Journal articles on the topic 'Banksia – New South Wales'

To see the other types of publications on this topic, follow the link: Banksia – New South Wales.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Banksia – New South Wales.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Greenwood, David R., Peter W. Haines, and David C. Steart. "New species of Banksieaeformis and a Banksia 'cone' (Proteaceae) from the tertiary of central Australia." Australian Systematic Botany 14, no. 6 (2001): 871. http://dx.doi.org/10.1071/sb97028.

Full text
Abstract:
Silicified leaf impressions attributed to the tribe Banksieae (Proteaceae) are reported from a new Tertiary macroflora from near Glen Helen, Northern Territory and from the Miocene Stuart Creek macroflora, northern South Australia. The fossil leaf material is described and placed in Banksieaeformis Hill & Christophel. Banksieaeformis serratus sp. nov. is very similar in gross morphology to the extant Banksia baueri R.Br. and B. serrata L.f. and is therefore representative of a leaf type in Banksia that is widespread geographically and climatically within Australia and that is unknown in Dryandra or other genera of the Banksieae. The leaf material from Stuart Creek and Woomera represents the lobed leaf form typical of Paleogene macrofloras from southern Australia, but one species,B. langii sp. nov., is closely similar in gross form to Banksieaephyllum taylorii R.J.Carpenter, G.J.Jordan & R.S.Hill et al. from the Late Paleocene of New South Wales and similarly may be sclerophyllous. Also reported are impressions of Banksia infructescences, or ‘seed cones’, in Neogene sediments near Marree and Woomera, South Australia. These fossils demonstrate the presence of Banksiinae in central Australia in the mid-Tertiary, potentially indicating the former existence of linking corridors between now widely separated populations of Banksia.
APA, Harvard, Vancouver, ISO, and other styles
2

Hackett, Damian J., and Ross L. Goldingay. "Pollination of Banksia spp. by non-flying mammals in north-eastern New South Wales." Australian Journal of Botany 49, no. 5 (2001): 637. http://dx.doi.org/10.1071/bt00004.

Full text
Abstract:
Despite the accumulating evidence that non-flying mammals are effective pollinators, further research is required to clarify how widespread this phenomenon is. The role of non-flying mammals as pollinators of four species of Banksia was investigated in north-eastern New South Wales. Nine species of non-flying mammals were captured amongst flowering Banksia and all carried variable amounts of Banksia pollen on their fur or in their faeces. Although not captured, feathertail gliders (Acrobates pygmaeus) were observed foraging at Banksia inflorescences. Squirrel gliders (Petaurus norfolcensis) visiting B. integrifoliaand pale field-rats (Rattus tunneyi) visiting B. ericifolia, carried substantial loads of pollen. Fur pollen loads for these species were of a magnitude similar to those of nectarivorous birds that were sampled closer to the time of foraging. Assessment of newly opened flowers indicated that considerable amounts of pollen were removed at night. The results of a pollinator exclusion experiment were inconclusive but B. ericifolia inflorescences exposed to nocturnal pollinators had consistently high fruit-set. This study lends additional support to the notion that pollination of Banksia by non-flying mammals is widespread.
APA, Harvard, Vancouver, ISO, and other styles
3

Thiele, K., and PY Ladiges. "The Banksia integrifolia L.f. species complex (Proteaceae)." Australian Systematic Botany 7, no. 4 (1994): 393. http://dx.doi.org/10.1071/sb9940393.

Full text
Abstract:
The Banksia integrifolia (Proteaceae : Grevilleoideae) species complex currently comprises three varieties: var. aquilonia from northern Queensland; var. integrifolia from coastal Victoria and New South Wales; and var. compar, which is polymorphic and comprises two forms, a coastal form from southern Queensland and a montane form from north-eastern New South Wales and south-eastern Queensland. Ordination analysis of morphological characters of adults and seedlings indicates that the montane populations of var. compar comprise a separate taxon, which is phenetically closer to var. integrifolia than it is to typical var. compar. Banksia integrifolia var. aquilonia is phenetically quite distinct from the remaining taxa. The new names and combinations Banksia integrifolia subsp. monticola K.R. Thiele, B. integrifolia subsp. aquilonia (A.S. George) K.R. Thiele and B. integrifolia subsp. compar (R.Br.) K.R. Thiele are published.
APA, Harvard, Vancouver, ISO, and other styles
4

Woodside, D. P., and G. H. Pyke. "A Comparison of Bats and Birds as Pollinators of Banksia integrifolia in Northern New South Wales, Australia." Australian Mammalogy 18, no. 1 (1995): 9. http://dx.doi.org/10.1071/am95009.

Full text
Abstract:
We captured Queensland Blossom Bats (Syconycteris australis) feeding at the flowers of Banksia integrifolia during the night and several honeyeater species feeding at the same flowers during the day. Nearby were flowering Melaleuca quinquenervia and various forested areas including littoral rainforest. Honeyeaters appear to be more frequent visitors to the Banksia flowers than Blossom Bats but less effective at transporting pollen. When they are feeding at Banksia flowers both birds and bats carry pollen on the parts of their bodies that contact successive inflorescences. Hence, both honeyeaters and bats are likely to be pollinators of B. integrifolia in our study area. However, the flowers produce nectar and dehisce pollen primarily at night, suggesting that Blossom Bats are more important than honeyeaters as pollinators of this plant. Banksia pollen was the most common item in the diet of the Blossom Bats during our study and the bats were able to digest the contents of this pollen. Interestingly, the diet of these animals also included relatively small amounts of Melaleuca pollen, fruit and arthropods. The spatial and temporal patterns of capture of the Blossom Bats suggested that Blossom Bats prefer to forage at Banksia flowers that are near to the forested areas and that adult bats may influence where and when younger bats feed. Banksia integrifolia appears to produce nectar mostly during the night and/or early morning in two different locations, one coastal and one on the tablelands, but shows different daily patterns of pollen anthesis in these locations.
APA, Harvard, Vancouver, ISO, and other styles
5

Stimpson, Margaret Leith, JEREMY J. BRUHL, and PETER H. WESTON. "Could this be Australia’s rarest Banksia? Banksia vincentia (Proteaceae), a new species known from fourteen plants from south-eastern New South Wales, Australia." Phytotaxa 163, no. 5 (March 31, 2014): 269. http://dx.doi.org/10.11646/phytotaxa.163.5.3.

Full text
Abstract:
Possession of hooked, distinctively discolorous styles, a broadly flabellate common bract subtending each flower pair, and a lignotuber place a putative new species, Banksia sp. Jervis Bay, in the B. spinulosa complex. Phenetic analysis of individuals from all named taxa in the B. spinulosa complex, including B. sp. Jervis Bay, based on leaf, floral, seed and bract characters support recognition of this species, which is described here as Banksia vincentia M.L.Stimpson & P.H.Weston. Known only from fourteen individuals, B. vincentia is distinguished by its semi-prostrate habit, with basally prostrate, distally ascending branches from the lignotuber, and distinctive perianth colouring. Its geographical location and ecological niche also separate it from its most similar congeners.
APA, Harvard, Vancouver, ISO, and other styles
6

Carpenter, RJ, GJ Jordan, and RS Hill. "Banksieaephyllum taylorii ( Proteaceae) from the late paleocene of New South Wales and its relevance to the origin of Australia's scleromorphic flora." Australian Systematic Botany 7, no. 4 (1994): 385. http://dx.doi.org/10.1071/sb9940385.

Full text
Abstract:
Leaf specimens from Late Paleocene sediments in New South Wales are assigned to a new species of Banksieaephyllum, B. taylorii. In gross morphology the leaves are indistinguishable from those of extant Dryandra formosa, and similar to a few other species of Dryandra and Banksia. These species have pinnately lobed leaves and are now confined to south-western Australia. In cuticular morphology, B. taylorii is most similar to Banksia species from subgenus Banksia, section Oncostylis. One species in this section, B. dryandroides, also has pinnately lobed leaves. The fossil specimens demonstrate that subtribe Banksiinae had differentiated by the Late Paleocene and represent the earliest record of angiosperm scleromorphy in Australia to date. The superficial placement of the stomates compared with most modem Banksiinae supports the hypothesis that xeromorphy in this group generally increased in response to the development of less mesic climates in the Late Tertiary.
APA, Harvard, Vancouver, ISO, and other styles
7

Smith, A. P., and M. Murray. "Habitat requirements of the squirrel glider (Petaurus norfolcensis) and associated possums and gliders on the New South Wales central coast." Wildlife Research 30, no. 3 (2003): 291. http://dx.doi.org/10.1071/wr01115.

Full text
Abstract:
One of the largest known populations of the threatened squirrel glider occurs in the Wyong and Lake Macquarie regions of the New South Wales central coast. A study of the habitat requirements and density of this population was undertaken as a component in a broader study to develop a regional conservation strategy for the species. The squirrel glider was found to be widespread at an estimated average density of 0.39 animals ha–1. It was most abundant in forests and woodlands with an overstorey of winter-flowering eucalypts (Corymbia maculata, Eucalyptus robusta, Eucalyptus tereticornis) or an understorey of winter-flowering banksias (Banksia spinulosa) or pinnate-leaved acacias (Acacia irrorata). The highest estimated density (0.7 ha–1) occurred in associations of scribbly gum (Eucalyptus haemastoma or racemosa), smooth-barked apple (Angophora costata) and red bloodwood (Corymbia gummifera) with an understorey of Banksia spp and Xanthorrhoea spp. The lowest estimated densities occurred in forests with an understorey dominated by casuarinas or non-pinnate acacias and in stunted, low (<17 m high) forest and woodland close to the coast. The abundance of all possums and gliders increased significantly with canopy height, canopy cover, the number of mature and old-growth trees and the number of trees with hollows. Preferred habitat of the squirrel glider in this region occurs predominantly on freehold land where it is threatened by clearing for coastal development. Implementation of planning provisions to protect squirrel glider habitat on private land will be necessary to maintain the existing regional population.
APA, Harvard, Vancouver, ISO, and other styles
8

Griffith, S. J., C. Bale, and P. Adam. "The influence of fire and rainfall upon seedling recruitment in sand-mass (wallum) heathland of north-eastern New South Wales." Australian Journal of Botany 52, no. 1 (2004): 93. http://dx.doi.org/10.1071/bt03108.

Full text
Abstract:
Wallum heathland is extensive on coastal sand masses in north-eastern New South Wales and south-eastern Queensland. Here the climate is subtropical, although monthly rainfall is highly variable and unreliable. We examined the influence of fire and rainfall on seedling recruitment in bradysporous dry-heathland [Banksia aemula R.Br., Melaleuca nodosa (Sol. ex Gaertn.) Sm.] and wet-heathland [Banksia oblongifolia Cav., B.�ericifolia L.f. subsp. macrantha (A.S.George) A.S.George, Leptospermum liversidgei R.T.Baker and H.G. Sm.] species. Two specific questions were addressed: (1) do elevated levels of soil moisture facilitate seedling recruitment; (2) is the post-fire environment superior for seedling recruitment? Field experiments demonstrated that heathland species studied here are capable of successful recruitment in atypical habitat, and this proceeds irrespective of fire and unreliable rainfall. Conditions for growth and reproduction were found to be adequate if not more favourable in dry heathland, and this outcome included species usually associated with wet heathland. Spatial and temporal trends in seedling emergence and survival were examined in relation to post-fire predation and plant resource availability. Existing ideas about wallum management and conservation are evaluated, in particular the role of fire.
APA, Harvard, Vancouver, ISO, and other styles
9

van Tets, I. G. "Can Flower-Feeding Marsupials Meet Their Nitrogen Requirements on Pollen in The Field?" Australian Mammalogy 20, no. 3 (1998): 383. http://dx.doi.org/10.1071/am98383.

Full text
Abstract:
Two arboreal marsupials, the eastern pygmy possum (Cercartetus nanus) and the sugar glider (Petaurus breviceps) have exceptionally low maintenance nitrogen requirements on pollen diets. This study compares their nitrogen requirements with the density of Banksia pollen that is available in the Barren Grounds Nature Reserve, New South Wales, a site where both species are known to forage on Banksia inflorescences. The pollen density was sufficiently high that both species were capable of meeting their maintenance nitrogen requirements on pollen whenever Banksia spp. were in flower. C. nanus required a smaller proportion of its home range than P. breviceps to do so and pollen was likely to be of much greater nutritional significance to both species in winter than in summer. This corresponds closely with the results of field studies comparing the diets of these mammals at different times of the year. Pollen is an important source of nitrogen for flower-feeding marsupials but its importance will vary between species depending on the marsupial&apos;s requirements, its body size and on the quantity of pollen that is available.
APA, Harvard, Vancouver, ISO, and other styles
10

Bowen, M., and R. Goldingay. "Distribution and Status of The Eastern Pygmy Possum (Cercartetus nanus) in New South Wales." Australian Mammalogy 21, no. 2 (1999): 153. http://dx.doi.org/10.1071/am00153.

Full text
Abstract:
The eastern pygmy possum (Cercartetus nanus) has a wide distribution in New South Wales (NSW), but is infrequently detected in fauna surveys. We collated available information on the distribution, habitat and detection rates for C. nanus in NSW from results of published and unpublished fauna surveys. These data, and those from the National Parks and Wildlife Service and Australian Museum databases, suggest that C. nanus populations are concentrated in south-eastern NSW and are sparsely distributed throughout the rest of the state. Several records extend the distribution of this species further west than currently shown by published distribution maps. Records show differences in habitat types occupied by C. nanus between south-eastern and north-eastern NSW. In south-eastern NSW, C. nanus occupies a range of habitats including heath, woodland and open forest, at a range of altitudes. In north-eastern NSW, C. nanus appears to be associated mainly with rainforest at high altitudes. Of the range of techniques available, nest boxes and Elliott traps positioned against flowering Banksia species are most effective at capturing C. nanus. Given the large survey effort and the small number of surveys detecting &gt;I0 C. nanus, it appears that this species is rare throughout most of NSW. We recommend that C. nanus be considered for listing as a vulnerable species in NSW.
APA, Harvard, Vancouver, ISO, and other styles
11

Parry-Jones, K., and ML Augee. "Food Selection by Grey-headed Flying Foxes (Pteropus poliocephalus) Occupying a Summer Colony Site near Gosford, New South Wales." Wildlife Research 18, no. 1 (1991): 111. http://dx.doi.org/10.1071/wr9910111.

Full text
Abstract:
A colony site occupied by grey-headed flying-foxes (Pteropus poliocephalus) from October to May on the central coast of N.S.W. was monitored over a 48 month period (1986-1990). Faecal and spat-out material was collected for microscopic determination of contents. Comparison of food items in the droppings with the array of possible food sources present in the vicinity of the colony at the same time showed a marked preference for certain foods, in particular blossoms of the family Myrtaceae and of the genus Banksia. Cultivated orchard fruits were not a preferred food and were only taken at times when preferred food items were scarce.
APA, Harvard, Vancouver, ISO, and other styles
12

Carpenter, Raymond J., Matthew P. Goodwin, Robert S. Hill, and Karola Kanold. "Silcrete plant fossils from Lightning Ridge, New South Wales: new evidence for climate change and monsoon elements in the Australian Cenozoic." Australian Journal of Botany 59, no. 5 (2011): 399. http://dx.doi.org/10.1071/bt11037.

Full text
Abstract:
Diverse Cenozoic (possibly latest Oligocene to mid–late Miocene) macrofossils from the Lightning Ridge opal fields are illustrated and discussed. Specimens identified to, or closely comparable with, extant taxa include ferns (Lygodium, Gleichenia and others), conifers now extinct in Australia (Dacrydium, Retrophyllum and Papuacedrus), Lauraceae (Cryptocarya/Cinnamomum), sclerophyllous Proteaceae (Banksia, Lomatia and Grevillea), Cunoniaceae/Elaeocarpaceae and Eucalyptus (and/or other Myrtaceae). Overall, at least four fern, three conifer and 30 angiosperm taxa are recognised. The climate supported many species with close relatives in wet Australasian habitats, including rainforests. However, a drier or more seasonal (?monsoonal) aspect is especially indicated by the presence of lobed leaves that resemble extant species of Brachychiton (Malvaceae), Erythrina (Fabaceae) and tribe Cercideae (Fabaceae). A degree of water stress is also suggested by the prevalence of narrow, toothed and/or deeply lobed angiosperm leaves.
APA, Harvard, Vancouver, ISO, and other styles
13

Dalgleish, Elizabeth. "Effectiveness of Invertebrate and Vertebrate Pollinators and the Influence of Pollen Limitation and Inflorescence Position on Follicle Production of Banksia aemula (Family Proteaceae)." Australian Journal of Botany 47, no. 4 (1999): 553. http://dx.doi.org/10.1071/bt97070.

Full text
Abstract:
Follicle development of Banksia aemula (R.Brown 1810) was studied in northern New South Wales, Australia, after exposure of inflorescences to different combinations of pollinator type and pollen quantity. When inflorescences within plants were exposed to all pollinators and provided with additional cross-pollen, follicle development was increased, suggesting that B. aemula was pollen-limited. The addition of cross-pollen did not increase follicle development when inflorescences within plants were exposed to invertebrate pollination only. Nor did exclusion of vertebrates significantly reduce follicle development of plants relative to that of others which were exposed to all pollinators. The vegetation surrounding plants influenced the follicle development of inflorescences, and inflorescences in peripheral positions had more follicles than inflorescences that were internal.
APA, Harvard, Vancouver, ISO, and other styles
14

Dodson, JR, and CA Myers. "Vegetation and Modern Pollen Rain From the Barrington Tops and Upper Hunter River Regions of New South Wales." Australian Journal of Botany 34, no. 3 (1986): 293. http://dx.doi.org/10.1071/bt9860293.

Full text
Abstract:
Vegetation was mapped using existing maps, Landsat interpretation, aerial photograph interpretation and some verification by ground surveys. Twenty-five moss cushions were collected to identify pollen rain and pollen indicators of vegetation for use in fossil pollen interpretation. Eucalyptus (10%), Poaceae (4-10%), Casuarina (4-5%), Asteraceae (Tubuliflorae) (0-4%) and Cyperaceae (0-2%) were the general components in the pollen rain of the region. Subtropical rain forest was characterized by Backhousia and low values of a wide range of taxa. Cool temperate rain forest had Nothofagus values above 40% and Eucalyptus values below 20%. Subalpine grasslands had Poaceae values above 45%, Eucalyptus values below 15% and small quantities of Epacridaceae and Goodeniaceae pollen. Subalpine forest and wet eucalypt forest formations had very similar pollen representation and could be confused in pollen spectra. However, Monotoca, Banksia, Leptospermum pollen and fern spores were more common in the wet eucalypt forests. Dry eucalypt formations were characterized by 2-20% values of Bursaria, Callitris and Dodonaea as well as eucalypt values.
APA, Harvard, Vancouver, ISO, and other styles
15

Bladon, R. V., C. R. Dickman, and I. D. Hume. "Effects of habitat fragmentation on the demography, movements and social organisation of the eastern pygmy-possum (Cercartetus nanus) in northern New South Wales." Wildlife Research 29, no. 1 (2002): 105. http://dx.doi.org/10.1071/wr01024.

Full text
Abstract:
A population of eastern pygmy-possums (Cercartetus nanus) was studied in northern New South Wales for almost 3 years. A total of 98 pygmy-possums was captured, of which 52 were captured only once. The sex ratio of the population did not differ significantly from parity. Mid-way through the study, 1.4 ha of the 4.0-ha study site was cleared. Pre-clearing capture rates in nest boxes averaged 33.5 ± 5.8 captures per 100 box checks per month, and the population was estimated by three methods to be at least 15–20 animals. There was no significant difference in body mass between adult males (23.7 ± 6.3 g) and adult females (27.1 ± 7.7 g). Males had significantly larger short-term home ranges (0.35 ± 0.14 ha) than females (0.14 ± 0.06 ha) and tended to move over greater distances each night. Breeding occurred from summer to early winter, and juveniles and sub-adults entered the population in autumn and winter. The mean number of pouch young was 3.9. The most likely minimum size at which juveniles left their mother was 9–11 g. Adult body mass and condition were highly variable over time, and did not appear to be related to either the breeding season or Banksia flowering. Fourteen pygmy-possums were found torpid during the study. Population troughs occurred in late winter and spring and were associated with low survival and/or seasonal migration, possibly linked to the cessation of Banksia flowering in July and the lack of alternative food sources at this time and/or increased use of nest boxes by Antechinus stuartii during late winter. Post-clearing, capture rates fell to 7.8 ± 1.6 captures per 100 box checks per month, the estimated population size fell to 5–8 animals, and there was an almost total lack of juvenile/sub-adult recruitment into the population. The results support concerns that the long-term survival of the eastern pygmy-possum in New South Wales is threatened by continued land clearing throughout much of its present range.
APA, Harvard, Vancouver, ISO, and other styles
16

Drury, Rebecca L., and Fritz Geiser. "Activity patterns and roosting of the eastern blossom-bat (Syconycteris australis)." Australian Mammalogy 36, no. 1 (2014): 29. http://dx.doi.org/10.1071/am13025.

Full text
Abstract:
We quantified activity patterns, foraging times and roost selection in the eastern blossom-bat (Syconycteris australis) (body mass 17.6 g) in coastal northern New South Wales in winter using radio-telemetry. Bats roosted either in rainforest near their foraging site of flowering coast banksia (Banksia integrifolia) and commuted only 0.3 ± 0.1 km (n = 8), whereas others roosted 2.0 ± 0.2 km (n = 4) away in wet sclerophyll forest. Most bats roosted in rainforest foliage, but in the wet sclerophyll forest cabbage palm leaves (Livistonia australis) were preferred roosts, which likely reflects behavioural thermoregulation by bats. Foraging commenced 44 ± 22 min after sunset in rainforest-roosting bats, whereas bats that roosted further away and likely flew over canopies/open ground to reach their foraging site left later, especially a female roosting with her likely young (~4 h after sunset). Bats returned to their roosts 64 ± 12 min before sunrise. Our study shows that S. australis is capable of commuting considerable distances between appropriate roost and foraging sites when nectar is abundant. Bats appear to vary foraging times appropriately to minimise exposure to predators and to undertake parental care.
APA, Harvard, Vancouver, ISO, and other styles
17

Sharpe, D. J., and R. L. Goldingay. "Feeding behaviour of the squirrel glider at Bungawalbin Nature Reserve, north-eastern New South Wales." Wildlife Research 25, no. 3 (1998): 243. http://dx.doi.org/10.1071/wr97037.

Full text
Abstract:
The diet of the squirrel glider (Petaurus norfolcensis) was described by qualitative observations of feeding behaviour at a floristically rich site on the north coast of New South Wales. Twelve gliders from six groups were examined over a 10-month period. Flowering and bark-shedding data were also collected. Nectar and pollen were the most important food resources and accounted for 59% of all observations. Banksia integrifolia was the most important source of these foods, but eucalypts were used heavily when in flower and several other genera were also visited. Feeding on arthropods constituted 26% of all feeding observations. Arthropods were harvested in all months of the study from a variety of substrates. Feeding on arthropods was relatively unimportant in May and June when pollen ingestion was presumed to be high. Honeydew was used but was absent from the diet during winter. Acacia gum was obtained from two species in autumn and one, Acacia irrorata, was incised to promote gum production. Corymbia intermedia and Angophora woodsiana were incised for sap in autumn and winter. Sap flows resulting from insect (borer) damage on other species were also used. Fruit, Acacia seeds and arils, and lichens were consumed on a few occasions. The squirrel glider displayed seasonal trends in feeding behaviour that, in part, accorded with observed phenological patterns. The foods used by the squirrel glider during this study were similar to those previously reported for the genus. However, few studies have documented such a diversity of dietary items at one site. Management of the squirrel glider appears to require the maintenance of floristic diversity, and particularly the persistence of midstorey species.
APA, Harvard, Vancouver, ISO, and other styles
18

Harris, Jamie M., Ross L. Goldingay, and Lyndon O. Brooks. "Population ecology of the eastern pygmy-possum (Cercartetus nanus) in a montane woodland in southern New South Wales." Australian Mammalogy 36, no. 2 (2014): 212. http://dx.doi.org/10.1071/am13044.

Full text
Abstract:
The population dynamics of nectar-feeding non-flying mammals are poorly documented. We investigated aspects of the population ecology of the eastern pygmy-possum (Cercartetus nanus) in southern New South Wales. We captured 65 individuals over a 4-year period during 5045 trap-nights and 1179 nest-box checks. The body mass of adult males (mean ± s.e. = 24.6 ± 1.0 g) was marginally not significantly different (P = 0.08) from that of non-parous adult females (28.2 ± 1.9 g). Females gave birth to a single litter each year of 3–4 young during February–May. No juveniles were detected in spring of any year. Mark–recapture modelling suggested that survival probability was constant over time (0.78) while recapture probability (0.04–0.81) varied with season and trap effort. The local population (estimated at ~20–25 individuals) underwent a regular seasonal variation in abundance, with a decline in spring coinciding with the cessation of flowering by Banksia. A population trough in spring has been observed elsewhere. This appears to represent some local migration from the study area, suggesting a strategy of high mobility to track floral resources. Conservation of this species will depend on a more detailed understanding of how flowering drives population dynamics.
APA, Harvard, Vancouver, ISO, and other styles
19

Harris, J. M., R. L. Goldingay, L. Broome, P. Craven, and K. S. Maloney. "Aspects of the ecology of the eastern pygmy-possum Cercartetus nanus at Jervis Bay, New South Wales." Australian Mammalogy 29, no. 1 (2007): 39. http://dx.doi.org/10.1071/am07004.

Full text
Abstract:
A variety of ecological data were collected on the eastern pygmy-possum Cercartetus nanus at Jervis Bay, in south-eastern New South Wales between March 2006 and January 2007. Elliott traps, pitfall traps, nest-boxes and spotlighting were used to survey for the species. Data on habitat suitability including abundance of food plants (flowering trees and shrubs) and potential nest sites were also collected. Home range data were gathered via radio telemetry. Three individuals were caught in 2150 trap-nights and one animal was re-trapped once. Radio-collars were attached to one animal of each sex and tracked for 11 days during March 2006. These possums used areas (using minimum convex polygons) of 0.85 ha (male) and 0.19 ha (female). The average overnight distance moved was 44 m for the male (range = 4-81 m) and 19 m for the female (range = 0-56 m). Nest-sites included hollows in the proteaceous shrubs Banksia serrata and B. ericifolia, and in the myrtaceous trees Corymbia gummifera, Eucalyptus sclerophylla, and Syncarpia glomulifera. Cercartetus nanus captures were confined to two sites that had the most prolific flowering of potential food plants and the highest availability of potential nest-sites. A review of literature and previous surveys of the surrounding area was a necessary precursor to field study and produced 57 records. Greater understanding of the impacts of development and fire are needed for conservation and management of this species.
APA, Harvard, Vancouver, ISO, and other styles
20

Phillips, S., D. Coburn, and R. James. "An Observation Of Cat Predation Upon An Eastern Blossom Bat Syconycteris Australis." Australian Mammalogy 23, no. 1 (2001): 57. http://dx.doi.org/10.1071/am01057.

Full text
Abstract:
WITH a body weight of 15 - 19 g and a mean headbody length of just over 60 mm (Churchill 1998), the eastern blossom bat Syconycteris australis is one of the smallest members of the sub-order Megachiroptera. Within Australia S. australis is restricted in distribution to the east coast from Cape York in Queensland to near Forster on the mid-north coast of New South Wales (NSW) (Law 1994a). Habitat requirements include both rainforest and/or wet sclerophyll forest for roosting purposes and proximal areas of heathland for foraging (Law 1993). The species survives on a diet of nectar and pollen and is heavily dependent upon Banksia integrifolia inflorescences during the winter months (Law 1994b, 1996; Coburn 1995). Blossom bats are generally regarded as solitary and exhibit strong fidelity to their feeding areas (Law 1993), although movements of up to 30 km have been reported (Law 1996).
APA, Harvard, Vancouver, ISO, and other styles
21

O'Brien, Eleanor K., Lucia A. Aguilar, David J. Ayre, and Robert J. Whelan. "Genetic tests of the isolation of rare coastal dwarf populations of Banksia spinulosa." Australian Journal of Botany 58, no. 8 (2010): 637. http://dx.doi.org/10.1071/bt10112.

Full text
Abstract:
In southern New South Wales, a suite of widespread plant species exhibit short-statured ‘dwarf’ growth forms on coastal headlands. It is unclear whether such populations are genetically distinct or whether dwarfism is a plastic response to the environment. We used four microsatellite markers to assess genetic differentiation among populations from coastal and inland sites for Banksia spinulosa var. spinulosa. We sampled plants from six locations, including from three ‘dwarf’ and three ‘normal’ populations. Mean levels of genetic diversity were slightly higher in the forest (Na = 7.07 ± 0.25; He = 0.80 ± 0.09) than on the coast (Na = 5.92 ± 0.70; He = 0.72 ± 0.10). In general, populations displayed genotypic diversity expected for outcrossed sexual reproduction, with 161 of 172 individuals displaying unique genotypes and mean values of FIS close to zero. However, we found evidence of at least limited clonal replication in each of four populations and, within one coastal population, 11 of 27 individuals displayed one of three replicated genotypes, implying that the effective population size may be considerably smaller than would be inferred from the number of plants at this site. Relative to studies with other Proteaceae, this set of populations showed low, although significant, levels of differentiation (global FST = 0.061; P < 0.001), with extremely low, although significant, divergence of forest and coastal populations (FRT = 0.009; P < 0.001). There was no evidence of isolation by distance. These data imply that coastal dwarf populations are genetically similar to more extensive inland populations but in at least one case, may be at a greater risk of extinction because of low effective population size.
APA, Harvard, Vancouver, ISO, and other styles
22

Rutherford, Susan, Stephen J. Griffith, and Nigel W. M. Warwick. "Water relations of selected wallum species in dry sclerophyll woodland on the lower north coast of New South Wales, Australia." Australian Journal of Botany 61, no. 4 (2013): 254. http://dx.doi.org/10.1071/bt13037.

Full text
Abstract:
The present study examined the water relations of wallum dry sclerophyll woodland on the lower north coast of New South Wales (NSW). Wallum is the regionally distinct vegetation of Quaternary dunefields and beach ridge plains along the eastern coast of Australia. Wallum sand masses contain large aquifers, and previous studies have suggested that many of the plant species may be groundwater dependent. However, the extent of this dependency is largely unknown, despite an increasing reliance on the aquifers for groundwater extraction. Fifteen species from five growth-form categories and seven plant families were investigated. The pre-dawn and midday xylem water potential (ψx) of all species was monitored over a 20-month period from December 2007 to July 2009. Pressure–volume curve traits were determined for each species in late autumn 2008, including the osmotic potential at full (π100) and zero (π0) turgor, and bulk modulus of elasticity (ε). Carbon isotope ratios (δ13C) were determined in mid-autumn 2008 to measure water use efficiency (WUE). Comparative differences in water relations could be loosely related to growth forms. A tree (Eucalyptus racemosa subsp. racemosa) and most large shrubs had low midday ψx, π100 and π0, and high ε and WUE; whereas the majority of small and medium shrubs had high midday ψx, π100 and π0, and low ε and WUE. However, some species of similar growth form displayed contrasting behaviour in their water relations (e.g. the herbs Caustis recurvata var. recurvata and Hypolaena fastigiata), and such differences require further investigation. The results suggest that E. racemosa subsp. racemosa is likely to be groundwater dependent, and large shrubs such as Banksia aemula may also utilise groundwater. Both species are widespread in wallum, and therefore have the potential to play a key role in monitoring ecosystem health where aquifers are subject to groundwater extraction.
APA, Harvard, Vancouver, ISO, and other styles
23

Law, BS. "Roosting and foraging ecology of the Queensland blossom bat (Syconycteris australis) in north-eastern New South Wales: flexibility in response to seasonal variation." Wildlife Research 20, no. 4 (1993): 419. http://dx.doi.org/10.1071/wr9930419.

Full text
Abstract:
Radiotelemetry was used to track blossom bats (Syconycteris australis) at Iluka and Harrington in northern New South Wales. A total of 31 bats was tracked to 110 roosts. Bats foraged on nectar and pollen in Banksia integrtfolia heathland, but roosted 50-4000m away in littoral rainforest. Bats showed a strong fidelity to their feeding area (about 13ha), returning to their original capture point each night and spending a large proportion of their foraging time there. After leaving their roost, adults spent, on average, 45% of their time active and remained in heathland throughout the night. All age-sex classes roosted solitarily during the day amongst rainforest foliage, usually in the subcanopy layer. Most roosts were occupied for one day only and adults were more roost-mobile than juveniles. Mean movements between roosts were greater at Harrington (125m), where the rainforest is fragmented, than at Iluka (42m), where rainforest is intact. Bats shifted their roosts seasonally, from the rainforest edge in winter to the rainforest interior in spring/autumn. This behaviour allows for avoidance of cold temperatures inside the forest in winter and of hot temperatures of the forest exterior in spring/autumn. A further possible response to the seasonal climate prevailing at the study area was a reduction in the commuting distance (from roosts to feeding areas) from autumn/spring (1.4km) to winter (0.8km). Such flexible roosting and foraging strategies may be effective in allowing S. australis to exploit subtropical and temperate areas of Australia.
APA, Harvard, Vancouver, ISO, and other styles
24

Griffith, Stephen J., Susan Rutherford, Kerri L. Clarke, and Nigel W. M. Warwick. "Water relations of wallum species in contrasting groundwater habitats of Pleistocene beach ridge barriers on the lower north coast of New South Wales, Australia." Australian Journal of Botany 63, no. 7 (2015): 618. http://dx.doi.org/10.1071/bt15103.

Full text
Abstract:
This study examined the water relations of sclerophyllous evergreen vegetation (wallum) on coastal sand barriers in eastern Australia. Many wallum species may be groundwater dependent, although the extent of this dependency is largely unknown. Twenty-six perennial tree, shrub and herb species were investigated in three groundwater habitats (ridge, open depression, closed depression). Pre-dawn and midday shoot xylem water potentials (ψx) were measured monthly between late autumn 2010 and late summer 2011. Pressure–volume curve traits were determined in mid- to late spring 2009, including the osmotic potential at full (π100) and zero (π0) turgor, and bulk modulus of elasticity (ε). Carbon isotope ratios (δ13C) were also determined in mid- to late spring 2009, to measure water-use efficiency (WUE). The species displayed a range of physiological strategies in response to water relations, and these strategies overlapped among contrasting growth forms and habitats. Linear relationships between osmotic and elastic adjustment were significant. A strong correlation between δ13C and distribution along the hydrological gradient was not apparent. Banksia ericifolia subsp. macrantha (A.S.George) A.S.George, Eucalyptus racemosa Cav. subsp. racemosa and Eucalyptus robusta Sm. displayed little seasonal variation in ψx and maintained a comparatively high pre-dawn ψx, and are therefore likely to be phreatophytic. Wetland vegetation in the lowest part of the landscape appeared to tolerate extreme fluctuations in water availability linked to a prevailing climatic pattern of variable and unreliable seasonal rainfall.
APA, Harvard, Vancouver, ISO, and other styles
25

Bradstock, RA, and M. Bedward. "Simulation of the Effect of Season of Fire on Post-Fire Seedling Emergence of Two Banksia Species Based on Long-Term Rainfall Records." Australian Journal of Botany 40, no. 1 (1992): 75. http://dx.doi.org/10.1071/bt9920075.

Full text
Abstract:
Simulations were used to investigate the effect of season of fire on seedling emergence in Banksia ericifolia and B. serrata in the Sydney region, New South Wales. The simulations were based on models of soil-surface moisture in response to rainfall, seedling emergence response to soil moisture and post-fire seed release from fruits as determined by fire intensity, derived from field and laboratory studies. Fires were modelled on the first day of each calendar month for a 50-year period. Levels of post-fire seedling emergence were calculated using rainfall data from the Sydney Observatory for the corresponding period (1931-1980). Trends in seedling emergence as a function of month of fire were examined. Alternative sets of simulations were performed to assess the effect of variations in post-dispersal seed mortality, fire intensity and induced summer dormancy. In both species, mean emergence was affected by season of fire only when a 10% per month level of post-dispersal seed mortality was simulated (there was no fire-season effect at lower mortality levels). Highest predicted emergence occurred after summer fires and lowest emergence after winter fires. A reduction in rate of seed release (lower intensity fire) and induced seed dormancy in summer also had a minor effect with respect to fire-season in B. ericifolia. Reported levels of post-dispersal seed mortality in Banksia species are often high, and therefore, the simulations suggested that there will be an effect of fire season on seedling emergence. However, given the high level of year to year variation in seasonal rainfall in the Sydney region, fire-season effects are not predictable in the short term. In the longer term, the timing of fire relative to sequences of wet and dry years may be of equal importance to season of fire in its effect on populations of these species.
APA, Harvard, Vancouver, ISO, and other styles
26

Tulloch, Ayesha I., and Chris R. Dickman. "Floristic and structural components of habitat use by the eastern pygmy-possum (Cercartetus nanus) in burnt and unburnt habitats." Wildlife Research 33, no. 8 (2006): 627. http://dx.doi.org/10.1071/wr06057.

Full text
Abstract:
The eastern pygmy-possum (Cercartetus nanus) occurs broadly but patchily in south-eastern Australia. It is a small, difficult-to-trap marsupial with poorly known resource and habitat preferences. This study investigated the structural and floristic habitat resources used and selected by C. nanus in Royal National Park (which was heavily burnt by bushfire in 1994) and Heathcote National Park (most of which had remained unburnt for over two decades at the time of study), in central-coastal New South Wales. Three different sampling methods were used – pitfall traps, Elliott traps and hair tubes – with pitfall trapping being by far the most effective method for detecting C. nanus. Live-trapping in different habitats revealed higher numbers of C. nanus in unburnt and burnt woodland, burnt heathland and burnt coastal complex than in unburnt coastal complex and burnt and unburnt rainforest. To identify the components of habitat contributing to this pattern, we first scored floristic and structural features of vegetation around trap stations and then quantified habitat components further by using spool- and radio-tracking. We found little evidence that C. nanus responded to any structural components of habitat, although arboreal activity was greater, not surprisingly, in wooded than in burnt heathland habitats. C. nanus was associated most strongly with the abundance of certain plants in the Proteaceae and Myrtaceae. In particular, the species prefers Banksia spp. (probably for food) and Eucalyptus and Xanthorrhoea spp. (probably for shelter).
APA, Harvard, Vancouver, ISO, and other styles
27

Law, Bradley S. "Residency and Site Fidelity of Marked Populations of the Common Blossom Bat Syconycteris australis in Relation to the Availability of Banksia Inflorescences in New South Wales, Australia." Oikos 77, no. 3 (December 1996): 447. http://dx.doi.org/10.2307/3545934.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Bradstock, RA, and PJ Myerscough. "The Survival and Population Response to Frequent Fires of Two Woody Resprouters Banksia serrata and Isopogon anemonifolius." Australian Journal of Botany 36, no. 4 (1988): 415. http://dx.doi.org/10.1071/bt9880415.

Full text
Abstract:
Plants of B. serrata and I. Anemonifolius resprout after fire, although the species differ in morphology (single-stemmed small tree, multistemmed low shrub respectively). If fires occur before newly established plants are fire-tolerant, populations will decline. The age of first fire tolerance was found to be lower in B. serrata (6 years) than in I. Anemonifolius (about 13 years). Rates of survival between and during fires were measured in the field along with rates of stem regrowth in fire-tolerant B. serrata juveniles. These results were used to predict rates of population decline under repeated fires sufficiently closely spaced to prevent the survival of newly established genets. In both species, juveniles were more prone to death than adults in fires and high-intensity fires caused most deaths. In B. serrata, adult stems (>2.0 cm d.b.h.) are mostly fire-tolerant, but fires less than 10 years apart can prevent many juveniles which survive from reaching adulthood. This restriction is less likely in I. Anemonifolius. As a result stands of B. serrata may decline more rapidly than I. Anemonifolius under 5-year fire cycles. I. Anemonifolius populations, however, may be more prone to decline when the interval between fires is slightly longer (e.g. 10 years) because lignotubers in young juveniles develop at a slower rate than in B. serrata. Extinction or substantial depletions of adult numbers may be approached in stands of either species after 50 years under some repeated 5- or 10-year fire cycles. The rate of such declines will depend directly on the structure of populations (proportions of adults and juveniles). Declines in populations of these resprouters may be likely under current fire regimes within the Sydney region of New South Wales, although these species are more likely to persist through long runs of frequent fire (<lo year interval) than some cohabiting species of obligate seeders.
APA, Harvard, Vancouver, ISO, and other styles
29

Harris, JM, and RL Goldingay. "Detection of the eastern pygmy-possum Cercartetus nanus (Marsupialia: Burramyidae) at Barren Grounds Nature Reserve, New South Wales." Australian Mammalogy 27, no. 1 (2005): 85. http://dx.doi.org/10.1071/am05085.

Full text
Abstract:
THE eastern pygmy-possum (Cercartetus nanus) has an extensive distribution, from south-eastern Queensland to south-eastern South Australia, and also into Tasmania (Strahan 1995). Despite this it is rarely detected in fauna surveys (Bowen and Goldingay 2000). This rarity in detection suggested that the species may be characterised by small and isolated populations, and therefore vulnerable to extinction. Consequently, it became listed as a 'Vulnerable' species in New South Wales (NSW) in 2001. Unless resolved, the low rate of detection of C. nanus will continue to hinder the acquisition of basic ecological information that is needed to more clearly define its conservation status and that is fundamental to the development of a recovery plan. An extensive body of survey data for NSW involving C. nanus has been reviewed by Bowen and Goldingay (2000). Among a range of survey methods aimed at detecting this species, trapping within flowering banksias and checking installed nest-boxes had the highest rates of detection. Indeed, one study in northern NSW captured 98 individuals over a 3- year period from within nest-boxes (Bladon et al. 2002). All other studies detected fewer than 15 C. nanus. It is clear that further research is required to investigate the effectiveness of a range of detection methods.
APA, Harvard, Vancouver, ISO, and other styles
30

Vadala, Anthony J., and Andrew N. Drinnan. "Elaborating the fossil history of Banksiinae: a new species of Banksieaephyllum (Proteaceae) from the late paleocene of New South Wales." Australian Systematic Botany 11, no. 4 (1998): 439. http://dx.doi.org/10.1071/sb97021.

Full text
Abstract:
Leaf fragments from Late Paleocene sediments at Cambalong Creek in the Southern Highlands of New South Wales are assigned to a new species ofBanksieaephyllum Cookson & Duigan,B. praefastigatum. A study of leaf form andmicromorphological characters of extant Banksieae was carried out to identify possible affinities for the new taxon, and a compendium of the architecturaland micromorphological characters of leaves of all described species ofBanksieaephyllum andBanksieaeformis Hill & Christophel is presented.Banksieaephyllum praefastigatum has characteristicstomatal and trichome features of both extinct and extant species ofBanksiinae, but is dissimilar in leaf morphology to any extant species ofBanksia L.f., Dryandra R.Br., orthe oldest previously described species ofBanksieaephyllum, B. tayloriiCarpenter, Hill & Jordan, with which it was contemporaneous.Banksieaephyllum praefastigatum, with its stronglydeveloped areolation and superficial stomates, is different from extantspecies of Banksiinae and Musgraveinae, and may represent a now-extinct sistertaxon to these subtribes in Banksieae, one which had not changed substantiallyfrom hypothesised mesophytic ancestral Proteaceae. Leaf morphology ofB. praefastigatum indicates that serrate-, lobed-andentire-margined forms of Banksieaephyllum were coeval in many localities throughout southern Australia during the Tertiary, and that Banksiinae had diversified significantly by the Early Tertiary, reflecting diversification of at least several other subtribes of subfamily Grevilleoideae by that time.
APA, Harvard, Vancouver, ISO, and other styles
31

Quin, DG. "Population ecology of the squirrel glider (Petaurus norfolcensis) and the sugar glider (P. breviceps) (Maruspialia : Petauridae) at Limeburners Creek, on the central north coast of New South Wales." Wildlife Research 22, no. 4 (1995): 471. http://dx.doi.org/10.1071/wr9950471.

Full text
Abstract:
The population ecology of the squirrel glider (Petaurus norfolcensis) and the sugar glider (P. breviceps) was studied at Limeburners Creek Nature Reserve, on the central north coast of New South Wales. The study was undertaken between July 1986 and November 1988. Sugar and squirrel gliders at Limeburners Creek exhibited similar home-range sizes (2.54 ha) despite considerable differences in mean mass between the species (squirrel glider 192-213 g; sugar glider, 104-119 g). Squirrel gliders existed at higher densities (049-1.54 ha-') than did sugar gliders (0.24454 ha-') and populations of both species exhibited male-biased sex ratios. The timing of births was not consistent between years and, at least in the squirrel glider, occurred in almost all months of the year over the 2.5-year study. Usually a winter peak in births that extended into spring was apparent, sometimes following an autumn peak. Mean litter size for both species (1.8-1.9) was similar to that recorded for the sugar glider in Victoria. Most adult females of both species exhibited the capacity to raise two litters in a year. Hence, natality rates (2.3-2.4 young per year) at Limeburners Creek were high relative to those recorded for the sugar glider in Victoria. Recruit persistence time (3.0-3.5 seasons) was similar between the species and recruitment appeared to be most successful during years when heavy eucalypt and Banksia flowering was recorded. Populations of both species were characterised by high rates of juvenile dispersal and mortality. Young gliders dispersed at a mean age of 10.9 and 12.5 months in the sugar glider and the squirrel glider, respecti~e1y;SquirreI gliders nested in colonies of 2-Bindividuals. Usually at least one male and two females nested together, suggesting a polygynous mating strategy. The mating system of the sugar glider at Limebumers Creek was less clear, but colonies appeared to comprise at least monogamous pairs with or without a surplus of males. Sugar glider colonies at Limebumers Creek varied in size from two to seven individuals. The larger squirrel gliders were clearly dominant to the smaller sugar gliders in interspecific behavioural interactions. Inconsistencies in body-weight fluctuations between years for both species were thought to be a consequence of the unpredictable nature of the aseasonal, coastal climates and resulting food resource abundances.
APA, Harvard, Vancouver, ISO, and other styles
32

Denham, Andrew J., and Robert J. Whelan. "Reproductive ecology and breeding system of Lomatia silaifolia (Proteaceae) following a fire." Australian Journal of Botany 48, no. 2 (2000): 261. http://dx.doi.org/10.1071/bt98039.

Full text
Abstract:
Lomatia silaifolia (Smith) R.Br. (Proteaceae) is a common shrub in southeastern Australian bushland that generally flowers only once after each fire. However, little is known of the details of this post-fire flowering. Lomatia silaifolia has flowers well-spaced along its conflorescence axis, unlike many other Proteaceae species, thus allowing close examination of the influence of flower position, mate choice and flowering sequence on fruit production. We examined breeding system, flowering phenology and spatial patterns of fruit set in the species, after a fire in September 1992 at Bulli Tops, New South Wales. Flowering occurred in December–January, but only in the first and second summers after the fire. The species is partially self-compatible: only 25% of self-pollinated conflorescences initiated fruits compared with 100% of cross-pollinated conflorescences. Fewer flowers initiated fruits after hand self-pollination (3.0%) than after cross-pollination (35.4%). Self-pollinated flowers produced fewer viable seeds (22.2% viable) than cross-pollinated flowers (62.9% viable). Pollen tubes were found in 72% of the self-pollinated flowers examined, suggesting that there is no early stylar self-incompatibility in the species. There was a high level of herbivory on flowering and fruiting branches, with 69% of unbagged branches completely destroyed. This may significantly affect recruitment in the species, given the limited opportunities for reproduction in the post-fire environment. In this study, fruits were evenly distributed along the conflorescence axis unlike some other species in the Proteaceae with more compressed, spike-like flowering structures where fruits are typically concentrated in some parts of the conflorescence axis (e.g. Telopea and some Banksia spp.). No bird visitors to flowers were observed, but a variety of insect visitors were identified.
APA, Harvard, Vancouver, ISO, and other styles
33

Underwood, A. J. "Grazing and disturbance: an experimental analysis of patchiness in recovery from a severe storm by the intertidal alga Hormosira banksii on rocky shores in New South Wales." Journal of Experimental Marine Biology and Ecology 231, no. 2 (December 1998): 291–306. http://dx.doi.org/10.1016/s0022-0981(98)00091-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Baehr, Barbara, Mark Harvey, H. M. Smith, and R. Ott. "The goblin spider genus Opopaea in Australia and the Pacific islands (Araneae: Oonopidae)." Memoirs of the Queensland Museum - Nature 58 (May 31, 2013): 107–338. http://dx.doi.org/10.17082/j.2204-1478.58.2013.2013-11.

Full text
Abstract:
The widespread and highly diverse goblin spider genus Opopaea Simon is a pantropical genus with biodiversity hotspots in Africa, Asia and Australia. We revise the Australian and Pacific species of the genus, provide redescriptions of the Australian species O. banksi (Hickman) and the Micronesian species O. foveolata Roewer, and new records of the pantropical O. deserticola Simon and O. concolor (Blackwall), as well as O. apicalis (Simon) which is newly transferred from Epectris, after the new synonymy of Epectris with Opopaea. The following species are provisionally transferred from Epectris to Opopaea, pending investigations into their generic affinities: O. conujaingensis (Xu), new combination from China; and O. mollis (Simon), new combination from Sri Lanka. Most Pacific Islands are inhabited by the four above-mentioned species but the following 15 newly described species are most likely native to the islands: from Fiji (O. fiji), Hawaii (O. hawaii), Palau (O. palau), New Caledonia (O. amieu, O. bicolor, O. burwelli, O. calcaris, O. goloboffi, O. monteithi, O. ndoua, O. platnicki, O. raveni, O. striata, O. touho, O. tuberculata). We treat the Australian Opopaea fauna and recognise 84 species including 71 new and 13 previously described species. The new Australian species include 21 species from New South Wales (O. acuminata, O. addsae, O. bushblitz, O. gerstmeieri, O. lebretoni, O. linea (also occurs in Queensland), O. magna, O. margaretehoffmannae, O. martini, O. michaeli, O. milledgei, O. nitens, O. ottoi, O. plana, O. simplex, O. sturt, O. suelewisae, O. sylvestrella, O. tenuis, O. ursulae, O. yorki); six from Northern Territory (O. ephemera, O. fishriver, O. gilliesi, O. johardingae, O. preecei, O. wongalara); 13 from Queensland (O. ameyi, O. brisbanensis, O. broadwater, O. carnarvon, O. carteri, O. chrisconwayi, O. douglasi, O. lambkinae, O. leichhardti, O. mcleani, O. proserpine, O. stanisici, O. ulrichi); three from South Australia (O. millbrook, O. mundy, O. stevensi); and 28 from Western Australia (O. aculeata, O. aurantiaca, O. billroth, O. callani, O. cowra, O. durranti, O. exoculata, O. flava, O. fragilis, O. framenaui, O. gracilis, O. gracillima, O. harmsi, O. johannae, O. julianneae, O. marangaroo, O. millstream, O. nadineae, O. pallida, O. pannawonica, O. pilbara, O. rixi, O. robusta, O. rugosa, O. subtilis, O. triangularis, O. wheelarra, O. whim). New records are provided for O. sown Baehr. Seven area-based keys to species are provided.
APA, Harvard, Vancouver, ISO, and other styles
35

Mast, Austin R., and Kevin Thiele. "The transfer of Dryandra R.Br. to Banksia L.f. (Proteaceae)." Australian Systematic Botany 20, no. 1 (2007): 63. http://dx.doi.org/10.1071/sb06016.

Full text
Abstract:
Phylogenies inferred from both chloroplast and nuclear DNA regions have placed the south-west Australian genus Dryandra R. Br. (93 spp.) among the descendents of the most recent common ancestor of the more widespread Australian genus Banksia L.f. (80 spp.). Here we consider the alternative solutions to maintaining monophyly at the generic rank and choose to make new combinations and replacement names for Dryandra in Banksia. We make the new combination Banksia ser. Dryandra in Banksia subgen. Banksia for 108 of the 109 new combinations at the ranks of species, subspecies, and variety and all 18 of the replacement names. We treat Banksia subgen. Banksia as the most inclusive clade that includes the type of Banksia (B. serrata) but not B. integrifolia. We erect Banksia subgen. Spathulatae to accommodate the species in the most inclusive clade that includes B. integrifolia but not B. serrata. These two subgenera of Banksia are equivalent to the clades informally called /Cryptostomata and /Phanerostomata elsewhere. We treat one of the new combinations, Banksia subulata, as incertae sedis within Banksia subgen. Banksia.
APA, Harvard, Vancouver, ISO, and other styles
36

Scalmer, Sean. "New South Wales." Australian Journal of Politics & History 50, no. 2 (June 2004): 257–64. http://dx.doi.org/10.1111/j.1467-8497.2004.247_2.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Thompson, Elaine. "New South Wales." Australian Cultural History 27, no. 2 (October 2009): 135–42. http://dx.doi.org/10.1080/07288430903164827.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Jones, P. A. "New South Wales." Australian Endodontic Newsletter 14, no. 2 (February 11, 2010): 6–7. http://dx.doi.org/10.1111/j.1747-4477.1988.tb00782.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Ferguson, Lorraine. "New South Wales." Australian Critical Care 5, no. 2 (June 1992): 23. http://dx.doi.org/10.1016/s1036-7314(92)70046-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Ferguson, Lorraine. "New South Wales." Australian Critical Care 5, no. 3 (September 1992): 23–24. http://dx.doi.org/10.1016/s1036-7314(92)70057-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Ferguson, Lorraine. "New South Wales." Australian Critical Care 5, no. 4 (December 1992): 25–26. http://dx.doi.org/10.1016/s1036-7314(92)70070-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Robertson, Sally. "New South Wales." Australian Critical Care 6, no. 1 (March 1993): 33. http://dx.doi.org/10.1016/s1036-7314(93)70101-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Robertson, Sally. "New South Wales." Australian Critical Care 6, no. 2 (June 1993): 34. http://dx.doi.org/10.1016/s1036-7314(93)70121-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Ferguson, Lorraine. "New South Wales." Australian Critical Care 6, no. 3 (September 1993): 33–34. http://dx.doi.org/10.1016/s1036-7314(93)70156-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Robertson, Sally. "New South Wales." Australian Critical Care 6, no. 4 (December 1993): 30. http://dx.doi.org/10.1016/s1036-7314(93)70180-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

"NEW SOUTH WALES." Australian Journal of Politics & History 3, no. 1 (April 7, 2008): 99–101. http://dx.doi.org/10.1111/j.1467-8497.1957.tb00371.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

"NEW SOUTH WALES." Australian Journal of Politics & History 3, no. 2 (April 7, 2008): 231–34. http://dx.doi.org/10.1111/j.1467-8497.1958.tb00386.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

"NEW SOUTH WALES." Australian Journal of Politics & History 4, no. 2 (April 7, 2008): 247–50. http://dx.doi.org/10.1111/j.1467-8497.1958.tb00402.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

"NEW SOUTH WALES." Australian Journal of Politics & History 10, no. 1 (April 7, 2008): 104–6. http://dx.doi.org/10.1111/j.1467-8497.1964.tb00736.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

"NEW SOUTH WALES." Australian Journal of Politics & History 10, no. 2 (April 7, 2008): 229–31. http://dx.doi.org/10.1111/j.1467-8497.1964.tb00752.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography