Sie sind auf Seite 1von 12

Classification and geochemistry of arid and semi-arid paleosols

Veronika Geiler, TU Bergakademie Freiberg

Abstract. Many authors (Mack et al. 1993, Retallack 1998) argued, that for the classification of paleosols it is useful to operate with the existing soil taxonomy, because then analogies between paleosols and recent soils can be drawn. Comparison is especially helpful for identifying conditions of formation, which are precipitation, temperature, parent material, morphology of landscape and time. These information can be employed for the reconstruction of paleo-environments. Classification is needed for soil types (paleosols respectively), but today the huge amount in use complicates communication and comprehension. The attempt of this paper is to give an overview of the most important soil types (after Soil Survey Staff 1975) for semi-arid and arid climates, adding common synonyms applied in other classifications.

Introduction
There are several approaches to paleosols. From a sedimentological point of view, they represent a gap in sedimentation. Furthermore paleosols can be employed as lithostratigraphic markers or to define paleo-morphology. In this paper special emphasis is laid on the possibilities paleosols offer for the reconstruction of paleo-climate. The formation of soils is mainly influenced by agents as precipitation, temperature, parent material, morphology of landscape and time. Therefore paleosols provide an excellent reservoir of information on weathering and climate conditions.

Classification
Not all current classifications are suitable for the application on paleosols. It is necessary, that classification is based on criteria, which can be observed in paleosols. Therefore U.S. Soil Taxonomy seems to be an appropriate choice. The twelve orders differentiated by U.S. Soil Survey Staff are not directly defined due to their climatic occurrence, but can be related to it (see Table 1). The marked soil types will be discussed in detail in the following paragraphs.
Table 1: Soil types with description and relation to climate Soil type Description (Soil Survey Staff 1975) (Retallack 1998) Aridisol desert soil Mollisol grassland soil Vertisol swelling clay soil Alfisol fertile forest soil Ultisol base-poor forest soil Spodosol sandy forest soil Oxisol tropical deeply weathered soil Histosol peaty soil Gelisol permafrost soil Entisol incipient soil Inceptisol young soil Andisol volcanic ash soil Climate (Retallack 1998) arid semi-arid - subhumid seasonal dry semi-arid - subhumid humid humid humid humid nival no indication no indication no indication

Veronika Geiler, TU Bergakademie Freiberg

Aridisol
Common synonyms for Aridisols are Cryids, Salids, Durids, Gypsids, Argids, Calcids, Cambids (all United States Department of Agriculture, in the following abbreviated by USDA), Yermosol (Food and Agricultural Organisation, in the following FAO), Xerosol (FAO), Calcixeralfs (USDA), Calcisol and Gypsisol.

a)

Fig. 1: Aridisol: a) photography of typical Aridisol (from www.wikipedia.de ) b) class sketch, for legend see Fig. 2, climate is represented by sun/ cloud symbols, time of formation by an hourglass (from Retallack 1997) c) sketch of petrography and soil texture of horizon A in thin section (from Retallack 1997)

Fig. 2: Legend for class sketches (from Retallack 1997)

Aridisols are desert soils. Depending on the parent material, a relative short time of development is required. They are common in arid and semi-arid climates, where as a consequence of the lack of rainfall easy soluble salts can endure. The surface horizon shows a light color. Its texture is soft and often vesicular (McFadden et al. 1998). Salinization can occur, if salts accumulate on the surface (www.wikipedia.de). Characteristic for the subsurface layers are shallow calcareous (Bk), gypsiferous (or anhydritic) (By) or salty (Bz) strata (see Fig. 1), which represent the depth of wetting during occasional rainfall and can therefore be correlated to the amount of mean annual precipitation (see also Depth to calcic horizon) (Retallack 1997). The appearance of cemented horizons varies from needle-fibre calcite over nodules to distinct hardpan calcretes (Wright and Tucker 1991) (See also Calcic horizon as indicator for the maturity of paleosols). Not cemented or clayey subsurface layers are also common. The material for the formation of either clayey, calcareous or salty horizons respectively derives from weathering or eolian deposition. After the latter extremely fine grained dust of easily weatherable minerals flushes down the profile and accumulates. On Aridisols only sparse vegetation grows, which includes prickly shrubs and cacti. Therefore root traces are not particularly common in paleosols. In general Aridisols have a very poor concentration of organic matter. Geomorphical settings for Aridisols are low lying areas, since on slopes erosion would prevent soil maturation (Retallack 1990).

Classification and geochemistry of arid and semi-arid paleosols

Mollisol
For Mollisols the following synonyms are used: Albolls, Aquolls, Rendolls, Cryolls, Xerolls, Ustolls, Udolls (all USDA), Chernozem, Solonetz (Canadian system of soil classification) Kastanozem, Phaeozem, Rendzinas, Greyzem and Chernozem (all FAO).

a)

Fig. 3: Mollisol: a) photography of typical Mollisol (from www.wikipedia.de ) b) class sketch, for legend see Fig. 2, climate is represented by sun/ cloud symbols, time of formation by an hourglass (from Retallack 1997) c) sketch of petrography and soil texture of horizon Bk in thin section (from Retallack 1997)

Mollisols form in grassland environments. Therefore their occurrence in the geological history is connected to the evolution of grass in Tertiary. Their development takes place in relative short time, of course in dependence of their host material. For Mollisols base-rich parent materials such as clay, marl or basalt are common. They are frequent under subhumid - semi-arid climate conditions. The surface layer is base-rich and consists of a mixture of clay and organic matter, which is not carbonaceous or coaly (Retallack 1990). This A horizons usually shows granular or crumb bed structures (Mack et al. 1993). Other significant criteria of Mollisols are abundant fine root traces of grassy vegetation and burrows of diverse populations of invertebrates. The subsurface can be either argillic (Bt), calcareous (Bk) or gypsiferous (By), but mostly an A-Bk-profile is found (See Fig.3). Mollisols form under a wide range of temperature in lowland to high mountain meadows (Retallack 1990).

Vertisol
For Vertisols a lot of local names exist as synonyms: cracking clays (Australia), Adobe (Philippines), Shachiang (China), Black Cotton Soils (India), Smolnitza (Bulgaria, Rumania), Tirs (Marocco), Makande (Malawi), Vleigrond (Southafrica), Sonsosuite (Nicaragua), Margalite soils (Indonesia), Densinegra soils (Angola), Grumusols (United States), in FAO also named Vertisol, Aquerts, Cryerts, Xererts, Torrerts, Usterts and Uderts (all USDA) Vertisols are defined as swelling clay soils. Depending to their parent material, which is commonly of intermediate or basaltic composition, the time span for formation varies extremely between hundreds and thousands of years. But they still require less time for generation than Aridisols and Mollisols. Vertisols occur in subhumid semi-arid climates with a pronounced dry season. Precipitation must be enough to form clays, which means about 180-1520mm mean annual rainfall. Vertisols are characterized by a uniform, at least half meter thick clayey horizon, which shows desiccation cracks during the dry season (Retallack 1990). The shrinking and swelling results in constant mixing, which causes Vertisols to have an extremely deep A horizon and no B horizon (Soil Survey Stuff 1999). The

Veronika Geiler, TU Bergakademie Freiberg

a)

Fig. 4: Vertisol: a) photography of typical Vertisol (from www.wikipedia.de ) b) class sketch, for legend see Fig. 2, climate is represented by sun/ cloud symbols, time of formation by an hourglass (from Retallack 1997) c) sketch of petrography and soil texture of horizon Bt in thin section (from Retallack 1997)

desiccation cracked surface forms a typical hummock-and-swale topography, also known as gilgai microrelief (See Fig.9). The corresponding subsurface also displays the disrupted, festoon-shaped texture (Mack et al. 1993). Slickensided profiles with internal deformations of horizons are frequent (Retallack 1990) (See Fig.4). Responsible for this appearance is the high content of clays (50-70%) and its swelling properties respectively. The clays consist predominantly of 2:1 and 2:2 layer clay minerals. Their color ranges from grey to red or the more familiar black (Soil Survey Staff 1999). Because of the dry climate and the sparse vegetation alkaline reactions and good reserves of exchangeable cations can be maintained (Driese et al. 2000). As geomorphological setting mainly flat regions and the feet of gentle slopes are possible (Retallack 1997).

Alfisol

Fig. 5: Alfiisol: a) photography of typical Alfisol (from www.wikipedia.de ) b) class sketch, for legend see Fig. 2, climate is represented by sun/ cloud symbols, time of formation by an hourglass (from Retallack 1997) c) sketch of petrography and soil texture of horizon Bt in thin section (from Retallack 1997)

Classification and geochemistry of arid and semi-arid paleosols

Synonyms for Alfisols are: Luvisols, Nitosols, Solonetz (all FAO), Aqualfs, Cryalfs, Udalfs, Ustalfs and Xeralfs (all USDA). Alfisols are defined as fertile forest soils. The prefix alf derives in this case not from the word pedalfer (Marbut, 1935, see also Pedocals and pedalfers), but refers to aluminium (Al) and iron (Fe). For generation Alfisols require relative short time. Their parent material is mainly base-rich as the soil itself. Alfisols are common in subhumid semi-arid climates. The surface layer is of light color. The subsurface horizon is mostly argillic (Bt) (See Fig. 5). Especially base-rich clays (smectites) are abundant (Retallack 1990). Alfisols represent the youngest forest soil order. Because of this they are less leached, and have a greater than 35% base saturation (www.wikipedia.de). Therefore the soil contains lots of exchangeable cations. A geochemical characteristic of Alfisols are molecular weathering ratios of alumina/ bases of less than two. Vegetation of Alfisols ranges from grassy woodland to open forest. The topography, in which Alfisols form, varies (Retallack 1990).

Common subordinate modifiers of paleosols (from Mack et al. 1993)


A finer classification of paleosols is not only possible by employment of subgroups, but also by application of subordinate modifiers (see Table 2).
Table 2: Common subordinate modifiers of paleosols (Mack et al. 1993) Presence of eluvial (E) horizon albic Presence of allophone or other Si and Al compounds allophanic Presence of illuvial clay argillic Presence of pedogenic carbonate calcic Presence of dark organic matter, but not coal carbonaceous concretionary Presence of glaebuls with concentric fabrics Low base status as indicated by the paucity of chemically unstable grains such as feldspar dystric and volcanic rock fragments High base status as indicated by the abundance of chemically unstable grains such as eutric feldspar and volcanic rock fragments Presence of iron oxides ferric Subsurface horizon that was hard at the time of soil formation (for example root traces and fragic burrows terminate or are diverted at this horizon) Evidence of periodic waterlogging, such as drap hues; mottles of drap color and yellow, red gleyed or brown; or presence of pedogenic pyrite or siderite Presence of vadose gypsum or anhydrite gypsic Presence of glaebuls with an undifferentiated internal fabric nodular Presence of light-colored A horizon ochric Presence of pedogenic salts more soluble than gypsum salic Presence of pedogenic silica silicic Presence of decimeter-scale desiccation cracks, wedge-shaped peds, hummock and swale vertic structures, slickensides, or clastic dikes Vitric Presence of relict or actual glass shards of or pumice

Calcic Horizon as indicator for the maturity of paleosols


Soil profiles of arid and semi-arid climates often contain pedogenic calcretes (Bk). They typically occur within Aridisols, Vertisols and Mollisols. In the following table pedogenic calcretes are classified according to their morphology, with the system devised by Netterberg (1967, 1980), which was developed for geotechnical surveys. The morphology of calcretes relates to stages seen in the development of calcrete profiles (Netterberg, 1980). Manchette has provided the most comprehensive sequence and has recognized six stages of

Veronika Geiler, TU Bergakademie Freiberg

development for gravelly parent material. Retallack extended this sequence for sandy and finer grained parent material (see Fig. 6).
Table 3: Morphological classification of calcretes based on Netterberg (1967, 1980) and Goudie (1983) with classification of pedogenic calcretes based on stages of development after Machette (1985). Modified after Quast (2003) Calcrete types Maturation Description (after Netterberg 1980 and Goudie 1983) stages (after Machette 1985) Calcareous soil Very weakly cemented or uncemented soil with small carbonate accumulations as grain coatings, paches of powdery carbonate including needle-fibre calcite (pseudomycelia), carbonate-filled fractures and small nodules Calcified soil A firmly cemented soil, just friable; few nodules. 10-50% carbonate Stage 1 Powder A fine, usually loose powder of calcium carbonate as a continuous body with little calcrete or no nodule development Pedotubule All, or nearly all, the secondary carbonate forms encrustations around roots calcrete or fills root or other tubes (tubules) Nodular calcrete Honeycomb calcrete Hardpan calcrete Laminar calcrete Boulder/ cobble calcrete Stage 2 Stage 3 Stage 4 Stage 5 Stage 6 (syn. glaebular calcrete of Netterberg, 1980) Discrete soft to very hard concretions of carbonate-cemented and/ or replaced soil. Concentrations may occur as laminated coatings to form pisoids Partly coalesced nodules with interstitial areas of less indurated material between (syn. petrocalcic horizon) An indurated horizon, sheet-like. Typically with a complex internal fabric, with sharp upper surface, gradational lower surface Indurated sheets of carbonate, typically undulose. Unsually, but not always, over hardpans or indurated rock substrates Disrupted hardpans due to fracturing, dissolution and rhizobrecciation (including tree-heave). Not always boulder grade. (Clasts are rounded due to dissolution)

Fig. 6: Stages of the morphology of carbonate accumulations in soils a) carbonate stages of Machette for gravelly parent material b) soil development stages of Retallack for sandy and finer grained parent material (from Retallack 1997)

Classification and geochemistry of arid and semi-arid paleosols

Implications of paleoclimate (after Retallack 1990)


Indicators of precipitation
The availability of water plays a major role in many soil-forming chemical reactions, which proceed in dilute solutions. Reacted products are removed or concentrated. Therefore it is no wonder that freely drained soils of humid climates show more profound chemical weathering than those of dry climates. Many chemical and mineralogical characteristics of soils can be associated to precipitation, but here only the most reliable and important are presented.

Pedocals and pedalfer


In soils of dry climates, calcium carbonate is an abundant component, whereas in humid climate it is dissolved, transported away and therefore not found. Marbut (1935) used this observation to divide soils into two main groups. Pedocals have free carbonate, while pedalfers lack it. Although calcareous paleosols are a reliable indicator of dry climate and noncalcareous paleosols of wet climates, it is not possible to specify a dividing line accurately without information on other aspects of paleo-climate (mean annual temperature, season of main rainfall, etc.).

Depth to calcic horizon


Free carbonate in soils usually forms a distinct calcareous layer or calcic (Bk) horizon. The position of this horizon within the soil profile reveals the depth of wetting of the soil by available water. Therefore in dry climates the calcic layer is closer to the surface than in wetter ones (see Fig. 7). But this rule applies only for soils of moderate development, which include nodules of carbonate rather than wisps or layers, in unconsolidated parent material and for soils of seasonal warm climates. And here still some difficulties appear. Erosion before burial falsifies the depth of the calcic horizon, as well as compaction does. A third problem are higher CO2 levels in the past, which lead to a deeper calcic horizon. But only in extreme greenhouse periods of the Jurassic-Cretaceous, OrdovicianSilurian and perhaps also early Precambrian this is a severe problem (Ekart et al. 1999).

Fig. 7: The relationship between mean annual rainfall and depth of the calcic horizon in 317 surface soils from all continents (from Retallck 1998)

Chemical composition
Weathering and soil formation leave characteristic imprints in the chemical composition of soils. Molecular weathering ratios reveal changing chemical proportions as a result of processes or properties like salinization (Na2O/ K2O), calcification ((CaO + MgO)/ Al2O3), clayeyness (Al2O3/ SiO2), base loss (Al2O3/(CaO + MgO + Na2O + K2O)), leaching (Ba/ Sr) and gleization (FeO/ Fe2O3). Fig. 8 shows an example how molecular weathering ratios can contribute to a more precise image of the conditions during soil fomation. Conclusions for climate can be drawn from the degree of depletion of alkali (Na+, K+) and alkali earth (Ca2+, Mg2+) elements normalized to silica and alumina, since in dry climates a lack of water prevents chemical weathering and transport of these elements.

Veronika Geiler, TU Bergakademie Freiberg

Fig. 8: Example of selected molecular weathering ratios for four paleosols in the Upper Hell Creek Formation, Montana (from Retallck 1994)

Strong correlations between the molecular ratio of bases/alumina and mean annual rainfall can be observed in soil subsurface (Bt) horizons (Marbut 1935). But this relationship doesnt hold for paleosols of either very wet (with common kaolinite) or dry climates (with evaporites) and for very strong or very weak developed paleosols, or those altered during burial by illitization.

Clay minerals
Clays, regarded as products of hydrolysis of weatherable minerals, can be linked to the amount of precipitation available for soils (Folkoff & Meetenmeyer 1987). Clay composition in soils is controlled by the grain size and mineral composition of parent materials, temperature, seasonality of rainfall and time for formation of a soil. For paleosols additional problems have occur, that is to say whether the clay formed during soil diagenesis or is inherited from parent material or altered after burial. With theses cautions in mind some generalized statements can be made. In wet climates clay are more likely to possess a 1:1 rather than a 1:2 layer structure. They obtain fewer cations and are lower in general weathering sequence of clay sized minerals. Especially in Vertisols, or generally in dry climates, swellable clays like sodium smectites produce a distinctive soil structure of domed columnar peds extending throughout much of the solum (natric horizon: McCahon & Miller 1997). In Table 4 indicative clay minerals are presented.
Table 4: Indicative clay minerals Dominant clay mineral Indicative for Palygorskitea, sepiolitea Smectite Kaolinite Iron oxides and alumina
a

Very arid climate Mean annual rainfall < 1000mm Mean annual rainfall between 1000 - 2000mm Mean annual rainfall > 2000mm

These clay minerals are easily dissolved and not stable during burial (Botha, Hughes 1992)

Classification and geochemistry of arid and semi-arid paleosols

Evaporate minerals
Evaporate minerals are widespread in climates, in which evaporation exceeds precipitation. Due to the lack of water, evaporate minerals and mainly salts can accumulate respectively, when a source of salts is present. Gypsum is very abundant, but also halite, sylvite or mirabilite. Like in calcic strata, the depth of the gypsic horizons gives a hint to the rate of rainfall. Salts as gypsum are easily soluble in groundwater. As residue of many ancient evaporate horizons only pseudomorphs of crystals or zones of breccia are left, where the overlying rock has collapsed into the dissolved layer zone (Bowles, Braddock 1963).

Desert pavement
When climates is too dry for constant vegetation soils are covered by a natural plaster of stones, developed due to eolian deflation. In such environments windsculpted stones (ventifacts), carbonate collars (rims of calcite around stones at ground level), rock varnish (armorphous iron-manganese crusts on top of stones) and vesicular structure are common (McFadden et al. 1998).

Indicators of temperature
Even though water is the agent for most soil-forming processes, the rate of reactions depends on the temperature. Refering to Van tHoffs rule for temperature, for every 10C increase in temperature the rate of reaction is doubled or even tripled. This coincides well with the observation that in warm or hot regions soils are much more deeply weathered than in cold areas (Birkeland 1992). With rising temperature the activity of the heavy oxygen isotope 18O increases relative to the light isotope 16O. In soil minerals or fossil biominerals the ratio of these isotopes (18O) is preserved. But the temperature of formation can only be concluded from the 18O-proxy, if the isotopic composition of the water from which the mineral precipitated is known. This original water composition is conserved in aragonite shells of bivalve or gastropod. But even if isotopic composition does not reveal absolute temperatures, it can still be useful to detect relative changes in temperature through time. Still caution is necessary, since the oxygen isotopic composition is not stable in the face of elevated temperatures and fluid migration during burial and metamorphism.

Oxygen isotopic composition

Indicators of seasonality
Seasonality produces varying amounts of rainfall, of heat, of dust influx and other agencies of soil formation that are essential to clayeyness, redness and base saturation of soils. But also other factors contribute to these features, therefore it is difficult to tease out the role of seasonality. Here only the most diagnostic features are reviewed.

Mukkara and gilgai


Gilgai microrelief and mukkara subsurface structure (see Fig.9) occur in swelling clay soils as indicators of a climate with pronounced dry and wet season. These structures are not found in extremely arid or humid climates, since at least some moisture is needed to provoke clay formation and provide a contrasting wet season. In humid climates clays become too deeply weathered and lose their swelling properties.

Fig. 9: Sketch of gilgai microrelief and mukkara subsurface structure (Retallack 1997)

10

Veronika Geiler, TU Bergakademie Freiberg

Concretions and argillans


Seasonal differences in the chemical condition of the soil may be reflected in concentric bands of concretions and strong banding of clay skins (argillans) (Nahon 1991). Especially in well-drained soils of subtropical, monsoonal climates calcareous nodules, which develop under alkaline conditions of dry climates, and ferruginous concretions occur, which in contrast formed in the wet season (Sehgal, Stoops 1972).

Patterns of root traces


A characteristic root pattern for seasonally dry climates is recognizable in paleosols. Grasses and trees possess a profuse surficial network of roots that supplies them with water during the wet part of the year. In the dry period grasses weather back to their root stock, whereas some trees obtain moisture through especially stout, deeply penetrating roots, termed sinkers (van Donselaar-ten Bokkel Huinink 1966). Also soils of dry swamps show a similar bimodal distribution pattern.

Carbon isotopic composition


In soils two different carbon isotope compositions are present. The first derives from soil organic matter, which reproduces the annual average composition of vegetative biomass (13Corg). The second reflects the amount of CO2 dissolved during dry-season from root respiration and microbial decay of organic matter (13Ccarb). The difference between both values (13Corg - 13Ccarb) within one soil profile shows the extent of seasonal change in precipitation, especially in strongly seasonal monsoonal climates. The difference is also connected to (i) diffusion of atmospheric CO2 in correlation to atmospheric partial pressure and (ii) equilibrium fractionation of CO2 for the temperature of formation. Seasonal changes can be detected as relative changes in isotopic values from succession of paleosols or modeled by assuming rates for diffusion and fractionation.

Indicators of greenhouse atmosphere


Methane and CO2 can be used as indicators for greenhouse atmosphere. Both gases were found in bubbles within amber (Landis et al. 1996) in paleosols, but these observations are rare due to the scarcity of amber. But their distinctive isotopic trace in the carbon isotopic composition of paleosols is detectable. Minerals like calcite, etc. conserve the isotopic composition, therefore it is possible to reconstruct former atmospheric contents of CO2 by assuming typical isotopic values of atmospheric and soil gases. Atmospheric methane usually leaves an isotopic trace in paleosols that is extraordinary depleted or light in isotopic values (less than -36 to as low as -80).

Alteration after burial (after Retallack 1990)


Some of the features used for interpreting paleo-environments from paleosols such as the amount of organic carbon, the oxidation state and the composition of clay minerals, are prone to alteration after burial. Fortunately, there remains an impressive residue of soil features that can be used for paleoenvironmental reconstruction, e.g. physical features such as nodules, root traces and burrows are particularly robust in the face of alteration after burial. In any case, for a correct interpretation it is necessary to distinguish between primary soil forming processes and those which occurred during burial. The most important alterations after burial are: Burial decomposition of organic matter, burial gleization of organic matter, burial reddening of iron oxides and hydroxides, cementation of primary porosity, compaction by overburden, illitization of smectite, zeolitization and celadonation of volcanic rocks, coalification of peat, kerogen maturation and cracking, neomorphism of carbonate and metamorphism

Conclusion
Paleosols are found in many terrestrial setting. Their generation was mainly influenced by agents as precipitation, temperature, parent material, morphology of landscape, time and later possibly alteration after burial. Therefore particular caution is necessary to distinguish primary and secondary features of soil formation. But as long as we are able to differentiate between both and interpret soil forming factors, a lot of information for the reconstruction of paleo-climate can be gained from the study of paleosols.

Classification and geochemistry of arid and semi-arid paleosols

11

References
Birkeland, P.W. (1992) Quartary soil chronosequences in various environments extremely arid to humid tropical. In: Weathering, soils and paleosols (Eds I.P. Martini &W.Chesworth). Elsevier, Amsterdam, 261-279 Botha, G.A. & Hughes, J.C. (1992) Pedogenic palysgorskyite and dolomite in a late Neogene sedimentary succession, northwest Transvaal, South Africa. Geoderma 53, 139-154 Bowles, C.G. & Braddock, W.A. (1963) Solution Breccias of the Minnelusa Formation in the Black Hills, South Dakota. US Geological Survey,Proffesional Paper 475C,C91-C95 Driese, S.G., Mora, C.I., Stiles, C.A., Joeckel, R.M. & Nordt, L.C. (2000) Mass-balance reconstruction of a modern Vertisol: implication for interpreting the geochemistry and burial alteration of paleo-Vertisols. Geoderma 95, 119-204 Ekart, D.P., Cerling, T.E., Montaez, I.P. & Tabor, N.J. (1999) A 400 million year carbon isotope record of pedogenic carbonate: implication for paleoatmospheric carbon dioxide. American Journal of Science 299, 805-827 Folkoff, M.E. & Meetenmeyer, V. (1987) Climatic controls of the geography of clay mineral-genesis. Association of American Geographers Annals 77, 635-650 Goudie, A.S. (1983) Calcretes. In: Goudie, A.S. & Pye, K. (eds): Chemical sediments and geomorphology: precipitates and residua in the near-surface environment, Academic Press, London, 93 132 Gyllenhaal, E.D. (1991) How accuretly can paleoprecipitation and paleo- climate change can be interpreted from subarial disconformities? PhD Thesis, University of Chicago Landis, G.P., Rigby, J.K., Sloan, R.E., Hengston, R.& Smee, L.W. (1996) Pele hypothesis: ancient atmospheres and geologic-geochemical controls in evolution, survival and extinction. In: Cretaceous-Tertiary Mass Extinction: Biotic and Environmental Changes (eds.N.MacLeod & G.Keller), W.W.Norton, New York, 519-556 Machette, M.N.(1985) Calcic Soils of the south-western United States. Geol. Soc. Amer., Spec. Pap., 203, 1- 21, Boulder Mack et al. (1993) Classification of paleosols. Geological Society of America Bulletin 105: 129-136 Marbut, C.F. (1935) Atlas of American Agriculture. Part III. Soils of the United States. US Department of Agricultural Advance Sheets 8. US Government Printing Office, Washington, DC McCahon, T.J.& Miller, K.B. (1997) Climatic significance of natric horizons in Permian (Asselian) paleosols of north-cental Kansas, U.S.A.. Sedimentology 44, 113-125 McFadden, L.D., McDonald, E.V., Wells, S.G., Anderson, K., Quade, J., & Forman, S.L. (1998) The vesicular layer and carbonate collars of desert soils and pavements: formation, age, and relation to climate change. Geomorphology 24, 101-145 Nahon, D.B. (1991) Introduction to the Petrology of Soils and Chemical Weathering. Wiley, New York Netterberg, F. (1967) Some road making properties of South African calcretes. Proc. 4th Reg. Conf. Soil Mechan. Found. Eng., Cape Town, 1, 77 81

12

Veronika Geiler, TU Bergakademie Freiberg

Netterberg, F. (1980) Geology of southern African calcretes. I. Terminology, description, macrofazies and classification. Trans. Geol. Soc. S. Afr., Cape Town, 83, 255 - 283 Quast, A. (2003): Calcretes aus jungpalozoischen Bodenbildungen: ein mglicher Proxy fr die CO2Konzentration der Paloatmosphre. Diss. Universitt Gttingen. 147pp Parrish, J.T. (1998) Interpreting Pre-Quartary Climate from the Geologic Record. Columbia University Press, New York. Retallack, G. (1990) Soils of the past: an introduction to paleopedology. 2nd Edition. Blackwell Science, Oxford. 404pp Retallack, G. (1994) A pedotype approach to latest Cretaceous and earliest Tertiary paleosols in eastern Montana. Geological Society of America Bulletin, 106, 1377 - 1397 Retallack, G. (1997) A color guide to paleosols. John Wiley & Sons, Chichester. 176pp Retallack, G. (1998) Core concepts of paleopedology. Quartary International 51/52: 203 212 Scheffer, F., Schachtschabel, P. (2002) Lehrbuch der Bodenkunde. 15. Auflage. Spektrum Akademischer Verlag, Heidelberg. 594pp Sehgal, J.L., Stoops, G. (1972) Pedogenic calcite accumulation in arid and semi-arid regions of the Indo-Gangetic alluvial plain of the erstwhile Punjab (India). Their morphology and origin. Geoderma 8, 59-72 Soil Survey Staff (1975) Soil taxonomy: A basic system of soil classification for making and interpreting soil surveys. US Dep. Agricult., NewYork. 754pp Soil Survey Staff (1999): Keys to Soil Taxonomy. US Dep. Agricult., Pocahontas Press, Blacksburg. 600pp van Donselaar-ten Bokkel Huinink, W.A.E. (1966) Structure, Root System and Periodicity of Savanna Plants and Vegetation in northern Surinam. North-Holland, Amsterdam Wright, V., Tucker, M. (1991) Calcretes: an Introduction. In: Calcretes: Reprint Series Volume 2 of the International Association of Sedimentologists. Blackwell Scientific Publications, Oxford. 1- 22 Websites
www.wikipedia.de

Das könnte Ihnen auch gefallen