Download Word - 154 KB - Department of the Environment

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Occupancy–abundance relationship wikipedia , lookup

Theoretical ecology wikipedia , lookup

Habitat conservation wikipedia , lookup

Biodiversity action plan wikipedia , lookup

Ecological fitting wikipedia , lookup

Molecular ecology wikipedia , lookup

Bifrenaria wikipedia , lookup

Introduced species wikipedia , lookup

Habitat wikipedia , lookup

Island restoration wikipedia , lookup

Transcript
Draft risk assessment report to amend the Live Import List to include Cane Toads, for the purpose of scientific research only
1.
Provide information on the
taxonomy of the species.
2.
Provide information on the
status of the species under
the Convention on
International Trade in
Endangered Species of Fauna
and Flora (CITES). For
example, is the species listed
on CITES Appendix I, II or III,
and if so, are there any
specific restrictions on the
movement of this species?
Include information on the
conservation value of the
species.
3.
Provide information about the
ecology of the species.
Include, but do not restrict
your response to:
- lifespan of the species
- size and weight range
- the natural geographic
range
a) Family Bufonidae (Class Amphibia, Order Anura)
b) Genus Bufo, also attributed to (sub-)genus Rhinella and genus Chaunus
c) Species marinus (i.e. Bufo marinus, also referred to as Rhinella marina, Chaunus marinus)
d) No subspecies described
e) Linnaeus 1758 (ITIS Global 2012)
f) Cane Toad, also known as Marine Toad, Giant Toad
g) This species is not a GMO (genetically-modified organism)
This species is not listed in CITES Appendices I, II or III (CITES 2012).
Cane toads are common and widespread throughout their native range, as well as within Australia (Lever 2001).
a) Both male and female toads have an average lifespan of 5-15 years in the wild (Tyler, 1989) with some animals surviving
>30 years in captivity (Grenard 2007).
b) Average size for an adult toad is 100-150 mm (snout-urostyle length), although adult females may exceed 200 mm (Lever
2001). Average weight is 500-800 g, but can exceed 1.5 kg in adult females (Lever 2001).
c) Cane toads are among the largest of the toad species, and adult females are larger than males. Both have a short, squat
body with short legs. Coloration is grey, brown, olive-brown or reddish-brown above, adult females often with mottled
brownish pattern, whereas males have uniform colour on the dorsal surface. Juveniles have a series of narrow dark bars,
-
habitat
diet, including potential to
feed on agricultural plants
social behaviour and
groupings
territorial and aggressive
behaviours
natural predators
characteristics that may
cause harm to humans and
other species
dashes or lines on the head and back. Adults are whitish or yellow below, usually speckled or mottled with dark brown.
The skin is very dry and warty above, rugose in adult males, granular on ventral side for both sexes. A prominent pair of
parotoid (shoulder) glands is located behind each eye. The supraocular region is warty and is separated from the smooth
interorbital region by a high bony ridge, the ridge on each side continuing forward to meet on the snout between the
nostrils. The toes on front limbs free and not webbed, toes on hind limbs have a tough leather-like webbing. Tadpoles
are less than 35 mm long, are jet black above and silvery white with black spots below (Cameron 2012; Cogger 2000;
Lever 2001). Male and female toads can be difficult to distinguish.
Similar native amphibian species are the Eastern Pobblebonk Frog (Limnodynastes dumerilii) and Giant Burrowing Frog
(Heleioporus australiacus) (Cameron 2012).
Adult male cane toad
(photos: http://www.canetoadsinoz.com)
Adult female cane toad
Metamorph toadlet
Cane toad tadpoles and metamorphlings (photo: CSIRO)
(photo:http://www.canetoadsinoz.com)
d) The cane toad is common and widespread throughout central America and northern South America, extending from
approx. 27°N (southern Texas, USA) southwards to approx. 30°S in Brazil and Argentina (Lever, 2001). Cane toads are
principally a lowland nocturnal species, rarely occurring at high altitudes (>1000 m asl) with the maximum elevation
for viable survival probably determined by minimal thermal tolerance limits (Lever 2001). Minimal thermal tolerances
likely determine the northerly and southerly limits of the toads’ distribution. Cane toads distinctly prefer disturbed
habitats modified by human activity, usually near water, but can persist (although at much lower densities) in
forested areas (Zug & Zug 1979). As toads generally prefer moist environments, increasing aridity may also limit the
toads’ distribution (Zug & Zug 1979). Adult toads have few natural predators, although juveniles and tadpoles are
highly susceptible to predation by a wide range of organisms (birds, fish, amphibians, reptiles). Toads have a
voracious appetite, and will attempt to eat almost anything that moves (Zug & Zug 1979) yet can survive for
extended periods without food. Thus, food availability is not a limiting factor in the natural range of the toad.
e)
Cane toads are capable of moving long distances (relative to body size); in Australia, rapid evolution of increased leg
length in adult toads enables dispersing animals to move >1 km in a single night, with some animals capable of
moving 10-50 km per year (Phillips et al 2006). In dry seasons, toads move to wetter shelter sites to prevent
dehyration (e.g. Cohen & Alford 1996).
f)
Cane toads can tolerate a wide thermal range (approx. 5-40°C). In cold temperatures (5-10°C), adult toads can remain
dormant in a sheltered and moist habitat for a considerable period (i.e. overwinter) (Easteal & Floyd 1986). In very
hot temperatures (>35°C), animals behaviourally thermoregulate (Krakauer 1970) - toads find shelter (under logs, in
burrows) and huddle together to minimise evaporative water loss so avoiding desiccation (Tingley & Shine 2011).
There is no evidence of cane toads aestivating in summer to avoid high temperatures. Larvae (tadpoles) can tolerate
high water temperatures (42°C), although many succumb to dehydration/desiccation when transient water bodies
evaporate (Zug and Zug 1979, Easteal & Floyd 1986).
g) Juvenile (i.e. post-metamorphosis from tadpole stage) and adult cane toads breathe atmospheric air. Larval stages
(tadpoles) are fully aquatic (Lever 2001).
h) Habitat requirements: B. marinus is essentially an inland species and breeds mostly in fresh water (Neill 1958).
However, it is also found in brackish water and is able to breed in water with salinity 15% of sea water (Neill 1958).
Adult toads can survive in salinities of up to 40% sea water (Liggins & Grigg 1985), and have been recorded swimming
in sea water (Myers 1948).
Breeding toads appear to prefer water with pH of 7 - 7.5, of low turbidity and high dissolved O2 (Evans et al 1996,
Semeniuk et al 2007). The mean preferred temperature of cane toads over all stages of larval development is 28.5°C
(range 26.7 - 30.4°C) (Floyd 1984). Adult toads are capable of tolerating a thermal range of 5-40°C.
Climate: B. marinus is a tropical and sub-tropical lowland species. It requires a moist environment to survive.
Nest sites: Adult toads are nocturnal. During the day, toads shelter beneath logs, river banks, in burrows (excavated
by other animals), where moisture in soil is retained providing a damp environment in which toads can avoid
desiccation and thermal extremes. Toads prefer disturbed habitats which have been modified by human activity,
such as grasslands, drainage ditches, cultivated areas, gardens etc., usually near water (Lever 2001).
Adult toads require access to suitable waterbodies to breed, such as marshes, swamps, lakes, ponds, dams, rivers
(lentic pools), channels, streams, mangrove flats.
i)
Social behaviour or groupings: outside of the breeding season, adult toads exist in loose aggregations and are often
found huddling together in shelters to conserve moisture. During the breeding season, male toads sometimes gather
in substantial aggregations numbering several hundred individuals (Easteal and Floyd 1986), calling to attract
females. As toads produce up to 30 000 eggs from a single clutch, densities of tadpoles in waterbodies can be very
high (Lever 2001). Prior to metamorphosis, tadpoles begin to aggregate in the margins of the water bodies in which
they hatched (Easteal and Floyd 1986). Metamorphosed toadlets tend to form closely packed aggregations around
the edge of their water body, prior to dispersal to other water bodies (Easteal and Floyd 1986). Intraspecific
competition between adult toads can be intense in the breeding season. Toads are cannibalistic (bigger toads may
eat smaller toads), but are not overtly aggressive to other species.
j)
Toads do not exhibit aggressive territorial behaviour. Some adult toads remain near the same water body for lengthy
periods of time, while others disperse long distances to new habitats. Toads are not aggressive to humans or other
animals (outside of usual predator-prey interactions). However, if placed under significant stress, adult toads can
spray toxins from their parotid glands up to 1 m at their aggressor.
k) Cane toads are highly toxic to many vertebrate (and some invertebrate) species. The parotoid (shoulder) glands of
cane toads contain potent bufogenins, bufotoxins, and other substances (Zug and Zug 1979). All life stages of the
cane toad (eggs, tadpoles, metamorphs and adults) are toxic, although toxin types and content change markedly
during toad ontogeny (Hayes et al 2009). Toxin levels and thus, danger to vertebrate predators, are high in eggs,
decrease through tadpole life, are lowest at around the time of metamorphosis, and increase rapidly thereafter
(Hayes et al 2009).
4.
Provide information on the
reproductive biology of the
species, including
- the age at maturity (first
breeding)
- how frequently breeding
occurs
- if the female can store
sperm
- how many eggs or liveborn young are produced
at each breeding event
a) B. marinus reaches sexual maturity at 6-18 months, with animals in warmer climates reaching sexual maturity earlier
(faster growth rate) (Lever 2001).
b) In tropical regions, toads breed throughout the year, but not continuously. In sub-tropical regions, toads breed in the
warmer months only (Krakauer 1968). In both cases, breeding is stimulated by the onset of the wet season (Lever
2001). However, breeding can continue into the dry season (Krakauer 1968). Males aggregate at a suitable
waterbody (e.g. still waterbody, with thick vegetation absent at waters’ edge) and call to attract females. Females
usually mate only once per breeding season, and spawn within 12 hours of mating (Easteal & Floyd 1986). Females
can produce up to 30,000 black eggs, about 1 mm in diameter, in long jelly strings about 3 mm thick, which are
usually attached to submergent or emergent aquatic vegetation. The eggs hatch in 48-72 hours into tadpoles. They
develop into toadlets in 4-8 weeks (Easteal & Floyd 1986).
-
if the species has
hybridised with other
species (both in the wild
and in captivity) or has the
potential to hybridise with
any other species
- if the species can
hybridise, are the progeny
fertile
c) Breeding can occur twice a year in sub-tropical regions, and throughout the year in tropical regions.
d) Not relevant (aquatic invertebrates)
e) Male toads can ‘change sex’ if testes are surgically removed (Bidder’s organ will then start to produce eggs).
Agricultural chemicals may cause male toads to adopt female characteristics (McCoy et al 2008).
f)
There is no record of B. marinus being able to hybridise with congeneric bufonid species
g) There are no native Australian bufonids (toads), thus it is impossible for Bufo marinus to hybridise with any native
species in Australia.
h) Individual toads are single-sexed, i.e. not hermaphroditic.
5.
Provide information on
whether this species has
established feral populations,
and if so, where those
populations are. Include
information on whether this
species has been introduced
to other countries, even if it
has not established feral
populations.
a) In a misguided attempt to control agricultural invertebrate pests, particularly sugar cane pests, the cane toad has
been introduced to many neo-tropical countries outside of the toad’s native range. Feral breeding populations of
cane toads have become established in southern US (Florida), and many islands throughout the Caribbean (Antigua,
Barbados, Carriacou, Grand Cayman, Grenada, Guadeloupe, Hispaniola, Jamaica, Martinique, Montserrat, Nevis,
Puerto Rico, St Kitts, St Lucia, St Vincent, US Virgin Islands), the Pacific (Samoa, Federated States of Micronesia,
Hawaiian islands, Fiji, Marianas Islands, Palau, Solomon Islands, Tuvalu),the Atlantic Ocean (Bermuda), the Indian
Ocean (Diego Garcia in the Chagos Archipelago), Asia (Japan, the Philippines), Oceania (Papua New Guinea) and
mainland Australia (Lever 2001).
b) The cane toad is considered an ecological pest in Australia. It has had an extreme adverse impact on some native
fauna (mostly predator species such as snakes, goannas, crocodiles, quolls) in northern Australia and the control of
this pest is the focus of extensive management and research effort (Shine 2010). Since its introduction to eastern
Queensland in 1935, the cane toad has adapted rapidly to Australian conditions and its range (still extending) now
covers a very large area of northern and western Australia, with extralimital (stowaway-founded) populations as far
south as Sydney.
c) In some of the countries to which the cane toad has been introduced, it has not established feral populations. This
has occurred in Taiwan, Thailand, Egypt, Mauritius, Cuba, Cook Islands, Marshall Islands.
6.
Provide information on, and
the results of any other
environmental risk
assessments undertaken on
the species both in Australia
and overseas, including any
Import Risk Analyses
undertaken by Biosecurity
Australia.
Risk assessments have been undertaken by both state and federal agencies in Australia. They include:
Markula A., Csurhes, S. and Hannan-Jones, M. (2010). Pest animal risk assessment: Cane Toad. Biosecurity Queensland
(Department of Employment, Economic Development and Innovation)
http://www.daff.qld.gov.au/documents/Biosecurity_EnvironmentalPests/IPA-Cane-Toad-Risk-Assessment.pdf

Identified cane toads in Queensland as ‘extreme’ threat species (using assessment model developed by Bomford
2008).
Van Dam, R.A., Walden, D.J. and Begg, G.W. (2002). A preliminary risk assessment of cane toads in Kakadu National Park.
Scientist Report No. 164, Supervising Scientist, Environment Australia, Darwin NT.
http://www.environment.gov.au/ssd/publications/ssr/pubs/ssr164.pdf

Identified cane toads in Kakadu NP as being a serious pest species, negatively affecting fauna (particularly predators),
tourism and cultural activities.
Page, A., Kirkpatrick, W. and Massam, M. (2008). Risk assessments for exotic reptiles and amphibians introduced to
Australia: Cane Toad (Bufo marinus). Department of Agriculture and Food Western Australia. http://www.feral.org.au/wpcontent/uploads/2010/11/Cane_toad_DAFWA_220410.pdf

Identified cane toads in WA as ‘extreme’ threat species (using assessment model developed by Bomford 2008).
Threat Abatement Plan for the biological effects, including lethal toxic ingestion, caused by cane toads. Australian Govt .
Dept. Sustainability, Environment, Water, Population and Communities (2011).
http://www.environment.gov.au/biodiversity/threatened/publications/tap/pubs/tap-cane-toads.pdf

Addresses key threatening process identified by Threatened Species Scientific Committee.
7.
8.
Assess the likelihood that the
species could establish a
breeding population in the
Australian environment
should it ever be released
from effective human control.
Include at least the following
factors:
- ability to find food sources
- ability to survive and adapt
to different climatic
conditions (e.g.
temperatures, rainfall
patterns)
- ability to find shelter
- rate of reproducing
- any characteristics that the
species has which could
increase its chance of
survival in the Australian
environment
The cane toad has already become established in Australia. Introduced to eastern Queensland in 1935, feral populations now
occur throughout north-eastern NSW, eastern and northern Queensland, northern NT and north-eastern WA.
Provide a comprehensive
assessment of the potential
impact of the species should it
establish feral population/s in
Australia. Include, but do not
restrict your assessment to
the impact of this species on:
- similar niche species (i.e.
competition with other
Feral populations of cane toads have been present in Australia since 1935. The ecological impact of the cane toad in Australia
has been well-documented, and is the subject of on-going research. For a comprehensive review, see Shine (2010).
Cane toads have adapted rapidly to Australian climatic conditions, and are able to withstand much drier conditions than in
their native range. Toads at the invasion front have evolved longer legs, enabling toads to colonise previously toad-free areas
at a faster rate (invasion front now moves approx. 50 km per year (Phillips et al 2007)). Toads in Australia occur at much
higher densities than in any other country, due to the absence of co-evolved predators (i.e. predators that have evolved
immunity to toad toxins: Shine 2010) and the presence of favourable environmental conditions (to toads).
Northern Australia provides ideal conditions for the successful breeding of cane toads and their continued dispersal and
colonization into new territory. The area lies within the toads’ ideal climatic range and provides an abundance of suitably
shallow waterbodies in which toads can spawn. Invertebrate food is plentiful and thus non-limiting to toad populations,
shelter sites are numerous, and predators occur at comparatively low densities and are highly susceptible to the toads’
chemical defences (toxins) (Shine 2010).
a) Toads have similar ecological requirements to those of native amphibian species. Toads can occur in very high
densities and so will have a correspondingly high demand for resources such as food and shelter, to the detriment of
native species (Schwarzkopf & Alford 1996). However, most negative effects on native amphibians (at both larval and
adult stages) are likely as a result of toad toxicity and behavioural interference than direct competition for resources
(Williamson 1999, Greenlees et al 2007, Pizzatto and Shine 2009; but see Crossland et al 2009).
-
-
-
species for food, shelter
etc.)
is the species susceptible
to, or could transmit any
pests or disease
probable prey/food
sources, including
agricultural crops
habitat and local
environmental conditions
any control/eradication
programs that could be
applied in Australia if the
species was released or
escaped
any characteristic or
behaviour of the species
which may cause land
degradation i.e. soil
erosion from hooves,
digging
any potential threat to
humans
b) Adult cane toads are generalist, opportunistic predators of invertebrates and small vertebrates. They are
predominantly arthropodivorous, preying on beetles, termites and ants (Cameron & Cogger 1992), and may occur in
large enough numbers that local populations of prey species become depleted (Greenlees et al 2006), although many
do not (Catling et al 1999). Toads also prey on molluscs, particularly native land snails. Toads feed opportunistically
on vertebrates, but analysis of stomach contents have shown that vertebrates constitute a very small part of the toad
diet, thus the impact of toad predation on vertebrate populations is negligible (Shine 2010).
c) Cane toads do not appear to have an adverse physical affect on their habitat or local environment. They do not
burrow or cause damage by digging, they do not damage or deplete vegetation, nor do they cause damage to
waterways by causing erosion on banks. Toads prefer disturbed habitats where vegetation has been removed or
modified, but are not capable of creating this matrix themselves. There is no evidence that toads are able to spread
weeds.
d) Cane toads do not appear to exhibit any behaviours that result in environmental degradation (e.g. digging etc.).
There is no evidence of cane toads polluting waterways, although the toxins in toad eggs kill native amphibian
tadpoles when eaten (Crossland et al 2008)
e) Cane toads have very little impact on primary industries. Contrary to popular belief, toads to do not cause
widespread mortality in chickens (Beckmann & Shine 2010) or in other vertebrate stock. However, toads do feed on
honey bees (Apis mellifera) and can cause substantial localised losses (Goodacre 1947). It is possible that high
densities of cane toads can have an indirect adverse effect on crops by causing the local depletion of invertebrates
that are beneficial to pollination and seed dispersal of crop plants. However, there is little evidence of this in
Australia. Cane toads are not effective in controlling insect pests (Shine 2010).
f)
Cane toads do not cause widespread damage to property. However, property owners have incurred costs when
attempting to deter toads. Beekeepers have had to modify hives to prevent access by cane toads, and some land
managers have attempted to restrict toad access to reserves and waterbodies by erecting fencing and planting
vegetation around ponds.
g) Cane toads are a social nuisance, and are generally disliked by the public. Recognised by communities throughout its
range in Australia as a pest, the cane toad has been targeted by community groups in attempted eradication
programs. Local concentrations of toads can cause problems in urban and suburban areas, where artificial lighting
attracts flying insects; toads aggregate often in large numbers to feed. Toads are toxic and pose a threat to people
and companion animals (see (h) below) and are thus unwelcome in backyards, sporting ovals, golf courses etc.
Additionally, aggregations of cane toads can block drains and foul swimming pools.
h) Cane toads are harmful to humans and many other vertebrate species. They secrete a cocktail of toxins from the
parotid glands (located on shoulder) which include bufotoxins and bufogenins (steroid derivatives that are cardioactive and/or cardio-toxic) that are life-threatening in humans and often fatal in other vertebrate species. Toxins can
be squirted up to 1 m from the parotid glands if the toad is subjected to acute physical or mental stress. Toxins are
also secreted from skin glands. Most severe cases of human poisoning from toad toxins have been when the toad (or
eggs) has been ingested. Many native predator species have died as a result of ingesting toad toxins (e.g. many frogeating snake species, goannas, crocodiles, northern quolls (Shine 2010)). Toads are toxic at all life stages; tadpoles
from native frog species are poisoned when they predate on toad tadpoles.
As cane toads are remarkably placid animals, cases of toad toxicity in humans are rare. A sensible precaution is to
avoid subjecting toads to perceived threats that may elicit a toxin-ejecting response, such as rough handling,
exposure to predation threat (e.g. from dogs), blunt trauma etc. Ingestion of any part of the cane toad should be
avoided. Medical treatment for toad poisoning is similar to that of digitalis poisoning (Brubacher et al 1996).
Cane toads are a known vector of Salmonella bacteria (excreted in toad faeces (Speare 1990)) and may be a reservoir
for Leptospira bacteria (which causes Weil’s disease (Babudieri et al 1973)), both of which may cause infections in
humans.
Many (but not all) Australian cane toads are infected with a parasitic lungworm, a nematode that arrived in Australia
with infected cane toads in 1935. Some species of native Australian frogs have been infected with this parasite, with
some species being adversely affected while others remain healthy (Pizzatto & Shine 2011). The cane toad has been
shown to be susceptible to infection with chytrid fungus under laboratory conditions (Berger et al 1998), but has not
been implicated in the spread of chytridiomycosis in native frogs.
9.
What conditions or
restrictions, if any, could be
applied to the import of the
species to reduce any
potential for negative
environmental impacts (e.g.
single sex imports, desexing
animal prior to import etc.)
Hundreds of millions of cane toads already occur in Australia; the only effect of listing cane toads for importation would be to
allow genetic strains of toads to be brought in from Hawaii (the site from which Australian toads were sourced in the original
importation in 1935) and from their native range (from which the toads were brought to Hawaii). By comparing these animals
to those in Australia already, we can understand how the process of transportation to a new country, and adaptation within
Australia, has modified the attributes of this invasive species. The importation of additional cane toads into Australia should
only be allowed to obtain animals for scientific research purposes. As such, animals should not be restricted to a single sex or
desexed. Toads should be kept in secure, licensed premises and euthanized once the research program has concluded.
Currently, there are no effective programs to eradicate populations of free-living cane toads in Australia. Some control
measures can depress local populations in the short-term, but no effective long-term control has yet been developed.
10. Provide a summary of the
types of activities that the
specimen may be used for if
imported into Australia (e.g.
pet, commercial, scientific).
You must discuss:
- the benefit of this species
for these activities
- potential trade in the
species
- why this species has been
chosen
Cane toads will be imported into Australia for scientific research purposes only. The aim of this research is to compare
characteristics (physical, physiological) of toads that have adapted to Australian conditions with those from their original
‘source’ population (i.e. from toads that were introduced into Hawaii), and from their native range in Venezuela (from the
original source population of Hawaiian toads). Thus, toads will be imported from Hawaii and Venezuela only.
Number of animals to be imported; 10 male and 10 female toads from each of the two source areas (i.e., a total of 40 toads:
20 from Hawaii and 20 from Venezuela.
Males of this species do not need to be kept segregated from each other. They will co-habit without displaying aggressive
behaviour.
Animals will be used for breeding in secure facilities in Australia. Breeding is easily controlled, because toads will not breed
unless they have access to a large waterbody in which to lay their aquatic eggs.
After the imported toads have bred, they will be euthanased; the scientific work will be conducted on their progeny, not on
the imported animals. This will minimise any chances of transmission of parasites or diseases, because the eggs, tadpoles and
metamorphs will have no contact with the adults. At the conclusion of the research program, all progeny of the imported
toads and will be euthanized and the carcasses destroyed.
Animals will be sourced from free-living wild populations in Hawaii and Venezuela.
11. Provide detailed guidelines on
the way in which the species
should be kept, transported
and disposed of in accordance
with the types of activity that
the species may be used for if
imported into Australia. You
must include:
- the containment (e.g.
cage, enclosure) and
management standards for
this species to prevent
escape or release. This
should also talk about the
security standards for this
specimen
- the disposal options for
surplus specimens
Cane toads will be collected from free-living populations in Venezuela (native populations that were the ultimate source of
the Australian cane toad population) and Hawaii (feral populations that were the immediate source of the Australian toad
population). Animals will be transported by air to Darwin, NT. Toads will be transported in IATA approved containers,
consisting of secure wooden or plastic crates with adequate ventilation, and packed with straw or wood wool to retain
moisture so preventing dehydration. Toads are gregarious animals so animals will not be segregated during transportation.
Animals will be in transit for less than 48 hours (door to door).
Once in Australia, the toads will be transported by road to a research station with suitable housing facilities. At all times,
toads will be kept in secure conditions to prevent escape. Toads will be housed in secure rooms so only approved personnel
will have access. Toads will be kept communally (up to 10 toads per 80 L plastic tub, a density comparable to that found in
free-living populations). Breeding will be induced by injection of synthetic hormones (lucrin), a standard practice in
laboratory studies of amphibians. After eggs have been laid and fertilised, the adult toads will be humanely euthanased. The
offspring will be raised in secure facilities.
Cane toads are very easy to keep in captivity, and colonies have been maintained in several research centres for >5 years.
Within each housing tub, toads will be provided with a pool and dry surfaces for perching to allow behavioural
thermoregulation and hydration. Toads will be kept within their optimal thermal range, and maintained via an appropriate
feeding regimen to sustain good health and body condition. Animals will be kept under conditions compliant with housing
and husbandry requirements as approved by the University of Sydney Animal Ethics Committee.
At the conclusion of the experiments, all toads will be euthanized by injection of a lethal dose of sodium pentabarbitone, the
method of euthanasia for amphibians sanctioned by the University of Sydney Animal Ethics Committee. Toad tissue will
either be retained and preserved (frozen, or fixed in formaldehyde or ethanol) or disposed of by incineration.
12. Provide information on all
other Commonwealth, state
and territory legislative
controls on the species,
including:
- the species’ current
quarantine status, or
- pest or noxious status, or
- whether it is prohibited or
controlled by permit or
licence in any state or
territory
Federal controls and assessments
The biological effect of the cane toad has been listed as a key threatening process under the Environment Protection and
Biodiversity Conservation Act 1999 (Threatened Species Scientific Committee 2005). A TAP (Threat Abatement Plan) has been
developed accordingly (SEWPaC 2011).
State controls
Queensland
Dept. Agriculture, Forestry and Fisheries places no controls on keeping cane toads in captivity, nor in importing from
overseas or interstate. The cane toad is not a declared species under the Land Protection (Pest and Stock Route
Management) Act 2002.
New South Wales
Dept. Industry and Innovation requires import permits and licences to keep non-indigenous animals (Non-Indigenous Animals
Act 1987), i.e. cane toads, in NSW research institutions. Annual returns accounting for all animals must be submitted.
Northern Territory
Dept. Primary Industries and Fisheries do not require permits for keeping cane toads in captivity. No record of import
restrictions on cane toads.
Victoria
The cane toad is listed as a restricted (controlled) species under the Catchment and Land Protection Act (1994). Permits from
Dept. Primary Industries are required to import or keep cane toads.
Western Australia
The cane toad is a declared (prohibited) species and can only be kept in captivity for scientific or educational purposes
(Agriculture and Related Resources Protection Act 1976). Permits from Dept. Primary Industries are required to import and
keep cane toads for research purposes.
South Australia
Importing and keeping cane toads is prohibited without a permit issued by Dept. Primary Industries and Regions (Natural
Resource Management Act 2004).
Tasmania
The importation of live cane toads is prohibited under section 32 of the Nature Conservation Act 2002 (Dept. Primary
Industries).
Biosecurity
An Import Risk Analysis may be required by DAFF (Biosecurity Australia). We are awaiting advice from Biosecurity Australia
on quarantine and import conditions.
13. Conditions/restrictions
Feral populations of cane toads have been established in Australia since 1935, all originating from animals imported in a
single shipment (although toads were bred in captivity and deliberately released over a wide area of eastern Queensland).
Today, the speed at which populations are colonising previously toad-free areas is increasing, and the total area inhabited by
cane toads is projected to increase substantially (Kearney et al 2008). Therefore, it is sensible that any imports of additional
cane toads from overseas are restricted as follows:




toads are imported for scientific research purposes only
toads are kept in a secure environment, with access restricted to authorized research personnel only
toads are euthanized at completion of experiments
toad tissue is either preserved in formaldehyde/ethanol, frozen or incinerated.
These restrictions will ensure that imported animals will not contribute to the feral cane toad population in Australia.
References
Babudieri, B., Carlos, E.R., Carlos, E.T. (1973). Pathogenic leptospira isolated from toad kidneys. Tropical Geographical Medicine 25: 297-299.
Beckmann, C. and Shine, R. (2010). The power of myth: the (non) impact of invasive cane toads (Bufo marinus) on domestic chickens (Gallus gallus). Animal
Production Science 50: 847-851.
Berger, L., Speare, R., Daszak, P., Green, D.E., Cunningham, A.A., Goggin, C. L., Slocombe, R., Ragan, M.A., Hyatt, A.D., McDonald, K.R., Hines, H.B., Lips, K.R.,
Marantelli, G. and Parkes, H. (1998). Chytridiomycosis causes amphibian mortality associated with population declines in the rainforests of
Australia and Central America. Proceedings of National Academy Science 95: 9031-9036.
Bomford, M. (2008). Risk assessment models for the establishment of exotic vertebrates in Australia and New Zealand: validating and refining risk
assessment models. Invasive Animals Cooperative Research Centre, Canberra.
Brubacher, J.R., Ravikumar, P.R., Bania, T., et al. (1996). Treatment of toad venom poisoning with digoxin-specific Fab fragments. Chest 110: 1282–1288.
Cameron, E. (2012). Cane toad (Australian Museum Online). http://australianmuseum.net.au/Cane-Toad (accessed 31/1/13)
Cameron E. E. and Cogger H. G. (1992). The herpetofauna of the Weipa region, Cape York Peninsula. Technical Report No. 7. Sydney: Australian Museum.
Catling P. C., Hertog A., Burt R. J., Wombey J. C., Forrester R. I. (1999). The short-term effect of cane toads (Bufo marinus) on native fauna in the Gulf
Country of the Northern Territory. Wildlife Research 26: 161–185.
CITES (2012). Appendices I, II & III. (Convention on International Trade in Endangered Species). http://www.cites.org/ (accessed 31/1/13)
Cogger, H. G. (2000). 'Reptiles & Amphibians of Australia.' Reed New Holland: Sydney.
Cohen, M.P. and Alford, R.A. (1996). Factors affecting diurnal shelter use by the cane toad, Bufo marinus. Herpetologica 52: 172-181.
Crossland, M.R., Brown, G.P., Anstis, M., Shilton, C.M. and Shine, R. (2008). Mass mortality of native anuran tadpoles in tropical Australia due to the invasive
cane toad (Bufo marinus). Biological Conservation 141: 2387-2394.
Crossland, M.R., Alford, R. A. and Shine, R. (2009). Impact of the invasive cane toad (Bufo marinus) on an Australian frog (Opisthodon ornatus) depends on
minor variation in reproductive timing. Oecologia 158: 625-632.
Easteal., S. And Floyd, R. B. (1986). The cane toad – an amphibian weed. In: Kitching, R.L. (ed) ‘The Ecology of Exotic Animals and Plants: some Australian
case histories’. Milton, Queensland: John Wiley. 27-42.
Evans, M., Yaber, C. and Hero, J. M. (1996). Factors influencing choice of breeding site by Bufo marinus in its natural range. Copeia 1996: 904–912.
Floyd, R.B. 1984. Variation in temperature preference with stage of development of Bufo marinus larvae. Journal of Herpetology 18: 153-158.
Goodacre, W. A. (1947). The giant toad (Bufo marinus) an enemy of bees. Agricultural Gazette of New South Wales 58: 374 – 375.
Greenlees M. J., Brown G. P., Webb J. K., Phillips B. L., Shine R. (2006). Effects of an invasive anuran (the cane toad, Bufo marinus) on the invertebrate fauna
of a tropical Australian floodplain. Animal Conservation 9: 431–438.
Greenlees, M. J., Brown, G. P., Webb, J. K., Phillips, B. L. and Shine, R. (2007). Do invasive cane toads (Chaunus marinus) compete with Australian frogs
(Cyclorana australis)? Austral Ecology 32: 900–907.
Grenard, S. (2007). ‘Frogs and Toads’. John Wiley & Sons.
Hayes R. A., Crossland M. R., Hagman M., Capon R. J., Shine R. (2009). Ontogenetic variation in the chemical defences of cane toads (Bufo marinus): toxin
profiles and effects on predators. Journal of Chemical Ecology 35:391–399.
ITIS Global: The Integrated Taxonomic Information System (version Sept 12). In: Species 2000 & ITIS Catalogue of Life, 28th January 2013 (Bisby F., Roskov
Y., Culham A., Orrell T., Nicolson D., Paglinawan L., Bailly N., Kirk P., Bourgoin T., Baillargeon G., Hernandez F., De Wever A., eds). Digital resource
at www.catalogueoflife.org/col/. Species 2000: Reading, UK.
Kearney, M., Phillips, B. L., Tracy, C. R., Christian, K. A., Betts, G. and Porter, W. P. (2008). Modelling species distributions without using species distributions:
the cane toad in Australia under current and future climates. Ecography 31: 423–434.
Krakauer, T. (1968). The ecology of the neotropical toad, Bufo marinus, in south Florida. Herpetologica 24: 214-221.
Krakauer, T. (1970). Tolerance limits of the toad, Bufo marinus, in south Florida. Comparative Biochemistry and Physiology 33:15-26.
Lever, C. (2001). The cane toad: the history and ecology of a successful colonist. Westbury Academic, UK.
Liggins, G.W. and Grigg, G.C. (1985). Osmoregulation of the cane toad Bufo marinus in salt water. Comparative Biochemistry and Physiology. 82A: 613-619.
McCoy,K.A., Bortnick, L.J., Campbell, C.M., Hamlin, H.J., Guillette, L.J., and St. Mary, C.M. (2008). Agriculture Alters Gonadal Form and Function in the
Toad Bufo marinus. Environmental Health Perspectives 116(11): 1526–1532.
Myers, G.S. (1948). The ability of amphibians to cross sea barriers, with special reference to Pacific zoogeography. Proceedings of the 7th Pacific Science
Congress 4: 19-27.
Neill, W.T. (1958). The occurrence of amphibians and reptiles in saltwater areas and a bibliography. Bulletin of Marine Science of the Gulf and Caribbean. 8:
1-97.
Phillips, B.L., Brown, G.P., Webb, J.K. and Shine, R. (2006). Invasion and the evolution of speed in toads. Nature 439: 803.
Phillips, B.L., Brown, G.P., Greenlees, M., Webb, J.K. and Shine, R. (2007). Rapid expansion of the cane toad (Bufo marinus) invasion front in tropical
Australia. Austral Ecology 32(2): 169-176.
Pizzatto L., and Shine R. 2009. Native Australian frogs avoid the scent of invasive cane toads. Austral Ecology 34: 77–82.
Pizzatto, L., and R. Shine. 2011. Ecological impacts of invading species: do parasites of the cane toad imperil Australian frogs? Austral Ecology 36: 954–963.
Schwarzkopf L., Alford R. A. 1996. Desiccation and shelter-site use in tropical amphibians: comparing toads with physical models. Functional Ecology 10:
193–200.
Semeniuk, M., Lemckert, F., and Shine, R. (2007) Breeding-site selection by cane toads (Bufo marinus) and native frogs in northern New South Wales,
Australia. Wildlife Research 3: 59–66.
SEWPaC. (2011). Threat Abatement Plan for the biological effects, including lethal toxic ingestion, caused by cane toads. Australian Govt . Dept.
Sustainability, Environment, Water, Population and Communities.
http://www.environment.gov.au/biodiversity/threatened/publications/tap/pubs/tap-cane-toads.pdf.
Shine, R. (2010). The ecological impact of invasive cane toads (Bufo marinus) in Australia. The Quarterly Review of Biology 85(3): 253-291.
Speare, R. (1990). A review of the diseases of the cane toad, Bufo marinus, with comments on biological control. Australian Wildlife Research 17: 387-410.
Tingley, R. and Shine, R. (2011). Desiccation risk drives the spatial ecology of an invasive anuran (Rhinella marina) in the Australian semi-desert. PLoS ONE
6(10): e25979.
Tyler, M.J. (1989). ‘Australian Frogs’. Penguin Books.
Williamson, I. (1999). Competition between the larvae of the introduced cane toad Bufo marinus (Anura: Bufonidae) and native anurans from the Darling
Downs area of southern Queensland. Australian Journal of Ecology 24: 636-643.
Zug, G.R. and Zug, P.B. (1979). The Marine Toad, Bufo marinus: a natural history resume of native populations. Smithsonian Contributions to Zoology 284: 158.