basins, in the centre the Officer, Amadeus
and Ngalia Basins, and in the east the Georgina Basin and the Adelaide Geosyncline.

Several orogenic blocks have not been assigned to any province. The Arnhem and Litchfield Blocks in
northern Australia appear to be Archaean in age. The Arunta Block is poorly exposed and may be
associated with the North and Central Australian Orogenic Provinces. The Georgetown Block is Middle
Proterozoic and contains high-grade gneiss intruded by granite and a younger, less deformed sequence,
also intruded by granite.

The East Australian Orogenic Province or Tasman Geosyncline, of Cambrian–Late Triassic age,
consists of low to medium grade metamorphosed sediments and volcanics intruded by acid plutonic
rocks.

The Kanmantoo Belt is the oldest part of the province. Other belts within this province are the Lachlan
Belt and its age equivalent, the North Queensland Blocks, and the Holdgkinson and New England–
Yarrol Belts.

The Trans-Australian Platform Cover consists of Permian–Tertiary sedimentation, which was deposited
over much of the eastern third of Australia.

Geology related to uranium-bearing areas

Approximately 90% of Australia’s initial in-ground resources occur in two main types of deposit:
haematite breccia complexes and unconformity-related deposits. In addition, there are other types of
deposit, including sandstone hosted and surficial (calcrete) deposits, which also contain significant
resources.

Haematite breccia complex deposits

Approximately 70% of resources exist in Proterozoic haematite granitic breccias at Olympic Dam
(South Australia), the largest uranium resource in the world. Broadly similar haematite breccia
mineralization in the same geological province is being evaluated somewhere else at Acropolis,
Carrapateena, Oak Dam, Prominent Hill and Wirrda Well, and at some of the younger breccia-hosted
deposits in the Mount Painter area.

The Olympic Dam deposit, occurring in a haematite-rich granite breccia complex in the Gawler Craton,
underlies ~300 m of flat-lying sedimentary rocks of the Stuart Shelf geological province [2]. Formed in
a post-orogenic tectonic setting [2], the breccia complex is linked with a plutonic intrusion and co-
magmatic continental felsic volcanics of Mesoproterozoic age.

The Olympic Dam Breccia Complex occurs entirely within the Roxby Downs Granite intrusion of
Mesoproterozoic age. The margins of the Olympic Dam Breccia Complex are gradational with the
Roxby Downs Granite. The Olympic Dam Breccia Complex includes a complete gradation from granite
breccias through haematite–granite breccias to haematite-rich breccias. The zone of haematite-rich
breccias is enveloped by granitic breccias, which extend ~3 km beyond the margins of the haematite-
rich breccias. There is an extensive zonal distribution of the major types of rocks within the Olympic
Dam Breccia Complex, and almost all of the economic Cu–U mineralization is hosted by haematite-
rich breccias (haematitic breccias and heterolithic haematite breccias) [3].

Intrusive tuffs and dykes of felsic, mafic and ultramafic rock types intrude into the Olympic Dam
Breccia Complex, particularly its southern and eastern portions. These intrusive rocks are intimately
linked with volcanic diatreme structures. Localised zones of volcaniclastic rocks widen upwards, and
close the unconformity they comprise surficial volcaniclastic rocks consisting chiefly of conglomerate
(containing fragments of Gawler Range volcanics) and laminated ash together with reworked
hydrothermal breccias.
The Olympic Dam deposit contains Fe, Cu, U, Au, Ag and rare earth elements (mainly Ce and La).
Only Cu, U, Au and Ag are currently being recovered [2]. Uranium mineralization at Olympic Dam
occurs as disseminations, microveinlets and aggregates of fine-grained pitchblende intergrown with Cu-
sulphides within haematitic breccias and heterolithic haematite breccias. Pitchblende also occurs as
small aggregates intergrown with or replacing breccia material. Minor abundance of brannerite and
coffinite are intimately related with the pitchblende. Zones of narrow, higher-grade uranium frequently
occur in zones of bornite–chalcocite, particularly with haematite breccias [4]. Some zones of high-grade
uranium mineralization cut the bornite–chalcopyrite interface. As of 1 January 2017, total recoverable
reserves for the Olympic Dam deposit were reported as 918 256 tU at an average grade of 0.025 %U
[5].

Unconformity-related deposits

Approximately 19% of resources are linked with Proterozoic unconformities, especially in the Alligator
Rivers field in the Northern Territory (Ranger, Koongarra and Jabiluka). The Alligator Rivers Uranium
Field (ARUF), which lies within the Pine Creek Inlier, is located ~220 km east of Darwin. The area’s
mineral potential was recognised in 1967, when the Bureau of Mineral Resources published a revision
of the 1:500 000 scale geological map of the Katherine–Darwin region, which depicted probable
Archaean basement in the Alligator Rivers area [2, 6]. The Archaean rocks were portrayed to be overlain
unconformably by metamorphosed and deformed Palaeoproterozoic strata, which, in turn, were overlain
by Mesoproterozoic sandstones of the McArthur Basin. The map underscored similarities to the deposits
of uranium in the Archaean–Palaeoproterozoic–Mesoproterozoic setting at Rum Jungle. The cover
sandstones in the Alligator Rivers area were dated to Late Palaeoproterozoic age [7, 8].

The Koongarra, Jabiluka and Ranger 1 deposits and most of the major prospects of uranium occur
adjacent to the Archean Nanambu Complex forming a regional dome. They are hosted in the lower
member of the Palaeoproterozoic Cahill Formation, which is characterized by the existence of
metamorphosed carbonate rocks and includes interlayered chloritized feldspathic quartzite, quartz
schist, mica schist, para-amphibolite and calc-silicate rock. The Myra Falls metamorphics, which host
the Nabarlek deposit, are likely metamorphosed equivalents of the lower member of the Cahill
Formation [2]. The deposits of uranium were formed after the main period of regional metamorphism.
According to U–Pb isotope dating, the age of mineralization at Ranger is 1737 ± 20 Ma age whereas
that at Jabiluka is 1437 ± 40 Ma [9]. Rb–Sr, U–Pb, Pb–Pb and Sm–Nd analysis of uraninite and minerals
from the inner alteration zone produced ages of primary mineralization in the range 1650–1610 Ma at
Jabiluka and Nabarlek [10].

Unconformity-related deposits have also been found in the Pine Creek Uranium Province, in the South
Alligator River Valley Mineral Field (SARMF), 220 km south-east of Darwin and in the Rum Jungle
Mineral Field (RJMF), 90 km south of Darwin. Four deposits have been mined in the RJMF — Mount
Burton, White’s, Dyson’s, and Rum Jungle Creek South — two of which produced Cu as well [4].
Production of uranium from the was taken from 14 small deposits, including Coronation Hill, Rock
Hole, El Sherana and El Sherana West, Palette and Saddle Ridge [2, 6, 11].

The Kintyre deposit in the Rudall River area (Western Australia) exists in metasediments of the
Yandagooge Formation in the Rudall Complex next to the unconformity with the Neoproterozoic
Coolbro Sandstone. The deposit shares similar characteristics with the deposits in the Alligator Rivers
region. The favourable lithologies for mineralization are interbedded chlorite schist and chert.
Mineralization occurs within a system of narrow, closely spaced veins. These veins occur along the
cleavage of a major NW trending shear zone, which has faulted the Coolbro Sandstone. Ore zones are
formed by multiple sets of tightly spaced mineralized veins. The ore zones are divided into five ore
bodies, which together make up the Kintyre deposit and include the Pioner, Kintyre and East Kintyre,
Nerada, Whale and East Whale deposits [2, 12].

Sandstone-hosted uranium deposits


Sandstone-hosted uranium deposits account for 4% of resources. The known sandstone hosted uranium
deposits are located in Western Australia, South Australia, Northern Territory and Queensland. The
basins and uranium fields containing these deposits are the Canning Basin (Western Australia),
Carnarvon Basin (Western Australia), Gunbarrel Basin (Western Australia), Ngalia Basin (Northern
Territory) and Amadeus Basin (Northern Territory), Westmoreland–Pandanus Creek field
(Queensland), Eucla Basin (South Australia) and Frome Embayment field (South Australia).

The Frome Embayment is a lobe on the southern portion of the Callabonna Sub-basin in the south-
western part of the Lake Eyre Basin [13]. The Callabonna Sub-basin contains shallow water sediments
of Tertiary age. The Barrier Ranges, Olary and Flinders bordering the embayment are comprised chiefly
of Precambrian and Cambrian sedimentary and metamorphic rocks which comprise widespread
occurrences of disseminated uranium mineralization and many small uranium deposits [1].

In Early Tertiary, well-sorted sand (Eyre Formation) was deposited as a thin, laterally continuous
horizon occupying the full width of the northern part of the Sub-basin. The equivalents of the Eyre
Formation to the south (i.e., angular, poorly sorted fluvial sand and interbedded silt and clay) were
deposited in major stream channels of limited areal extent [14]. The channels were incised into
Precambrian basement and Late Cretaceous Marree Subgroup marine clay. Dolomite, sand and clay of
the Mioene Namba Formation form a continuous sequence disconformably overlying the channel
sediments. A thicker sequence of the Namba Formation was deposited near the Flinders Ranges,
forming the small Poontana Sub-basin.

The Goulds Dam, Yarramba, East Kalkaroo and Honeymoon deposits are hosted in palaeochannel sands
of the Palaeocene–Eocene Eyre Formation, whereas the Beverley deposit lies within the sands of the
overlying Miocene Namba Formation [2]. The palaeochannels in the southern part of the Frome
Embayment border a structural high in the underlying basement known as the Benagerie Ridge. The
Beverley North and Four Mile East deposits are in the Eyre Formation, whereas the Four Mile West is
believed to be hosted in sediments equivalent to the Cretaceous Bulldog Shale.

The Tertiary sands were sourced from Precambrian granitic and metamorphic rocks in the neighbouring
uplands and were deposited in the channels together with abundant plant matter. Uranium contained in
rock fragments and mineral detritus was deposited together with the channel sands [2, 14].

Two genetic models have been proposed. One model suggests that oxidizing groundwater, flowing
sluggishly through the channel sands, leached uranium and reprecipitated it, down-gradient, at the redox
interface. Roll front deposits developed at the redox interface, especially where groundwater flow was
hampered by lower permeability and thinner sand units towards the channel banks. The migration of
groundwater resulted in oxidation of pyrite and organic matter, imparting red and orange coloured iron
oxide stains on the sands [2]. The second model suggests that uranium was introduced in solution rather
than in rock fragments and mineral detritus, was transported through the palaeochannel and was
precipitated in favourable reducing settings [2].

In the Eucla Basin, uranium exists in the basal Eocene palaeochannel sediments overlying the Archaean
and Proterozoic gneiss, granites and volcanics of the Gawler Craton [2]. Low-grade mineralization of
significant extent is known in the Wynbring and Warrior palaeochannels. In addition, mineralization is
known to be present further south in the Narlaby and Yaninee palaeochannels.

The Westmoreland deposits are located in the Palaeoproterozoic–Mesoproterozoic McArthur Basin,


north-west Queensland, 400 km NNW of Mount Isa, in an area adjoining the Pandanus Creek area in
the Northern Territory. The McArthur Basin is a 6–10 km thick sequence of sedimentary and volcanic
rocks that were deposited between 1 800 and 1 575 Ma. The four main geological settings of the deposits
and occurrences of uranium in the Westmoreland–Pandanus Creek field are as follows [2, 6]:
(i) Stratabound mineralization in the uppermost sandstone unit of the basal Westmoreland
Conglomerate, subparallel to the contact with overlying basic volcanics of the Seigal Volcanics
(Redtree, Junnagunna, Long Pocket deposits, Southern Comfort occurrence);
(ii) Discordant, steeply dipping zones of mineralization contiguous to the contact with intrusive
basic dykes and sills coeval with the overlying mafic Seigal Volcanics (Wanigarango,
Oogoodoo occurrences, and Mageera Huarabagoo deposit);
(iii) Mineralization associated with fractures in altered basic volcanics (King’s Ransom, El Hussen,
Old Parr occurrences and Cobar 2 deposit); and
(iv) Mineralization related with shear zones within altered acid volcanics (Pandanus Creek
occurrence).

In the Amadeus Basin, the Angela and Pamela uranium deposits lie within the Undandita sandstone
member of the youngest unit (i.e., Brewer Conglomerate) in the Amadeus Basin. This sandstone
member is the uppermost unit of a thick sequence of terrigenous sediments (i.e., the Pertnjara Group)
of Late Devonian–Early Carboniferous age [15]. The Undandita Member consists of medium- to coarse-
grained lithic arkose interbedded with thin mudstone units, and fine- to coarse-grained lithic sandstones.
South of the MacDonnell Ranges, this sequence interfingers with the Brewer Conglomerate. In the
Missionary Syncline, 15 km south of Alice Springs, the sequence attains a maximum thickness of
3000 m. The sediments are mostly oxidised, although preserved within the sequence if a wedge-shaped
zone of reduced sandstone [1.16]. The area of the reduced sandstones within the Undandita Member is
defined by a redox boundary; the upper part of it is where the uranium deposits are located [15].

The Ngalia Basin is an elongate, intracratonic basin filled by continental and marine strata – chiefly
arenaceous with interbedded dolomite and shale – of Neoproterozoic and Palaeozoic age. The basement
is comprised of highly deformed metamorphics, granites and sediments of the Palaeoproterozoic Arunta
Block. The sequence is made up of 11 formations with maximum cumulative thickness of ~7500 m.
Most formations are separated by unconformities [17].

Uranium mineralization occurs in the lower part of the Mount Eclipse Sandstone, which is of Late
Devonian–Late Carboniferous age, and hosted by medium- to coarse-grained feldspathic sandstones
commonly cemented by carbonate [2]. Drilling in 1980–1988 identified three deposits, together called
the Mulga Rock deposits, hosted by Eocene palaeochannel sediments of the Gunbarrel Basin [2]. Peat
and clayey peat, which occur just beneath the redox limit at the bottom of the weathered zone, host the
uranium mineralization. The mineralised zones are horizontal and occur at a depth of 20–50 m,
dependent on variations in surface topography and changes in the level of the redox limit [2].

A major unconformity exists between the Carnarvon Basin shelf strata and the Archaean(?)–
Mesoproterozoic metasedimentary and granitic basement rocks [2]. Basal conglomerate in
palaeochannels is followed by other formations that cut the basement to the east. The Yarraloola
conglomerate is overlain by the Cretaceous shallow water/marine Birdrong Sandstone, which, in turn,
is overlain by marine shale and radiolarite. The Cretaceous strata are overlain by Cenozoic calcareous
gravel, clay and siltstone with an erosional hiatus. Uranium mineralization occurs in the basal parts of
the Yarraloola Conglomerate and the Birdrong Sandstone, and it is related with strong oxidation of the
erst while sediments are reduced by groundwater flowing along the confined aquifer beneath the
Muderong Shale [2].

Within the Yampi Embayment of the Canning Basin, the Early Carboniferous Yampi Sandstone hosts
the Oobagooma deposit [2]. The embayment is a NW-trending fault-controlled graben bordered on
three sides by Proterozoic metamorphics. The Yampi Sandstone was deposited in a deltaic environment
influenced by fluvial and tidal processes. Mineralization is hosted by sandstones rich in pyrite and
organic matter. Higher-grade mineralization occurs in two zones: a 1–6 m thick lower band lying at
depths of 65–85 m, and a 1–5 m thick upper band lying at depths of 48–55 m [2]. In the upper band,
mineralization forms a roll front deposit and it is controlled by a combination of structural,
sedimentological and redox factors [2].
Surficial (calcrete) deposits

Surficial (calcrete) deposits account for 4% of Australia’s uranium resources and occur mostly in the
Yilgarn Craton, at Centipede, Lake Maitland, Lake Way and Yeelirrie (Western Australia).

Calcretes have been forming since Pliocene under semi-arid to arid climatic conditions. Carnotite
mineralization is extensive in calcreted trunk valleys of the Tertiary drainage system that developed
over an area of 400 000 km² of south-western Australia [18]. However, the significant prospects and
known calcrete hosted deposits of uranium are restricted to the granitic rocks in the northern sector of
the Yilgarn Craton [2]. Anomalous concentrations of surficial uranium mineralization in calcreted
drainage channels spread north of the Yilgarn Craton and exist in the Proterozoic Bangemall Basin and
Gascoyne Complex, in the Archaean Pilbara Craton, and in portions of the Northern Territory and South
Australia. The Yilgarn Craton, the ‘Menzies line’ to the south and the ‘Meckering line’ to the west
controlled the extent of the distribution of the principal calcrete uranium deposits [19].

In the northern part of the Yilgarn Craton, the Yeelirrie deposit, located 650 km north-east of Perth, is
hosted within valley calcretes sitting on the drainage channel of a broad flat valley [2]. The Yeelirrie
catchment area is underlain nearly wholly by highly weathered granitic rocks. The uranium deposit is a
horizontal sheet ~1.5 km wide and ~9 km long. The bulk of the mineralization is restricted to the interval
4–8 m beneath the surface, with ~90% existing beneath the water table. The mineralised material
assaying 0.08% U or greater has an average thickness of 3 m [1.20]. Roughly 90% of the mineralization
exists at the 4 m thick transition zone between the clay–quartz and the calcrete [2].

The Lake Way deposit, 16 km south-east of Wiluna, is at the north-eastern margin of Lake Way.
Mineralization occurs in clay and earthy calcrete in the lower levels of a Tertiary drainage channel
where it enters the north-east margin of Lake Way. Carnotite exists in clay–gravel, on bedding planes,
on slickenside surfaces, and as coatings on broken calcrete blocks at the interface between water table
and air, extending down to 2 m below the interface and up to 1 m above it [21]. There exist four areas
of ore grade mineralization linked by areas of sub-economic mineralization. The average thickness of
the mineralization is 1.5 m, varying from a few centimetres to a maximum of 5 m [2].

There are several uranium calcrete deposits at Lake Maitland, 102 km south-east of Wiluna. The
deposits, extending in a N–S trending zone, underlie the northern part of Lake Maitland. The
mineralised zone is 300–600 m wide and ~6 km long, 0.2–2.0 m thick (maximum 3.75 m) and lies 1.5–
2.0 m below the surface [2]. The mineralization occurs as the mineral carnotite within ‘slabby’ calcrete,
but is also present in clay, silt, and sand [2].

In the Hinkler Well–Centipede drainage system, which enters the south-western side of Lake Way,
valley calcretes cover a distance of 33 km [22]. The valley calcrete is over 2 km wide in the western
part of the system, but narrows to 0.5 km before widening into a chemical delta inward to Lake Way.
The calcrete’s thickness increases from 5 m to 15 m up-drainage. The Centipede deposit consists of
three separate lenses of higher-grade mineralization, with carnotite existing in a carbonate matrix. The
thickness of mineralised zones varies from 1 to 5 m thick and they sit below overburden that is 0–6 m
thick [23].

Other deposit types

The remaining resources occur in a variety of deposits [24]:

 Metasomatite (Mount Isa uranium field with the Valhalla deposit (Queensland));
 Metamorphic (Mary Kathleen (Queensland));
 Volcanic (Maureen, Ben Lomond (Queensland));
 Intrusive (Crocker Well, Radium Hill (South Australia));
 Quartz-pebble conglomerates (Pilbara Craton, Halls Creek Orogen, Yerrida Basin and the
Hamersley Basin (Western Australia)).

6.1.1. Uranium exploration

The existence of uranium in Australia has long been known even before it became the target of any
systematic exploration. It was first noted in Australia from Carcoar (New South Wales) in 1894, where
torbernite was discovered associated with cobalt mineralization. Two important occurrences of uranium
were found in Mount Painter (South Australia) in 1906 and at Radium Hill (South Australia) in 1910.

Uranium exploration in Australia began in 1944 in reaction to appeals from the Governments of the
United States of America and the United Kingdom. The known deposits at Radium Hill and Mount
Painter were investigated by both State and Federal Government geologists. To encourage exploration,
tax incentives were introduced by the Federal Government in 1948 for the discovery of uranium
orebodies. Further stimuli to explore and develop resources of uranium were presented in 1949 when a
6-year uranium ore buying pool in Australia was approved, which guaranteed fixed prices for uranium
ore. Tax breaks were granted in 1952, especially around known mineral fields.

Significant discoveries of uranium were made at Rum Jungle (Northern Territory) in 1949, in the South
Alligator Valley (Northern Territory) in 1953, at Mary Kathleen (Queensland) in 1954 and at
Westmoreland (Queensland) in 1956. Small occurrences were discovered at several locations
throughout the continent. Prospectors using Geiger counters were the ones who made most of the
significant discoveries during those years. As the present sales contracts were met, there seemed small
hope for further sales and exploration practically ceased in the late 1950s.

Uranium exploration resumed in 1966, owing chiefly to the exceptionally strong world perception that
utilization of nuclear energy for the generation of electricity would intensify abruptly. To encourage
exploration, the Federal Government relaxed the existing export policy for uranium in 1967 and, as a
result, expenditures for exploration of uranium rose quickly during the period 1967–1972. However,
the costs of exploration uranium diminished during 1972–1975, as the policies of the then Labor
Government actively discouraged exploration for uranium by private companies. In the later part of
1972–1975, Government funded uranium exploration was conducted by the Australian Atomic Energy
Commission.

After the election of the Liberal–National Party Coalition to Government in late 1975, exploration
increased gradually to a record level of AU$94 million (in constant 2000 AU$) in 1980. Some of the
factors attributed to this revival of exploration for uranium were:

 Sharp rises in uranium spot market prices. Prices negotiated for sales under long term contracts
also rose since the mid-1970s;
 Release in 1976 and 1977 of findings from the environmental inquiry on Ranger uranium;
 The Federal Government’s announcement of Australia’s uranium policy in 1977, which
promoted the continued growth of Australia’s uranium mining industry under stringent
regulated conditions.

Unlike earlier uranium exploration by prospectors, exploration from 1966 onwards was carried out by
major companies employing state-of-art equipment and techniques, which had respectively huge
budgets. The advancement of multi-channel gamma ray spectrometers with large volume crystal
detectors amplified the efficacy of airborne radiometric surveys.

In the 1970s, mapping and exploration led to improved understanding of the distribution of uranium,
and so the search could be focused more effectively on those geological environments regarded most
likely to host deposits of uranium. The regional mapping carried out by the Bureau of Mineral Resources
and the State geological surveys was exploited effectively by exploration teams in target selection, and
companies carried out airborne radiometric surveys using multi-channel gamma ray spectrometers. This
phase of exploration was highly successful and practically all of the significant deposits in Australia
were discovered during 1969–1980. Important discoveries during this period were Ranger (1970),
Nabarlek (1970), Koongarra (1970) and Jabiluka (1973) in the Alligator Rivers area of the Northern
Territory; Beverley (1969) and Olympic Dam (1975) in South Australia; and Yeelirrie (1972) in
Western Australia [2].

From the peak level reached in 1980, costs of exploration for uranium dropped abruptly to AU$28
million (in constant 2000 AU$) in 1983. This was brought by the economic recession in the major
industrial nations, the execution of energy conservation policies in response to the oil shocks of the
mid-1970s and the abrupt decline in spot market prices of uranium from 1976 onwards. In 1983, the
then Labor Government presented the ‘Three mines’ policy, under which exports of uranium were
allowed only from the Olympic Dam, Ranger and Nabarlek mines. Likewise, in the early 1980s, the
State Governments of Victoria and New South Wales enacted legislation to ban exploration for, and
mining of, uranium. Therefore, no uranium exploration has been carried in these States since then.

Despite of the reducing effect of the ‘Three mines’ policy on exploration for uranium, the discovery of
the Kintyre deposit (Paterson Province, Western Australia) in 1985 led to an increase in expenditures
for exploration during 1985–1988. The goal of that was to locate similar deposits elsewhere in the
Paterson Province. Subsequently, exploration waned from 1989 onwards and reached an historic low
in 1994.

In January 1994, the Native Title Act 1993 was enacted by the Federal Government. The Act obliged
exploration companies inform and discuss with Native Title parties prior to granting of exploration
tenements over lands where Native Title exists, or which are subject to a registered Native Title claim.
The Act obliged the formulation of complementary State/territory legislation — a process that has
hindered the authorization of exploration licence applications for all minerals, including uranium. This
has had a significant impact on exploration for uranium in Western Australia, where large tracts of
Crown land are affected by Native Title land claims.

After the abolishment of the ‘Three mines’ policy by Liberal–National Party Coalition in 1996, uranium
exploration increased even as demand for uranium improved. However, during 1998–1999, exploration
later waned when numerous large companies ceased to explore for uranium in Australia in response to
continued low prices.

During the late 1990s, two factors strongly affected the focus of exploration for uranium. Firstly, the
economic success of both the Olympic Dam mine (haematite breccia complex-type deposit) and the
Ranger mine (unconformity-related deposit) proved that these deposit-types of uranium are significant
targets for exploration. Secondly, fruitful advance, in the past decade, of low cost in situ leach (ISL)
technology for developing sandstone-hosted uranium deposits rekindled the exploration for this deposit-
type of uranium.

The areas chiefly explored for uranium in the late 1990s are the following:

 Arnhem Land (Northern Territory): exploration targets are unconformity-related deposits in


Palaeoproterozoic metasediments overlain by thick sandstone cover of the Kombolgie
Subgroup;
 Paterson Province (Western Australia): exploration targets are unconformity related deposits in
Palaeoproterozoic metasediments of the Rudall Metamorphic Complex, which hosts the
Kintyre orebody;
 Gunbarrel, Canning and Carnarvon Basins (Western Australia) and Frome Embayment (South
Australia): exploration targets are sandstone-hosted uranium deposits;
 Olympic Dam area (South Australia): exploration targets are breccia complex type deposits in
Mesoproterozoic granitoids of the Gawler Craton, overlain by the Stuart Shelf sedimentary
sequence;
 Westmoreland area (Queensland): exploration targets are sandstone type deposits in
Proterozoic sediments of the McArthur Basin;
 Mount Isa Inlier (Queensland): exploration targets are metasomatite type deposits in
Proterozoic metasediments;
 Tertiary palaeochannel sediments overlying the Yilgarn Craton (Western Australia):
exploration targets are calcrete type deposits;
 Mount Gee area (South Australia): exploration targets are breccia complex-type deposits in
Palaeozoic haematite breccias.

In 2001–2002, the areas chiefly explored for uranium were the Gawler Craton/Stuart Shelf region, the
Frome Embayment and Arnhem Land. In November 2001, Minotaur Resources Ltd publicised the
discovery of gold, copper, rare earth and uranium mineralization in haematite breccias at the Prominent
Hill prospect (South Australia). This mineralization exists in Proterozoic basement, which is covered
by at least 100 m of younger sedimentary rocks. The geological setting, the overlapping magnetic and
gravity anomalies, and the style of mineralization are apparently generally similar to the Olympic Dam
deposit, located roughly 150 km to the south-east. Precious metal, copper, uranium and rare earth
assemblages are also apparently similar to those at Olympic Dam, but uranium concentrations are fairly
lower. Indeed, when the Prominent Hill mine was developed uranium was not recovered.

During 2003–2004, exploration activities in the Frome Embayment intensified. Drilling tested the target
areas, which had been delineated by airborne electromagnetic surveys that established the limits of
buried palaeochannels. Heathgate Resources publicised the discovery of a new uranium-mineralised
zone, called Deep South zone, roughly 3 km south of the Beverley deposit. The Deep South ore zone
occurs in sands, like those that host the Beverley deposit, and was mined together with the other deposits
there.

Southern Cross Resources has endured exploration in the region of the Goulds Dam, East Kalkaroo and
Honeymoon deposits. A new low–medium grade uranium-mineralised zone was found in an area of the
Yarramba palaeochannel, ~1.5 km north-west of the Honeymoon deposit. The new zone, called Brooks
Dam prospect, has been tested by drilling over a distance of 1 km along the palaeochannel, and the
company thinks that it may stretch farther to the south like the main Honeymoon deposit.

In 2004, WMC Resources conveyed that exploration drilling in the south-eastern part of the Olympic
Dam deposit has delineated considerable additional resources, close to 30% increase over the resources
to December 2003. Minotaur Exploration Pty Ltd has endured exploration drilling of the copper–gold–
uranium–rare earth mineralization at the Prominent Hill deposit (South Australia).

Important discoveries during 2005 and 2006 included: extensions of the Valhalla and Skal deposits
(Mount Isa region (Queensland)), major extensions of the Olympic Dam deposit (South Australia), and
Four Mile deposit (8 km north-west of the Beverley mine (South Australia)).

At Four Mile deposit, in the Frome Embayment, drilling has delineated an extensive mineralised area
measuring 5 km2 in Palaeogene (Tertiary) sands along the flanks of Proterozoic basement rocks of the
North Flinders Ranges. Within this extensive area are two deposits. As of 2012, the East and West
deposits were estimated to contain inferred resources of 11 000 tU and 16 000 tU, respectively [1.5]. A
third deposit, Four Mile North-East is being investigated.

The principal areas where exploration for uranium was undertaken during 2007 and 2008 were the
Mount Isa region (Queensland), the Alligator Rivers region (Northern Territory), the Gawler
Craton/Stuart Shelf region (South Australia) and the Frome Embayment (South Australia). Several
discoveries were announced, including the N-147 project in the Alligator Rivers region (Northern
Territory), the Thunderball deposit in the Pine Creek geosynclines (Northern Territory), the Blackbush
deposits (South Australia) and the Double 8 deposit in Tertiary palaeochannel sands (Western
Australia).

Drilling and exploration delineated considerable extra resources in the south-eastern part of the Olympic
Dam deposit. As of June 2010, total measured and indicated plus inferred resources (JORC code)
amounted to 9075 million t with average grades of 0.87% Cu, 0.023% U (2 078 000 tU), 0.32 g/t Au
and 1.5 g/t Ag [5].

In 2009, Heathgate Resources found two new sandstone-hosted deposits, Pepegoona and Pannikan, in
the Frome Embayment, near Beverley. Exploration drilling continued at Carrapateena, a haematite
breccia complex-type deposit similar to Olympic Dam. Total inferred resources are 203 million t with
average grades of 1.31% Cu, 0.027% U, 0.56 g/t Au and 6 g/t Ag [5].

In 2018, exploration expenditures for uranium decreased to AUD 12.3 million from AUD 19.8 million
in 2017. Main activities in 2017–2018 were [25]:

Mulga Rock: In January 2018, Vimy Resources released a definitive feasibility study for the Mulga
Rock project. It involved open-pit mining of four polymetallic deposits with uranium situated in
sandstone-hosted carbonaceous material. It has a 15-year mine life and is anticipated to produce 1
346 tU annually.

Yeelirrie: The project received environmental approval from the Western Australia Government in
January 2017 and from the Commonwealth Government in April 2019. It is estimated that production
from the Yeelirrie would be nearly 3 300 tU per year over 19 years.

Wiluna: Toro expanded the Wiluna project proposal, which encompasses the Lake Maitland and
Millipede resources, and received environmental approval from the Western Australian Government in
January 2017 and the Commonwealth in July 2017.

Historical exploration data are given in Figure 2, (4 208 536 metres of drilling within a total expenditure
of USS $1 751 397 000.

FIG. 2. Domestic uranium exploration data for Australia. Comparison of exploration expenditures, drilling and
uranium market price (US$ current).
Foreign exploration and development expenditure

In 1967–1971, expenditures reported by Australian companies engaged in exploration for uranium in


foreign countries amounted to US $210 000, nearly all of which was spent in New Zealand.

In 1999, Paladin Resources Ltd (an Australian exploration company) bought the Kayelekera uranium
deposit in Malawi, and in 2002, the Langer Heinrich uranium deposit in Namibia. At both projects,
engineering and feasibility studies were undertaken during 2001 and 2002.

A comprehensive drilling programme and feasibility study on the Langer Heinrich deposit during 1997
and 1998 was completed by another Australian exploration company, Aztec Resources Ltd, which is
the previous owner of the deposit. In 2004, a broad exploration-drilling programme delineated areas of
extra resources contiguous to the Langer Heinrich deposit. In May 2005, the company opted to ensue
with development of the deposit. During 2005 and 2006, Paladin finished the development of an open
cut mining operation at Langer Heinrich. Mine production commenced in early 2006 [25, 26]. Paladin
also conducted further exploration at the Kayelekera deposit in Malawi. During 2007 and 2008, Paladin
finished the development of an open cut mining operation at the Kayelekera deposit. Mine production
commenced in May 2009. Production reached 681 t in 2010 and 846 t in 2011. The project was put into
care and maintenance in 2015.

In 2015–2016, several Australian mineral companies conducted exploration activities in Canada,


Kyrgyzstan, Malawi, Namibia, Peru and Tanzania.

6.1.2. Uranium resources

6.1.2.1. Identified conventional resources

The historical variation in reasonably assured resources and inferred resources are shown in Figures 3
and 4 respectively [5]. Australia’s identified conventional resources recoverable at costs of <US
$260/kgU amounted to 2 054 800 tU as of 1 January 2017 [5]. The resources are summarised in Tables
1 – 3 and their locations shown in Fig. 6.1.

As of 1 January 2019, Australia’s total identified resources recoverable at a cost of less than
USD 130/kg U amounted to 1 183 861 tU of reasonably assured conventional resources and 508 807
tU of inferred conventional resources [25].

Most of Australia’s identified conventional resources, recoverable at <US $130/kgU, are contained
within the following six deposits:

(i) Olympic Dam: This is a haematite breccia-type deposit and is the largest uranium resource in
the world. Based on ore reserve and mineral resource data described by BHP Billiton, as of
June 2018, it was estimated that the total resources of the deposit amount to 2 064 480 tU at an
average grade of 192 ppm U [28]. Uranium is a co-product of copper mining at Olympic Dam;
silver and gold are likewise recovered;

(ii) Ranger: The orebodies (Ranger 1 n°1 and Ranger 1 n°3) were found in 1970 by a joint venture
of the Electrolytic Zinc Company of Australia Limited and the Peko Wallsend Operations Ltd
(Peko). As of 1 January 2017, ERA estimated that the total resources of the Ranger deposit
amounted to 47 463 tU [29] The mineralised zone has a strike length identified to date of about
1.2 km, lying at a depth of between 250 and 550 m, immediately east of the pit, and remaining
open to the north.

(iii) Jabiluka 1: This uranium deposit was discovered by Pancontinental Mining Limited in 1973
in the Northern Territory. Additional drilling outlined the uranium orebody of the larger
Jabiluka 2 deposit some 1 km to the east. Jabiluka is located 20 km north of Ranger and 230 km
east of Darwin. The deposit is located within the Kakadu National Park, although the area of
the mine lease is excluded from the National Park and is adjacent to the Ranger lease. In 2000,
after intensive drilling in the underground access to the Jabiluka orebody, the overall resource
was revised by the Energy Resources of Australia, with some decrease in actual reserves. The
project has been abandoned;

(iv) Koongarra: This is a relatively high grade but small uranium deposit in the Alligator Rivers
area of the Northern Territory. It is located roughly 30 km south of Ranger, the lease area being
on Aboriginal land. Koongarra is an unconformity-related type deposit. The project has been
abandoned;

(v) Kintyre: This deposit is a medium- grade uranium orebody with a limited surface outcrop in the
remote Rudall region of Western Australia. This is on the western edge of the Great Sandy
Desert in the eastern Pilbara region of Western Australia, roughly 1200 km north-east of Perth
and some 70 km south of Telfer. It was found in 1985 by Rio Tinto Exploration through surface
follow-up of several radiometric anomalies detected as the result of an airborne survey. Cameco
bought the project in 2008;

(vi) Yeelirrie: This deposit is located between Leinster and Wiluna, Western Australia, some
420 km north of Kalgoorlie. It is the largest and richest sedimentary calcrete deposit in the
world. Uranium is present as carnotite (hydrated potassium uranium vanadium oxide). The
deposit extends for 9 km; it is up to 1.5 km wide, up to 7 m thick and is overlain by overburden
with an average thickness of 7 m. The Yeelirrie uranium deposit was acquired by Cameco from
BHP Billiton in 2012.

The UDEPO database lists the most significant deposits for Australia as Olympic Dam, Ranger 1 n°3,
Carrapateena, Jabiluka 2, Ranger 1 n°1, Yeelirrie, Ranger Deeps, Valhalla, Mount Gee, and Kintyre.

FIG. 3. Historical variation of recoverable reasonably assured resources within various cost categories in
Australia. Periods where no resources are shown in any cost categories are periods where resources are not
reported, either by the Member State or as a secretariat estimate.
FIG. 4. Historical variation of recoverable inferred resources within various cost categories in Australia. Periods
where no resources are shown in any cost categories are periods where resources are not reported, either by the
Member State or as a secretariat estimate.

TABLE 1. KNOWN URANIUM RESOURCES IN OTHER DEPOSITS–DISTRICTS [2]

Deposit Grade (%U) Resource (tU)

Angela (Northern Territory) 0.11 11 860


Ben Lomond (Queensland) 0.193 4120
Beverley (South Australia) 0.152 18 320
Bigrlyi/Ngalia (Northern Territory) 0.082 10 346
Carapateena (South Australia) 0.0155 144 000
Crocker Well (South Australia) 0.0244 4 490
Four Mile deposits (South Australia) 0.27 46 603
Goulds Dam (South Australia) 0.043 9 625
Lake Maitland (Western Australia) 0.047 10 180
Manyingee (Western Australia) 0.072 9 960
Maureen (Queensland) 0.077 2 530
Mount Fitch (Northern Territory) 0.03 5 580
Mount Gee (South Australia) 0.052 26 590
Mulga Rock district (Western Australia) 0.048 34 690
Mullaquana district (South Australia) 0.025 15 770
New Well (Napperby) (Northern Territory) 0.032 3 092
Nolans Bore (Northern Territory) 0.017 9 300
Nyang, Carley Bore (Western Australia) 0.027 6 545
Oobagooma (Western Australia) 0.102 8 345
Ponton, Double 8 (Western Australia) 0.025 6 615
Prominent Hill (South Australia) 0.01 10 300
Skal, Andersons, Bikini, Watta (Queensland) 0.045 17 025
Thatcher Soak E and W (Western Australia) 0.024 8 925
Valhalla (Queensland) 0.069 29 340
Westmoreland district (Queensland) 0.07 19 980
Wiluna district (Western Australia) 0.048 14 010

TABLE 2. REASONABLY ASSURED RESOURCES (tU) [25]


(As of 1 January 2019)

Deposit type <US $80/kgU <US $130/kgU <US $260/kgU

Unconformity-related n.a. 110 249 114 044


Sandstone n.a. 63 893 72 776
Haematite breccia complex n.a. 898 546 967 737
Granite-related n.a. 322 322
Intrusive n.a. 13 439 18 761
Volcanic-related n.a. 2731 5125
Metasomatite n.a. 29 136 34 448
Surficial n.a. 65 545 71 632

Total n.a. 1 183 861 1 284 845

TABLE 3. INFERRED RESOURCES (tU) [25]


(As of 1 January 2019)

Deposit type <US $80/kgU <US $130/kgU <US $260/kgU

Unconformity-related n.a.a 37 491 55 230


Sandstone n.a. 42 542 105 375
Haematite breccia complex n.a. 428 506 532 127
Granite related n.a. 0 28
Intrusive n.a. 0 10 785
Volcanic-related n.a. 0 1 089
Metasomatite n.a. 0 11 515
Surficial n.a. 268 48 336

Total n.a. 508 807 764 575

a n.a.: not available.

Undiscovered conventional resources and nonconventional resources

Estimates are not made of Australia’s undiscovered conventional resources. Estimates are also not made
of Australia’s uranium resources in the categories of unconventional resources and other materials.
Potential for new discoveries

Australia does not report undiscovered resources (prognosticated or speculative) to the Red Book.
However, based on geological knowledge of previously discovered deposits and regional geological
information, there is significant (high) potential for additional uranium deposits to be found, including:

 Unconformity-related deposits, including high grade deposits at the contact and immediately
above the unconformity, particularly in Arnhem Land in the Northern Territory but also in the
Granites–Tanami region (Northern Territory and Western Australia), Paterson Province
(Western Australia) and in the Gawler Craton (South Australia);
 Haematite breccia deposits, particularly in the Gawler/Stuart Shelf region of South Australia;
 Sandstone-hosted deposits in sedimentary strata in various regions adjacent to uranium
enriched basement, in particular in the Frome Embayment region (South Australia);
 Metasomatite type deposits in the Mount Isa region of Queensland;
 Calcrete type deposits in Cenozoic palaeochannel sands in Western Australia.

Uranium production

Uranium was first recovered in Australia as a by-product of ore mined for radium at Radium Hill and
Mount Painter (South Australia). About 2 000 t of ore were treated and the uranium content had minor
commercial interest for use in ceramic glazes. As only a fraction of the uranium content of the ore was
recovered, this production can be considered insignificant. Between 1954 and 1971, Australia produced
some 7 732 tU from plants at five locations. The first phase of uranium production in Australia ceased
after the closure of the Rum Jungle plant in 1971

Uranium production details, including historical data, are given in Tables 4 – 5 and Figure 5 reflecting
a total production of 219 028 tU up to 2019. This is comprised of 129 684 tU, 75 169 tU, 13 337 tU
and 838 tU from open pit, co/by-product (Olympic Dam underground), in situ leach and other
underground mining respectively.

Total uranium mine production for 2018 from the three operating mines, Four Mile, Olympic Dam and
Ranger amounted to 6 526 tU.

TABLE 4. URANIUM PRODUCTION PHASE 1: 1954–1971 [2]

South Alligator South Alligator


Rum Jungle Radium Hill Mary Kathleen Valley (NT)
Mine Valley (NT)
(NTa) (SAb) (Qldc) (South Alligator
(United Uranium)
Uranium)

Mining method OPd UGe OP OP–UG UG


Operational
period 1954–1971 1954–1962 1958–1963 1959–1964 1959–1962

Average grade
0.24–0.34 0.09–0.13 0.13 0.30–0.58 0.95
(%U)
Production (tU) 2 993 721 3 460 441 117

a NT: Northern Territory.


b SA: South Australia.
c Qld: Queensland.
d OP: open pit.
e UG: underground mining.
TABLE 5. URANIUM PRODUCTION CENTRE TECHNICAL INFORMATION [25]
(As of 1 January 2019)

Production centre Ranger Olympic Dam Four Mile

Operational status Operating Operating Operating

Startup date 1981 1988 2014

Source of ore:
 Deposit name Ranger 3 Olympic Dam Four Mile
 Deposit type Unconformity Haematite breccia Sandstone
 Reserves (tU) 31 278 1 304 160 22 561
 Grade (%U) 0.24 0.041 0.29

Mining/milling operation:
OP & UG
 Type 85% UG&OPb ISLc
 Mining recovery (%) 4.5 M t/year 85% n.a.%
 Size 12 M t/year 1.62 M L/h

Processing plant:

 Acid/alkaline Acid Acid Acid

 Type SXe FLOTf, SX IXg

 Average process recovery (%) 87 68 85


 Nominal capacity (tU/year) 2 100 3 250 1 700

a OP: open pit.


b UG: underground mining.
c ISL: in situ leaching.
d CWG: crush wet grind.
e SX: solvent extraction.
f FLOT: flotation.
g IX: ion exchange.

Uranium production in Australia resumed in 1976. During this second phase, production was recorded
from the operations at Mary Kathleen (4 072 tU in 1976–1982), Nabarlek (9 208 tU in 1979–1988),
Ranger (since 1981), Olympic Dam (since 1988), Beverley (since 2000) and Honeymoon (since 2011).
From 1976 to the end of 2010, production from these operations totalled 162 546 tU.

The Ranger mine is about 230 km east of Darwin, in the Northern Territory, and is surrounded by the
Kakadu National Park. The mine (Ranger Pit 1) opened in 1981 at a production rate of approximately
2 800 tU/year, which has since been expanded to 4 660 tU/year capacity. Mining at the second pit
(Ranger Pit 3) commenced in 1997 and finished at the end of 2012, although production continued from
large ore stockpiles. Treatment is conventional acid leach.
FIG. 5. Historical uranium production in Australia (Data in light green are from the Red Book Retrospective, in
dark green from Red Books).

Uranium mining at Ranger will cease in January 2021. Final rehabilitation is to be completed by January
2026.

The Olympic Dam mine lies about 560 km north of Adelaide, in an arid part of South Australia.
Olympic Dam is the largest single uranium resource in the world and mining operations commenced in
1988. The deposit is mined underground at a depth of 350 m. The mine produces copper, with gold and
uranium as major by-products. Annual production capacity has been expanded from 1 500 to 3 820 tU.

About 80% of the uranium is recovered by conventional acid leach of flotation tailings derived from
copper recovery. Most of the remaining 20% is from acid leach of the copper concentrate, but this
concentrate still contains up to 0.15% U.

There are plans to greatly increase the mine’s size and output by accessing the orebody through
excavation of an open pit measuring about 4.1 km × 3.5 km and 1.0 km deep. As of 2009, BHP Billiton
is investigating the feasibility of expanding the capacity of the Olympic Dam operations to produce
16 100 tU (19 000 t U3O8) annually. It is proposed to mine the southern portion of the deposit by open
pit in conjunction with underground mining (sub-level stoping) in the northern portion of the deposit.

A draft environmental impact study was released on 1 May 2009. Following feedback, a supplementary
environmental impact study was submitted to the Federal Government in December 2010 and approval
of the proposal was announced in October 2011. BHP Billiton moved the project to the feasibility study
stage in March 2011. However, a decision to postpone the expansion and undertake further studies was
announced in late 2012.

In October 2019, BHP submitted a referral to the South Australian and Commonwealth governments for
a potential incremental increase in production and renewal of site infrastructure.

The Beverley uranium deposit is situated 520 km north of Adelaide, on the plains north-west of Lake
Frome. It is a relatively young sandstone deposit, with uranium mineralization leached from the Mount
Painter region, and was Australia’s first commercial ISL operation. Mine construction started in 1999.
The deposit consists of three mineralised zones (north, central and south) located within a buried
palaeochannel (the Beverley aquifer) in Tertiary sediments of the Frome Basin. Groundwater salinity
ranges from 3 000 mg/L total dissolved solids in the north to 12 000 mg/L total dissolved solids in the
south.

As the main orebodies became depleted, the Beverley North project was initiated in 2009 and in 2010
a field leach trial at Pepegoona was successful. This became a satellite operation, with loaded resin
being trucked to the treatment plant. In 2011, mining commenced at nearby Pannikin with the
installation of a second satellite plant [3, 27].

The Beverley and Beverley North mines have been in care and maintenance since 2012 and 2018
respectively. Current production comes from the Four Mile mine.

Honeymoon deposit is currently in care and maintenance. In 2018, Boss Resources Ltd completed a
program of field leach trials that successfully demonstrated the application of the ion-exchange process.

6.1.2.2. Future projects

Several future projects are summarised in Table 6.

Uranium production at Honeymoon, an ISL mine, started in late 2011 and it is anticipated to increase
progressively to 340 tU/year.

At Yeelirrie, BHP Billiton undertook drilling to upgrade the resource estimate (total current resources
are 44 500 tU at an average grade of 0.13% U) and started a feasibility study for development of the
deposit. The project was sold to Cameco in 2012. When the acquisition is complete, Cameco intends to
conduct its own mineral resource estimation to generate an NI 43–101 compliant value.

The Wiluna project comprises four shallow calcrete hosted deposits: Lake Way, Lake Maitland,
Centipede and Milipede. In 2010, Toro Energy excavated an evaluation pit at the Centipede deposit to
increase confidence both in the resource estimates and in the proposed mining method. Environmental
approvals were well advanced by the end of 2012. In 2018, the Wiluna Uranium Project has received
federal and state government environmental approvals for mining uranium at the four deposits.

At Kintyre, Cameco commenced an environmental impact assessment and, in 2012, it completed a


prefeasibility study and signed a mine development agreement with the relevant Aboriginal group. A
detailed feasibility study is yet to be undertaken. The deposit is intended to be mined by open pit. The
proposed annual production will be between 2 300 and 3 000 tU. The project received environmental
approval in 2015.

6.1.3. Environmental activities and sociocultural issues

6.1.3.1. Northern Territory

The Federal Government has responsibility for supervising the environmental management of uranium
mining in the Alligator Rivers region, which is Federal land. The Ranger mine, Jabiluka mine (on care
and maintenance after initial development work) and Nabarlek mine (mined out and nearing completion
of remediation work) lie within this region. The Northern Territory Government has responsibility for
the day-to-day regulation of mining activities, with the responsibilities determined by a suite of
legislation and agreements between the two Governments (Federal and State) to minimise any
environmental impacts arising from uranium mining. Environmental supervision and oversight of
operations in the Alligator Rivers region are provided by the Supervising Scientist, a statutory officer
of the Federal Government, who derives authority from the Environment Protection (Alligator Rivers
Region) Act.
The Federal Government’s Office of the Supervising Scientist has overseen the environmental aspects
of uranium mining operations in the Alligator Rivers region since mining commenced at Nabarlek and
Ranger. It is also discharging this role in relation to the potential development of Jabiluka. The
Supervising Scientist, supported by the Environmental Research Institute of the Supervising Scientist,
coordinates and supervises measures for the protection and restoration of the environment of the
Alligator Rivers region in response to the effects of uranium mining. The Office of the Supervising
Scientist measures environmental performance at the mines, including the rehabilitation of Nabarlek,
through twice yearly audit exercises.

TABLE 6. FUTURE URANIUM PRODUCTION CENTRES [25]


(As of 1 January 2019)

Name Honeymoon Mulga Rock Yeelirrie

Status Care and Maintenance Planned Planned

Startup date 2011 Unknown Unknown

Source of ore:
• Deposit name Honeymoon, Goulds Dam,
Yeelirrie
Jasons Princess, Shogun,
Ambassador,
Emperor
Calcrete
Sandstone Sandstone28 836
• Deposit type
39 409
20 732
• Reserves (tU)
0.08
0.13
0.15
• Grade (%U)

Mining operation:
• Type ISL OP OP
• Size (Mt ore/year) n.a. n.a. n.a.
• Average mining recovery 95 n.a.
(%)

Processing plant:
• Acid/alkaline Acid
Acid
• Type Alkaline
SX&IX
• Size n.a. n.a.
n.a.85
• Average process recovery 87.3 80
(%)

Nominal production capacity 769 1346 3 265


(tU/year)

Expansion plans n.a. n.a. n.a.

Other n.a. n.a. n.a.


Name Wiluna Kintyre

Status Planned Planned

Startup date Unknown Unknown

Source of ore: Centipede, Lake Way, Kintyre


• Deposit name Millipede,
Lake Maitland
• Deposit type
Calcrete Proterozoic unconformity
• Reserves (tU)
16 653 18 253
• Grade (%U)
0.09 0.53

Mining operation:
• Type OP OP
• Size (Mt ore/year) n.a. n.a.
• Average mining recovery n.a. n.a.
(%)

Processing plant:

• Acid/alkaline Alkaline Alkaline


• Type IX n.a.
• Size 1700
• Average process recovery 80 80
(%)

Nominal production capacity 577 2290


(tU/year)

Expansion plans n.a. n.a.

Other n.a. n.a.

6.1.3.2.South Australia

The South Australian Government has responsibility for regulating the Olympic Dam, Beverley and
Honeymoon projects. Olympic Dam is principally regulated under site-specific South Australian State
Government legislation — the Roxby Downs (Indenture Ratification) Act 1982 as amended (the
Indenture). Beverley and Honeymoon are regulated under a range of South Australian legislation
applicable to mining, including mining of radioactive substances.

The Department for Manufacturing, Innovation, Trade, Resources and Energy regulates day-to-day
mining activities, with the Environment Protection Authority responsible for radiation protection issues.

6.1.4. National policies relating to uranium

The Federal Government supports the development of a sustainable Australian uranium mining sector
in line with world’s best practice environmental and safety standards and allows the export of uranium
to those countries which observe the Treaty on the Non-Proliferation of Nuclear Weapons and which
are committed to non-proliferation and nuclear safeguards. Non-nuclear-weapon States must also have
in force an Additional Protocol.

In November 2008, the Government of Western Australia overturned the ban on uranium mining put in
place by the previous State Government. Mining of uranium in Western Australia is subject to strict
safety and security provisions, including meeting all the necessary international safeguards and rigorous
environmental approvals for mining and transporting uranium. A decade’s long ban on uranium mining
was reversed in Queensland in 2012.

References
[1] CENTRAL INTELLIGENCE AGENCY, The World Factbook: Australia, Washington, DC,
https://www.cia.gov/library/publications/the-world-factbook/index.html (accessed August 2012).
[2] McKAY, A., MIEZITIS, Y., Australia’s Uranium Resources, Geology and Development of Deposits, AGSO
Mineral Resource Rep. 1, Geoscience Australia, Canberra (2000).
[3] REEVE, J.S., CROSS, K.C., SMITH, R.N., ORESKES, N., “Olympic Dam copper–uranium–gold deposit”, Geology
of the Mineral Deposits of Australia and Papua New Guinea (HUGHES, F.E., Ed.), Australasian Institute of Mining
and Metallurgy, Melbourne (1990) 1009–1035.
[4] CROSS, K.C., DALY, S.J., FLINT, R.B., “Mineralization associated with the GRV and Hiltaba Suite granitoids —
Olympic Dam deposit”, The Geology of South Australia, Vol. 1: The Precambrian, Geological Survey of South
Australia, Adelaide (1993) 132–138.
[5] OECD NUCLEAR ENERGY AGENCY, INTERNATIONAL ATOMIC ENERGY AGENCY, Uranium 2018:
Resources, Production and Demand, OECD, Paris (2018).
[6] LALLY, J.H., BAJWAH, Z.U., Uranium deposits of the Northern Territory. Northern Territory Geological Survey,
Report 20 (2006).
[7] SWEET, I.P., BRAKEL, A.T., CARSON, L., The Kombolgie Subgroup — a new look at an old ‘formation’, AGSO
Research Newsletter 30, Geoscience Australia, Canberra (1999) 26–28.
[8] LUDWIG, K.R., et al., Age of uranium mineralisation in the Jabiluka and Ranger deposits, Northern Territory,
Australia, Econ. Geol. 82 (1987) 857–874.
[9] FOY, M.F., “South Alligator Valley uranium deposits”, Economic Geology of Australia and Papua New Guinea,
Vol. 1: Metals, Monograph 5, Australasian Institute of Mining and Metallurgy, Melbourne (1975) 301–303.
[10] POLITO, P.A., KYSER, T.K., THOMAS, D., MARLATT, J., DREVER, D., Re-evaluation of the petrogenesis of
the Proterozoic Jabiluka unconformity-related uranium deposit, Northern Territory, Australia. Miner. Deposita 40
(2005) 257-288.
[11] McKAY, A.D., Kintyre Deposit: Estimation of Identified Uranium Resources, Bur. Min. Res. Geol. Geophys,
Canberra (1992) (unpublished).
[12] CALLEN, R.A., ALLEY, N.F., GREENWOOD, D.R., “Lake Eyre Basin”, The Geology of South Australia, Vol. 2:
The Phanerozoic, Geological Survey of South Australia, Adelaide (1995).
[13] BRUNT, D.A., Uranium in Tertiary stream channels, Lake Frome area, South Australia, Proc. Austral. Inst. Min.
Metall. 266 (1978) 79–90.
[14] LINDSAY, J.F., KORSCH, R.J., “The evolution of the Amadeus Basin, Central Australia”, Geological and
Geophysical Studies in the Amadeus Basin, Central Australia, Bur. Min. Res. Geol. Geophys Bull. 236,
Canberra (1991) 7–32.
[15] BORSCHOFF, J., FARIS, I., “Angela and Pamela uranium deposits”, Geology of the Mineral Deposits of Australia
and Papua New Guinea (HUGHES, F.E., Ed.), Australasian Institute of Mining and Metallurgy, Melbourne
(1990) 1139–1142.
[16] WELLS, A.T., MOSS, F.J., The Ngalia Basin, Northern Territory: Stratigraphy and Structure, Bur. Min. Res. Geol.
Geophys Bull. 212, Canberra (1983).
[17] GASKIN, A.J., BUTT, C.R.M., DEUTSCHER, R.L., HORWITZ, R.C., MANN, A.W., “Hydrology of uranium
deposits in calcretes of Western Australia”, Energy Resources of the Pacific Region (Proc. Int. Conf. Honolulu,
1978), American Association of Petroleum Geologists, Tulsa (1981).
[18] BUTT, C.R.M., HORWITZ, R.C., MANN, A.W., Uranium Occurrences in Calcrete and Associated Sediments in
Western Australia, Rep. FP16, CSIRO Research Laboratories Division of Mineralogy, Commonwealth Scientific
and Industrial Research Organization, Canberra (1977).
[19] WESTERN MINING CORPORATION, Yeelirrie Uranium Project, Western Australia, Draft Environmental Impact
Statement and Environmental Review and Management Programme, WMC (1978).
[20] BRIAN LANCASTER & ASSOCIATES, Draft Environmental Impact Statement, Lake Way Joint Venture
Environmental Review and Management Programme, Melbourne (1981).
[21] CRABB, D., DUDLEY, R., MANN, A.W., “Hinkler Well and Centipede deposits”, Surficial Uranium Deposits,
IAEA-TECDOC-322, IAEA, Vienna (1985) 133–136.
[22] BRUNT, D.A., “Miscellaneous uranium deposits in Western Australia”, Geology of the Mineral Deposits of
Australia and Papua New Guinea (HUGHES, F.E., Ed.), Australasian Institute of Mining and Metallurgy, Melbourne
(1990) 1615–1620.
[23] INTERNATIONAL ATOMIC ENERGY AGENCY, World Distribution of Uranium Deposits (UDEPO), with
Uranium Deposit Classification, 2009 Edition, IAEA-TECDOC-1629, IAEA, Vienna (2009).
[24] OECD NUCLEAR ENERGY AGENCY, Forty Years of Uranium, Production and Demand in Perspective: The Red
Book Retrospective, OECD, Paris (2006) 280 pp.
[25] OECD NUCLEAR ENERGY AGENCY, INTERNATIONAL ATOMIC ENERGY AGENCY, Uranium 2009:
Resources, Production and Demand, OECD, Paris (2010).
[26] OECD NUCLEAR ENERGY AGENCY, INTERNATIONAL ATOMIC ENERGY AGENCY, Uranium 2005:
Resources, Production and Demand, OECD, Paris (2006).
[27] OECD NUCLEAR ENERGY AGENCY, INTERNATIONAL ATOMIC ENERGY AGENCY, Uranium 2007:
Resources, Production and Demand, OECD, Paris (2008).
[28] BHP, 2018 Annual report. https://www.bhp.com/-/media/documents/investors/annual-
reports/2018/bhpannualreport2018.pdf (2019)
[29] ERA, Annual Statement Of Reserves And Resources Reserves, https://www.energyres.com.au/operations/reserves-
resources/. January 2018

Updated from INTERNATIONAL ATOMIC ENERGY AGENCY, World Uranium Geology, Exploration, Resources and Production, IAEA, Vienna (2020) by J.R. Blaise & M. Fairclough (Dec.2020)

You might also like