Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-04-30T15:23:26.461Z Has data issue: false hasContentIssue false

Do Stable Isotope Data from Calcrete Record Late Pleistocene Monsoonal Climate Variation in the Thar Desert of India?

Published online by Cambridge University Press:  20 January 2017

Julian E. Andrews
Affiliation:
School of Environmental Sciences, University of East Anglia, Norwich, NR4 7TJ, United Kingdom
Ashok K. Singhvi
Affiliation:
Physical Research Laboratory, Ahmedabad, 380009, India
Ansu J. Kailath
Affiliation:
Physical Research Laboratory, Ahmedabad, 380009, India
Ralph Kuhn
Affiliation:
Forschungstelle Archaeometrice, Max Planck Institut fur Kernphysik, P.O. 103980, 69029, Heidelberg, Germany
Paul F. Dennis
Affiliation:
School of Environmental Sciences, University of East Anglia, Norwich, NR4 7TJ, United Kingdom
Sampat K. Tandon
Affiliation:
Department of Geology, University of Delhi, Delhi, 110007, India
Ram P. Dhir
Affiliation:
Central Arid Zone Research Institute, Jodhpur, 3420003, India

Abstract

Late Pleistocene terrestrial climate records in India may be preserved in oxygen and carbon stable isotopes in pedogenic calcrete. Petrography shows that calcrete nodules in Quaternary sediments of the Thar Desert in Rajasthan are pedogenic, with little evidence for postpedogenic alteration. The calcrete occurs in four laterally persistent and one nonpersistent eolian units, separated by colluvial gravel. Thermoluminescence and infrared- and green-light-stimulated luminescence of host quartz and feldspar grains gave age brackets for persistent eolian units I–IV of ca. 70,000–60,000, ca. 60,000–55,000, ca. 55,000–43,000, and ca. 43,000–∼25,000 yr, respectively. The youngest eolian unit (V) is <10,000 yr old and contains no calcrete. Stable oxygen isotope compositions of calcretes in most of eolian unit I, in the upper part of eolian unit IV, and in the nonpersistent eolian unit, range between −4.6 and −2.1‰ PDB. These values, up to 4.4‰ greater than values from eolian units II and III, are interpreted as representing nonmonsoonal18O-enriched “normal continental” waters during climatic phases when the monsoon weakened or failed. Conversely, 25,000–60,000-yr-old calcretes (eolian units II and III) probably formed under monsoonal conditions. The two periods of weakened monsoon are consistent with other paleoclimatic data from India and may represent widespread aridity on the Indian subcontinent during isotope stages 2 and 4. The total variation in δ13C is 1.7‰ (0.0–1.7‰), and δ13C covaries positively and linearly with δ18O. δ13C values are highest when δ18O values indicate the most arid climatic conditions. This is best explained by expansion of C4grasses at the expense of C3plants at low latitudes during glacial periods when atmospheric pCO2was lowered. C4dominance was overridingly influenced by global change in atmospheric pCO2despite the lowered summer rainfall.

Type
Original Articles
Copyright
University of Washington

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aitken, M.J. (1985). Thermoluminescence Dating. Academic Press, London.Google Scholar
Alam, M.S., Keppens, E., and Paipe, R. (1997). The use of oxygen and carbon isotope composition of pedogenic carbonates from Pleistocene palaeosols in NW Bangladesh, as palaeoclimatic indicators. Quaternary Science Reviews 16, 161168.CrossRefGoogle Scholar
Amundson, R.G., Chadwick, O.A., Sowers, J.M., and Doner, H.E. (1988). Relationship between climate and vegetation and the stable carbon isotope chemistry of soils in the eastern Mojave Desert, Nevada. Quaternary Research 29, 245254.Google Scholar
Andrews, J.E., Riding, R., and Dennis, P.F. (1993). Stable isotopic compositions of Recent freshwater cyanobacterial carbonates from the British Isles: Local and regional environmental controls. Sedimentology 40, 303314.Google Scholar
Bryson, R.A., and Swain, A.M. (1981). Holocene variations of monsoon rainfall in Rajasthan. Quaternary Research 16, 135145.CrossRefGoogle Scholar
Buczynski, C., and Chafetz, H.S. (1987). Siliciclastic grain breakage and displacement due to carbonate crystal growth: An example from the Lenders Formation, Permian of North Central Texas, USA. Sedimentology 34, 837843.CrossRefGoogle Scholar
Calvert, F., and Julia, R. (1983). Pisoids in the caliche profiles of Tarragona, north east Spain. Coated Grains Springer, Berlin.p. 456–473Google Scholar
Cerling, T.E. (1984). The stable isotopic composition of modern soil carbonate and its relationship to climate. Earth and Planetary Science Letters 71, 229240.CrossRefGoogle Scholar
Cerling, T.E., and Quade, J. (1993). Stable carbon and oxygen isotopes in soil carbonates.Swart, P.K., Lohmann, K.C., McKenzie, J., Savin, S. Climate Change in Continental Isotopic Records American Geophysical Union, Washington.136.Google Scholar
Cerling, T.E., Harris, J.M., MacFadden, B.J., Leakey, M.G., Quade, J., Eisenmann, V., and Ehleringer, J.R. (1997). Global vegetation change through the Miocene/Pliocene boundary. Nature 153158.Google Scholar
Chawla, S., Dhir, R.P., and Singhvi, A.K. (1992). Thermoluminescence chronology of sand profiles in the Thar Desert and their implications. Quaternary Science Reviews 11, 2532.Google Scholar
Clemens, S., Prell, W., Murray, D., Shimmield, G., and Weedon, G. (1991). Forcing mechanisms of the Indian Ocean monsoon. Nature 353, 720725.Google Scholar
Cole, D.R., and Monger, H.C. (1994). Influence of atmospheric CO2 4 . Nature 368, 533536.Google Scholar
Coplen, T.B. (1994). Reporting of stable hydrogen, carbon and oxygen isotopic abundances. Pure and Applied Chemistry 66, 273276.CrossRefGoogle Scholar
Courty, M.A., Dhir, R.P., and Raghavan, H. (1987). Microfabrics of calcium carbonate accumulations in arid soils of Western India.Fedoroff, M., Courty, M.A. Micromorphologie des Sols Assoc. Francaise pour l'Etude du Sol, 227234.Google Scholar
Dansgaard, W. (1964). Stable isotopes in precipitation. Tellus 16, 436468.Google Scholar
Deines, P. (1980). The isotopic composition of reduced organic carbon.Fritz, P., Fontes, J.C. Handbook of Environmental Isotope Geochemistry Elsevier, Amsterdam.329406.Google Scholar
Dhir, R.P., Kar, A., Wadhawan, S.K., Rajaguru, S.N., Misra, V.N., Singhvi, A.K., and Sharma, S.B. (1992). Thar Desert in Rajasthan. Geological Society of India, Bangalore.Google Scholar
Duller, G.A.T., and Botter-Jensen, L. (1993). Luminescence from potassium feldspars stimulated by infrared and green light. Radiation Protection Dosimetry 47, 683688.CrossRefGoogle Scholar
Duplessy, J.C. (1982). Glacial to interglacial contrasts in the northern Indian Ocean. Nature 494498.CrossRefGoogle Scholar
Ehleringer, J.R., Cerling, T.E., and Helliker, B.R. (1997). C4 2 . Oecologia 112, 285299.Google Scholar
Emeis, K.-C., Anderson, D.M., Doose, H., Kroon, D., and Shulz-Bull, D. (1995). Sea-surface temperatures and the history of monsoon upwelling in the northwest Arabian Sea during the last 500,000 years. Quaternary Research 43, 355361.Google Scholar
Fein, J.S., and Stephens, P.L. (1987). Monsoons. Wiley, New York.Google Scholar
Felix, C., and Singhvi, A.K. (1997). Study of non-linear luminescence dose growth curves for the estimation of paleodose in luminescence dating. Radiation Measurements 27, 599609.CrossRefGoogle Scholar
Fitzpatrick, E.A. (1984). Micromorphology of Soils. Chapman and Hall, London.Google Scholar
Goodfriend, G.A., and Magaritz, M. (1988). Palaeosols and Late Pleistocene rainfall fluctuations in the Negev Desert. Nature 332, 144146.CrossRefGoogle Scholar
Gupta, J.P. (1983). Some studies on hydrothermal regime and daytime heat fluxes in a desert sandy soil with and without vegetation. Archives for Meteorology, Geophysics and Bioclimatology Series B 32, 99107.Google Scholar
Gupta, J.P. (1986). Moisture and thermal regimes of the desert soils of Rajasthan, India, and their management for higher plant production. Journal des Sciences Hydrologiques 31, 347359.CrossRefGoogle Scholar
Hattersley, P.W. (1983). The distribution of C3 4 . Oecologia 57, 113128.Google Scholar
Hays, P.D., and Grossman, E.L. (1991). Oxygen isotopes in meteoric calcite cements as indicators of continental palaeoclimate. Geology 19, 441444.Google Scholar
Kim, S.T., and O'Neil, J.R. (1997). Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochimica et Cosmochimica Acta 61, 34613475.CrossRefGoogle Scholar
Krishnamurthy, R.V., and Bhattacharya, K. (1991). Stable oxygen and hydrogen isotope ratios in shallow ground waters from India and a study of the role of evapotranspiration in the Indian monsoon.Taylor, J.H.P., O'Neil, J.R., Kaplan, I.R. Stable Isotope Geochemistry: A Tribute to Samuel Epstein 187193.Google Scholar
Lang, A., Lindauer, S., Khun, R., and Wager, G.A. (1993). Procedures used for optically and infrared stimulated luminescence dating of sediments in Heidelberg. Ancient TL 14, 711.Google Scholar
Machette, M.N. (1985). Calcic soils of the south western United States.Weide, D.L. Soils and Quaternary Geology of the Southwest United States 121.Google Scholar
Maher, B.A., and Thompson, R. (1995). Paleorainfall reconstructions from pedogenic magnetic susceptibility variations in the Chinese loess and paleosols. Quaternary Research 44, 383391.Google Scholar
Manabe, S., and Hahn, D.G. (1977). Simulation of the tropical climate of an ice age. Journal of Geophysical Research 82, 38883911.Google Scholar
O'Neil, J.R., Clayton, R.N., and Mayeda, T.K. (1969). Oxygen isotope fractionation in divalent metal carbonates. Journal of Chemical Physics 51, 55475558.CrossRefGoogle Scholar
Pendall, E., and Amundson, R. (1990). The stable isotope chemistry of pedogenic carbonate in an alluvial soil from the Punjab, Pakistan. Soil Science 149, 199211.CrossRefGoogle Scholar
Prell, W., and Kutzbach, J.E. (1992). Sensitivity of the Indian monsoon to forcing parameters and implications for its evolution. Nature 360, 647652.Google Scholar
Prescott, J.R., Huntley, D.J., and Hutton, J.T. (1993). Estimation of equivalent dose in thermoluminescence dating—The Australian slide method. Ancient TL 11, 15.Google Scholar
Quade, J., Cerling, T.E., and Bowman, J.R. (1989). Systematic variations in the carbon and oxygen isotopic composition of pedogenic carbonate along elevation transects in the southern Great Basin, United States. Geological Society of America Bulletin 101, 464475.Google Scholar
Rajagopalan, G., Sukumar, R., Ramesh, R., Plant, R.K., and Rajagopalan, G. (1997). Late Quaternary vegetational and climatic changes from tropical peats in southern India—An extended record up to 40,000 years BP. Current Science 73, 6063.Google Scholar
Ramesh, R., Jani, R.A., and Bhushan, R. (1993). Stable isotopic evidence for the origin of salt lakes in the Thar desert. Journal of Arid Environments 25, 117123.Google Scholar
Rozanski, K., Araguas-Araguas, L., and Gonfiantini, R. (1993). Isotopic patterns in modern global precipitation.Swart, P.K., Lohmann, K.C., McKenzie, J., Savin, S. Climate change in continental isotopic records American Geophysical Union, Washington.136.Google Scholar
Sarnthein, M. (1978). Sand deserts during glacial maximum and climatic optimum. Nature 272, 4346.Google Scholar
Singh, G., Joshi, R.D., Chopra, S.K., and Singh, A.B. (1974). Late Quaternary history of vegetation and climate of the Rajasthan Desert, India. Philosophical Transactions of the Royal Society of London Series B 267, 467501.Google Scholar
Singh, G., Wasson, R.J., and Agrawal, D.P. (1990). Vegetational and seasonal climatic changes since the last full glacial in the Thar Desert, northwestern India. Review of Palaeobotany and Palynology 64, 351358.Google Scholar
Singhvi, A.K., Banerjee, D., Ramesh, R., Rajaguru, S.N., and Gogte, V. (1996). A luminescence method for dating ‘dirty’ pedogenic carbonates for paleoenvironmental reconstruction. Earth and Planetary Science Letters 139, 321332.CrossRefGoogle Scholar
Sirocko, F., Sarnthein, M., Erlenkeuser, E., Lange, H., Arnold, M., and Duplessy, J.C. (1993). Century-scale events in monsoonal climate over the past 24,000 years. Nature 364, 322324.Google Scholar
Sukumar, R., Ramesh, R., Pant, R.K., and Rajagopalan, G. (1993). A δ13 . Nature 364, 703706.CrossRefGoogle Scholar
Tandon, S.K., and Friend, P.F. (1989). Near surface shrinkage and carbonate replacement processes, Arran Cornstone Formation, Scotland. Sedimentology 36, 11131126.CrossRefGoogle Scholar
Tandon, S.K., Sareen, B.K., Rao, M.S., and Singhvi, A.K. (1997). Aggradation history and luminescence chronology of Late Quaternary semi-arid sequences of the Sabarmati basin, Gujarat, Western India. Palaeogeography, Palaeoclimatology, Palaeoecology 128, 339357.Google Scholar
Van Campo, E.V. (1986). Monsoon fluctuations in two 20,000-yr B.P. oxygen-isotope/pollen records off south west India. Quaternary Research 26, 376388.Google Scholar
Wasson, R.J., Smith, G.I., and Agrawal, D.P. (1984). Late Quaternary sediments, minerals and inferred geochemical history of Didwana Lake, Thar desert, India. Palaeogeography, Palaeoclimatology, Palaeoecology 46, 345372.Google Scholar
Wright, V.P. (1986). The role of fungal biomineralization in the formation of Early Carboniferous soil fabrics. Sedimentology 33, 831838.Google Scholar
Wright, V.P., and Tucker, M.E. (1991). Introduction.Wright, V.P., Tucker, M.E. Calcretes 122.Google Scholar