Journal articles on the topic 'Soil mechanics South Australia'

To see the other types of publications on this topic, follow the link: Soil mechanics South Australia.

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 'Soil mechanics South Australia.'

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

Baker, G. H., P. J. Carter, and V. J. Barrett. "Survival and biomass of exotic earthworms, Aporrectodea spp. (Lumbricidae), when introduced to pastures in south-eastern Australia." Australian Journal of Agricultural Research 50, no. 7 (1999): 1233. http://dx.doi.org/10.1071/ar98181.

Full text
Abstract:
The earthworm fauna of pastures in south-eastern Australia is dominated by exotic lumbricid earthworms, in particular the endogeic species, Aporrectodea caliginosa and A. trapezoides. Anecic species such as A. longa are very rare. All 3 species were introduced within cages in 10 pastures on a range of soil types within the region. Five months later, A. longa had generally survived the best and A. trapezoides the worst. The survivals and weights of individual worms varied between sites for all 3 species. The survivals of A. caliginosa and A. longa, and to a lesser extent A. trapezoides, were positively correlated with soil clay content. The weights of A. caliginosa and A. longa, but not A. trapezoides, were positively correlated with soil P content. The survivals and weights of A. longa and A. trapezoides and the weights only of A. caliginosa decreased with increasing inoculation density, suggesting increased intraspecific competition for resources, particularly in the first two species. A. longa reduced the abundance and biomass of the exotic acanthodrilid earthworm, Microscolex dubius, at one site, and the total biomass of 3 native megascolecid species at another, when these latter species occurred as contaminants in A. longa cages. The addition of lime had no effect on the survivals and weights of A. caliginosa, A. longa, and A. trapezoides, although the soils were acid at the sites tested. The addition of sheep dung increased the survival and weights of some species at some sites. Mechanical disturbance of the soil within cages reduced the survivals of A. longa and A. trapezoides. A. longa was released without being caged at 25 sites within one pasture in South Australia. Four years later, it was recovered at all release points. A. longa has the potential to colonise pastures widely throughout the higher rainfall regions of south-eastern Australia.
APA, Harvard, Vancouver, ISO, and other styles
2

Luck, Joanne E., Rosa Crnov, Barbara Czerniakowski, Ian W. Smith, and Jane R. Moran. "Investigating the Presence of Biotic Agents Associated with Mundulla Yellows." Plant Disease 90, no. 4 (April 2006): 404–10. http://dx.doi.org/10.1094/pd-90-0404.

Full text
Abstract:
The role of biotic agents in the dieback syndrome Mundulla Yellows (MY) was investigated by analysis of 40 Eucalyptus camaldulensis, E. leucoxylon, or E. cladocalyx trees and soil samples from South Australia and Victoria, Australia. No pathogenic fungi, bacteria, phytoplasmas, or insect pests or vectors were found to be associated with MY. However, nematode analysis identified Merlinius spp. to be associated with soil, but not roots, from symptomatic trees. Interveinal chlorosis symptoms were not transmissible by seed, mechanical inoculation, or grafting using plant material derived from symptomatic trees. Virus-like particles were detected at a single symptomatic study site using transmission electron microscopy. MY symptoms were induced in E. camaldulensis seedlings by sowing seed from asymptomatic trees into sterilized and unsterilized soil collected from underneath symptomatic trees. Significantly, sterilized soil induced more severe symptoms in seedlings than unsterilized soil. Soil collected from under asymptomatic trees did not induce MY symptoms. This preliminary investigation indicates that, with the exception of Merlinius spp., pathogenic organisms and pests were not consistently associated with MY symptoms.
APA, Harvard, Vancouver, ISO, and other styles
3

Pickering, Bianca J., Jamie E. Burton, Trent D. Penman, Madeleine A. Grant, and Jane G. Cawson. "Long-Term Response of Fuel to Mechanical Mastication in South-Eastern Australia." Fire 5, no. 3 (June 3, 2022): 76. http://dx.doi.org/10.3390/fire5030076.

Full text
Abstract:
Mechanical mastication is a fuel management strategy that modifies vegetation structure to reduce the impact of wildfire. Although past research has quantified immediate changes to fuel post-mastication, few studies consider longer-term fuel trajectories and climatic drivers of this change. Our study sought to quantify changes to fuel loads and structure over time following mastication and as a function of landscape aridity. Measurements were made at 63 sites in Victoria, Australia. All sites had been masticated within the previous 9 years to remove over-abundant shrubs and small trees. We used generalised additive models to explore trends over time and along an aridity gradient. Surface fuel loads were highest immediately post-mastication and in the most arid sites. The surface fine fuel load declined over time, whereas the surface coarse fuel load remained high; these trends occurred irrespective of landscape aridity. Standing fuel (understorey and midstorey vegetation) regenerated consistently, but shrub cover was still substantially low at 9 years post-mastication. Fire managers need to consider the trade-off between a persistently higher surface coarse fuel load and reduced shrub cover to evaluate the efficacy of mastication for fuel management. Coarse fuel may increase soil heating and smoke emissions, but less shrub cover will likely moderate fire behaviour.
APA, Harvard, Vancouver, ISO, and other styles
4

Crombie, DS, JT Tippett, and TC Hill. "Dawn Water Potential and Root Depth of Trees and Understorey Species in Southwestern Australia." Australian Journal of Botany 36, no. 6 (1988): 621. http://dx.doi.org/10.1071/bt9880621.

Full text
Abstract:
Water relations of selected tree and understorey species in the jarrah forest of south-western Australia were studied during summer drought and the results related to root morphology. Seasonal patterns of predawn water potential (Ψp) differed between species according to root depth and between sites according to average annual rainfall. Dawn water potentials fell most rapidly and by the greatest amount in plants with the shallowest roots. Dawn water potentials of medium and deep rooted species were not consistently different. Separation of Ψp between sites of different annual rainfall was less marked than was separation by root depth. Changes in Ψp, were consistent with a top-to-bottom drying of the soil profiles. We suggest that measurements of Ψp of plants of appropriate root depth can be used to monitor the drying of soils as an alternative to more expensive mechanical and electrical methods.
APA, Harvard, Vancouver, ISO, and other styles
5

Abdallah, Ahmed M., Hanuman S. Jat, Madhu Choudhary, Emad F. Abdelaty, Parbodh C. Sharma, and Mangi L. Jat. "Conservation Agriculture Effects on Soil Water Holding Capacity and Water-Saving Varied with Management Practices and Agroecological Conditions: A Review." Agronomy 11, no. 9 (August 24, 2021): 1681. http://dx.doi.org/10.3390/agronomy11091681.

Full text
Abstract:
Improving soil water holding capacity (WHC) through conservation agriculture (CA)-practices, i.e., minimum mechanical soil disturbance, crop diversification, and soil mulch cover/crop residue retention, could buffer soil resilience against climate change. CA-practices could increase soil organic carbon (SOC) and alter pore size distribution (PSD); thus, they could improve soil WHC. This paper aims to review to what extent CA-practices can influence soil WHC and water-availability through SOC build-up and the change of the PSD. In general, the sequestered SOC due to the adoption of CA does not translate into a significant increase in soil WHC, because the increase in SOC is limited to the top 5–10 cm, which limits the capacity of SOC to increase the WHC of the whole soil profile. The effect of CA-practices on PSD had a slight effect on soil WHC, because long-term adoption of CA-practices increases macro- and bio-porosity at the expense of the water-holding pores. However, a positive effect of CA-practices on water-saving and availability has been widely reported. Researchers attributed this positive effect to the increase in water infiltration and reduction in evaporation from the soil surface (due to mulching crop residue). In conclusion, the benefits of CA in the SOC and soil WHC requires considering the whole soil profile, not only the top soil layer. The positive effect of CA on water-saving is attributed to increasing water infiltration and reducing evaporation from the soil surface. CA-practices’ effects are more evident in arid and semi-arid regions; therefore, arable-lands in Sub-Sahara Africa, Australia, and South-Asia are expected to benefit more. This review enhances our understanding of the role of SOC and its quantitative effect in increasing water availability and soil resilience to climate change.
APA, Harvard, Vancouver, ISO, and other styles
6

Jettner, R., S. P. Loss, L. D. Martin, and K. H. M. Siddique. "Responses of faba bean (Vicia faba L.) to sowing rate in south-western Australia II Canopy development, radiation absorption and dry matter partitioning." Australian Journal of Agricultural Research 49, no. 6 (1998): 999. http://dx.doi.org/10.1071/a98003.

Full text
Abstract:
Sowing rate influences plant density, canopy development, radiation absorption, dry matter production and its partitioning, and seed yield. The canopy development, radiation interception, and dry matter partitioning of faba bean (cv. Fiord) were examined using 6 sowing rate treatments from 70 to 270 kg/ha in field experiments conducted over 3 years at Northam as part of a larger investigation of sowing rate responses in faba bean in south-western Australia. High sowing rates resulted in significantly earlier canopy closure, larger green area indexes, more radiation absorption, more dry matter accumulation particularly during the early vegetative stages, and greater seed yield than treatments where a low plant density was established. The results suggest that further increases in canopy development, radiation absorption, dry matter accumulation, and seed yield are possible by using sowing rates in excess of 270 kg/ha. The rate of node appearance was relatively constant within and across seasons (1 every 65·9 degree-days), whereas the number of branches per plant declined with increasing plant density, and less branches survived through to maturity at high density. The peak photosynthetically active radiation absorption (75-85%) measured at green area index of 2·9-3·8 in the highest sowing rate treatment in this study is similar to previous reports for other crops. The estimated radiation use efflciency (1·30 g/MJ) was constant across sowing rate treatments and seasons. High sowing rates produced tall crops with the lowest pods further from the soil surface than those at low plant density, and hence, mechanical harvesting was easier. The growth of individual plants may have been limited by the low growing season rainfall (266-441 mm) and/or low soil pH (5·0 in CaCl2) at the site, and competition between plants for radiation was probably small even at the highest sowing rate. Early canopy closure and greater dry matter production with high sowing rates may also cause greater suppression of weeds and aphids.
APA, Harvard, Vancouver, ISO, and other styles
7

Regan, K. L., K. H. M. Siddique, and L. D. Martin. "Response of kabuli chickpea (Cicer arietinum L.) to sowing rate in Mediterranean-type environments of south-western Australia." Australian Journal of Experimental Agriculture 43, no. 1 (2003): 87. http://dx.doi.org/10.1071/ea01200.

Full text
Abstract:
The effect of sowing rate (60–320 kg/ha) on the growth and seed yield of kabuli chickpea (cv. Kaniva) was assessed at 11 sites for 4 seasons in the cropping regions of south-western Australia. The economic optimum plant density and yield potential were estimated using an asymptotic model fitted to the data and calculating the sowing rate above which the cost of additional seed was equivalent to the revenue that could be achieved from the extra seed yield produced, assuming a 10 and 50% opportunity cost. On average for all sites and seasons, plant densities ranged from 10 plants/m2 when sown at 60 kg/ha to 43�plants/m2 when sown at 320 kg/ha. Assuming a mean seed weight of 400 mg and a germination of 80%, then on average 75% of viable seeds sown (or 60% of sown seeds) established as plants. The poor establishment rates are thought to be associated with reduced viability caused by mechanical damage, storage conditions, fungal infection in the soil, and unfavourable seed bed moisture and temperatures. In general, there was a positive relationship between sowing rate and seed yield. Seed yield increases at higher sowing rates were mainly associated with the greater number of plants per unit area. There were fewer pods per plant at higher sowing rates, but there were more pods per unit area. Changing the sowing rate had little effect on mean seed weight and the number of seeds per pod. The economic optimum plant density varied from 8 to 68 plants/m2, depending on the location, but the mean (27�plants/m2) was within the range currently recommended in southern Australia (25–35 plants/m2). Due to the low establishment rates observed in this study, we estimate a sowing rate greater (160–185 kg/ha) than currently suggested (110–160 kg/ha) to achieve this density. There was a strong relationship between economic optimum plant density and seed yield potential (r2 = 0.66, P<0.01), which allows an estimation of the most profitable sowing rate, depending on the seed yield potential of the site. For most crops yielding about 1.0 t/ha in southern Australia, a plant density of 25 plants/m2 is most profitable, while in higher-yielding situations (>1.5 t/ha) plant densities >35�plants/m2 will produce the most profit.
APA, Harvard, Vancouver, ISO, and other styles
8

SINGH, B., and R. J. GILKES. "Properties of soil kaolinites from south-western Australia." Journal of Soil Science 43, no. 4 (December 1992): 645–67. http://dx.doi.org/10.1111/j.1365-2389.1992.tb00165.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Aldaoud, R., W. Guppy, L. Callinan, S. F. Flett, K. A. Wratten, G. A. Murray, T. Cook, and A. McAllister. "Occurrence of Phytophthora clandestina in Trifolium subterraneum paddocks in Australia." Australian Journal of Experimental Agriculture 41, no. 2 (2001): 187. http://dx.doi.org/10.1071/ea00048.

Full text
Abstract:
In 1995–96, a survey of soil samples from subterranean clover (Trifolium subterraneum L.) paddocks was conducted across Victoria, South Australia, New South Wales and Western Australia, to determine the distribution and the prevalence of races of Phytophthora clandestina (as determined by the development of root rot on differential cultivars), and the association of its occurrence with paddock variables. In all states, there was a weak but significant association between P. clandestina detected in soil samples and subsequent root rot susceptibility of differential cultivars grown in these soil samples. Phytophthora clandestina was found in 38% of the sampled sites, with a significantly lower prevalence in South Australia (27%). There were significant positive associations between P. clandestina detection and increased soil salinity (Western Australia), early growth stages of subterranean clover (Victoria), mature subterranean clover (South Australia), recently sown subterranean clover (South Australia), paddocks with higher subterranean clover content (Victoria), where herbicides were not applied (South Australia), irrigation (New South Wales and Victoria), cattle grazing (South Australia and Victoria), early sampling dates (Victoria and New South Wales), sampling shortly after the autumn break or first irrigation (Victoria), shorter soil storage time (Victoria) and farmer’s perception of root rot being present (Victoria and New South Wales). Only 29% of P. clandestina isolates could be classified under the 5 known races. Some of the unknown races were virulent on cv. Seaton Park LF (most resistant) and others were avirulent on cv. Woogenellup (most susceptible). Race 1 was significantly less prevalent in South Australia than Victoria and race 0 was significantly less prevalent in New South Wales than in South Australia and Western Australia. This study revealed extremely wide variation in the virulence of P. clandestina. The potential importance of the results on programs to breed for resistance to root rot are discussed. in South Australia.
APA, Harvard, Vancouver, ISO, and other styles
10

St. Pierre, Tim G. "Mössbauer Spectra of Soil Kaolins from South-Western Australia." Clays and Clay Minerals 40, no. 3 (1992): 341–46. http://dx.doi.org/10.1346/ccmn.1992.0400315.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Doran-Browne, Natalie A., John Ive, Phillip Graham, and Richard J. Eckard. "Carbon-neutral wool farming in south-eastern Australia." Animal Production Science 56, no. 3 (2016): 417. http://dx.doi.org/10.1071/an15541.

Full text
Abstract:
Ruminant livestock production generates higher levels of greenhouse gas emissions (GHGE) compared with other types of farming. Therefore, it is desirable to reduce or offset those emissions where possible. Although mitigation options exist that reduce ruminant GHGE through the use of feed management, flock structure or breeding management, these options only reduce the existing emissions by up to 30% whereas planting trees and subsequent carbon sequestration in trees and soil has the potential for livestock emissions to be offset in their entirety. Trees can introduce additional co-benefits that may increase production such as reduced salinity and therefore increased pasture production, shelter for animals or reduced erosion. Trees will also use more water and compete with pastures for water and light. Therefore, careful planning is required to locate trees where the co-benefits can be maximised instead of any negative trade-offs. This study analysed the carbon balance of a wool case study farm, Talaheni, in south-eastern Australia to determine if the farm was carbon neutral. The Australian National Greenhouse Gas Inventory was used to calculate GHGE and carbon stocks, with national emissions factors used where available, and otherwise figures from the IPCC methodology being used. Sources of GHGE were from livestock, energy and fuel, and carbon stocks were present in the trees and soil. The results showed that from when the farm was purchased in 1980–2012 the farm had sequestered 11 times more carbon dioxide equivalents (CO2e) in trees and soil than was produced by livestock and energy. Between 1980 and 2012 a total of 31 100 t CO2e were sequestered with 19 300 and 11 800 t CO2e in trees and soil, respectively, whereas farm emissions totalled 2800 t CO2e. There was a sufficient increase in soil carbon stocks alone to offset all GHGE at the study site. This study demonstrated that there are substantial gains to be made in soil carbon stocks where initial soils are eroded and degraded and there is the opportunity to increase soil carbon either through planting trees or introducing perennial pastures to store more carbon under pastures. Further research would be beneficial on the carbon-neutral potential of farms in more fertile, high-rainfall areas. These areas typically have higher stocking rates than the present study and would require higher levels of carbon stocks for the farm to be carbon neutral.
APA, Harvard, Vancouver, ISO, and other styles
12

Ward, S. C. "Soil development on rehabilitated bauxite mines in south-west Australia." Soil Research 38, no. 2 (2000): 453. http://dx.doi.org/10.1071/sr99032.

Full text
Abstract:
Rehabilitation after bauxite mining in the jarrah forest aims to re-establish a self-sustaining forest. This implies that ecosystem processes will be re-established and soil nutrient stores will be similar to those of the unmined forest. This study determined the baseline levels of a number of soil properties in areas of jarrah forest typically cleared and mined for bauxite, and the effect of mining and rehabilitation processes on the vertical distribution of soil nitrogen. In addition, the changes in soil nitrogen, extractable phosphorus, extractable potassium, and pH that occurred up to 8.5 years after rehabilitation were investigated. The values of the soil properties in unmined forest found in this study were similar to those previously measured in the jarrah forest. Values of all parameters varied significantly with depth. Considerable mixing of soil horizons occurred when the topsoil was stripped, placed on an area undergoing rehabilitation, and ripped. Many soil parameters showed significant changes in the period 3.5–8.5 years after rehabilitation. Levels of total nitrogen in the top 10 cm of soil in rehabilitated areas increased from around 0.04–0.05% initially to levels approaching those found in good-quality jarrah forest (0.10–0.30%) after 8.5 years. Soil pH declined in the surface layer after rehabilitation. The rate of acidification is likely to decrease in future years. The levels of the soil chemical parameters were mostly within or moving towards the range of values found in areas typically mined for bauxite. The results indicate that ecosystem processes have been re-established on the rehabilitated areas and that there is no obvious soil chemical impediment to the establishment of a self-sustaining forest ecosystem on rehabilitated bauxite mines.
APA, Harvard, Vancouver, ISO, and other styles
13

Cock, GJ. "Moisture characteristics of irrigated Mallee soils in South Australia." Australian Journal of Experimental Agriculture 25, no. 1 (1985): 209. http://dx.doi.org/10.1071/ea9850209.

Full text
Abstract:
The soil moisture characteristics of undisturbed samples of Mallee soils, taken from typical profiles of the Riverland district neat Berri in South Australia, were determined. Samples were grouped according to texture and bulk density and, for each group, the moisture storage between matric potentials was derived. Over the usual range of soil moisture tensions (-0 to 40kPa) these showed only small variation between soil groups since, while moisture storage at field capacity and at wilting point does vary with texture; 50 to 60 mm/m is available between field capacity (-7 kPa) and the re-irrigation point (-30 to 40kPa) in all soils.
APA, Harvard, Vancouver, ISO, and other styles
14

Radke, J. K., A. R. Dexter, and O. J. Devine. "Tillage Effects on Soil Temperature, Soil Water, and Wheat Growth in South Australia." Soil Science Society of America Journal 49, no. 6 (November 1985): 1542–47. http://dx.doi.org/10.2136/sssaj1985.03615995004900060042x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Gardner, WK, RG Fawcett, GR Steed, JE Pratley, DM Whitfield, Hvan Rees, and Rees H. Van. "Crop production on duplex soils in south-eastern Australia." Australian Journal of Experimental Agriculture 32, no. 7 (1992): 915. http://dx.doi.org/10.1071/ea9920915.

Full text
Abstract:
The environment, duplex soil types and trends in crop production in South Australia, southern New South Wales, north-eastern and north-central Victoria, the southern Wimmera and the Victorian Western District are reviewed. In the latter 2 regions, pastoral industries dominate and crop production is curtailed by regular and severe soil waterlogging, except for limited areas of lower rainfall. Subsurface drainage can eliminate waterlogging, but is feasible only for the Western District where subsoils are sufficiently stable. The other regions all have a long history of soil degradation due to cropping practices, but these effects can now be minimised with the use of direct drilling and stubble retention cropping methods. A vigorous pasture ley phase is still considered necessary to maintain nitrogen levels and to restore soil structure to adequate levels for sustainable farming. Future productivity improvements will require increased root growth in the subsoils. Deep ripping, 'slotting' of gypsum, and crop species capable of opening up subsoils are techniques which may hold promise in this regard. The inclusion of lucerne, a perennial species, in annual pastures and intercropping at intervals is a technique being pioneered in north-central and western Victoria and may provide the best opportunity to crop duplex soils successfully without associated land degradation.
APA, Harvard, Vancouver, ISO, and other styles
16

Harper, R. J., R. J. Gilkes, M. J. Hill, and D. J. Carter. "Wind erosion and soil carbon dynamics in south-western Australia." Aeolian Research 1, no. 3-4 (January 2010): 129–41. http://dx.doi.org/10.1016/j.aeolia.2009.10.003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

FOULDS, W. "Nutrient concentrations of foliage and soil in South-western Australia." New Phytologist 125, no. 3 (November 1993): 529–46. http://dx.doi.org/10.1111/j.1469-8137.1993.tb03901.x.

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

Naidu, R., RH Merry, GJ Churchman, MJ Wright, RS Murray, RW Fitzpatrick, and BA Zarcinas. "Sodicity in South Australia - a review." Soil Research 31, no. 6 (1993): 911. http://dx.doi.org/10.1071/sr9930911.

Full text
Abstract:
The current knowledge of the nature and distribution of sodic soils in South Australia is reviewed. The agriculturally developed area of South Australia lies south of latitude 32-degrees-S. and is mainly used for low intensity grazing and dry land cereal/sheep production. A high proportion of the State, including much of the high rainfall area, has soils which are sodic (>6% ESP) through a significant proportion of the profile but information on the precise nature of sodicity in these soils is limited. Where exchangeable cation data axe available, the analytical techniques used often did not precisely delineate between soluble salts in the soil and ions on exchange sites. Therefore, many of the datasets have major weaknesses and may be unreliable. Since many soils with ESP <6 also show dispersive characteristics typical of sodic soils, there is an urgent need for new sodicity studies relating to distribution and the criteria (ESP) used to identify dispersive soils. Information on the effect of sodicity on nutrient requirements of plants, especially the modern varieties, is scarce both locally and internationally, making development of management strategies for economically sustainable crop production difficult. Further, many different grades of gypsum are available in South Australia. Preliminary studies show the presence of impurities drastically influences gypsum dissolution characteristics. More effort is needed to assess the quality and reactivity of South Australian gypsum. Some effort has been directed by land managers towards reclamation and management of sodic soils by using both gypsum and lime either separately or as mixtures. However, there is neither a scientific basis for the application of gypsum-lime mixtures nor crop production data to support such management strategies.
APA, Harvard, Vancouver, ISO, and other styles
19

Webb, Ashley A., Georgina L. Kelly, and Warwick J. Dougherty. "Soil governance in the agricultural landscapes of New South Wales, Australia." International Journal of Rural Law and Policy, no. 1 (March 29, 2015): 1–16. http://dx.doi.org/10.5130/ijrlp.i1.2015.4169.

Full text
Abstract:
Soil is a valuable natural resource. In the state of New South Wales, Australia, the governance of soil has evolved since Federation in 1901. Following rapid agricultural development, and in the face of widespread soil degradation, the establishment of the Soil Conservation Service marked a turning point in the management of soil. Throughout the 20th century, advances in knowledge were translated into evolving governance frameworks that were largely reactionary but saw progressive reforms such as water pollution legislation and case studies of catchment-scale land and vegetation management. In the 21st century, significant reforms have embedded sustainable use of agricultural soils within catchment- and landscape-scale legislative and institutional frameworks. What is clear, however, is that a multitude of governance strategies and models are utilised in NSW. No single governance model is applicable to all situations because it is necessary to combine elements of several different mechanisms or instruments to achieve the most desired outcomes. Where an industry, such as the sugar industry, has taken ownership of an issue such as acid sulfate soil management, self-regulation has proven to be extremely effective. In the case of co-managing agricultural soils with other landuses, such as mining, petroleum exploration and urban development, regulation, compliance and enforcement mechanisms have been preferred. Institutional arrangements in the form of independent commissioners have also played a role. At the landscape or total catchment level, it is clear that a mix of mechanisms is required. Fundamental, however, to the successful evolution of soil governance is strategic investment in soil research and development that informs the ongoing productive use of agricultural landscapes while preventing land degradation or adverse environmental effects.
APA, Harvard, Vancouver, ISO, and other styles
20

Angus, J. F., A. F. van Herwaarden, D. P. Heenan, R. A. Fischer, and G. N. Howe. "The source of mineral nitrogen for cereals in south-eastern Australia." Australian Journal of Agricultural Research 49, no. 3 (1998): 511. http://dx.doi.org/10.1071/a97125.

Full text
Abstract:
The relative importance of soil mineral nitrogen (N) available at the time of sowing ormineralised during the growing season was investigated for 6 crops of dryland wheat. The soil mineral N in the root-zone was sampled at sowing and maturity and the rate of net mineralisation in the top 10 cm was estimated by sequential sampling throughout the growing season, using an in situ method. Mineralisation during crop growth was modelled in relation to total soil N, ambient temperature, andsoil water content. Mineral N accumulated before sowing varied by a factor of 3 between the sites (from 67 to 195 kgN/ha), while the net mineralisation during crop growth varied by a factor of 2 (from 43 to 99 kgN/ha). The model indicated that 0·092% of total N was mineralised per day when temperature and water were not limiting, with rates decreasing for lower temperatures and soil water contents. When tested with independent data, the model predicted the mineralisation rate of soil growing continuous wheat crops but underestimated mineralisation of soil in a clover-wheat rotation. For crops yielding <3 t/ha, the supply of N was mostly from mineralisation during crop growth and the contribution from mineral N accumulated before sowing was relatively small. For crops yielding >4 t/ha, thesupply of N was mostly from N present in the soil at the time of sowing. The implication is that for crops to achieve their water-limited yield, they must be supplied with an amount of N greater than can be expected from mineralisation during the growing season, either from fertiliser or from mineral N accumulated earlier.
APA, Harvard, Vancouver, ISO, and other styles
21

Yang, Xihua, Jonathan Gray, Greg Chapman, Qinggaozi Zhu, Mitch Tulau, and Sally McInnes-Clarke. "Digital mapping of soil erodibility for water erosion in New South Wales, Australia." Soil Research 56, no. 2 (2018): 158. http://dx.doi.org/10.1071/sr17058.

Full text
Abstract:
Soil erodibility represents the soil’s response to rainfall and run-off erosivity and is related to soil properties such as organic matter content, texture, structure, permeability and aggregate stability. Soil erodibility is an important factor in soil erosion modelling, such as the Revised Universal Soil Loss Equation (RUSLE), in which it is represented by the soil erodibility factor (K-factor). However, determination of soil erodibility at larger spatial scales is often problematic because of the lack of spatial data on soil properties and field measurements for model validation. Recently, a major national project has resulted in the release of digital soil maps (DSMs) for a wide range of key soil properties over the entire Australian continent at approximately 90-m spatial resolution. In the present study we used the DSMs and New South Wales (NSW) Soil and Land Information System to map and validate soil erodibility for soil depths up to 100 cm. We assessed eight empirical methods or existing maps on erodibility estimation and produced a harmonised high-resolution soil erodibility map for the entire state of NSW with improvements based on studies in NSW. The modelled erodibility values were compared with those from field measurements at soil plots for NSW soils and revealed good agreement. The erodibility map shows similar patterns as that of the parent material lithology classes, but no obvious trend with any single soil property. Most of the modelled erodibility values range from 0.02 to 0.07 t ha h ha–1 MJ–1 mm–1 with a mean (± s.d.) of 0.035 ± 0.007 t ha h ha–1 MJ–1 mm–1. The validated K-factor map was further used along with other RUSLE factors to assess soil loss across NSW for preventing and managing soil erosion.
APA, Harvard, Vancouver, ISO, and other styles
22

Unkovich, Murray, Therese McBeath, Rick Llewellyn, James Hall, Vadakattu VSR Gupta, and Lynne M. Macdonald. "Challenges and opportunities for grain farming on sandy soils of semi-arid south and south-eastern Australia." Soil Research 58, no. 4 (2020): 323. http://dx.doi.org/10.1071/sr19161.

Full text
Abstract:
Sandy soils make up a substantial fraction of cropping land in low rainfall (&lt;450 mm p.a.) south and south-eastern Australia. In this paper we review the possible soil constraints to increased production on these soils in this region. Many of these soils have a very low (&lt;3%) clay content and suffer from severe water repellency, making crop establishment and weed control problematic. Crops which do emerge are faced with uneven soil wetting and poor access to nutrients, with crop nutrition constraints exacerbated by low fertility (soil organic matter &lt; 1%) and low cation exchange capacity. Zones of high penetration resistance appear common and have multiple causes (natural settling, cementation and traffic induced) which restrict root growth to &lt;40 cm. Crop water use and grain yield are therefore likely to be well below the water-limited potential. Water repellency is readily diagnosed and where apparent should be the primary management target. Repellency can be mitigated through the use of furrow and other sowing technologies, along with soil wetting agents. These techniques appear to be affected by site and soil nuances and need to be refined for local soils and conditions. Once crop establishment on water repellent soils has been optimised, attention could be turned to opportunities for improving crop rooting depth through the use of deep tillage or deep ripping techniques. The required ripping depth, and how long the effects may last, are unclear and need further research, as do the most effective and efficient machinery requirements to achieve sustained deeper root growth. Crop nutrition matched to the water-limited crop yield potential is the third pillar of crop production that needs to be addressed. Low soil organic matter, low cation exchange capacity, low biological activity and limited nutrient cycling perhaps make this a greater challenge than in higher rainfall regions with finer textured soils. Interactions between nutrients in soils and fertilisers are likely to occur and make nutrient management more difficult. While amelioration (elimination) of water repellency is possible through the addition of clay to the soil surface, the opportunities for this may be restricted to the ~30% of the sandy soils of the region where clay is readily at hand. The amounts of clay required to eliminate repellency (~5%) are insufficient to significantly improve soil fertility or soil water holding capacity. More revolutionary soil amelioration treatments, involving additions and incorporation of clay and organic matter to soils offer the possibility of a more elevated crop yield plateau. Considerable research would be required to provide predictive capacity with respect to where and when these practices are effective.
APA, Harvard, Vancouver, ISO, and other styles
23

Hopmans, P., N. Collett, and R. Bickford. "Effects of fire retardant on heathland soils in south-eastern Australia." Soil Research 45, no. 8 (2007): 607. http://dx.doi.org/10.1071/sr07040.

Full text
Abstract:
A study was undertaken to assess the effects of fire retardant application, unmodified by heat of fire, on soil properties in 2 fire-prone heathland communities at Marlo and the Grampians in south-eastern Australia. Fire retardant (Phos-Chek D75-R at 0.144 g/L) was applied at rates of 0.5, 1.0, and 1.5 L/m2 and compared with control treatments of nil and 1.0 L/m2 of water. Monitoring of surface soils showed that pH at both sites decreased while soil salinity increased immediately after application followed by a rapid decline to pre-treatment values within 12 months. The impact of retardant on total carbon and nitrogen was minor and within the range of natural variation of C and N in surface soils at both sites. Levels of readily available or labile forms of N increased at both sites but declined rapidly to background values after 12 months. Applications of retardant progressively increased extractable P in the surface soil at Marlo, in contrast to the Grampians where a rapid increase was observed after two months followed by a decline after 12 months. These results showed a significant increase in labile P in the surface soil after 12 months and also indicated that a large proportion of the phosphate applied had leached into the subsoil. Likewise, fire retardant applied at the highest rate caused increases in labile sulfate after 2 months at both sites, followed by a rapid decline to background levels. It is expected that the elevated levels of soil phosphate in particular could have a long-term impacts on growth and composition of heathland vegetation known to be sensitive to elevated levels of phosphate in soil.
APA, Harvard, Vancouver, ISO, and other styles
24

Hopmans, Peter. "Stem deformity inPinus radiata plantations in south-eastern Australia:." Plant and Soil 122, no. 1 (February 1990): 97–104. http://dx.doi.org/10.1007/bf02851915.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Hopmans, Peter, Matt Kitching, and George Croatto. "Stem deformity inPinus radiata plantations in south-eastern Australia." Plant and Soil 175, no. 1 (August 1995): 31–44. http://dx.doi.org/10.1007/bf02413008.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Fisher, JM, and TW Hancock. "Population dynamics of Heterodera avenae Woll. in South Australia." Australian Journal of Agricultural Research 42, no. 1 (1991): 53. http://dx.doi.org/10.1071/ar9910053.

Full text
Abstract:
Population changes of the cereal cyst nematode (Heterodera avenae Woll.) under various rotational regimes in the field were examined. A density of 5 eggs/g soils caused a loss of about 10% in yield of wheat cv. Bayonet under the experimental conditions. Maximum multiplication rate at low initial densities was about l0x, but this rapidly decreased as initial density increased. Equilibrium levels ranged from 15 eggs/g soil up to 40 eggs/g under different conditions. The resistant wheat, Aus 10894, maintained an equilibrium level of about 1 egg/g soil-a level low enough to avoid damage in a following crop. Annual percentage hatch varied from 70-90% with an average of 85%. Presence or absence of plant species had little effect on per cent hatch. About 7.5% of the nematodes successfully penetrated and became established in the root systems of seedlings, invading both seminal and nodal roots, but only about 0.5% became established in the principal axes of the seminal roots. The implications of these data for various rotational practices are discussed.
APA, Harvard, Vancouver, ISO, and other styles
27

Barrett-Lennard, Edward G., Geoffrey C. Anderson, Karen W. Holmes, and Aidan Sinnott. "High soil sodicity and alkalinity cause transient salinity in south-western Australia." Soil Research 54, no. 4 (2016): 407. http://dx.doi.org/10.1071/sr15052.

Full text
Abstract:
Transient salinity associated with increased dispersion of clays is arguably one of the most economically important soil constraints in Australia because it occurs on land that is regularly cropped. However, this issue is rarely studied. This paper examines the occurrence of transient salinity on agricultural land in the south-west of Western Australia and the factors causing it. We analysed four soil datasets from the region, collected at scales varying from the entire south-west to a single paddock. A variety of soil parameters were correlated with increased electrical conductivity (EC1:5). The most significant relationships were invariably with measures of exchangeable sodium (Na+; 53–85% of variance accounted for), and this factor appears to be most responsible for transient salinity. Another parameter correlated with increased EC1:5 was alkalinity. This has been associated with the increased dispersion of kaolinite and consequent decreases in soil hydraulic conductivity; kaolinite is the most common clay mineral in the south-west of Western Australia. Other factors correlated with increased EC1:5 were increasing clay, increasing depth in the soil profile and decreasing rainfall. These factors are environmental indicators of transient salinity. Affected soils might be ameliorated by application of agents to increase soil hydraulic conductivity, such as gypsum and/or elemental sulfur.
APA, Harvard, Vancouver, ISO, and other styles
28

Wang, T. H., Y. T. Tchan, A. M. M. Zeman, and I. R. Kennedy. "Presence of sodium dependent azotobacter in Australia (New South Wales)." Soil Biology and Biochemistry 25, no. 5 (May 1993): 637–39. http://dx.doi.org/10.1016/0038-0717(93)90205-p.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Nash, David, Craig Butler, Justine Cody, Michael St J. Warne, Mike J. McLaughlin, Dianne Heemsbergen, Kris Broos, et al. "Effects of Biosolids Application on Pasture and Grape Vines in South-Eastern Australia." Applied and Environmental Soil Science 2011 (2011): 1–11. http://dx.doi.org/10.1155/2011/342916.

Full text
Abstract:
Biosolids were applied to a pasture and a vineyard in south-eastern Australia. At both sites, soil Cd, Cu, and Zn concentrations linearly increased with biosolids application rates although not to the extent of exceeding soil quality guidelines. Biosolids marginally increased soil C and N concentrations at the pasture site but significantly increased P concentrations. With lower overall soil fertility at the vineyard, biosolids increased C, N, and P concentrations. At neither site did biosolids application affect soil microbial endpoints. Biosolids increased pasture production compared to the unfertilised control but had little effect on grape production or quality. Interestingly, over the 3-year trial, there was no difference in pasture production between the biosolids treated plots and plots receiving inorganic fertiliser. These results suggest that biosolids could be used as a fertiliser to stimulate pasture production and as a soil conditioner to improve vineyard soils in this region.
APA, Harvard, Vancouver, ISO, and other styles
30

Benyon, Richard G., S. Theiveyanathan, and Tanya M. Doody. "Impacts of tree plantations on groundwater in south-eastern Australia." Australian Journal of Botany 54, no. 2 (2006): 181. http://dx.doi.org/10.1071/bt05046.

Full text
Abstract:
In some regions dependent on groundwater, such as the lower south-east of South Australia in the Green Triangle, deep-rooted, woody vegetation might have undesirable hydrological impacts by competing for finite, good-quality groundwater resources. In other regions, such as the Riverina in south-central New South Wales, where rising watertables and associated salinisation is threatening the viability of agriculture, woody vegetation might have beneficial hydrological impacts. In response to a growing need to better understand the impacts of tree plantations on groundwater, annual evapotranspiration and transpiration were measured at 21 plantation sites in the Green Triangle and the Riverina. Sources of tree water uptake from rainfall and groundwater were determined by measurements of evapotranspiration and soil water over periods of 2–5 years. In the Green Triangle, under a combination of permeable soil over groundwater of low salinity (<2000 mg L–1) at 6-m depth or less, in a highly transmissive aquifer, annual evapotranspiration at eight research sites in Pinus radiata D.Don and Eucalyptus globulus Labill. plantations averaged 1090 mm year–1 (range 847–1343 mm year–1), compared with mean annual precipitation of 630 mm year–1. These plantation sites used groundwater at a mean annual rate of 435 mm year–1 (range 108–670 mm year–1). At eight other plantation sites that had greater depth to the watertable or a root-impeding layer, annual evapotranspiration was equal to, or slightly less than, annual rainfall (mean 623 mm year–1, range 540–795 mm year–1). In the Riverina, where groundwater was always present within 3 m of the surface, Eucalyptus grandis Hill ex Maiden trees at three sites with medium or heavy clay, alkaline, sodic, saline subsoils used little or no groundwater, whereas E. grandis and Corymbia maculata (Hook.) K.D.Hill and L.A.S.Johnson trees at a site with a neutral sandy soil and groundwater of low salinity used 380 and 730 mm year–1 of groundwater (respectively 41 and 53% of total annual evapotranspiration). We conclude that commonly grown Eucalyptus species and P. radiata are able to use groundwater under a combination of light- or medium-textured soil and shallow depth to a low-salinity watertable.
APA, Harvard, Vancouver, ISO, and other styles
31

Grimes, K. G. "The South-East Karst Province of South Australia." Environmental Geology 23, no. 2 (March 1994): 134–48. http://dx.doi.org/10.1007/bf00766987.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

McFarlane, JD, GJ Judson, and J. Gouzos. "Copper deficiency in ruminants in the South East of South Australia." Australian Journal of Experimental Agriculture 30, no. 2 (1990): 187. http://dx.doi.org/10.1071/ea9900187.

Full text
Abstract:
Pasture development in the South East of South Australia has depended upon trace element enriched fertiliser applications. Despite the wide usage of copper-enriched fertilisers, copper deficiency is still evident in livestock at pasture, particularly cattle. Serum collected from cows and heifers during the systematic sampling program of the Brucellosis and Tuberculosis Eradication Scheme was analysed for copper. Of the 3611 pooled herd samples analysed, approximately 9% had low serum copper concentrations (<7 �mol/L). Distribution of those herds identified to be at risk of copper deficiency appeared to be random, apart from areas of high risk on peat soils and the coastal fringe of calcareous sands. Analysis of pasture samples collected from paddocks with cattle having low serum copper concentrations showed that low serum copper was usually associated with raised molybdenum rather than low copper concentrations in pasture. In some instances, moderate concentrations of molybdenum and sulfur in pasture and soil ingestion associated with high iron concentrations may combine to cause hypocupraemia, especially when livestock graze stubbles and subterranean clover pastures in summer-autumn and short pastures in winter. Only 6% of pasture samples had less than 4 mg Cu/kg DM, a concentration which indicates possible copper deficiency in subterranean clover or strawberry clover.
APA, Harvard, Vancouver, ISO, and other styles
33

Reuter, DJ, CB Dyson, DE Elliott, DC Lewis, and CL Rudd. "An appraisal of soil phosphorus testing data for crops and pastures in South Australia." Australian Journal of Experimental Agriculture 35, no. 7 (1995): 979. http://dx.doi.org/10.1071/ea9950979.

Full text
Abstract:
Data from more than 580 field experiments conducted in South Australia over the past 30 years have been re-examined to estimate extractable soil phosphorus (P) levels related to 90% maximum yield (C90) for 7 crop species (wheat, barley, oilseed rape, sunflower, field peas, faba beans, potato) and 3 types of legume-based pasture (subterranean clover, strawberry clover, annual medics). Data from both single-year and longer term experiments were evaluated. The C90 value for each species was derived from the relationship between proportional yield responsiveness to applied P fertiliser rates (determined as grain yield in crops and herbage yield in ungrazed pastures) and extractable P concentrations in surface soils sampled before sowing. Most data assessments involved the Colwell soil P test and soils sampled in autumn to 10 cm depth. When all data for a species were considered together, the relationship between proportional yield response to applied P and soil P status was typically variable, particularly where Colwell soil P concentration was around C90. When data could be grouped according to common soil types, soil surface texture, or P sorption indices (selected sites), better relationships were discerned. From such segregated data sets, different C90 estimates were derived for either different soil types or soil properties. We recommend that site descriptors associated with the supply of soil P to plant roots be determined as a matter of course in future P fertiliser experiments in South Australia. Given the above, we also contend that the Colwell soil P test is reasonably robust for estimating P fertiliser requirements for the diverse range of soils in the agricultural regions of the State. In medium- and longer term experiments, changes in Colwell soil P concentration were measured in the absence or presence of newly applied P fertiliser. The rate of change (mg soil P/kg per kg applied P/ha) appeared to vary with soil type (or soil properties) and, perhaps, cropping frequency. Relatively minor changes in soil P status were observed due to different tillage practices. In developing P fertiliser budgets, we conclude that a major knowledge gap exists for estimating the residual effectiveness of P fertiliser applied to diverse soil types under a wide range of South Australian farming systems.
APA, Harvard, Vancouver, ISO, and other styles
34

Baker, G. H., P. J. Carter, and V. J. Barrett. "Influence of earthworms, Aporrectodea spp. (Lumbricidae), on lime burial in pasture soils in south-eastern Australia." Soil Research 37, no. 5 (1999): 831. http://dx.doi.org/10.1071/sr98106.

Full text
Abstract:
The relative abilities of 3 exotic lumbricid earthworms, the endogeic Aporrectodea caliginosa and A. trapezoides and the anecic A. longa, to bury surface-applied lime and help ameliorate soil acidity were measured in cages in 7 pasture soils in south-eastern Australia. All 3 species buried lime, mostly within the top 5 cm of the soil profile, but A. longa buried it deeper than A. caliginosa and A. trapezoides. A. longa significantly increased soil pH at 15–20 cm depth at some sites within 5 months (winter–spring, the earthworm ‘season’ in the Mediterranean climate of south-eastern Australia). Lime burial varied markedly between sites. These site differences were explained, at least in part, by variations in rainfall. Lime burial increased with earthworm density. A minimum density of 214 A. longa/m 2 was needed to significantly enhance lime burial within one season. Higher densities were required for the other two species. However, per unit of biomass, A. caliginosa and A. trapezoides were generally more able to bury lime in the upper soil layers (2 . 5–10 cm depth) than A. longa. Agricultural soils in south-eastern Australia are dominated by shallow burrowing species such as A. caliginosa and A. trapezoides. Deeper burrowers such as A. longa are rare. Introduction of A. longa to soils in high-rainfall regions of south-eastern Australia, where it does not presently occur, should enhance lime burial and help reduce soil acidity.
APA, Harvard, Vancouver, ISO, and other styles
35

Coventry, D. R., W. J. Slattery, V. F. Burnett, and G. W. Ganning. "Longevity of wheat yield response to lime in south-eastern Australia." Australian Journal of Experimental Agriculture 37, no. 5 (1997): 571. http://dx.doi.org/10.1071/ea96146.

Full text
Abstract:
Summary. A long-term experiment in north-eastern Victoria has been regularly monitored for wheat yield responses to a range of lime and fertiliser treatments, and the soil sampled for acidity attributes. Substantial grain yield increases have been consistently obtained over a period of 12 years with a single lime application. Lime applied at 2.5 t/ha in 1980 was still providing yield increases of 24% with an acid-tolerant wheat (Matong, 1992 season) and 79% with an acid-sensitive wheat (Oxley, 1993 season) relative to no lime treatment. The 2 wheat cultivars responded differently to phosphorus fertiliser, with the acid-sensitive wheat less responsive to phosphorus fertiliser in the absence of lime. The use of a regular lime application applied as a fertiliser (125 kg lime/ha) with the wheat seed gave only a small grain yield increase (8% Matong, 16% Oxley), despite 1 t/ha of lime applied over the 12-year period. Liming the soil at a rate of 2.5 t/ha (1980) initially raised the soil pH by about 1.0 unit and removed most soluble aluminium (0–10 cm). However, after 12 years of crop–pasture rotation after the initial 2.5 t lime/ha treatment the soil pH had declined by 0.7 of a pH unit and exchangeable aluminium was substantially increased, almost to levels prior to the initial application of lime. Given the continued yield responsiveness obtained following the initial application of lime, this practice, rather than regular applications of small amounts of lime, is recommended for wheat production on strongly acidic (pHw < 5.5) soils in south-eastern Australia.
APA, Harvard, Vancouver, ISO, and other styles
36

Gray, Jonathan, Senani Karunaratne, Thomas Bishop, Brian Wilson, and Manoharan Veeragathipillai. "Driving factors of soil organic carbon fractions over New South Wales, Australia." Geoderma 353 (November 2019): 213–26. http://dx.doi.org/10.1016/j.geoderma.2019.06.032.

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

Bakker, Derk, Grey Poulish, and Stephen Davies. "Water repellence and productivity of lateritic gravelly podosols in South West Australia." Geoderma Regional 17 (June 2019): e00223. http://dx.doi.org/10.1016/j.geodrs.2019.e00223.

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

Standish, Rachel J., and Richard J. Hobbs. "Restoration of OCBILs in south-western Australia: Response to Hopper." Plant and Soil 330, no. 1-2 (October 9, 2009): 15–18. http://dx.doi.org/10.1007/s11104-009-0182-z.

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

Lau, I., and M. Verrall. "Acid sulphate soil mapping with hyperspectral imagery at South Yunderup, Western Australia." ASEG Extended Abstracts 2009, no. 1 (2009): 1. http://dx.doi.org/10.1071/aseg2009ab100.

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

Brooker, Peter I. "Modelling spatial variability using soil profiles in the Riverland of South Australia." Environment International 27, no. 2-3 (September 2001): 121–26. http://dx.doi.org/10.1016/s0160-4120(01)00071-x.

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

Tighe, Matthew, Nick Reid, Brian Wilson, and Sue V. Briggs. "Invasive native scrub and soil condition in semi-arid south-eastern Australia." Agriculture, Ecosystems & Environment 132, no. 3-4 (August 2009): 212–22. http://dx.doi.org/10.1016/j.agee.2009.04.001.

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

Muyen, Zahida, Graham A. Moore, and Roger J. Wrigley. "Soil salinity and sodicity effects of wastewater irrigation in South East Australia." Agricultural Water Management 99, no. 1 (November 2011): 33–41. http://dx.doi.org/10.1016/j.agwat.2011.07.021.

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

Venn, Susanna E., and John W. Morgan. "Soil seedbank composition and dynamics across alpine summits in south-eastern Australia." Australian Journal of Botany 58, no. 5 (2010): 349. http://dx.doi.org/10.1071/bt10058.

Full text
Abstract:
Alpine soil seedbanks are generally regarded as small and unimportant to regeneration. Here, we investigate for the first time the composition of the readily germinable soil seedbank across alpine summits in south-eastern Australia. We aimed to compare the species in the seedbank with the standing vegetation, show seasonal variations in seedbank composition and identify regeneration strategies of alpine seedbank species. By using standard glasshouse and cold-stratification germination techniques, the germinable soil seedbank across the study region was found to comprise 39 species from 25 families, with species from the Asteraceae the most common. Persistent seedbanks were found across all eight alpine summits (1668–1970 m), comparable in seed density (150 ± 27 to 1330 ± 294 per m2) with those of other alpine areas in the northern and southern hemispheres. The density of germinable seeds varied widely among sites and between collection times (autumn, spring) and there were no trends in seed density with altitude. The qualitative and quantitative similarity between the seedbank species and the standing vegetation was low. Correlations between the proportions of species in regeneration categories (from obligate seeders, through to vegetative regenerators) in the standing vegetation and the seedbank were also poor. Our results indicate a divergence between the species in the current standing vegetation and those present in the readily germinable soil seed bank. The current patterns and predominance of seed-regenerating species in the seedbank indicate that these species may have an important role to play in regulating and contributing to future changes in the vegetation assemblage.
APA, Harvard, Vancouver, ISO, and other styles
44

Shi, Xian-Zhong, Mehrooz Aspandiar, and David Oldmeadow. "Reflectance spectral characterization of acid sulphate soil in South Yunderup, Western Australia." International Journal of Remote Sensing 35, no. 10 (April 24, 2014): 3537–55. http://dx.doi.org/10.1080/01431161.2014.907938.

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

Maliki, Ali Al, David Bruce, and Gary Owens. "Spatial distribution of Pb in urban soil from Port Pirie, South Australia." Environmental Technology & Innovation 4 (October 2015): 123–36. http://dx.doi.org/10.1016/j.eti.2015.05.002.

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

Jankowski, Nathan R., Zenobia Jacobs, and Paul Goldberg. "Optical dating and soil micromorphology at MacCauley's Beach, New South Wales, Australia." Earth Surface Processes and Landforms 40, no. 2 (August 5, 2014): 229–42. http://dx.doi.org/10.1002/esp.3622.

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

Singh, B., and RJ Gilkes. "Phosphorus sorption in relation to soil properties for the major soil types of South-Western Australia." Soil Research 29, no. 5 (1991): 603. http://dx.doi.org/10.1071/sr9910603.

Full text
Abstract:
The P sorption characteristics of 97 soils that are representative of the agricultural areas of Western Australia were described using Langmuir and Freundlich equations. The Langmuir P maximum (xm) ranged from 11 to 2132 �g g-1 soil and the Freundlich k coefficient ranged from 1 to 1681. Clay content, DCB Fe and Al, oxalate Fe and AL, and pyrophosphate Al were positively related to xm and k. By using stepwise regression analysis, the combination of DCB and oxalate-soluble A1 predicted more than 75% Of the variation in the P sorption coefficients. Reactive Al compounds may thus be responsible for much of the P sorption by these soils. Soil pH in 1 M NaF (pH 8.2), which is normally used for the detection of allophanic material, was strongly related to the P sorption coefficients and might therefore be used as a quick test for predicting the P sorption capacity of soils.
APA, Harvard, Vancouver, ISO, and other styles
48

Rabbi, S. M. Fazle, Matthew Tighe, Annette Cowie, Brian R. Wilson, Graeme Schwenke, Malem Mcleod, Warwick Badgery, and Jeff Baldock. "The relationships between land uses, soil management practices, and soil carbon fractions in South Eastern Australia." Agriculture, Ecosystems & Environment 197 (December 2014): 41–52. http://dx.doi.org/10.1016/j.agee.2014.06.020.

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

Odgers, Nathan P., Karen W. Holmes, Ted Griffin, and Craig Liddicoat. "Derivation of soil-attribute estimations from legacy soil maps." Soil Research 53, no. 8 (2015): 881. http://dx.doi.org/10.1071/sr14274.

Full text
Abstract:
It is increasingly necessary to apply quantitative techniques to legacy soil polygon maps given that legacy soil maps may be the only source of soil information over large areas. Spatial disaggregation provides a means of extracting information from legacy soil maps and enables us to downscale the original information to produce new soil class maps at finer levels of detail. This is a useful outcome in its own right; however, the disaggregated soil-class coverage can also be used to make digital maps of soil properties with associated estimates of uncertainty. In this work, we take the spatially disaggregated soil-class coverage for all of Western Australia and the agricultural region of South Australia and demonstrate its application in mapping clay content at six depth intervals in the soil profile. Estimates of uncertainty are provided in the form of the 90% prediction interval. The work can be considered an example of harmonisation to a common output specification. The validation results highlighted areas in the landscape and taxonomic spaces where more knowledge of soil properties is necessary.
APA, Harvard, Vancouver, ISO, and other styles
50

Tighe, M., N. Reid, B. R. Wilson, and M. T. McHenry. "High soil acidity under native shrub encroachment in the Cobar Pediplain, south-eastern Australia." Rangeland Journal 40, no. 5 (2018): 451. http://dx.doi.org/10.1071/rj17124.

Full text
Abstract:
This study investigated the chemical characteristics of shallow (0–30 cm) soil profiles under shrubs in areas of dense encroachment and compared them with shallow soil profiles under nearby large trees. Consistent patterns of high soil acidity were found under shrubs, as well as lower litter alkalinity, lower relative concentrations of calcium (Ca2+), lower effective cation exchange capacity, and higher aluminium (Al3+) and sodium (Na+) in the soil profile compared with under trees. Soil pH (CaCl2) was strongly correlated with the Ca content of surface litter. These findings suggest that shrubs (which at most sites included the shrub form of tree species) cycle alkalinity differently from large and mature trees, resulting in high acidity in the shallow soil profile acidity, and possible loss of alkalinity via surface movement of material from areas of dense encroachment.
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