首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Five variants of the distribution of clay (<0.001 mm) and physical clay (<0.01 mm) fractions along the vertical profiles of Vertisols (slitozems) and vertic soils (slitic subtypes of different soil types) from the European part of Russia are distinguished: (1) accumulative, (2) even, (3) regressive, (4) with a maximum in the middle-profile horizon and with their approximately equal contents in the upper and the lower horizons, and (5) eluvial–illuvial. These distribution patterns are related to the lithological specificity of sedimentation and formation of parent materials composed of swelling clays of different geneses and ages. Solonetzic, eluvial- gley, and solodic processes contribute to the development of the eluvial–illuvial and, partly, regressive variants of clay distribution. All the five variants with a predominance of the even distribution pattern can be found in Vertisols. Most of Vertisols in the European part of Russia have a medium clayey or a heavy clayey texture in the entire profile. The regressive distribution pattern is typical of the group of vertic soils. In the upper horizons of Vertisols, where slickensides do not form, the texture is usually heavier than that in the analogous horizons of vertic soils. The middle-profile and lower horizons with slickensides have similar statistical distributions of particle-size fractions in Vertisols proper and in vertic soils. However, in Vertisols, a tendency for a more frequent occurrence of the soils with a higher content of the clay fraction and with a higher portion of this fraction in the physical clay fraction is observed (as compared with the vertic soils).  相似文献   

2.
An analysis of the spatial distribution of the sodium adsorption ratio (SAR) values in the upper subsurface horizons of Vertisols made it possible to estimate the degree of sheet erosion on the dissected coastal plain composed of the sodium-bearing montmorillonite clay in Israel. The depth of the appearance of slickensides in the profile of the Vertisols was used as an additional criterion of the disturbance of these soils by erosion. The approximate rates of erosional dissection of the local landscapes after the deposition of montmorillonite clay are discussed.  相似文献   

3.
《CATENA》2001,43(3):231-249
The Purna Valley, a part of the Payanghat Plain, is an oval basin covering an area of about 1.9 Mha in central India. The soils are deep, calcareous, clayey and very dark greyish brown to dark brown in colour, and meet criteria for the Vertisols order of Soil Taxonomy. Cracks extend down to the slickenside zones in soils of the northeastern part of the valley, but cut through the slickenside zones in the soils of the southwestern part. Soils of the southwestern part are strongly alkaline with exchangeable sodium percentages (ESPs) of 5–26, whereas those of the northeastern part are moderately alkaline and have ESP values <5. Despite their similar coefficients of linear extensibility, volumetric shrinkage potentials, clay contents and amounts of fine clay smectite, the plasmic fabric of the slickenside horizons in soils of the northeastern part is porostriated, whereas in soils of the southwestern part, it is stipple-speckled to mosaic-speckled, indicating weak plasma separation. The soils have both pedogenic and nonpedogenic calcium carbonate, but those of the southwestern part have more pedogenic carbonate than those in the northeastern part. The semiarid climate removes Ca2+ ions from the soil solution by precipitating carbonate, and also causes the ESP and sodium absorption ratio to increase with depth. The lack of soil water in soils of the southwestern area is thought to be the reason for weak swelling of smectite, for larger amounts of pedogenic carbonate and for cracks cutting through the slickenside zones. The Vertisols of the southwestern part of the of the Purna Valley are, therefore, polygenetic because they exhibit distinctly different soil properties resulting from increasing aridity in the valley during the late Holocene.  相似文献   

4.
5.
Earlier studies on soil degradation in Vertisols of the Purna Valley of central India indicated that the semiarid climate characterized by a mean annual rainfall (MAR) of 875 mm and a tropustic moisture regime is responsible for the development of calcareous sodic soils. Recent observations, however, indicate that in the adjacent east upland of the Purna Valley, namely in the Pedhi Watershed, Vertisols have drainage problems, although the area receives a higher MAR than the Purna Valley, the total MAR being 975 mm. The Pedhi Watershed covers an area of 44 321 ha, and is characterized by a tropustic moisture regime and a hyperthermic temperature regime. Vertisols of the Pedhi Watershed are deep, calcareous, clayey and very dark greyish‐brown to dark yellowish‐brown in colour. Vertisols occur on both microhigh (MH) and microlow (ML) positions. The distance between the MH and ML positions is approximately 6 km and the elevation difference is 0.5–5 m. Cracks > 1 cm wide extend down to the slickenside zones in soils of ML whereas they cut these zones in some soils in MH positions. The soils of the MH positions are strongly alkaline and those of ML are mildly alkaline. The present study attempts to relate the distinctly different morphological and chemical properties of Vertisols in the MH and ML positions to pinpoint the prime factor responsible for the impairment of drainage on the basis of physical, chemical, mineralogical and micromorphological data. These data were obtained from 13 Vertisol pedons of methodically selected sites in the Pedhi Watershed. Despite their similar coefficient of linear extensibility (COLE), volumetric shrinkage potential (VSP), clay contents and amounts of fine smectite clay, the plasmic fabric of the slickenside horizons in soils of ML is porostriated, whereas in soils of MH it is stipple speckled to mosaic speckled, indicating weak plasma separation. The soils have both pedogenic and non‐pedogenic calcium carbonates (CaCO3). The semiarid climate induces the precipitation of CaCO3 with a concomitant development of subsoil sodicity. The degree of development of sodicity (Exchangeable sodium percentage (ESP) ≥ 5) is more in soils of MH as evidenced by the higher amount of pedogenic CaCO3 (PC). The lack of water in soils of MH position is the reason for weak swelling of smectite, for larger amounts of PC, for higher alkalinity and sodicity, and for cracks cutting through the slickenside zones. Formation of sodic Vertisols in MH alongside non‐sodic Vertisols in ML positions is a unique phenomenon. It develops because of microtopographic differences which modify distribution of water across the landscape and facilitate greater penetration of rainwater in ML positions. The development of sodicity due to microtopographic differences assumes a great importance when a future land resource management programme on Vertisols in the higher MAR zone of overall a semiarid climate is considered. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

6.
Moralistic soils (vertic and non-vertic black clays) were sampled along a 200 m top sequence in the Transvaal Highveld, South Africa. The Milk wood soil on the upper part of the margalitic top sequence has an Al horizon which lacks both slickensides and self-mulching properties but has strong blocky structure, the others are all self-mulching (i.e. Vertisols). The Vertisols differ in some morphological properties such as the absence (Mngazi series) and presence (Arcadia series) of pedogenic calcite accumulation, both overlying decomposed dolerite, whereas the third member from a red-black centenary sequence is calcareous with a gleyed subsoil horizon. Chemical and XRD analysis suggest that the clays of the Vertisols comprise mainly a partly chloritized iron-rich smectite with Ie > Mg + Al in the octahedral layer (and interlayer positions), while in the upland Milkwood soil intergradient halloysitic clay occurs in addition to a more‘open’structured chloritized iron smectite. The mineralo-chemical results and the geomorphic data suggest that the genesis of these soils is controlled primarily by the internal soil water regime.  相似文献   

7.
Using an aluminum (Al)‐sensitive maize cultivar, we investigated the influence of Al on the apoplastic solute bypass flow and its relationship with Al‐induced (1 h, 50 μM) callose formation and root growth. We selected the fluorescent probes 8‐hydroxypyrene‐1,3,6‐trisulfonic acid, trisodium salt (MW 524) (HPTS) and dextran‐Texas Red (TR) conjugates (MW 3,000, 10,000, and 40,000) to monitor their apoplastic transport. Confocal laser–scanning microscopy (CLSM) analysis and spectrofluorometric quantification showed Al‐induced callose formation in peripheral root cells within 1 h. Pretreatment of plants with the callose synthesis inhibitor 2‐deoxy‐D‐glucose (DDG) reduced the callose formation by half. Uptake experiments with both HPTS and dextrans showed uniform dye distribution in control root apices. After Al treatment for 1 or 2 h, which inhibited root growth by 32% or 50%, respectively, the dyes accumulated in the epidermal and outer cortical cell layers, especially in the 1–2 mm apical root zone. Al treatment reduced the export of the dyes out of the apical 1 cm treatment zone. This was due to strong sorption of HPTS but not of dextrans by Al‐loaded cell walls. Aluminum treatment reduced loading into the xylem sap particularly of higher–molecular weight dextrans. Pretreatment of roots with DDG and presence of 50 mM mannitol during the Al treatment partially forestalled the inhibitory effect of Al on the dye transport, but only slightly reduced the Al‐induced growth inhibition. Exudation experiments revealed that xylem water flow remained unaffected by the Al treatment of the root tips. The results with dextran suggest that Al binding in cell walls of the root apex inhibits apoplastic bypass flow of higher–molecular weight solutes, which might contribute to Al‐induced inhibition of root growth.  相似文献   

8.
The Vertisols of the Purna Valley, which cover the districts of Amravati, Akola and Buldhana in the state of Maharashtra, India, lack any perceptible evidence of salt efflorescence on the soil surface which would indicate the presence of salt, but the drainage conditions are poor. The limited data available indicate that the adverse physical condition of the soils is due to their poor hydraulic conductivity (HC), which is impaired by sodium in the exchange complex. However, the factors and processes that are inherently related to the development of sodicity in these shrink-swell soils are not yet understood. In order to establish the cause-effect relationship, eight Vertisol pedons from methodically selected sites were studied morphometrically in the field, and for their sodicity-related physical and chemical properties in the laboratory. The soils are deep, calcareous, clayey and very dark greyish brown to dark brown in colour. Cracks extend up to the slickenside zones in soils of Pedons 1–3 in the northeastern area of the valley, while they cut through the slickenside zones in soils of Pedons 4–8 in the southwest. The slickenside faces were larger in the soils of the southwest than in those of the northeast. All these soils meet the specifications of the Vertisols order of soil taxonomy. Saturation extracts of the soils had very low electrical conductivity (ECe ⩽ 2 ds m−1). They ranged from moderately alkaline in the northeast, Pedons 1–3, to strongly alkaline in the southwest, pedons 4–8. In soils from the northeast the exchangeable sodium percentage (ESP) was less than 5 throughout the depth of the pedons, whereas in other soils it was up to 6 in the surface horizons and between 7 and 26 in the subsoil horizons; four of these soils qualified as sodic according to the criteria of the United States Salinity Laboratory. The inherently low hydraulic conductivity was due to the dispersion of clay particles caused by a high percentage of exchangeable magnesium (EMP) in the highly smectitic soils, and also to a slight increase in ESP (⩾5). The results of this study suggest that ESP 5 should be used as the lower limit for sodic subgroups of Vertisols, rather than ESP 15 as given in Keys to Soil Taxonomy (Soil Survey Staff, 1994). This is because there are severe limitations to the use of such soils owing to the development of adverse physical conditions even at such a low ESP. The authors emphasize the need to keep this fact in mind during future land resource management programmes on the soils of this valley as well as on similar soils occurring elsewhere. The development of sodicity in the soils of the southwestern part of the valley has been attributed to the semi-arid climatic conditions that have induced the pedogenetic process of depletion of calcium ions from the soil solution in the form of calcium carbonate, thereby resulting in an increase of both the sodium adsorption ratio (SAR) and the ESP with pedon depth. This chemical degradation, which affects the sodicity of Vertisols, appears to be a basic process that needs to be recognised in the future along with those already described as natural processes of soil degradation.  相似文献   

9.
10.
The attachment of Escherichia coli and Streptococcus bovis of bovine faecal origin to soil clay particles has been examined using scanning electron microscopy. Both organisms produce extracellular polymers through which the cells form bridges to clay surfaces. Differences in the extent of cell-to-particle bridging between E. coli-and S. bovis-soil complexes are explained in terms of the charge characteristics of the cells and the properties of the exopolymers.  相似文献   

11.
At present precise information on the b-fabric of Vertisols is inadequate for gaining a comprehensive knowledge about such soils formed in alluvium of basic igneous and metamorphic rocks and spread in varying agro-climatic zones of peninsular India. The aim of the present study, on five benchmark Vertisol series (Sarol, Aroli, Nimone, Bellary and Kovilpatti), was to assess the possible reasons for differences in the b-fabric of the subsoils despite their common shrink-swell properties, as evidenced by the presence of sphenoids and/or slickensides, and similar clay contents and amounts of fine clay smectite. In Sarol and Aroli soils from the sub-humid region, the b-fabric is porostriated, whereas in Nimone and Kovilpatti soils of the semi-arid and Bellary soils of the arid regions the fabric is either mosaic-speckled or granostriated, indicating weak plasma separation, a result of restricted swelling of clays. The related distribution pattern of the s-matrix in all these soils was open porphyric. Recent literature has suggested that weak plasma separation is an effect of dissolution and crystallization of calcite. However, generally low concentrations of soluble calcium (c. 1 mmol dm?3) in soil solutions of Vertisols in arid and semi-arid agroclimates suggest that the presence of calcite does not guarantee the presence of sufficient soluble calcium to impair swelling of clays by contracting the diffuse double layer. Weak plasma separation in Bellary and Kovilpatti soils compared with Sarol and Aroli soils is explained in terms of a decrease in the internal surface area of fine smectite, as evident from a very high degree of chloritization in smectite interlayers, and smaller proportions of coarse smectite. In Nimone soils the weak plasma separation is primarily due only to a decreased internal surface area of fine smectite.  相似文献   

12.
Structured subsoil horizons are characterized by biopores and shrinkage cracks, which may serve as preferential flow paths. The surfaces of cracks and biopores may be coated by clay‐organic material. The spatially‐distributed organic matter (OM) composition at such structural surfaces was studied at the millimetre scale using diffuse reflectance infrared Fourier transform (DRIFT) spectroscopy in the mid‐infrared range (MIR). Intact biopores such as earthworm burrows and root channels, and crack surfaces of nine subsoil horizons were analysed. The samples were from arable and forest Luvisols, one Regosol, one Stagnosol and Cambisols developed from loess, till, mudstone and limestone. For better comparison between soils, the DRIFT signal intensities were corrected for the particle‐size effects. The OM was characterized by the ratio between alkyl‐ (C–H) and carbonyl (C=O) functional groups (C–H/C=O), which represent an index of the potential wettability (PWI) of the OM. The PWI was larger for biopores than for crack surfaces and the soil matrix, indicating a smaller potential wettability of OM at biopore surfaces. The millimetre‐scale spatial variability of OM was especially large for the surfaces of root channels. Samples from till‐derived Luvisols had smaller PWI (with greater potential wettability than surfaces from loess‐derived Luvisols) than other soil types. The mean PWI of the arable Luvisol crack surfaces was less than that of the forest Luvisol samples. The results suggest that the spatial distribution of OM properties at intact structural surfaces may be important for describing sorption and mass transfer processes during preferential flow.  相似文献   

13.
引入分维理论评价土壤对优势流的敏感性   总被引:5,自引:0,他引:5  
Food dye Brilliant Blue was introduced as the tracer in a dye-tracing experiment to obtain dye profile patterns of sandy loam soil,aeolian sandy soil,percolating paddy soil and permeable paddy soil.The dyed soil profiles were then photographed and the photos were scanned into a computer.Edited with certain software,only the dyed areas were left on the profile photos,which indicted the preferential flow paths for water and solute transport.Fractal dimensions of the dye patterns were calculated according to Arnold‘s function.Soil particle size distribution was analyzed by pipette method.The regression analysis showed that there was significant relationship between soil clay content and fractal dimension D of the dye pattern of soil profile.Based on the experiment results,the possibility of introducing fractal dimension to estimation of soil sensitivity to preferential flow is discussed.  相似文献   

14.
综述了植物根孔的研究方法,并从植物根孔对水分和溶质运移的影响、对土壤吸附和解吸过程的影响和对土壤微生物的影响等方面详细阐述了植物根孔功能研究的进展情况,指出了植物根孔研究方面存在的问题和今后应加强的研究领域。  相似文献   

15.
On the basis of soil studies along routes and on key plots, 35 new areas of soils with definite features of vertigenesis have been identified in Belgorod and Voronezh oblasts and in the northern part of Volgograd oblast (in the Don River basin). Earlier, vertic soils were not noted for these areas. In the studied region, their portion in the soil cover is much less than 1%. All the delineated areas of vertic soils are confined to the outcrops of swelling clay materials of different origins (marine, lacustrine, glacial, and colluvial sediments) and ages (Quaternary or Tertiary) that may be found in four landscape positions: (1) in the deep closed depressions within vast flat watersheds; (2) in the bottoms of wide hollows on interfluvial slopes and, sometimes, on steeper slopes of local ravines; (3) in the hydromorphic solonetzic soil complexes, and (4) on step-like interfluvial surfaces with the outcrops of Tertiary clays. Within the studied areas, soils with different degrees of expression (six grades) of vertic properties are present. These soils belong to the type of dark vertic soils proper and to vertic subtypes of different soil types according to the Russian soil classification system; according to the WRB system, they belong to Vertisols proper and to reference soil units with a Vertic prefix in the groups of Chernozems, Phaeozems, and Solonetzes. Statistical data on the morphometric indices of the vertic properties (the depth and thickness of the soil horizons with slickensides, a wedge-shaped structure, and cracks filled with material from the upper horizons) and the depth and thickness of the Vertic horizon are analyzed.  相似文献   

16.
福建省变性土特性的研究   总被引:6,自引:4,他引:6  
在福建沿海低丘、台地玄武岩风化物上发育一种暗色粘质土壤,历史上归属于砖红壤性红壤。本研究证明,它明显区别于砖红壤性红壤,主要表现于:风化--淋溶作用和脱硅富铝化作用较弱,通体暗色,层次分异不明显,粘粒含量高,粘土矿物以蒙脱石占优势,有明显膨胀收缩性,干季出现一定深度的开裂,阳离子交换量大,盐基饱和度和pH值较高。据此,这种土壤有必要从砖红壤性红壤中划分出来。按美国土壤系统分类,本土壤定名为细质、蒙脱石型、高热、典型艳色湿润变性土。可以认为,变性土在我国也应作为一个独立的分类单元。  相似文献   

17.
L.A. Sullivan  A.J. Koppi 《Geoderma》1987,40(3-4):317-332
The in-situ distribution and morphology of organic materials located on, or near, structural surfaces within some soils is examined on a submicroscopic scale by comparing scanning electron microscope images of the same areas of soil samples before and after low temperature ashing. Electron-translucent organic matter coatings up to 1/2 μm thick, and thinner, electron-opaque organic matter coatings were found on structural surfaces within these soils. Oribatid faecal pellets in one of the soils were found to contain aluminosilicate clay minerals. Fine-clay sized spheres of biogenic opaline silica were found to be contained within the epidermis of a decaying root. These studies show that the combined use of low temperature ashing and scanning electron microscopy will be a valuable technique for in-situ investigations of submicroscopic organic matter within soils.  相似文献   

18.
Specific features of a gley soil developed on varved lacustrine clay have been studied by scanning electron microscopy and electron probe microanalysis. Individual clay platelets can be clearly observed in the red-brown parent material but they cannot be resolved in the gleyed surfaces which have a smooth skin-like texture. Ochreous mottles are shown to have a porous blocky structure, the surfaces of the blocks being coated with globular iron oxide. Areas of manganese staining have a characteristic sponge-like morphology.  相似文献   

19.
黄淮海平原三种土壤中优势流现象的试验研究   总被引:14,自引:0,他引:14  
在黄淮海平原选择三种不同质地类型的土壤(砂壤土、淤土、风沙土),各设1.5m×1.5m两个小区,其中一个小区在试验开始前两癸灌水约56kg,以获得不同的初始含水量。将100mm含有染色剂亮蓝的水灌入小区,一天之后,开挖1m×1m的剖面拍照,进行图像分析处理,计算剖面中染色百分比随深度的变异,研究所形成的优势流的状况与土壤类型、土壤初始含水量的关系。结果表明:不同土壤发生优势流的程度不同,结构发育好  相似文献   

20.
土壤优先流运动的活动流场模型模拟和敏感性分析   总被引:3,自引:1,他引:2  
在壤土和砂土条件下分别采用碘-淀粉染色示踪方法和亮蓝染色示踪方法各开展了2个染色示踪试验,分别采用活动流场模型和二域模型模拟计算了各试验入渗后染色区内的土壤含水率和溶质浓度分布,通过相对均方根误差分析评价了两个模型模拟预测优先流发展的有效性;此外,通过敏感性分析研究了不同入渗条件(土壤质地、入渗水量和土壤初始含水率)下活动流场模型模拟预测结果(入渗深度)对活动流场模型分形特征参数变化的敏感度。模型检验分析结果显示活动流场模型对土壤水入渗深度、入渗后染色区内土壤含水率和溶质浓度分布的预测精度要明显高于二域模型的模拟预测精度;活动流场模型较好的捕捉到了优先流运动整体的非均匀特征。敏感性分析结果显示,当降雨入渗水量和土壤初始含水率相同时,入渗深度对活动流场模型分形特征参数(γ)的敏感度随着γ的增大而增大;相同活动流场模型分形特征参数(γ)值条件下(即流动非均匀程度相同),入渗深度对活动流场模型分形特征参数(γ)的敏感度随着入渗水量的增大和土壤初始含水率的升高而减小。  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号