Journal articles on the topic 'Eastern Banksia'

To see the other types of publications on this topic, follow the link: Eastern Banksia.

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 'Eastern Banksia.'

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

USHER, A. V., D. J. AYRE, and R. J. WHELAN. "Microsatellites for eastern Australian Banksia species." Molecular Ecology Notes 5, no. 4 (December 2005): 821–23. http://dx.doi.org/10.1111/j.1471-8286.2005.01075.x.

Full text
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

Ward, SJ. "Life-History of the Eastern Pygmy-Possum, Cercartetus-Nanus (Burramyidae, Marsupialia), in South-Eastern Australia." Australian Journal of Zoology 38, no. 3 (1990): 287. http://dx.doi.org/10.1071/zo9900287.

Full text
Abstract:
Populations of Cercartetus nanus were investigated in three areas of Victoria: two areas of Banksia woodland at Wilsons Promontory National Park and an area of mixed eucalypt forest with an under- storey of B. spinulosa at Nar Nar Goon North, east of Melbourne. Most births occurred between November and March, but in areas where the dominant Banksia sp. flowered in winter they took place year-round. Most females produced two litters in a year, but some produced three. Males were reproductively active throughout the year. Litter sizes ranged from two to six, with a modal size of four. Pouch life lasted 30 days and weaning occurred at 65 days. Growth was rapid, young became independent immediately after weaning, and matured as early as 4.5-5.0 months old. Maximum longevity in the field was at least 4 years.
APA, Harvard, Vancouver, ISO, and other styles
4

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
5

Perkins, Ian, John Diamond, Georgina SanRoque, Lyn Raffan, Bettina Digby, Peter Jensen, and Daniel Hirschfeld. "Eastern Suburbs Banksia Scrub: Rescuing an endangered ecological community." Ecological Management & Restoration 13, no. 3 (September 2012): 224–37. http://dx.doi.org/10.1111/emr.12002.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Mast, Austin R. "Molecular systematics of subtribe Banksiinae (Banksia and Dryandra; Proteaceae) Based on cpDNA and nrDNA sequence data: Implications for taxonomy and biogeography." Australian Systematic Botany 11, no. 4 (1998): 321. http://dx.doi.org/10.1071/sb97026.

Full text
Abstract:
Despite considerable research interest in the subtribe Banksiinae (Banksia L.f. and Dryandra R.Br.), no strongly supported phylogenetic hypothesis for the relationship between the genera exists, nor have molecular characters been sampled for phylogenetic reconstruction at any level. In this study, DNA sequence characters were sampled from chloroplast DNA (cpDNA; the trnL intron, the trnL 3′ exon, and the spacer between the trnL 3′ exon and trnF) and nuclear ribosomal DNA (nrDNA; both internal transcribed spacers) of 18 species of Banksia and five of Dryandra, with six outgroup taxa from the subfamily Grevilleoideae. The molecular characters provided the opportunity to code taxa outside of Banksia for cladistic comparison with the genus—an opportunity not previously provided by morphological characters. Cladistic analyses, using parsimony, explored the effects of various weightings of transition to transversion events and base substitution to insertion and deletion events to determine which relationships in the cladograms were robust. The trnL/trnF and ITS characters strongly supported a paraphyletic Banksia with respect to a monophyletic Dryandra. The molecular results supported a single root for Thiele and Ladiges’(1996) unrooted morphological cladogram along the branch between the Isotylis to B. fuscolutea clade and the Grandes to B. tricuspis clade. George’s (1981) subgenus Banksia and section Banksia appeared dramatically non-monophyletic. The distribution of eastern taxa at derived positions on the molecular cladograms suggested considerable cladogenesis in the the genus prior to the formation of the Nullarbor Plain during the Tertiary.
APA, Harvard, Vancouver, ISO, and other styles
7

Mccredie, TA, KW Dixon, and K. Sivasithamparam. "Variability in the Resistance of Banksia L.f. Species to Phytophthora cinnamomi Rands." Australian Journal of Botany 33, no. 6 (1985): 629. http://dx.doi.org/10.1071/bt9850629.

Full text
Abstract:
Resistance of Banksia species to Phytophthora cinnamomi was determined under plantation conditions for 39 Western Australian and 10 eastern Australian Banksia spp. Plants were inoculated twice using millet seed inoculum, at the start of the study and 95 days later. To test for intraspecific variants or escape from disease, surviving individuals were stem-inoculated with P. cinnamomi. Infection in plant tissues was confirmed by reisolation with selective media. Horticulturally exploited species including B. hookerana, B. coccinea, B. prionotes, B. occidentalis, B. baxteri, B. sceptrum, B. speciosa, B. grandis, B. menziesii and B. victoriae were found to be suscep- tible to P. cinnamomi. Eight eastern Australian species showed resistance (95% survival). Seven West- ern Australian species including all prostrate species tested were found to have low susceptibility. There was a poor correlation between levels of susceptibility and taxonomic series in the genus, particularly in eastern and western component species of the pan-Australian series Orthostylis and Spicigerae. Banksia spp. from uniform or summer maximum rainfall regions were resistant or of low susceptibility while those from dominantly winter rainfall areas, especially with free draining soils, were highly sus- ceptible. Stem inoculation confirmed that one plant each of B. coccinea and B. hookerana was resistant among the P. cinnamomi-susceptible, horticulturally exploited species. Thus stem inoculation proved a useful diagnostic tool, and girdling rather than longitudinal fungus growth through stem tissue is more appropriate as a measure of species' resistance.
APA, Harvard, Vancouver, ISO, and other styles
8

Law, Bradley, Mark Chidel, Alf Britton, and Caragh Threlfall. "Comparison of microhabitat use in young regrowth and unlogged forest by the eastern pygmy-possum (Cercartetus nanus)." Australian Mammalogy 40, no. 1 (2018): 1. http://dx.doi.org/10.1071/am16041.

Full text
Abstract:
We describe microhabitat use and response to disturbance by the eastern pygmy-possum (Cercartetus nanus) in heathy dry sclerophyll forest using spool-and-line-tracking. We compared unlogged forest with forest regenerating four years after selective logging. Structural and floristic attributes were scored along spool lines and compared with a random line for each possum. We found that possums (n = 23) selected areas based on both structural and floristic attributes. Possums selected dense understorey, especially that comprising flowering hairpin banksia (Banksia spinulosa) and Gymea lily (Doryanthes excelsa). Fallen logs were not selected by nocturnally active possums. Spool lines in regrowth forest had less eucalypt cover and more understorey cover (e.g. D. excelsa and B. spinulosa) than unlogged forest. Conversely, cover of Banksia serrata was less in regrowth than unlogged forest. Spool lines were commonly found both at ground level (mean = 52–57% of lengths) and above the ground (43–48% of lengths). There was no difference in the mean spool height between the logging treatments (regrowth: 0.47 ± 0.14 m; unlogged: 0.49 ± 0.10 m; ± s.e.). Overall, our results suggest that the dense, flowering understorey that regenerates after selective logging is suitable for use and is the primary attribute selected by active pygmy-possums.
APA, Harvard, Vancouver, ISO, and other styles
9

Thiele, K., and PY Ladiges. "A cladistic analysis of Banksia (Proteaceae)." Australian Systematic Botany 9, no. 5 (1996): 661. http://dx.doi.org/10.1071/sb9960661.

Full text
Abstract:
Banksia is a genus of more than 90 taxa, many of which are common and characteristic in sclerophyll communities in eastern and south-western Australia. Cladistic analyses based on morphological and anatomical characters were used to resolve relationships in the genus. An initial analysis of 35 terminal taxa, including 9 infrageneric taxa assumed to be monophyletic on the basis of one or more synapomorphies, allowed resolution of basal nodes. Subsequent analyses of the putatively monophyletic infrageneric taxa allowed resolution of distal nodes. Some of these lower-level analyses used a mixture of qualitative characters and coded morphometric characters. Together, the analyses afforded a high degree of resolution within the genus, although relationships of some taxa were not well supported. A new infrageneric classification, in which Banksia is divided into 2 subgenera, 12 series and 11 subseries, is proposed. The classification is broadly similar to previously published classifications of the genus, but discards a number of taxa shown to be para- or poly-phyletic. The following new names are published: Banksia series Lindleyanae K.Thiele, series Ochraceae K.Thiele, subseries Leptophyllae K.Thiele, subseries Longistyles K.Thiele, subseries Nutantes K.Thiele, subseries Sphaerocarpae K.Thiele, subseries Cratistyles K.Thiele, subseries Acclives K.Thiele, subseries Integrifoliae K.Thiele, subseries Ericifoliae K.Thiele, subseries Occidentales K.Thiele and subseries Spinulosae K.Thiele. New combinations are provided for Banksia penicillata (A.S.George) K.Thiele, B. brevidentata (A.S.George) K.Thiele, B. hiemalis (A.S.George) K.Thiele and B. dolichostyla (A.S.George) K.Thiele.
APA, Harvard, Vancouver, ISO, and other styles
10

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
11

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's requirements, its body size and on the quantity of pollen that is available.
APA, Harvard, Vancouver, ISO, and other styles
12

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
13

Wills, Timothy J. "Using Banksia (Proteaceae) node counts to estimate time since fire." Australian Journal of Botany 51, no. 3 (2003): 239. http://dx.doi.org/10.1071/bt01074.

Full text
Abstract:
In Australia, numerous methods have been used to determine the time since the last fire at a given site. One method involves counting the number of annual growth nodes on Banksia spp. that are either killed by fire, or regenerate from surviving rootstocks, to determine above-ground plant age. Although a number of studies have used the Banksia node-count method to estimate plant and therefore site age, no published data assess the reliability of this method. This study attempted to determine the accuracy of the method, with shrub-form B. marginata individuals from five sites of different age, in a south-eastern Australian sand heath. A significant relationship was found between modal node (internode) counts and the known time since fire, with internode counts at four of the five sites accurate to within 1 year of actual site age up to 21 years (n > 50 individuals per site). The results suggest that the Banksia node-count method is a useful tool for determining site age up to 21 years. However, sample sizes need to be appropriate for the area sampled, given the potential error in counting nodes and the inherent site variability in age classes.
APA, Harvard, Vancouver, ISO, and other styles
14

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
15

Evans, KM, and A. Bunce. "A comparison of the foraging behaviour of the eastern pygmy-possum (Cercartetus nanus) and nectarivorous birds in a Banksia integrifolia woodland." Australian Mammalogy 22, no. 1 (2000): 81. http://dx.doi.org/10.1071/am00081.

Full text
Abstract:
The foraging behaviour of a non-flying mammal, the eastern pygmy-possum (Cercartetus nanus) and nectarivorous birds was compared in a Banksia integrifolia woodland at Wilson's Promontory National Park, Victoria, Australia. Exclusion experiments performed previously in this woodland indicate that both non-flying mammals and nectarivorous birds are important pollinators of B. integrifolia (Cunningham 1991: Oecologia 87: 86-90). In this study it is shown that C. nanus and nectarivorous birds employ different foraging tactics. Nectarivorous birds tended to move further between trees (Χ = 8.16 ± 1.06 m) than C. nanus (Χ = 5.64 ± 0.75 m), although these differences were not significant. Nectarivorous birds were attracted to trees with a significantly larger number of inflorescences (Χ = 36.55 ± 2.84) than C. nanus (Χ = 18.65 ± 2.95), and visited a significantly greater number of inflorescences per tree (Χ = 4.24 ± 0.35) than C. nanus (Χ = 2.33 ± 0.22). Although the two pollinator groups were attracted to banksia plants by different cues, once in the plants they visited an equal proportion of the available inflorescences.
APA, Harvard, Vancouver, ISO, and other styles
16

Goldingay, RL, SM Carthew, and RJ Whelan. "Transfer of Banksia-Spinulosa Pollen by Mammals - Implications for Pollination." Australian Journal of Zoology 35, no. 4 (1987): 319. http://dx.doi.org/10.1071/zo9870319.

Full text
Abstract:
Native mammals have been implicated by various authors as visitors to flowers of Australian plants in both eastern and western Australia, but few data are available to allow an estimation of their potential as pollinators. In the present study, Antechinus stuartii, Petaurus breviceps and Rattus fuscipes were regularly trapped in flowering Banksia spinulosa. A few Cercartetus nanus were also captured. Individuals of all species carried pollen on their fur. Pollen loads were greater on mammals which had been in traps for short periods, which suggests that mammals will groom pollen from their fur if left longer in traps. Therefore, pollen loads on foraging mammals have probably been dramatically underestimated by previous authors, and their potential to effect pollination may have been greatly underestimated. Furthermore, significantly more pollen was removed from flowers of B. spinulosa at night than during the day, suggesting the importance of nocturnal pollinators at this site.
APA, Harvard, Vancouver, ISO, and other styles
17

Lambert, Geoff, and Judy Lambert. "Progress with restoration and management of Eastern Suburbs Banksia Scrub on North Head, Sydney." Ecological Management & Restoration 16, no. 2 (May 2015): 95–105. http://dx.doi.org/10.1111/emr.12160.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Blair, John, and Paul Osmond. "Employing Green Roofs to Support Endangered Plant Species: The Eastern Suburbs Banksia Scrub in Australia." Open Journal of Ecology 10, no. 03 (2020): 111–40. http://dx.doi.org/10.4236/oje.2020.103009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

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
20

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 >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
21

Goldingay, Ross L., and Jo Keohan. "Population density of the eastern pygmy-possum in a heath–woodland habitat." Australian Journal of Zoology 65, no. 6 (2017): 391. http://dx.doi.org/10.1071/zo18026.

Full text
Abstract:
The eastern pygmy-possum (Cercartetus nanus) has posed a challenge in attempts to describe its population density due to low rates of capture, preference for patchy habitats and periodic influxes of subadult individuals. We conducted a mark–recapture study of this species using a grid of nest boxes in a 9-ha patch of banksia heath–woodland. We captured 54 adults across the two years of our study. We estimated the density of adult pygmy-possums to be 1.5–4.2 ha–1 from different population models. This is substantially lower than previous estimates in equivalent habitat because we focussed on adults and recognised that they were not confined to the area bounded by our grid. We captured 36 subadults over the two years but they could not be reliably modelled due to extremely low recapture rates, which reflect high rates of dispersal and also mortality. For this reason, only the number of adults should be used to characterise populations of this species. Further study is required to investigate population dynamics over time and to describe the density of eastern pygmy-possums in other habitats.
APA, Harvard, Vancouver, ISO, and other styles
22

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
23

Stimpson, Margaret L., Peter H. Weston, Ralph (Wal) D. B. Whalley, and Jeremy J. Bruhl. "A morphometric analysis of the Banksia spinulosa complex (Proteaceae) and its complex taxonomic implications." Australian Systematic Botany 29, no. 1 (2016): 55. http://dx.doi.org/10.1071/sb15030.

Full text
Abstract:
Specimens of all known taxa and putative entities belonging to the Banksia spinulosa complex were collected from Kuranda in northern Queensland, western to central Queensland and down the eastern coast of Australia to Wilsons Promontory in southern Victoria. These specimens were used to investigate morphological variation in habit, stems, leaves, inflorescences, fruits and seeds in the complex. Phenetic analysis (unweighted pair-group method with arithmetic mean, UPGMA, clustering and semi-strong hybrid multi-dimensional scaling, SSH–MDS, ordination) was performed on the full dataset of 233 entities using 33 characters (18 quantitative, two binary and 13 multistate). To facilitate visualisation of patterns in both clustering and ordination, we also analysed subgroups based on the results of the phenogram from the full dataset. The results showed that the five known and described taxa are phenetically distinct, viz. B. collina sens. str., B. cunninghamii sens. str., B. neoanglica, B. spinulosa and B. vincentia, and provided support for a further 12 morphometrically diagnosable entities, four of which could not be diagnosed with simple combinations of character states and require further investigation. The present study has highlighted that there is much more hidden morphological diversity in the B. spinulosa complex than has previously been recognised in any of the current competing taxonomies.
APA, Harvard, Vancouver, ISO, and other styles
24

Bell, Tina, Kevin Tolhurst, and Michael Wouters. "Effects of the fire retardant Phos-Chek on vegetation in eastern Australian heathlands." International Journal of Wildland Fire 14, no. 2 (2005): 199. http://dx.doi.org/10.1071/wf04024.

Full text
Abstract:
The effects of the fire retardant Phos-Chek D75R on species composition, survival and growth of eastern Australian heathland vegetation are described. Two sites in Victoria were selected, Victoria Valley in the Grampian Ranges and Marlo in East Gippsland. Both areas supported heathland vegetation that was long, unburnt and relatively undisturbed. Plots were subjected to single applications of increasing concentrations of retardant (0.5, 1.0 and 1.5 L fire retardant m−2) or no additional fire retardant (‘Control’ and ‘Water only’ treatments). A single application of Phos-Chek did not appear to significantly change species composition or projected foliage cover of the major life forms of native heathland vegetation. However, it did cause whole plant and shoot death of key species Allocasuarina paludosa, Banksia marginata, Leptospermum continentale and L. myrsinoides, and was observed to affect other species. The fertilising effect of the fire retardant generally increased shoot growth of the key species but did not significantly increase the overall height of these species. The application of fire retardant enhanced weed invasion, particularly when supplied at higher concentrations. A number of research recommendations are made from this preliminary investigation.
APA, Harvard, Vancouver, ISO, and other styles
25

Broadhurst, Linda, David Bush, and Jim Begley. "Managing Genetic Diversity and Representation in Banksia marginata (Proteaceae) Seed Production Areas Used for Conservation and Restoration." Diversity 13, no. 2 (January 21, 2021): 39. http://dx.doi.org/10.3390/d13020039.

Full text
Abstract:
Landscape degradation is a major threat to global biodiversity that is being further exacerbated by climate change. Halting or reversing biodiversity decline using seed-based restoration requires tons of seed, most of which is sourced from wild populations. However, in regions where restoration is most urgent, wild seed sources are often fragmented, declining and producing seed with low genetic diversity. Seed production areas (SPAs) can help to reduce the burden of collecting native seed from remnant vegetation, improve genetic diversity in managed seed crops and contribute to species conservation. Banksia marginata (Proteaceae) is a key restoration species in south-eastern Australia but is highly fragmented and declining across much of its range. We evaluated genetic diversity, population genetic structure and relatedness in two B. marginata SPAs and the wild populations from which the SPA germplasm was sourced. We found high levels of relatedness within most remnants and that the population genetic structure was best described by three groups of trees. We suggest that SPAs are likely to be important to meet future native seed demand but that best practice protocols are required to assist land managers design and manage these resources including genetic analyses to guide the selection of germplasm.
APA, Harvard, Vancouver, ISO, and other styles
26

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
27

Groves, RH, PJ Hocking, and A. Mcmahon. "Distribution of Biomass, Nitrogen, Phosphorus and Other Nutrients in Banksia marginata and B. ornata Shoots of Different Ages After Fire." Australian Journal of Botany 34, no. 6 (1986): 709. http://dx.doi.org/10.1071/bt9860709.

Full text
Abstract:
The heathland form of Banksia marginata Cav. regenerates rarely from seed but commonly by resprout- ing from buds on lateral roots, whereas Banksia ornata F. Muell. regenerates only from seed, usually released after fire. The two species co-occur in heath vegetation on nutrient-poor soils in south-eastern South Australia and western Victoria. Shoots were sampled from stands of B. marginata aged from 1 to 25 years and of B. ornata aged from 1 to 50+ years after fire in the Little Desert National Park, western Victoria. B. marginata, the resprouter, distributed a greater proportion of the total shoot dry matter and content of all nutrients to vegetative growth over its shorter life span than B. ornata, the non-sprouter. About 50% of the total phosphorus in B. ornata shoots at 50+ years was present in cones (including seeds) compared with only about 20% in B. marginata shoots at a comparable stage of senescence (25 years). This difference between the species was also true to a lesser degree for nitrogen. There were considerable differences between other nutrients in their distribution patterns in shoots. Nutrients could be grouped together on the basis of distribution in shoots more satisfactorily than on presumed physio- logical roles. Stems were major sites of nutrient accumulation in both species. The content of a particular nutrient in seeds as a proportion of the content in the living parts of the shoot ranged from 0.03% (Na, Mn) to 2.0% (P) in B. marginata, and from 0.3% (Na) to as high as 31% (P) in B. ornata. Concen- trations of all nutrients except sodium were much higher in seeds than in the woody cones or vegetative organs of both species; seeds of B. ornata were particularly rich in calcium and manganese. We conclude that the different patterns of distribution of biomass and nutrients, especially nitrogen and phosphorus, within shoots of the two species reflect their different regenerative modes after fire. Introduction Phosphorus and, to a lesser extent, nitrogen limit the growth of sclerophyllous shrubs on nutrient-poor soils in southern Australia
APA, Harvard, Vancouver, ISO, and other styles
28

Muir, Annette M., Peter A. Vesk, and Graham Hepworth. "Reproductive trajectories over decadal time-spans after fire for eight obligate-seeder shrub species in south-eastern Australia." Australian Journal of Botany 62, no. 5 (2014): 369. http://dx.doi.org/10.1071/bt14117.

Full text
Abstract:
Intervals between fires are critical for the persistence of obligate-seeding shrubs, and are often used in planning fires for fuel reduction and biodiversity conservation in fire-prone ecosystems worldwide. Yet information about the trajectories of reproductive performance for such species is limited and information is often qualitative. To test existing assumptions about reproductive maturity periods for eight obligate-seeding shrubs (with both canopy and soil seedbanks) in foothill forests of south-eastern Australia, we used a chronosequence approach, with sites from 2 years to >40 years post-fire. Quantitative measurements of flowering and fruiting were used to fit models of reproductive response in relation to time-since-fire for each species. Inferred reproductive maturity for each species, based on modelled times to reach 80% of maximum flower production, varied from 5 to 18 years post-fire. For a subset of three species, models predicted 80% maximum seed production occurring 1–7 years later than flowering. Our results confirmed or extended assumptions about post-fire reproductive maturity for these species, and provided a basis for improved incorporation of plant life-history in ecological fire planning. We infer that increased fire frequency makes one of our study taxa, Banksia spinulosa var. cunninghamii (Sieber ex Rchb.) A.S.George, vulnerable to decline because of its long reproductive maturity period and serotinous seed storage.
APA, Harvard, Vancouver, ISO, and other styles
29

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
30

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
31

Turner, James M. "The interrelationship between torpor expression and nest site use of western and eastern pygmy-possums (Cercartetus spp.)." Australian Mammalogy 42, no. 1 (2020): 85. http://dx.doi.org/10.1071/am19005.

Full text
Abstract:
Physiology and behaviour are closely linked, making knowledge of the interaction between species’ energetics and activities important when attempting to understand how animals function in the wild. I examined torpor use by western pygmy-possums (Cercartetus concinnus) and eastern pygmy-possums (C. nanus) in relation to nest site characteristics and movement patterns. In coastal mallee heath in winter, C. concinnus nested beneath leaf litter at the base of dead Banksia ornata, where they employed torpor on 69% of observed days. In warm temperate sclerophyll forest, C. nanus nested in tree hollows of Eucalyptus spp. and used torpor on 64% of days in winter and 10% in summer. Torpor was used in nest sites that were buffered from outside temperature extremes. Both species frequently reused nest sites and while C. nanus was more likely to employ torpor in a previously used site, site familiarity did not influence torpor use for C. concinnus. Additionally, C. nanus was more likely to use torpor in hollows with a higher relative thickness in both seasons. No relationship was found between range size and the number of tracking days or capture body mass, though sample sizes were small. I suggest that the thermal attributes of nest sites influence torpor use for both species and this is likely vital for maintaining a positive energy balance, stressing the importance of preserving habitat with ample potential nest sites for conservation management.
APA, Harvard, Vancouver, ISO, and other styles
32

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
33

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
34

Bennett, LT. "The Expansion of Leptospermum laevigatum on the Yanakie Isthmus, Wilson's Promontory, Under Changes in the Burning and Grazing Regimes." Australian Journal of Botany 42, no. 5 (1994): 555. http://dx.doi.org/10.1071/bt9940555.

Full text
Abstract:
The distribution of selected vegetation types on the Yanakie Isthmus, Wilson's Promontory National Park, was mapped from aerial photographs from 1941, 1972 and 1987. The main changes in the vegetation dynamics were: (1) an expansion of Leptospevmum laevigatum into grasslands and into Banksia integrifolia woodlands with herbaceous understoreys, and (2) a stabilisation of dunes by shmbs dominated by Leptospennum laevigatum. The total area of L. laevigatum shrubland and scrub increased from 2179 ha in 1941 to 3436 ha in 1972 and 4516 ha in 1987. Land-use changes in this period included the exclusion of fire in the early 1970s, after a history of regular burning, and an increase in grazing pressure primarily due to population expansions of the rabbit and the eastern grey kangaroo. Fire was not a prerequisite of the L. laevigatum expansion on the Isthmus because the spread continued after fire was excluded; nor was fire the primary cause of the expansion because the percentage yearly increase in the area of L. laevigatum was, on average, similar before and after 1972. An increase in grazing pressure was identified as the probable cause of the L. laevigatum expansion due to: (1) the exposure of bare ground, and (2) the restriction of the feeding range of cattle (known to graze both L. laevigatum and Acacia sophorae on the Isthmus).
APA, Harvard, Vancouver, ISO, and other styles
35

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
36

McLay, Todd G. B., Michael J. Bayly, and Pauline Y. Ladiges. "Is south-western Western Australia a centre of origin for eastern Australian taxa or is the centre an artefact of a method of analysis? A comment on Hakea and its supposed divergence over the past 12 million years." Australian Systematic Botany 29, no. 2 (2016): 87. http://dx.doi.org/10.1071/sb16024.

Full text
Abstract:
Lamont et al. (2016) concluded that the Australian sclerophyllous genus Hakea (Proteaceae) arose 18million years ago in the South West of Western Australia (SWA) and dispersed 18 times to eastern (EA) and central Australia (CA) only 12million years ago (mid-Miocene). Their explanation of the biogeographic history of Hakea was based on the following: accepting a fully resolved molecular phylogenetic tree, although ~40% of nodes had posterior probability values below 0.95; using all nodes including geographically paralogous nodes to determine ancestral area probabilities; and applying a strict clock to estimate clade divergence times. Our re-analyses of the same dataset using a relaxed clock model pushes the age of Hakea to 32.4 (21.8–43.7) million years ago relative to its nearest outgroups, and the age of the divergence of two major clades (A and B) to 24.7 (17.2–33.7) million years ago. Calibration based on a new finding of Late Cretaceous fossil Banksia pushes these dates to 48.0 (24.3–75.2) million years ago and 36.6 (18.5–55.4) million years ago respectively. We confirm that each of the two main clades includes lineages in SWA, CA and EA. At the basal node of Clade A, two eastern Australian species form the sister group to three SWA scrub–heath–Eremaean species. These two groups together are sister to a large, mostly unresolved clade of SWA, CA and EA taxa. Similarly, at the base of Clade B is a polytomy of lineages from the SWA, CA and EA, with no resolution of area relationships. There is no evidence of a centre of origin and diversification of the genus is older than the mid-Miocene, being at least Oligocene, and probably older, although calibration points for molecular dating are too far removed from the ingroup to provide any great confidence in the methodology. Consideration should be given to the possibility of vicariance of multiple, widespread ancestral lineages as an explanation for lineages now disjunct between EA and SWA.
APA, Harvard, Vancouver, ISO, and other styles
37

Burgin, Shelley, Danny Wotherspoon, Dennis John Hitchen, and Peter Ridgeway. "Habitat use by the jacky lizard Amphibolurus muricatus in a highly degraded urban area." Animal Biology 61, no. 2 (2011): 185–97. http://dx.doi.org/10.1163/157075511x566515.

Full text
Abstract:
AbstractOver time native vegetation remnants in urban areas are typically eroded in size and number due to pressures from urban expansion and consolidation. Such remnants, frequently neglected and invaded by weeds, may constitute the last remaining habitat for some species' populations in urban areas. In the restoration of remnants for biodiversity, weed removal is often a high priority but there is a dearth of information on the role that exotic vegetation plays as habitat for fauna such as small reptiles. We investigated the vegetation type preference of urban remnants at the edge of a Sydney golf course by Amphibolurus muricatus, the native jacky lizard. The three vegetation types present were Eastern Suburbs Banksia Scrub (an Endangered Ecological Community) with sparse groundcover, dense stands of the introduced Eragrostis curvula African love grass, and open fairways of lawn: three structurally different habitats. Captured jacky lizards were spooled and their movements traced by following the thread left as they moved through their home range. Jacky lizards preferred areas that afford them most cover. While they foraged throughout the stands of love grass, they tended to avoid the edge of native vegetation remnants. They also basked on the lawn close to the vegetation where they had recently foraged, or traversed it to enter natural vegetation or grass. We concluded that introduced love grass offered additional habitat because of the relatively dense vegetation cover, and that areas should not be managed with the assumption that invasive weeds are detrimental to native species without appropriate assessment.
APA, Harvard, Vancouver, ISO, and other styles
38

Shearer, BL, and JT Tippett. "Distribution and Impact of Armillaria luteobubalina in the Eucalyptus marginata Forest of South-Western Australia." Australian Journal of Botany 36, no. 4 (1988): 433. http://dx.doi.org/10.1071/bt9880433.

Full text
Abstract:
Armillaria luteobubalina is a widespread primary pathogen in the Eucalyptus marginata forest of south-western Australia. Over 200 infection centres were identified during the 5-year period between 1981 and 1985. The fungus sporulated during June and July, usually from roots but sometimes from stems (e.g. E. calophylla). Armillaria luteobubalina basidiomes were found originating from roots of 34 plant species, with greatest incidence on roots of E. marginata. Root systems were excavated and patterns of A . Luteobubalina invasion recorded. Rhizomorphs were not found and fungal spread between hosts was via root to root contacts, Variation in host species' susceptibility to the fungus was reflected in different patterns of xylem compartmentalisation and variable amounts of cambial damage. The degree of resistance expressed at the collar or lower stem determined the fate of individuals of the various species. Lack of resistance in Eucalyptus wandoo to tangential spread of A. luteobubalina often resulted in death by the time columns of decay had advanced into the lower stem or butt. Banksia grandis, E. calophylla, E. gomphocephala, and E. marginata resisted to varying degrees. Inverted V-shaped lesions, often mis- taken for fire scars, were evidence of the ability of E. gomphocephala and E. marginata individuals to resist tangential spread and prevent girdling of stems. In stems of E. calophylla, lesions did not have a definite V shape, decay penetrated deeper and the fungus persisted longer than in those of E. marginata. Host mortality following infection was greater in the intermediate- and low-rainfall zones of the eastern E. marginata forest than in the high-rainfall zone to the west.
APA, Harvard, Vancouver, ISO, and other styles
39

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
40

Kwiatkowska, Monika, Justyna Żabicka, Grzegorz Migdałek, Piotr Żabicki, Marlena Cubała, Jerzy Bohdanowicz, Aneta Słomka, et al. "Comprehensive characteristics and genetic diversity of the endemic Australian Viola banksii (section Erpetion, Violaceae)." Australian Journal of Botany 67, no. 2 (2019): 81. http://dx.doi.org/10.1071/bt18233.

Full text
Abstract:
Viola banksii, the type species of section Erpetion, is endemic in eastern mainland Australia. In this paper we characterise morphological and anatomical features and assess genome size and genetic diversity in combination with the breeding system. V. banksii develops exclusively chasmogamous flowers. Ovules are anatropous, crassinucellate and bitegmic, the female gametophyte is of the Polygonum type, and the embryo is of Asterad type surrounded by nuclear endosperm. Pollen is non-heteromorphic, 3-aperturate, and highly viable. V. banksii grows in shade on moist, well drained, often sandy soils, and this is reflected in the anatomy of its organs, which includes a lack of subepidermal collenchyma in aerial parts, large leaf epidermal cells with thin cell walls, a narrow cuticle layer, and vascular bundles with xylem that are not rich in vessels. V. banksii is tolerant to zinc and lead based on phytotoxicity test. The high chromosome number (2n = 10x = 50) does not correspond to a small genome size (2C DNA = 1.27 pg). Low mean intra-populational gene diversity (HS = 0.077) detected by ISSR markers confirms the strong influence of selfing and clonal propagation by pseudostolons. Unique morphological traits of V. banksii include nyctinastic petal movement, the lack of a floral spur, the presence of gland-like protuberances on two stamens, and the presence of pseudostolons, which could be a synapomorphy for the whole section.
APA, Harvard, Vancouver, ISO, and other styles
41

Salman, Alp, and F. Saadet Karakulak. "Cephalopods in the diet of albacore, Thunnus alalunga, from the eastern Mediterranean." Journal of the Marine Biological Association of the United Kingdom 89, no. 3 (September 19, 2008): 635–40. http://dx.doi.org/10.1017/s0025315408002555.

Full text
Abstract:
In this study, the stomach contents of 116 albacore specimens, Thunnus alalunga were examined from the eastern Mediterranean Sea. Fifty-five of the 116 stomachs examined were empty. The occurrence of major prey categories in stomachs were 95.1% cephalopods, 47.5% teleosts and 39.3% crustaceans with a total of 633 individuals belonging to 14 species identified. Heteroteuthis dispar from the order Sepiolida constituted 56.40% of the main cephalopod prey followed by Onychoteuthis banksii from the order Teuthida.
APA, Harvard, Vancouver, ISO, and other styles
42

HILL, ROBERT S., and DAVID C. CHRISTOPHEL. "Tertiary leaves of the tribe Banksieae (Proteaceae) from south-eastern Australia." Botanical Journal of the Linnean Society 97, no. 2 (June 1988): 205–27. http://dx.doi.org/10.1111/j.1095-8339.1988.tb02462.x.

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

Gemelli, F., C. R. Johnson, and J. T. Wright. "Gastropod communities associated with different morphologies of the intertidal seaweed Hormosira banksii." Marine and Freshwater Research 70, no. 2 (2019): 280. http://dx.doi.org/10.1071/mf18159.

Full text
Abstract:
Hormosira banksii is an important intertidal habitat-forming seaweed in southern Australia that shows large variation in morphology. We examined the relationship between morphological variation in Hormosira and associated gastropod community structure, abundance and diversity in Tasmania, southern Australia. We sampled both Hormosira and gastropods from sites in two habitats (coast and estuary), two times (February–March and October–November) at two regions (northern and eastern Tasmania). There were distinct Hormosira morphs on the north coast (small individuals), east coast (intermediate sized individuals) and in estuaries (large individuals). Multivariate analysis showed that gastropod communities varied among the different algal morphologies, and suggest an influence of morphological traits, specifically thallus length, vesicle number and size, on the distribution patterns of gastropod species. Despite the finding of distinct gastropod communities associated with the different Hormosira morphs, because the different morphs occur in different locations with different physical conditions, we cannot unequivocally attribute the differences in gastropod communities to seaweed morphology per se. Nonetheless, our results confirm that H. banksii is an important foundation species in the intertidal zone and suggest a habitat-specific effect of algal morphological traits on gastropods.
APA, Harvard, Vancouver, ISO, and other styles
44

Dede, Ayhan, Alp Salman, and Arda M. Tonay. "Stomach contents of by-caught striped dolphins (Stenella coeruleoalba) in the eastern Mediterranean Sea." Journal of the Marine Biological Association of the United Kingdom 96, no. 4 (September 28, 2015): 869–75. http://dx.doi.org/10.1017/s0025315415001538.

Full text
Abstract:
Stomach contents of six striped dolphins taken as by-catch in the swordfish fishery in the eastern Mediterranean Sea off the Turkish coast were examined. In total, 29 taxa were identified to species or family and 1777 individual food items (1394 bony fishes, 289 cephalopods, 94 crustaceans) were counted.Diaphusspp. andCeratoscopelus maderensiswere the most remarkable ones, as they accounted for 70.45% of the total number of fishes.Onychoteuthis banksii, on the other hand, was the only cephalopod species found in all stomach content analyses and represented 38.06% of the total number of cephalopods. Bony fish species:Myctophum punctatum, Notoscopelus elongatus, Electrona risso, Sudis hyalina,Moridae sp., Phycidae sp., Sternoptychidae sp. and cephalopods:Pterygioteuthis giardiandChtenopteryx siculawere reported the first time in the stomach contents of striped dolphin in the Mediterranean Sea.
APA, Harvard, Vancouver, ISO, and other styles
45

Weber, M. G., C. E. Van Wagner, and Monte Hummel. "Selected Parameters of Fire Behavior and Pinus banksiana Lamb. Regeneration in Eastern Ontario." Forestry Chronicle 63, no. 5 (October 1, 1987): 340–46. http://dx.doi.org/10.5558/tfc63340-5.

Full text
Abstract:
Fire behavior variables were quantified in eastern Ontario jack pine (Pinus banksiana Lamb.) ecosystems and used to interpret observed fire impacts and effects. A series of seven fires, ranging in frontal fire intensity from 70 to 17 000 W/m, were documented. Forest floor moisture content prior to burning was negatively correlated with weight of forest floor consumed per unit area (r2 = 0.97) and per cent mineral soil bared (r2 = 0.95). Frontal fire intensity was positively correlated with per cent tree mortality (r2 = 0.98) and mean height of char (r2 = 0.76). Frontal fire intensities of 17 000 kW/m resulted in seedling numbers of 30 000 to over 50 000 ha−1 considered to be more than adequate for establishing the next generation of crop trees. Jack pine mean seedling height, 13 to 16 years after fire, was also positively correlated with frontal fire intensity (r2 = 0.82), ranging from 0.5 to 3.8 m on lowest and highest intensity burns, respectively. Similar relationships were found when seedling height was regressed against per cent tree mortality (r2 = 0.62) and forest floor consumption (r2 = 0.79).Results are discussed in terms of ecological requirements of the species, particularly during the regeneration phase, and it is concluded that quantification of fire behavior observations is mandatory if burning conditions are to be understood and/or duplicated by the land manager for the attainment of a given forest management objective.
APA, Harvard, Vancouver, ISO, and other styles
46

Rochon, Caroline, David Paré, Nellia Pélardy, Damase P. Khasa, and J. André Fortin. "Ecology and productivity of Cantharellus cibarius var. roseocanus in two eastern Canadian jack pine stands." Botany 89, no. 10 (October 2011): 663–75. http://dx.doi.org/10.1139/b11-058.

Full text
Abstract:
Despite the economic importance of chanterelles, much remains to be known about their habitat requirements. Cantharellus cibarius var. roseocanus Redhead, Norvell & Danell sporocarp productivity was measured during three growing seasons in two Pinus banksiana Lamb. stands of boreal forest. The objective was to determine how the variability in stand, plant association, edaphic, and meteorological conditions was related to sporocarp productivity. DNA of this species was detected in organic and mineral soil horizons. Sporocarp productivity was similar for both stands, but the absence of colonies on trails at one of the sites likely reflects microenvironmental conditions that are unsuitable for chanterelle growth. Under the prevailing site conditions, preferred microhabitats were characterized by high stand density, high C:N ratio, and frequent moss presence. The Solidago puberula Nutt. – Comptonia peregrina (L.) Coulter – Pinus banksiana association, lichen presence, and as much clay and silt content as can possibly be found on this moderately acidic sandy soil favoured the productivity of this chanterelle, whereas ericaceous species presence was negatively correlated with chanterelle productivity. Positive correlations were found between total rainfall 1 week prior to fructification, air temperature 2 weeks prior to fructification, and sporocarp productivity. Results highlight the specific conditions favourable to Cantharellus cibarius var. roseocanus fructifications within these stands.
APA, Harvard, Vancouver, ISO, and other styles
47

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
48

Sharma, Mahadev, and S. Y. Zhang. "Stand Density Management Diagram for Jack Pine Stands in Eastern Canada." Northern Journal of Applied Forestry 24, no. 1 (March 1, 2007): 22–29. http://dx.doi.org/10.1093/njaf/24.1.22.

Full text
Abstract:
Abstract A stand density management diagram was developed for jack pine (Pinus banksiana Lamb.) stands using the data obtained from 125 permanent sample plots (PSPs) established in Ontario and 232 PSPs in Quebec, Canada. The diagram was evaluated using data from 40 PSPs established in Ontario. Recently developed and efficient models have been used in constructing the diagram to estimate diameters and heights for the trees for which no diameters or heights were recorded at the time of stand inventory. Relative density indices of 0.15, 0.40, and 0.55 were used, corresponding to the line of approximate crown closure, the limit of productive zone, and the lower limit of competition-related mortality, respectively. If two stand characteristics are known, including mean total tree volume, quadratic mean diameter, trees per hectare, and average dominant height, the others can be readily obtained using the diagram. The consequences of various thinning scenarios can be plotted and visualized in the field without the need for computer simulation.
APA, Harvard, Vancouver, ISO, and other styles
49

Duchesne, Luc C., and Renée Tellier. "Effect of site preparation on the nutrient content of non-crop vegetation in a jack pine clear-cut." Forestry Chronicle 73, no. 6 (December 1, 1997): 711–14. http://dx.doi.org/10.5558/tfc73711-6.

Full text
Abstract:
The nutrient (N, P, K, Mg, and Ca) content of the aboveground living non-crop vegetation of a jack pine (Pinus banksiana Lamb.) clear-cut in eastern Ontario was investigated for two years after site preparation, which consisted of prescribed burning under different fire intensities and disk trencher scarification. Total plant nutrient content generally followed biomass accumulation with higher levels of plant N, P, Ca, and K in clear-cuts and scarified sites than in burned-over sites. In the first growing season, concentrations of N, P, and K were higher in the vegetation of burned-over plots than in scarified and clear-cut plots. Mg concentrations were greater in burned-over and scarified plots than in the clear-cut plots. Ca concentrations did not differ among the treatments. Concentration of P and K and the total amount of N, P and K in aboveground non-crop vegetation were correlated well with fire intensity at the end of the first growing season whereas only K concentration and quantities were correlated to fire intensity within two years after treatment. Key words: prescribed burning; disk trenching scarification; fire; Pinus banksiana; nutrients
APA, Harvard, Vancouver, ISO, and other styles
50

Burridge, TR, T. Portelli, and P. Ashton. "Effect of sewage effluents on germination of three marine brown algal macrophytes." Marine and Freshwater Research 47, no. 8 (1996): 1009. http://dx.doi.org/10.1071/mf9961009.

Full text
Abstract:
Inhibition of germination of zygotes of the fucoid macroalgae Hormosira banksii and Phyllospora comosa and zoospores of the laminarian Macrocystis angustifolia was used as an end-point to assess the toxicity of three sewage effluents of differing quality. For each species, between-assay variation was low and results of tests with the reference toxicant 2,4-dichlorophenoxyacetic acid suggested that results are reproducible, especially in R. comosa. Each species showed a greater sensitivity to primary-treated effluent than to secondary-treated effluent, and higher variability in response to the primary effluent. High variation in response for each species when exposed to the primary effluent (compared with that for the secondary effluent) is presumably indicative of variation in quality of the primary effluent. The capacity to reproduce these assays, the sensitivity of species employed, and the ecological relevance of germination as a toxicological end-point suggest that germination tests of this nature may be useful in biological testing of effluent quality at discharge sites in south-eastern Australia.
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