Sunteți pe pagina 1din 467

SCOPE 43

Stable Isotopes

Scientific Committee on Problems of the Environment SCOPE Executive Committee, elected 10 June 1988 Officers President: Professor F. di Castri, CEPE/CNRS, Centre L. Emberger, Route de Mende, BP 5051, 34033 Montpellier Cedex, France. Vice-President: Academician M. V. Ivanov, Institute of Microbiology, USSR Academy of Sciences, GSP-7 Prospekt 60 letija Oktjabrja 7, 117811 Moscow, USSR. Vice-President: Professor C. R. Krishna Murti, Scientific Commission for Continuing Studies on Effects of Bhopal Gas Leakage on Life Systems, Cabinet Secretariat, 2nd floor, Sardar Patel Bhavan, New Delhi 110001, India. Treasurer: Doctor T. E. Lovejoy, Smithsonian Institution, Washington, DC 20560, USA. Secretary-General: Professor J. W. B. Stewart, Saskatchewan Institute of Pedology, University of Saskatchewan, Saskatoon, S7NOWO Saskatchewan, Canada.

Members

Professor M. O. Andreae (IUGG representative), Max-Planck-Institut fUr Chemie, Postfach 3060, D-6500 Mainz, FRG. Professor M. A. Ayyad, Faculty of Science, Alexandria University, Moharram Bey, Alexandria, Egypt. Professor R. Herrera (lUBS representative), Centro de Ecologia y Ciencias Ambientales (IVle), Carretera Pan american a km. 11, Apartado 21827, Caracas, Venezuela. Professor M. Kecskes, Department of Microbiology, University of Agricultural Sciences, Pater K. utca 1, 2103 G6d6116, Hungary. Professor R. O. Slatyer, School of Biological Sciences, Australian National University, PO Box 475, Canberra, ACT 2601, Australia.

SCOPE 43

Stable Isotopes: Natural and Anthropogenic Sulphur in the Environment


Edited by H. R. Krouse Department of Physics and Astronomy, The University of Calgary, Canada and
V. A. Grinenko

Vernadsky Institute of Geochemistry and Analytical Chemistry, Moscow, USSR f,f-<~' ~

~~

~~ ,

'" 1ff

Published on behalf of the Scientific Committee on Problems of the Environment (SCOPE) of the International Council of Scientific Unions (ICSU) in collaboration with the' United Nations Environment Programme by JOHN WILEY & SONS Chichester. New York. Brisbane. Toronto. Singapore

WileyEditorial Offices
John Wiley & Sons Ltd, Baffins Lane, Chichester, West Sussex P019 IUD, England John Wiley & Sons Inc., 605 Third Avenue, New York, NY 10158-0012, USA Jacaranda Wiley Ltd, G.P.O. Box 859, Brisbane, Queensland 4001, Australia John Wiley & Sons (Canada) Ltd, 22 Worcester Road, Rexdale, Ontario MpW Ill, Canada John Wiley & Sons (SEA) Pte Ltd, 37 Jalan Pemimpin #05-04, Block B, Union Industrial Building, Singapore 2057 Copyright @ 1991 by the Scientific Committee on the Problems of the Environment (SCOPE) All rights reserved. No part of this book may be reproduced by any means, or transmitted, or translated into a machine language without the written permission of the publisher.
A catalogue record for this book is available from the British Library

Printed and bound in Great Britain by Biddies Ltd, Guildford, Surrey

SCOPE 1: SCOPE 2: SCOPE 3:

Global Environment Monitoring, 1971, 68 pp (out of print) Man-Made Lakes as Modified Ecosystems, 1972, 76 pp (out of print) Global Environmental Monitoring System's (GEMS): Action Plan for Phase 1, 1973, 132 pp (out of print) SCOPE 4: Environmental Sciences in Developing Countries, 1974, 72 pp (out of print) Environment and Development, proceedings of SCOPE/UNEP Symposium on Environmental Sciences in Developing Countries, Nairobi, February 11-23, 1974, 418 pp (out of print) SCOPE 5: Environmental Impact Assessment: Principles and Procedures, Second Edition, 1979, 208 pp SCOPE 6: Environmental Pollutants: Selected Analytical Methods, 1975, 277 pp (out of print) SCOPE 7: Nitrogen, Phosphorus and Sulphur: Global Cycles, 1975, 129 pp (out of print) SCOPE 8: Risk Assessment of Environmental Hazard, 1978, 132 pp (out of print) SCOPE 9: SCOPE SCOPE SCOPE SCOPE SCOPE 10: 11: 12: 13: 14: Simulation Modelling of Environmental Problems, 1978, 128 pp (out of print) Environmental Issues, 1977, 242 pp (out of print) Shelter Provision in Developing Countries, 1978, 112 pp (out of print) Principles of Ecotoxicology, 1978, 372 pp (out of print) The Global Carbon Cycle, 1979, 491 pp (out of print) Saharan Dust: Mobilization, Transport, Deposition, 1979,320 pp (out of print) Environmental Risk Assessment, 1980, 176 pp (out of print) Carbon Cycle Modelling, 1981,404 pp (out of print) Some Perspectives of the Major Biogeochemical Cycles, 1981, 175 pp (out of print) The Role of Fire in Northern Circumpolar Ecosystems, 1983, 344 pp (out of print) The Global Biogeochemical Sulphur Cycle, 1983,495 pp (out of print) Methods for Assessing the Effects of Chemicals on Reproductive Functions, 1983,568 pp (out of print) The Major Biogeochemical Cycles and Their Interactions, 1983,554 pp Effects of Pollutants at the Ecosystem Level, 1984, 460 pp The Role of Terrestrial Vegetation in tbe Global Carbon Cycle: Measurement by Remote Sensing, 1984, 272 pp Noise Pollution, 1986, 466 pp Appraisal of Tests to Predict the Environmental Behaviour of Chemicals, 1985, 400 pp Methods for Estimating Risks of Chemical Injury: Human and NonHuman Biota and Ecosystems, 1985, 712 pp

SCOPE 15: SCOPE 16: SCOPE 17: SCOPE 18: SCOPE 19: SCOPE 20: SCOPE 21: SCOPE 22: SCOPE 23: SCOPE 24: SCOPE 25: SCOPE 26:

SCOPE 27: Climate Impact Assessment: Studies of the Interaction of Climate and Society, 1985, 650 pp

SCOPE 28: Environmental Consequencesof Nuclear War


SCOPE 29: SCOPE 30:
SCOPE 31: SCOPE 32: SCOPE 33: SCOPE 34: SCOPE 35: SCOPE 36: SCOPE 37: SCOPE 38: SCOPE 39: SCOPE 40: SCOPE 41: SCOPE 42: SCOPE 43: SCOPE 44: SCOPE 45: Volume I Physical and Atmospheric Effects, 1986, 400 pp Volume II Ecological and Agricultural Effects, 1985, 563 pp The Greenhouse Effect, Climatic Change, and Ecosystems, 1986, 574 pp Methods for Assessing the Effects of Mixtures of Chemicals, 1987, 928 pp Lead, Mercury, Cadmium and Arsenic in the Environment, 1987, 384 pp Land Transformation in Agriculture, 1987, 552 pp Nitrogen Cycling in Coastal Marine Environments, 1988, 478 pp Practitioner's Handbook on the Modelling of Dynamic Change in Ecosystems, 1988, 196 pp Scales and Global Change: Spatial and Temporal Variability in Biospheric and Geospheric Processes, 1988, 376 pp Acidification in Tropical Countries, 1988, 424 pp Biological Invasions: A Global Perspective, 1989, 528 pp Ecotoxicology and Climate with Special Reference to Hot and Cold Climates, 1989, 432 pp Evolution of the Global Biogeochemical Sulphur Cycle, 1989, 242 pp Methods for Assessing and Reducing Injury from Chemical Accidents, 1989, 320 pp Short-Term Toxicity Tests for Non-genotoxic Effects, 1990, 360 pp Biogeochemistry of Major World Rivers, 1991, 380 pp Stable Isotopes: Natural and Anthropogenic Sulphur in the Environment, 1991, 464 pp Ecosystem Experiments, 1991 Methods for Assessing Exposure of Human and Non-Human Biota, 1991

SCOPE 46: Long-Term Ecological Research, 1991

International Council of Scientific Unions (ICSU) Scientific Committee on Problems of the Environment (SCOPE) SCOPE is one of a number of committees established by a nongovernmental group of scientific organizations, the International Council of Scientific Unions (ICSU). The membership of ICSU includes representatives from 74 National Academies of Science, 20 International Unions, and 26 other bodies called Scientific Associates. To cover multidisciplinary activities which include the interests of several unions, ICSU has established 10 scientific committees, of which SCOPE is one. Currently, representatives of 35 member countries and 20 international scientific bodies participate in the work of SCOPE, which directs particular attention to the needs of developing countries. SCOPE was established in 1969 in response to the environmental concerns emerging at that time; ICSU recognized that many of these concerns required scientific inputs spanning several disciplines and ICSU Unions. SCOPE's first task was to prepare a report on Global Environmental Monitoring (SCOPE 1, 1971) for the UN Stockholm Conference on the Human Environment. The mandate of SCOPE is to assemble, review, and assess the information available on man-made environmental changes and the effects of these changes on man; to assess and evaluate the methodologies of measurement of environmental parameters; to provide an intelligence service on current research; and by the recruitment of the best available scientific information and constructive thinking to establish itself as a corpus of informed advice for the benefit of centres of fundamental research and of organizations and agencies operationally engaged in studies of the environment. SCOPE is governed by a General Assembly, which meets every three years. Between such meetings its activities are directed by the Executive Committee. Funds to meet SCOPE expenses are provided by contributions from SCOPE Committees, an annual subvention from ICSU (and through ICSU, from UNESCO), contracts from the French Ministere de I'Environment, contracts with International Agencies, particularly UNEP, and grants from Foundations and industrial enterprises.

Executive Secretary: V. Plocq

Secretariat: 51Bid deMontmorency


75016 PARIS

R.E.Munn Editor-in-Chief SCOPE Publications

Contents

Preface Workshop Participants and Contributors to this Volume Chapter 1 Sulphur Isotopes in Nature and the Environment: An Overview H.G. Thode 1.1 Introduction 1.2 Theory and measurement of isotope effects 1.2.1 Equilibrium isotope effects 1.2.2 Kinetic isotope effects-isotope fractionation in unidirectional process 1.2.3 Fractionation factors 1.3 Sulphur isotope variations in nature 1.3.1 Biological sulphur cycle 1.3.1.1 Isotope fractionation by assimilatory reduction of sulphate 1.3.1.2 Isotope effects in the dissimilatory reduction of sulphate 1.3.1.3 Isotope fractionation during oxidation of sulphide 1.3.2 The marine sedimentary cycle 1.3.3 Fossil fuels 1.4 Sulphur isotopes in the atmosphere 1.5 Rain water 1.5.1 Ocean sulphate in rain water 1.6 Lakes and rivers 1.7 Vegetation 1.8 Applications

XIX XXI

1 1 2 2

5 6 8 9 10 11 12 14 15 16 19 19 20 20 21

Contents

Chapter 2 Oxygen Isotope Fractionation for Understanding the Sulphur Cycle B.D. Holt and R. Kumar 2.1 Introduction 2.2 Bases for the use of oxygen isotopy 2.2.1 Sulphate-water isotope exchange 2.2.2 SOrwater isotope exchange 2.2.2.1 Aqueous systems 2.2.2.2 Systems with no liquid water 2.2.3 8180 values of S02 oxidants and of associated water 2.2.4 Rapid hydration of S03 2.2.5 Sulphide oxidation 2.2.6 Bacterial S0422.2.7 Sulphate crystallization 2.3 Applications of 180 studies to environmental sulphates 2.3.1 Atmosphere 2.3.1.1 Precipitation sulphates 2.3.1.2 Aerosol sulphates 2.3.1.3 Sulphur dioxide 2.3.2 Hydrosphere 2.3.2.1 Oceans 2.3.2.2 Streams and lakes 2.3.2.3 Ground waters 2.3.3 Lithosphere 2.4 Conclusions Chapter 3 The Isotopic Analysis of Sulphur and Oxygen C.E. Rees and B.D. Holt 3.1 The isotopic analysis of sulphur 3.1.1 Introduction 3.1.2 The chemical pretreatment of sulphur samples 3.1. 2.1 Introduction 3.1.2.2 The conversion of barium sulphate to silver sulphide 3.1.2.3 Sulphur in the atmosphere and in water 3.1.2.4 Sulphur in vegetation, soils, and sediments 3.1.2.5 Sulphur in petroleum, coal, and sulphur ores 3.1.3 The preparation of sulphur samples for mass spectrometry 3.1.3.1 Introduction

27 27 27 28 28 28 29 32 32 32 33 33 34 34 34 36 37 37 37 38 38 38 39 43 43 43 43 43 44 44 45 47 48 48

Contents 3.1.3.2 Combustion of sulphides in oxygen 3.1.3.3 Oxidation of sulphides using a solid oxygen donor 3.1.3.4 Direct conversion of barium sulphate to sulphur dioxide 3.1.3.5 The purification and storage of sulphur dioxide 3.1.3.6 The preparation of sulphur hexafluoride 3.1.4 The mass spectrometric analysis of sulphur 3.2 Oxygen isotopes 3.2.1 Introduction 3.2.2 Sample collection 3.2.3 Sample preparation 3.2.3.1 Introduction 3.2.3.2 Sulphate 3.2.3.3 Water 3.2.3.4 S02 and O2 3.2.4 The mass spectrometric analysis of oxygen

XI

48 50 51 52 53 54 55 55 55 55 55 56 58 59 59

Chapter 4 Lithospheric Sources of Sulphur 65 H. Nielsen, J. Pilot, L.N. Grinenko, V.A. Grinenko, A. Yu. Lein, 1. W. Smith, and R. G. Pankina 65 4.1 Sulphur in the cosmos and the crust-mantle system 4.2 Sulphur and oxygen isotope composition of marine 68 evaporite and barite strata 68 4.2.1 Introduction 4.2.2 Compilation of 034S and 0180 data for marine 69 evaporites 69 4.2.2.1 General remarks on the data 4.2.2.2 Description of the 034S- and 0180-age 82 curves 4.2.2.3 Minima and abrupt rises in the 034S-age 85 curve 86 4.2.2.4 Non-marine evaporites 86 4.2.3 Data from marine barites 86 4.2.4 Reservoir of evaporite sulphur 87 4.3 Isotopic composition of sulphur in sulphide ores 91 4.4 Native elemental sulphur deposits 95 4.5 Sulphur isotope composition of fossil fuels 95 4.5.1 Peat and lignite 96 4.5.2 Coal 4.5.3 Oil 100 4.5.3.1 Introduction 100

Xli

Contents 4.5.3.2 Summary of 634S data for oil 4.5.3.3 Estimate of the mean (j34Svalue of oil 4.5.3.4 Use of sulphur isotope analyses to determine petroleum contamination in sediments 4.5.4 HzS in hydrocarbon deposits 4.6 Flux of volcanogenic sulphur to the atmosphere and isotopic composition of total sulphur 4.6.1 Volcanogenic sulphur flux to the atmosphere 4.6.2 Isotopic composition of volcanogenic sulphur 4.6.2.1 Isotopic composition of gases emitted to the atmosphere during eruption 4.6.2.2 Isotopic composition of ash emitted to the atmosphere during eruptions 4.6.2.3 The sulphur isotope composition of gaseous emissions from fumaroles and solfataras 4.6.3 The isotopic composition of total volcanic sulphur entering the atmosphere 100 107

108 109 116 116 119 120 120 121 124 133 133 133 133 134 134 135 135 137 137 138 138 140 141 142 142 143

Chapter

5 Sulphur Isotope Variations in the Atmosphere

L. Newman, H.R. Krouse, and V. A. Grinenko 5.1 Introduction 5.2 Sulphur species in the atmosphere 5.2.1 Atmospheric S compounds and their concentrations 5.2.2 The atmospheric chemistry of sulphur 5.2.2.1 HzS oxidation 5.2.2.2 So dust 5.2.2.3 SOz chemistry 5.3 Analytical techniques 5.3.1 Sampling of atmospheric S compounds 5.3.2 Further characterization of atmospheric compounds for sulphur isotope analyses 5.3.2.1 Wind direction 5.3.2.2 Particle size 5.3.2.3 Photographic examination of particles 5.3.2.4 Endogenous versus adsorbed sulphur 5.3.3 The use of sulphur isotope data from vegetation and soils 5.4 Sulphur isotope abundance variations in atmospheric compounds 5.4.1 Fractionation of sulphur isotopes during transformations of atmospheric sulphur compounds

143

Contents 5.4.2 Natural sources 5.4.2.1 Sea spray and marine aerosols 5.4.2.2 Biogenic emissions 5.4.2.3 Volcanic activity 5.4.3 Anthropogenic sources 5.4.3.1 Combustion of coal 5.4.3.2 Combustion and refining of oil and gas 5.4.3.3 Sulphide ores 5.4.3.4 Gypsum mining and processing 5.4.4 Summary 5.5 Delineation of anthropogenic and natural fluxes of atmospheric S compounds 5.5.1 Use of 334S and concentration data to identify sources of atmospheric gases 5.5.2 Use of 334S and concentration data to identify sources of SOi- in precipitation 5.5.3 Use of 3180 and concentration data to identify sources of SOi- in precipitation 5.5.4 Isotopic delineation of sources of sulphur emissions in local situations 5.5.5 Construction of regional balances of atmospheric sulphur 5.5.6 Estimate of the isotopic composition of biogenic sulphur in the atmosphere 5.6 Potential of using enriched sulphur isotopes for studying transport, transformations, and removal of atmospheric sulphur 5.6.1 Introduction 5.6.2 Background fluctuations 5.6.3 Dispersion and dilution 5.6.4 Economic evaluation

Xlll

144 144 146 147 148 148 148 148 149 149 152 152 157 157 158 159 163

168 168 168 169 170

Chapter 6 Hydrosphere 177 J.O. Nriagu, C.E. Rees, V.L. Mekhtiyeva, A. Yu. Lein, P. Fritz, R.J. Drimmie, R. G. Pankina, B. W. Robinson, and H. R. Krouse 177 6.1 Introduction 178 6.2 Oceans 178 6.2.1 Dissolved sulphate 183 6.2.2 Sulphur of aquatic organisms 189 6.3 Modern world sediments 189 6.3.1 Intensity of sulphate reduction 6.3.2 Rate of bacterial sulphate reduction and other factors influencing the distribution of sulphur

XIV

Contents compounds and their 5348 valuesin recent sediments 6.3.3 Mass and isotopic balance of sulphur in recent sediments 6.3.4 Influence of the rate of sulphate reduction on sulphide mineral formation 6.4 Lakes 6.4.1 Lakes: water column 6.4.1.1 Isotopic changes due to lake pollution 6.4.1.2 Oligotrophic lakes 6.4.1.3 Eutrophic lakes 6.4.1.4 Meromictic lakes 6.4.2 Lakes: sediments 6.4.2.1 Anthropogenic influence on the isotopic composition of sulphur in lake sediments of eastern North America 6.5 Isotopic composition of sulphur in continental seas 6.5.1 The Black Sea 6.5.2 The Azov Sea 6.5.3 The Caspian Sea 6.5.4 The Baltic Sea 6.5.5 The Red Sea 6.5.6 Conclusion 6.6 Rivers 6.7 Ground water 6.7.1 Introduction 6.7.2 The recharge environment 6.7.3 Ground water 6.8 Isotopic composition of sulphate in gas- and oilfield formation waters 6.9 Geothermal areas 6.10 Warm and cold springs 6.10.1 Introduction 6.10.2 Springs containing primarily dissolved sulphate 6.10.3 Springs containing primarily dissolved sulphide 6.10.4 Sulphur geochemistry of springs in relation to anthropogenic sulphur 6.11 Summary and key areas for further research

193 196 197 198 198 198 204 205 207 211

211 217 217 219 220 222 223 224 225 229 229 230 236 242 245 247 247 247 249 250 250 267 267 267 267

Chapter

7 Pedosphere and Biosphere H.R. Krouse, J. W.E. Stewart and V.A. Crinenko 7.1 Pedosphere 7.1.1 Forms of sulphur in soil 7.1.1.1 Inorganic sulphur

Contents 7.1.1.2 Organic sulphur 7.1.1.3 Transformations of sulphur in soils 7.1.2 Sources of sulphur compounds in soil 7.1.3 Sulphur and oxygen isotope variations in soil 7.1.4 Relationships among 034Svalues of different soil extracts 7.1.5 Transfer of sulphur into the soil as revealed by sulphur isotope data 7.2 Vegetation 7.2.1 Forms of sulphur in vegetation 7.2.2 Sulphur isotope distribution in vegetation 7.2.3 Assimilation of sulphate by plants and algae 7.2.4 Uptake of sulphide by plants and algae 7.2.5 Relative incorporation of soil and atmospheric sulphur by vegetation 7.2.6 The influence of biological parameters on the sulphur isotope composition of plants 7.2.7 Emission of reduced sulphur compounds by vegetation 7.3 Food webs and higher animals

xv 268 271 271 272 277 280 282 282 283 285 289 290 296 297 299

Chapter 8 Case Studies and Potential Applications 307 H.R. Krouse, L.N. Grinenko, V.A. Grinenko, L. Newman, 1. Forrest, N. Nakai, Y. Tsuji, T. Yatsumimi, V. Takeuchi, B. W. Robinson, M.K. Stewart, A. Gunatilaka, L.A. (Chambers) Plumb, J. W. Smith, F. Buzek, 1. Cerny, J. Sramek, A.G. Menon, G. V.A. lyer, V. S. Venkatasubramanian, B.E.C. Egboka, M.M. lrogbenachi, and C.A. Eligwe 307 8.1 Introduction 8.2 Sulphur isotope tracing of the fate of emissions from 309 sour gas processing in Alberta, Canada 8.2.1 Overview 309 309 8.2.2 The sour gas industry of Alberta 8.2.3 Sulphur isotope composition of the atmosphere in Alberta 312 312 8.2.3.1 Regional pattern 8.2.3.2 Relationship of 034S to concentrations of 314 atmospheric S compounds 8.2.3.3 Dependence of 034Svalues of atmospheric 315 S on wind direction 8.2.3.4 Isotopic composition of atmospheric particulate matter 316

XVI

Contents

8.2.4 Sulphur isotope composition of surface waters


8.2.5 Sulphur isotope abundances in soils and vegetation of Alberta 8.2.6 Isotopic modelling of uptake of industrial S by the environment near sour gas processing plants in Alberta 8.2.7 Discussion and recommendations 8.3 The isotopic composition and content of sulphur in soils of Kansk-Achinsk fuel power generation complex 8.3.1 Introduction 8.3.2 Results 8.3.3 Conclusion 8.4 Sulphur isotope measurements relevant to power plant emissions in the northeastern United States 8.4.1 Introduction 8.4.2 Fuels in use 8.4.3 Sulphur isotope compositions of ambient ground-level SOz and sulphate 8.4.4 Available fuels and other sources 8.4.5 Conclusions 8.5 Environmental sulphur isotope studies in Japan 8.5.1 Introduction 8.5.2 The quantity and sulphur isotopic composition of hydrogen sulphide released from tidal flats 8.5.2.1 Sampling 8.5.2.2 Observations and discussion 8.5.3 Secondary production in sulphuric acid from volcanic ash in surface water after a volcanic eruption 8.5.3.1 Introduction 8.5.3.2 Results and discussion 8.5.4 Sources of atmospheric sulphur compounds based on the sulphur isotopic composition of S04Z~ in precipitation in Japan, 1960-79 8.5.4.1 Introduction 8.5.4.2 Atmospheric SOz and HzS and pH of precipitation in Japan 8.5.4.3 Sulphur isotopic composition of SO/- in rain and snow in Japan 8.5.5 Estimation of the contribution of anthropogenic sulphur to the atmosphere in Japan 8.6 The use of sulphur and other stable isotopes in environmental studies of regional groundwater flow and sulphate mineral formation in Kuwait

317 318

320 323

327 327 327 331 331 331 332 333 338 343 343 343 344 344 344

347 347 348

352 352 353 355 358

361

Contents 8.6.1 Introduction 8.6.2 Geological setting 8.6.3 Sampling and analytical techniques 8.6.4 Isotopic composition of the waters 8.6.5 Isotopic composition of the sulphate 8.6.6 Conclusions 8.7 Natural sulphur isotope distributions for sulphate and sulphides in semi-arid, marginal marine environments in Australia 8.7.1 Introduction 8.7.2 Study sites 8.7.2.1 Spencer Gulf, South Australia 8.7.2.2 Shark Bay, Western Australia 8.7.3 Isotope data for water and sulphates 8.7.3.1 Continental waters, Spencer Gulf 8.7.3.2 Pertidal waters, Spencer Gulf 8.7.3.3 Shark Bay 8.7.4 Isotope distribution in peritidal sulphides 8.7.5 Summary 8.8 Contributions to sulphur to the atmosphere from utilization of fossil fuels, Australia 8.8.1 Coals 8.8.2 Crudes and condensates 8.8.3 Natural gases 8.8.4 Oil shales 8.8.5 Isotopic composition and fate of sulphur mined in Australia with copper, lead, zinc, and nickel ores 8.9 Sulphur isotope studies of atmospheric S and the corrosion of monuments in Prague, Czechoslovakia 8.9.1 Introduction 8.9.2 Analytical 8.9.3 Results and discussion 8.9.3.1 Atmospheric studies 8.9.3.2 Stone monuments 8.10 Determination of total sulphur 034S values in ore deposits of the Precambrian of India 8.10.1 Overview 8.10.2 Introduction 8.10.3 Sulphur isotope data and estimation of ore solution parameters 8.10.4 Trace elements in the evaluation of o34Stotal 8.10.5 Conclusions 8.11 Potential research areas for sulphur isotope geochemistry in Nigeria

XVll

361 363 365 365 369 371

371 371 372 372 372 373 377 377 379 382 385 385 385 387 391 391

391 399 399 400 400 400 403 405 405 405 406 408 409 409

XVIl1

Contents
8.11.1

Introduction

8.11.2 Research areas in Nigeria 8.11.3 Stable sulphur isotopes in basic geochemical research 8.11.3.1 Basement complex rocks 8.11.3.2 Lead-zinc mineralization and salt deposits 8.11.3.3 Agulu-Nanka gully complex sulphur deposits 8.11.4 Fossil fuel deposits 8.11.4.1 Coal and mine drainage 8.11.4.2 Petroleum and natural gas 8.11.4.3 Oil shales and tar sands 8.11.5 Emission of sulphur 8.11.6 Other research areas
Appendix

409 411 411 411 413 413 413 414 415 416 416 416 423 427

Chapter 9 Summary H.R. Krouse and V.A. Grinenko Index

431

Preface

This synthesis draws on the results of a major international scientific workshop held in Pushchino (Moscow Region) in summer 1983, organised jointly by the United Nations Environment Programme (UNEP) and SCOPE, with major financial support from UNEP. Today one should not discuss the sulphur cycle without reference to isotope data. Isotopes are different nuclear forms of the same element. The number of protons in the nucleus is constant for a given element-16 in the case of sulphur. Different numbers of neutrons in the nucleus correspond to different isotopes: 16 for 32S, 17 for 33S, etc. Some isotopes are not stable and undergo radioactive decay. Radioactive isotopes such as 35Shave been used as tracers in biological and agricultural studies. However, variations in the abundances of stable isotopes provide a more universal label for monitoring the global cycling of sulphur. This concept is the basis of this book. Stable isotope studies are based on the fact that isotopes of an element differ in their masses, e.g. 34S is heavier than 32S. Many physical and chemical processes in nature are mass dependent and hence molecules with different isotopes participate at different rates. When conversions are incomplete, the remaining reactant does not have the same isotope composition as the product, i.e. the isotopes have been 'fractionated'. By measuring isotope abundances of a natural sample, information can be gained about its past geochemistry. In controlled laboratory studies, isotope fractionation can be measured and the data used to interpret natural isotope variations. There are two extremes in stable isotope investigations. On the one hand, isotopic selectivity provides information about a process. On the other hand, if one wants to use the isotopic composition of an element to follow its fate in the environment, isotope fractionation must either be minimal or capable of being evaluated. Fortunately, there are a number of processes in which sulphur isotope selectivity is small or non-existent. These include certain high-temperature industrial processes, oxidation of H2S, S02, etc. in the atmosphere, solid phase reactions proceeding layer by layer, and assimilation of S042- by bacteria and plants. XIX

xx

Preface

A considerable body of stablesulphur isotopedata was introdu\;edin SCOPE 19: The Global Biogeochemical Sulphur Cycle. The current volume differs from this previous presentation in that it strives to show how stable isotope abundances may be used to differentiate between natural and anthropogenic sulphur in the environment. It is intended to serve as a handbook to anyone interested in this investigative tool. To this end, basic principles and analytical techniques are included as well as documentation of relevant research in different parts of the world. It is most fitting that the first chapter is an overview by Dr H.G. Thode, the primary driving force in the history of sulphur isotope investigations. Since oxygen isotope abundances in sulphates are also informative for deciphering the sulphur cycle, the second chapter by Dr B.D. Holt reviews the fundamentals of this topic. In turn, the lithosphere, atmosphere, hydrosphere, and biosphere are discussed by teams of international experts. Since sulphur leaving one 'sphere' enters another, it is difficult to avoid duplication. Indeed, a strict adherence to non-overlap would leave serious gaps in individual sections. The final chapter is devoted to case studies which range from situations where the expertise is well developed to the identification of potential research areas. It is anticipated that the broad spectrum of applications in this chapter will stimulate both the novice and experienced investigator. Regrettably, the Editors had to delete many detail~ from some manuscripts. Material from contributors who did not attend the workshop were also included to achieve a better balance. The contributions of the late Dr Charles Edward Rees (1940-84) are especially acknowledged. 'Ted' was a close friend of the editors and well known to the global community of stable isotope researchers for his competent science and gentle manner. His contributions were manifold but he is perhaps best remembered for his numerical modelling of stable isotope phenomena. He not only wrote part of Chapter 3 on preparation techniques for sulphur isotope analyses but he also served as a very effective member of the Editorial Advisory Committee during its meeting at Kananaskis. In latter stages of preparing this book, the editors and participants of UNEP/SCOPE meetings were also saddened to hear of the passing (1989) of Professor E.T. Degens. As Chairman of the Carbon Unit, he was a welcomed stimulus at Sulphur Unit meetings and provided invaluable input. The hospitality extended by the Soviet Academy of Science, The Institute of Microbiology, Moscow, and the Kananaskis Centre for Environmental Studies at The University of Calgary during meetings of members of the Editorial Board is gratefully acknowledged.

Workshop Participants and Contributors to this Volume

BAIBAKOV, S.N.

Deputy Chairman of the USSR Commission of UNEP, USSR State Committee for Science and Technology, PO Box 438, Moscow, 107053, USSR. Institute of Biochemistry and Physiology of Microorganisms, Pushchino, Moscow Region, 142292, USSR. Vernadsky Institute of Geochemistry and Analytical Chemistry, Kosygin Street, 19, Moscow, USSR. Institute of Biochemistry and Physiology of Microorganisms, Pushchino, Moscow Region, 142292, USSR. Geological Survey of Czechoslovakia, Prague, Czechoslovakia. Geological Survey of Czechoslovakia, Prague, Czechoslovakia. SCOPE/UNEP Carbon Unit, Geology-Paleontology Institute, Hamburg University, 0-2000 Hamburg, Bundesstrasse 55, FRG. Department of Earth Sciences, University of Waterloo, Waterloo, Ontario, N2L 3G 1, Canada. Department of Earth Sciences, Anambra State University of Technology, PMB 5025, Awka, Nigeria.
XXI

BELYAEv, S.S.

BELY!, V.M.

BONDAR, V.A.

BUZEK, F. CERNY, J. DEGENS, E.T. (deceased)

DRIMMIE, R.J.

EGBOKA, B.E.C.

XXII

Workshop

Participants and Contributors

ELiGWE, C.A.

FORREST, J.

Department of Metallurgical/Materials Engineering, Anambra State University of Technology, Awka Campus, PMB 5025, Awka, Nigeria. Environmental Chemistry Division Applied Science Department, Brookhaven National Laboratory, Upton, NY 11973, USA. CSIRO, Division of Plant Industry, PO Box 1600, Canberra, ACT 2601, Australia. GSF Institut fUr Hydrologie, Neuherberg, D-8042, FRG. Department of Biology, Jordan Hall 138, Indiana University, Bloomington, IN 47405, USA. Vernadsky Institute of Geochemistry and Analytical Chemistry, Kosygin Street, 19, Moscow, 117975, USSR. Manager, SCOPE/UNEP International Sulphur Unit, Institute of Biochemistry and Physiology of Microorganisms, Pushchino, Moscow Region, 142292, USSR. Institute of Biochemistry and Physiology of Microorganisms, Pushchino, Moscow Region, 142292, USSR. Faculty of Geochemistry, Department of Geology, Moscow State University, Lenin Hills, Moscow, USSR. Vernadsky Institute of Geochemistry and Analytical Chemistry, Kosygin Street, 19. Moscow, 117975, USSR. Department of Geology, Kuwait University, PO Box 5969, Kuwait. Lublin, ul. Novotki, Instytut Fizyki UMCS, Poland. Chemical Technology Division, Environmental Chemistry Group,

FRENEY, J.R.

FRITZ, P. FRY, B.D.

GALIMOV, E.M.

GOROKHOV, V.D.

GRIGORYAN, D.I.

GRINENKO, L.N.

GRINENKO, V.A.

GUNATILAKA, A.

HALAS, S. HOLT, B.D.

Workshop Participants and Contributors

XXll1

Argonne National Laboratory, IL 60439, USA.


IROGBENACHI, M.M.

Department of Earth Sciences, Anambra State University of Technology, Awka Campus, PMB 5025, Awka, Nigeria. Organic Geochemistry Laboratory, Geology-Paleontology Institute, Hamburg University, Bundesstrasse 55, D-2000, Hamburg, 13, FRG. Institute of Microbiology, USSR Academy of Sciences, GSP-7 Prospekt 60 Letija Oktjabrja 7, 117811 Moscow, USSR. Department of Physics, Indian Institute of Science, Bangalore, 560012, India. Department of Physics and Astronomy, The University of Calgary, Calgary, Alberta, T2N 1N4, Canada. Geomicrobiology Division, University of Oldenburg, D-2900, Oldenburg, FRG. Chemical Technology Division, Environmental Chemistry Group, Argonne National Laboratory, IL 60439, USA. Vernadsky Institute of Geochemistry and Analytical Chemistry, Kosygin Street, 19. Moscow, 117975, USSR. Institute of Oceanology, Krasikov Street, 24, Moscow, USSR. Scientific Secretary of the National SCOPE Committee, USSR, Academy of Sciences, Vavilov Street, 40, Moscow, USSR. All Union Research Institute of Geology and Prospecting of Oil Enthusiasts, Highway 38, Moscow, USSR. Department of Physics, Indian Institute of Science, Bangalore, 560012, India.

ITTEKKOT, V.A.

IVANOV,M.V.

IYER, G.V.A.

KROUSE, H.R.

KRUMBEIN, W.E.

KUMAR, R.

LEIN, A. YU.

LISTISYN, M.A.P.
LUKYANOV, N.K.

MEKHTIYEVA, V.L.

MENON, A.G.

XXIV

Workshop Participants and Contributors Vernadsky Institute of Geochemistry and Analytical Chemistry, Kosygin Street, 19, Moscow, 117975, USSR. Department of Earth Sciences, Nagoya University, Nagoya 464, Japan. Environmental Chemistry Division, Applied Science Department, Brookhaven National Laboratory, Upton, NY 11973, USA. Centre des Faibles Radioactivites, Physico-chimie de l'atmosphere Dept., Centre Scientifique 91190 Gif-sur- Yvette, France. Stable Isotope Laboratory Geochemical Institute, University of G6ttingen, G6ttingen, D-3400, FRG. Department of the Environment, National Water Research Institute, Burlington, Ontario, L7R 4A6, Canada. All Union Research Institute of Geology and Prospecting of Oil Enthusiasts, Highway 38, Moscow, USSR. Department of Geological Sciences, Bergakademie Freiberg, 9200 Freiberg, GDR. Executive Secretary of SCOPE, SCOPE, 51 Bid de Montmorency, 75016 Paris, France.
L.A.

MIGDlSOV, A.A.

NAKAI, N. NEWMAN, L.

NGUYEN, B.C.

NIELSEN, H.N.

NRIAGU, J.O.

PANKINA, R.G.

PILOT, J.

PLOCQ, V.

PLUMB, (Chambers)

Baas Becking Geobiological Laboratory, Canberra City, ACT 2601, Australia. Department of Chemistry, McMaster University, Hamilton, Ontario, L8S 4K1, Canada. Institute of Nuclear Science, DSIR, Lower Hutt, New Zealand. Department of Microbiology, Swedish University of Agricultural Sciences, S-750 07, Uppsala, Sweden.

REES, C.E. (deceased)

ROBINSON, B.W. ROSSWALL, T.

Workshop Participants and Contributors


RYABOSHAPKO, A.G.

xxv

SASAKI, A.

SMITH, J.W.

Institute of Applied Geophysics, Glebovskaya Street 20-B, Moscow, USSR. Department of Mineral Deposits, Geological Survey of Japan, 1-1-3 Higashi, Yatabe, Ibaraki 305, Japan. CSIRO Division of Fossil Fuels, PO Box 136, North Ryde, NSW 2113, Australia. State Restoration Studies, Pohorelec Square 22, CS-11800, Prague, Czechoslovakia. Saskatchewan Institute of Pedology, University of Saskatchewan, Saskatoon, Saskatchewan, S7N OWO, Canada. Institute of Nuclear Science, DSIR, Lower Hutt, New Zealand. Department of Earth Sciences, Nagoya University, Nagoya 464, Japan. Department of Chemistry, Nuclear Research Building, McMaster University, 1280 Main Street West, Hamilton, Ontario, L8S 4Kl, Canada Department of Earth Sciences, Nagoya University, Nagoya 464, Japan. Department of Physics, Indian Institute of Science, Bangalore, 560012, India. Department of Earth Sciences, Nagoya University, Nagoya 464, Japan. Vice-President of SCOPE, Institute of Microbiology, Moscow, 117312, USSR. Institute of Biochemistry and Physiology of Microorganisms, Pushchino, Moscow Region, 142292, USSR.

SRAMEK, J.

STEWART,J.W.B.

STEWART, M.K.

TAKEUCHI, V. THODE, H.G.

TSUJI, Y.
VENKATASUBRAMANIAN,

V.S.
Y ATSUMINI, T.

ZAVARZIN, M.G.A.

ZYAKUN, A.M.

CHAPTER 1

Sulphur Isotopes in Nature and the Environment: An Overview


H.G. THODE

1.1

INTRODUCTION

By 1931 it had long been suspected that there were differences in the chemical properties of the isotopes of an element, but it was the discovery of the rare heavy isotopes of the light elements, 13C, 15N, 170, and 180, in 1929 and particularly that of deuterium by Urey and his coworkers in 1931 that led to the detection of such differences. It was correctly felt that the 100% difference in the masses of the hydrogen isotopes would give rise to appreciable chemical dissimilarities, and in the succeeding decade, Urey's group and others predicted and measured such differences. An extension of their studies to isotopic compounds of carbon, nitrogen, and oxygen showed that small differences in the chemical properties of the isotopes of these elements existed as well. The laboratory measurement of chemical isotope effects, both equilibrium and kinetic, has developed into an extremely useful tool for the investigation of the mechanisms of chemical reactions. In addition, it has been recognized that isotope effects occur in nature, so that samples of an element will have variable isotope compositions or ratios reflecting differences in their chemical, biological, and geological histories. Over the past 40 years, extensive measurements of the isotope abundance variations of hydrogen, boron, carbon, oxygen, and sulphur have aided in the determination of geochemical properties of these elements and in the solution of many geochemical problems. Sulphur isotope geochemistry has been a particularly rewarding field of investigation because of the relatively large percentage mass difference between the two principal isotopes, the variety of chemical forms of sulphur, and their widespread occurrences in the earth's lithosphere, hydrosphere, and atmosphere. Sulphur isotope geochemistry studies began in the late 1940s (Thode et al., 1949; Trofimov, 1949). Sulphur isotope ratio studies have been concerned with such problems as isotope fractionation in the biological sulphur cycle, the sulphur-bearing
Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko <9 1991 SCOPE, Published by John Wiley & Sons LId

Stable Isotopes

gases of volcanoes, the isotopiccomposition of present-day and ancient


oceans, isotope distribution patterns in recent and ancient sediments and in coal and petroleum, the evolution of early life, and modes of formation and depositional histories of sulphide mineral deposits, etc. It is not surprising that sulphur isotope ratio measurements are now finding increased application in environmental studies. 1.2 THEORY AND MEASUREMENT OF ISOTOPE EFFECTS

Sulphur has four stable isotopes (32S, 33S, 34S, 3oS) whose percentage abundances are approximately 95.0, 0.75, 4.20, and 0.017 respectively (MacNamara and Thode, 1950). Isotope abundance variations are generally considered in terms of the abundance ratio 34Sp2S of the two principal isotopes. Isotope effects are generally small and this makes it convenient to refer to fractional differences in isotope ratios ('del' values (8) of samples relative to a standard, and to express these differences in parts per thousand (per mil):
(34SP2S)samplc
{ (34SP2S

834S(%0)=

- I }

x 1000

)standard

The generally accepted standard is troilite from the Canyon Diablo meteorite for which 34Sp2S is assigned the value 1/22.22. There are several reasons for expressing sulphur isotope ratio variations in this manner. Firstly, the mass spectrometer techniques used for sulphur isotope abundance measurements make the determination of differences in 34Sp2S ratios between samples much more precise than the determination of the absolute values of the ratios for either sample. Using approximately 1 mg of sulphur, as sulphur dioxide, or a few tenths of a milligram of sulphur, as sulphur hexafluoride, it is possible to compare samples with a precision of the order of 0.1%0 (McKinney et at., 1950; Rees, 1978). Secondly, the use of Canyon Diablo troilite (which is extremely uniform in isotopic content) as a standard makes interlaboratory comparison of results much easier than if a number of personal standards are used. Finally, there is strong evidence that the ratio 34Sf32Sin the troilite of iron nickel meteorites is the primordial value for the solar system and is the initial value of the earth's crust and mantle. It is therefore the natural base line to which terrestrial variations should be related. 1.2.1 Equilibrium isotope effects Isotope fractionation is a function of the relative masses of reacting molecules and in chemical systems may result during either isotopic exchange or

Sulphur Isotopes in Nature and the Environment: An Overview

unidirectional processes. Consider, for example, isotope exchange between sulphur dioxide and hydrogen sulphide. The exchange reactions are:
HZ3ZS ~ 3ZS0Z

and
HZ34S ~ 34S0Z

These equations can be combined to give the overall isotope exchange equation:
K

Hz34S + 3ZS0z ~ HZ3zS + 34S0Z

The equilibrium constant K may be expressed in terms of concentration or of partition functions (Q) for the molecules involved (Urey and Greiff, 1935; Urey, 1947; Bigeleisen and Mayer, 1947):
[34S0Z] / [3ZS0Z] Q34S0Z / Q3ZS0Z K = [Hl4S] /[HlZS] = QHz34S / QHlzS

The expression involving ratios of partition functions for isotopic molecules is convenient for the prediction of K since such ratios may be evaluated solely in terms of the fundamental vibrational frequencies of the molecules concerned. The expression for K involving ratios of concentration for isotopic molecules shows that when K is not unity the ratio 34SPZSwill not be the same in the two equilibrated phases. Thus the extent to which K differs from unity is a measure of the equilibrium isotope effect. For the HzS and SOz system above, the isotope equilibrium constant is 1.0064 at 800 K (Thode et al., 1971). Thus at 800 K under equilibrium exchange conditions, the 034S value of SOz will be 6.4%0 heavier than that of HzS (Figure 1.1). The decrease in the equilibrium isotope effects (K - 1) with increasing temperature, as seen for the SOz-HzS exchange reaction in Table 1.1, is the basis for isotope thermometry. Exchange reactions between sulphur compounds theoretically predict enrichment of 34S in the more oxidized species, the largest fractionation factor occurring in the exchange between HzS and SOi- , the two extreme oxidation states. It must be noted that although an isotope equilibrium constant may be calculated for a pair of sulphur compounds, there is no guarantee that isotope equilibrium will in fact be set up. If conditions are such that no reaction paths are possible, then the isotopic compositions of the two compounds will necessarily remain unaltered. In the case of the SOz and

4
+4 ----------------

Stable Isotopes

+3

+2

-L--------0 ~ + I -;~ (/) a .,. \ ro \ :> \

-- - -- - ----

~ ~

6.4%01/

T:

t
11200 K
I

TEMPERATURE CHANGE

2.6 %0 TIME

-I

~ -\~,- - - - - - - - - - - - - - - - - _

-2

\ ,,

- -: - - - - - - - - -=-::.:.: .: .= ---~--- - - --I


I ,,-

,,--

,
, "
,

- 31::- - - - - - - - - - - - - -= =-:- :: ~ -4 Figure 1.1 Changes

'--

I '"

/
/

".
'HYDROGEN SULPHIDE

1 '"

-~

I/

in (')34Svalues as isotopic equilibrium

is approached

Table 1.1 Equilibrium constants for some sulphur isotope exchange reactions (from Tudge and Thode, 1950) Isotope exchange reactions
Hl4S + 3ZS0/- ~ HZ3zS + 34S0/Hl4S + 3ZS03 ~ Hz'zS + 34S03 Hz'4S + 3ZS0Z ~ HZ3ZS + 34S0Z HZ34S + 3ZS0Z ~ HZ3ZS + 34S0Z HZ34S + 3ZS0Z ~ Hz'zS + 34S0Z 34S02(g) + WZS03 (aq) ~ 3ZS0Z(g) + W4S03 (aq) HZ3zS + H34S- ~ Hz'4S + H3zSPZS8 + HZ34S ~ 34S8 + HZ3zS " From Thode et al. (1971). h From Thode et al. (1945).

TCOe) 25 25 0 25 527 25 25 25

K (calc.) 1.074 1.070 1.037" 1.032" 1.0064" 1.003

K (exp.)

1.006" 1.019h 1.006

H2S exchange system, for example, it is necessary to have water present and to go to high temperatures so that reaction paths may be established between two sulphur compounds. In other cases, the time required to establish equilibrium may be long even on the geological time scale. The

Sulphur Isotopes in Nature and the Environment: An Overview

system would then have an isotopic distribution between that of its prior condition and equilibrium (Figure 1.1). 1.2.2 Kinetic isotope effects-isotope fractionation in unidirectional process

For systems undergoing unidirectional chemical reactions, the situation is somewhat more complicated than that for equilibrium processes and more attention must be paid to the reaction mechanism and possible intermediates involved in the formation of final products. Isotope fractionation in a unidirectional process results from differences in reaction rates of the different isotopic species. In a single-step, zero, or firstorder reaction, the isotope fractionation factor between the instantaneously generated product and remaining reactant is simply given by the ratio of rate constants for the two competing isotopic reactions. For example, in the chemical reduction of sulphate the ratio of rate constants k3z/k34 for the competing reactions:
k32

3ZS0/-

Hz'zS

and
k34 ~ HZ34S

3ZS0/-

has been measured and is 1.022 at ron:t temperature (Harrison and Thode, 1957). Since the 3zS0i- species reacts 1.022 times faster than the 34S0i-, the HzS produced at any instant is depleted in 34Sby about 22%0relative to the remaining SOl-. Predictions of rates of chemical reactions are made within the framework of the transition state theory (Eyring, 1935; Evans and Polanyi, 1935) where it is assumed that reactant molecules are in equilibrium with molecules known as 'activated complexes' in a transition state through which the reaction proceeds to products. Using the framework of this theory, the ratio of rate constants for the above two competing isotopic reactions may be written (Bigeleisen, 1947; Bigeleisen and Wolfsberg, 1958) as
~

k1

Q34S0i{ Q32S01-

k2

/ (Q32S0i-)*

(Q34S01-)*

~ } "2

The ratio of rate constants for the reaction of light and heavy isotopic speciesis therefore expressed,asin the caseof equilibrium constants,simply in the form of two partition function ratios, one for the two isotopic

Stable Isotopes

reactants 50/- and one for the two isotopic species of 'activated complexes'. The factor V/U2 in the expression is a mass term ratio for the two isotopic species, where v may be related to a vibrational mode in the activated complex. By making certain assumptions concerning the nature of the transition state, estimates for the ratio of rate constants for two competing isotopic reactions can be made. Therefore relative rate constants determined experimentally when combined with theoretical calculations provide information on the nature of the activated complex. For example, the large kinetic isotope effect obtained in the chemical reduction of sulphate can only be understood if we assume the breaking of a S-O bond in the initial rate-determining step, e.g. SO4
2-

~ (s

0 3 2- -0 ) * ~ SO 3 2-

In a multistep reaction, the situation may be more complex. In general, isotope fractionation occurs in the initial rate-determining step. However, in the case of a rapid equilibrium step followed by a unidirectional step, the equilibrium isotope effect in step 1 and the kinetic isotope effect in step 2 are additive. 1.2.3 Fractionation factors For purposes of understanding isotope abundance variations, it is important to distinguish between simple process factors as given by equilibrium constants or by ratios of rate constants' for a given isotopic reaction and fractionation observed for systems where the simple process factors may be multiplied many times. Commercial processes for the separation of isotopes depend on this kind of multiplication (Urey and Greiff, 1935). Where such systems occur in nature, apparent isotope fractionation factors will be observed that are significantly larger than those calculated or measured for simple processes. A batch process is one such system that can give rise to large fractionation. In the case of kinetic isotope effects, the changes in isotope ratios of both reactants and products with time depends on the extent of reaction (Bigeleisen and Wolfsberg, 1958). In Figure 1.2, reduction of sulphate to hydrogen sulphide is assumed to have a ratio of rate constants for the 32S and 34S molecular species of 1.024. For small fractions of reaction, the sulphide product will be depleted in 34Sby approximately 24%0with respect to the initial sulphate. As the reaction proceeds this depletion becomes smaller until at 100% reaction, the sulphide produced will have the same isotopic composition as the initial sulphate. For a small fraction of reaction, the sulphate isotopic composition will be virtually unaltered, but as the reaction proceeds, because of the continuing preferential reduction of

Sulphur Isotopes in Nature and the Environment:

An Overview

32S0/-, the remaining sulphate will become more and more enriched in 34Sand will so attain higher and higher 034Svalues (Figure 1.2). The curves in Figure 1.2 are consistent with Rayleigh distillation equations. In natural batch systems the ultimate extent of fractionation depends on the supply of reactant, and the terms 'open' and 'closed' systems have been used to differentiate between 'infinite' and 'limited' reactant reservoirs (Nakai and Jensen, 1964). In an 'open' system the isotopic composition of the reactant remains essentially unchanged and the product will be depleted in 34Sby one simple process factor given by the ratio of rate constants for the two isotopic species. In a 'closed' system the isotopic composition of both reactant and product change with time as the reaction proceeds, as seen in Figure 1.2. Clearly batch processes of this type can occur in nature and give rise to a variety of observable fractionation factors which will depend on the

80

70

60

50
0 ,,!? 0

40

" '"

U)

(.()

30

~TED

SULPHATE

20

10

00

30

EXTENT OF REACTION ("'.) 40 50 60 70

80

90

...
"0 ,,!? ~ (j) -10

100 0

PRODUCT SULPHIDE

---

" '"

(.() -20

~---------------

- 30
Figure 1.2 Changes in &34Svalues of remaining reactant and accumulated product where 32S0l- is being reduced 1.024 times faster than 34S042-

Stable Isotopes

completeness of reaction, Thismustbetakenintoaccount in theinterpretation of natural isotope distribution patterns.


1.3 SULPHUR ISOTOPE VARIATIONS IN NATURE

Figure 1.3 shows the ranges of 034Svalues found in nature for a number of different forms of sulphur. The zero point of the 034S scale is troilite from the Canyon Diablo meteorite (Thode, 1963). Some terrestrial samples fall outside the range of Figure 1.3, such as barite concretions from the ocean floor at +87%0 (Sakai, 1971) and marcasite associated with petroleum at
-

53.1%0(Austin, 1970). However, 98% of all samples analysed fall in the

range -40 to +40%0) (Nielsen, 1979).

834 S(%o) +50 +40 +30 +20 +10 -10

-20
I

-30
I

-40
I

-50
I

METEORITES

- -- .
.
-

..

BAS I C SI LLS

. . ----IGNEOUS ROCKS

s'
SEA

of VOLCANIC ORIGIN

WATER

EVAPORITES

-- - . .

..

RAIN

and

SNOW SULPHIDES

SEDIMENTARY

PETROLEUM

COAL

Figure 1.3 Sulphur isotope distribution in nature

Sulphur Isotopes in Nature and the Environment: An Overview

It can be seen that a broad distinction may be made between sulphur which has and sulphur which has not become involved in the sedimentary cycle. Sulphur in meteorites, ultra basic and basic sills, granitic intrusions, and igneous rocks of primary origin have 834S values in a narrow range close to zero (Shima et al., 1963; Thode et al., 1962; Ryznar et al., 1967; Schwarcz, 1973; Grinenko et al., 1970; Mitchell and Krouse, 1975). Volcanic gases and rocks tend to have a wider range of 834S values distributed symmetrically about 834S = O.This significantrange of values is due to a variety of inorganic chemical reactions and equilibria which tend to enrich oxidized forms of sulphur and deplete reduced forms of sulphur in 34S (Sakai, 1957; Rafter et al., 1960; Heald et al., 1963; Oana and Ishikawa, 1966; Grinenko and Thode, 1970; Puchelt et al., 1971; Hoefs, 1973). Also some volcanoes eject not only primary material (834S = 0) but act as pumps for sea water and other sources of sulphur compounds. Inorganic reactions with isotopic equilibria in hydrothermal fluids theoretically lead to oxidized and reduced forms of sulphur which are respectively enriched and depleted in 34S (Ohmoto, 1972). In comparison, sulphur in sediments, in sea water, and in materials such as coal and petroleum that have participated in the biological sulphur cycle, displays a much wider range of 834Svalues. This distinction provides a basic diagnostic technique in sulphur isotope studies. When an igneous rock is found that displays a wide range of 834S values or a narrow range well removed from zero, it is possible to state with some certainty that the rock is not of primary origin but is a reworked sediment or at least has a component of sedimentary sulphur. 1.3.1 Biological sulphur cycle In general, evaporites (sulphates) tend to be enriched and sedimentary sulphides depleted in 34S, as one might expect from thermodynamic considerations. However, there is no evidence of isotopic exchange between

sulphate and sulphide below temperatures of ~ 100C. Further,

although

the chemical reduction of sulphate to hydrogen sulphide takes place readily at room temperature in the presence of strong reducing agents and acids, chemical reduction under more realistic geochemical conditions has not been

demonstrated below temperatures of

~ 100C.

It is now well established that these large deviations of 834S from zero are due largely to the fractionation of the sulphur isotopes in biologically mitigated reactions at low temperatures. The biological cycle of sulphur is in part characterized by the activity of sulphur oxidizing and reducing bacteria which together account for the bulk of contemporary turnover rates in the biosphere (Baas-Becking, 1925; Postgate, 1959; Alexander, 1971). Whereas photosynthetic!photolithotrophic sulphur bacteria oxidize reduced

10

Stable Isotopes

sulphur, desulphovibrio desulphuricans or 'sulphate respirers' reduce sulphate to sulphide in anaerobic ecosystems in the presence of organic matter. Sulphate reduction in the sulphur cycle may be assimilatory or dissimilatory. During assimilatory reduction, ,such as occurs in the plant metabolism of sulphate, the sulphur is reduced from a valence of +6 to a valence of -2 in the synthesized products (amino acids and proteins). On the other hand, dissimilatory reduction of sulphate with the release of hydrogen sulphide often occurs in the bacterial reduction of sulphate. The turnover rates of sulphur in these latter dissimilatory processes exceed those during assimilatory reduction by several orders of magnitude. Accordingly the decisive biological control of the sulphur cycle is exercised by the sulphate reducing bacteria. It is during this dissimilatory reduction of sulphate to HzS carried out by a few extant genera (viz. desulphovibrio, desulphotomaculum and desulphomotas) that relatively large isotope effects occur. The bacterial reduction which takes place under anaerobic conditions in the presence of organic matter is therefore a major factor in the geochemical cycle of sulphur and accounts for the large fluxes of sulphur depleted in 34S into the lithosphere from the hydrosphere (Thode et al., 1951; Jones and Starkey, 1957; Kaplan and Rittenberg, 1964). 1.3.1.1 Isotope fractionation by assimilatory reduction of sulphate Assimilatory reduction of sulphate by plants and animals in both freshwater and marine environments results in only slight isotope fractionation, the 834S of reduced organic sulphur ranging from +0.5 to -4.4%0 relative to the sulphate in the external milieu (Ishii, 1953; Kaplan et al., 1963; Mekhtiyeva and Pankina, 1968; Mekhtiyeva, 1971). Similar small fractionations between cellular sulphur and sulphate have been noted during growth of a number of microorganisms (Kaplan and Rittenberg, 1964; Chambers and Trudinger, 1978). These fractionations varied from -0.9 to 2.8%0, the reduced cellular sulphur being isotopically light relative to the sulphate nutrient, whereas Ishii (1953), quoted by Kemp and Thode (1968), reported that metabolism of sulphate by plants (algae) involved little or no isotope fractionation < 1.2%0. Assimilatory organisms rarely produce significant amounts of free hydrogen sulphide from sulphate. There is evidence, however, that abnormal nutritional factors may modify sulphur metabolic pathways and thus lead to traces of hydrogen sulphide formed and isotope fractionation. Krouse et al. (1984) have shown that vegetation under stress from high concentrations of sulphur in atmospheric plumes from sour gas wells show such effects (Section 7.2.7).

Sulphur Isotopes in Nature and the Environment: An Overview

11

1.3.1. 2 Isotope effects in the dissimilatory reduction of sulphate Laboratory culture experiments performed with Desulphovibrio desulphuricans have shown that isotope effects associated with dissimilatory sulphate reduction vary between 834Svalues of + 3 and -46%0 depending in part on experimental conditions and possibly the organism (Ford, 1957; Harrison and Thode, 1958; Kaplan and Rittenberg, 1964; Kemp and Thode, 1968; Chambers and Trudinger, 1978). In general, the degree of fractionation is an inverse function of the rate of reaction. The highest fractionations are obtained at low metabolic rates of sulphate reduction, whereas the minimum effect at high metabolic rates ranges close to 0%0. The influence of temperature and electron donors on isotope fractionation can generally be interpreted in terms of their effect on the rate of reduction. Also at very low sulphate concentrations 0.01 mM), the magnitude of the fractionation decreases and the direction of the effect is apparently reversed (Harrison and Thode, 1958). Various models of isotope fractionation during dissimilatory sulphate reduction have been proposed to account for the isotope effects obtained under various conditions in terms of the biochemistry and the reduction pathways (Harrison and Thode, 1958; Kaplan and Rittenberg, 1964; Kemp and Thode, 1968; Rees, 1973; see Chambers and Trudinger, 1978 review). The pathway of dissimilatory sulphate reduction is composed of four principal enzyme catalysed steps as follows:
2
External sulphate ::;::::: internal
~ ~

3 adenasine ::;::::: sulphite phosphosulphate (APS)

4 -0> hydrogen sulphide

sulphate

An essential feature of the Rees model is that steps 1, 2, and 3 are reversible so that the system, external sulphate -0>sulphite, represents a pool allowing the kinetic isotope effect to be expressed during sulphite reduction. Rees assumed that only the forwards steps 1, 3, and 4 involved fractionation so that the overall isotope effect given by his steady-state model reduced to
a = al + a3 Xl X2 + a4 Xl X2 X3

where ai, a3, and a4 are the fractionation factors for the forward reactions 1, 3, and 4, and Xl> X2, and X3 are the ratios of the backward to forward flows associated with steps 1, 2, and 3 respectively. The sizableisotope shifts in the negativedirection associatedwith reaction series both in culture experiments and in natural environments are accordingly

12

Stable Isotopes

composites of the isotope effects inherent in the four numbered reaction steps depicted above. Attempts to account for the observed fractionation in principle consider four boundary conditions with one of the component reactions becoming rate controlling under specific conditions. Thus when the reaction proceeds unidirectionally, the overall isotope effect is simply that in the first step al (+3%0). In contrast, a multistep, unidirectional, first-order sequence exhibits the isotope effect in the ratedetermining step or slowest step. The other extreme experimental value, -46 to -50%0, occurs when the parameters XI, X2 and X3 approach unity. This means that the reduction of sulphite to hydrogen sulphide has been slowed down to the point where there is a build-up of the back flows from sulphite to APS so that the value of X3 is no longer zero. Although there is considerable uncertainty as to the kinetic isotope effects to be assigned to the various steps in the reaction sequence ai, a2, a3, and a4, it is generally agreed that the isotope effects associated with steps 1 and 2 are in the order of a few per mil only, the value for step 1 even becoming positive +3%0 at sulphate concentrations < 0.01 mM (Harrison and Thode, 1958). On the other hand, relatively large isotope effects are associated with reaction steps 3 and 4 which involve breaking of sulphur-oxygen bonds. Values proposed for step 3 vary between -25%0 (Rees, 1973) and -10 to -15%0 (Chambers and Trudinger, 1978), whereas the effect linked to sulphite reduction (step 4) should also be about -25%0. 1.3.1.3 Isotope fractionation during oxidation of sulphide In the biological sulphur cycle, several groups of bacteria produce sulphate by oxidation of reduced inorganic sulphur compounds. These include photosynthetic and chemosynthetic autotrophic organisms as well as heterotrophic organisms. Chemosynthetic sulphur bacteria (thiobacilli) utilize the energy released by the reactions of sulphide with free oxygen. Hence, thiobacilli are aerobes, whereas photosynthetic sulphur bacteria are strict anaerobes. Isotope effects in the oxidation steps of the biological sulphur cycle in autotrophic processes involving both chemosynthetic and photosynthetic sulphur bacteria have been studied (see review by Chambers and Trudinger, 1978). Studies of isotope effects during sulphide oxidation by photosynthetic organisms have produced conflicting results, perhaps due to different techniques for measuring the isotope effects and to the formation of intermediates in some cases. However, in general these isotope effects are small and inverse. Early laboratory experiments with Chromatium sp. photosynthetic oxidizing bacteria indicated isotope effects of less than 1%0in the oxidation of hydrogen sulphide and sulphur and to sulphate by either

Sulphur Isotopes in Nature and the Environment: An Overview

13

growing cultures or resting cell suspensions (Ford, 1957). Later, Kaplan et ai. (1960) using Chromatium sp. and Mekhtiyeva and Kondrateva (1966) using Rhodopseudomonas sp. reported that sulphate became slightly reduced in 34S(inverse effect) and that there was little or no isotope effect associated with sulphur formation. On the other hand, D.P. Kelly et ai., as reported by Chambers and Trudinger (1978), found that sulphur produced during sulphide oxidation by Chiorobium thiosuiphatophiium became enriched in 34Sby up to 5%0,a result similar to that obtained during abiogenic oxidation of sulphide to sulphur by Kaplan and Rafter (1958). Both sulphur and sulphate became enriched in 34Sduring sulphide oxidation by photosynthetic bacterium (E. Shaposhnikovii, reported by Ivanov et ai., 1976). Finally, Kaplan and Rittenberg (1964) showed that in contrast to their earlier results, significant enrichment of 32S in sulphur resulted during sulphide oxidation by Chromatium sp., whereas no significant fractionation was found in sulphate. Recently Fry et ai. (1983) reported isotopic enrichment factors for both photoheterotrophic or photoautotrophic oxidation of sulphide to elemental sulphur by Chromatium vinosum. Their results showed the product sulphur to be enriched in 34S (inverse isotope effect) by 2.4 0.7%0. This is in reasonable agreement with the average for the earlier work referred to above. Fry et ai. (1983) point out that the small inverse isotope effect observed during sulphide oxidation is likely to be associated with proton equilibria in the sulphide system. The equilibrium isotope effect in the exchange reaction
::':::

H232S + H34S~ ~

Hl4S

+ H32S-

is 6%0(Table 1.1), favouring 34S in the H2S. Thus if the 34S enriched H2S is the active component utilized by the cell, in the photosynthetic oxidation of sulphide, small inverse isotope effects will result. The actual magnitude will be of the order of 2-3%0 depending on the pH and proportions of H2S and HS in the medium. Chemosynthetic oxidation of sulphide in the presence of oxygen may involve larger isotope effects. This is suggested by the data of Kaplan and Rittenberg (1964), where large normal isotope effects C2S species reacting faster) of up to -18%0 were observed with aerated cultures of Thiobacillus concretivorous. Extensive formation and accumulation of polythionates occurred in the cultures of these experiments enriched consistently in heavier isotopes by 0.6-19.0%0. In the geological record, therefore, oxidized forms of sulphur thought to be derived from particular sulphide phases may be expected to differ isotopically from the precursor only if the mechanism of oxidation has been aerobic.

14 1.3.2 The marine sedimentary cycle

Stable Isotopes

Clearly the major isotope fractionation in the sedimentary cycle occurs in the bacterial reduction of seawater sulphate to H2S in anaerobic bottom waters and shallow sediments. The reduced sulphur in marine sediments, largely fixed as FeS and FeS2, is therefore isotopically light or depleted in 34S relative to seawater sulphate by up to 70%0(SCOPE 19, 1983). The actual 334S value of the sedimentary sulphides relative to sea water will depend on several factors. Firstly, it will depend on the isotope effect in the bacterial reduction of sulphate under the environmental conditions prevailing, such as available organic matter, etc. Secondly, it will depend on the fraction of the sulphides that may be considered formed in an 'open' system at the sea sediment interface with an infinite pool of sulphate and on the fraction that may be considered in a 'closed' system after burial. In an 'open' system, the full isotope effect will be realized, whereas in the 'closed' system, there is no net isotope fractionation after complete reduction of residual sulphate. The extent to which sulphate reduction takes place in an 'open' or 'closed' system will depend largely on reducing conditions and sedimentation rates. Two extreme cases are Black Sea sediments, in which 334S values for sulphide are displaced by up to 55%0relative to the sea sulphate (sea anoxic and sedimentation rates are low-at 0.1 mm yr-1; Vinogradov et al., 1962), and sediments off the coast of Venezuela, in which 334S values of the sulphides are displaced by only 15%0relative to the sea sulphate (water column oxygenated sedimentation rate is 25 cm yr-I; Thode et al., 1960). Some reduced sulphur in sediments is derived from plant decay. Plant metabolism of marine sulphate does not involve sulphur isotope selectivity. However, in view of the low sulphur content of plant organic matter (generally 0.1-1 %), this component of reduced sulphur in marine sediments is usually relatively small. The fractionation of sulphur isotopes in the bacterial reduction of marine sulphate and the flux of reduced sulphur depleted in 34Sinto the sediments has led to a build-up of 34Sin the oceans, beginning with the rise of sulphate
reducers possibly in the late Archean starting with 334SoccanS042-

=0

(Thode and Goodwin, 1983). Except near freshwater contributaries,

present-day ocean sulphate is

remarkably uniform with 334Svalues near +21%0 (this value is based on the
latest findings using SF6 by Rees (1978). Earlier workers using S02 reported values around +20.3%0.) However, data from evaporite deposits attest to significant temporal isotopic variations in the ancient oceans (Thode and Monster, 1964; Nielsen and Ricke, 1964; Thode and Monster, 1965; Holser and Kaplan, 1966; Davies and Krouse, 1975). Figure 4.2, Section 4.2.2.2 shows the spread in 334Svalues for evaporites of a given age. Both the lowest and mean values have been considered as the closest approach to the value for the

Sulphur Isotopes in Nature and the Environment: An Overview

15

contemporaneous ocean. The higher 834S values result for evaporites deposited from restricted or inland seas where the time constant for change in 834S is small relative to the oceans. The 834Svalues for marine sulphate, uniform at +21%0 today, has varied between a minimum of about + 10%0and a maximum of about + 30%0over the last 400 million years (see Section 4.2). Rees (1970) and Holland (1973) have modelled these changes which have been caused by biological activity and changes in relative fluxes of sulphide and sulphate mineral precipitation. Further, biological isotope fractionations superimposed on the 834S variations of the ancient oceans accounts for the large spread in 834Svalues found for sulphides in sedimentary rocks. Deposits related to ancient oceans such as Pb-Zn ores (Campbell et at., 1968; Sasaki and Krouse, 1969; Sangster, 1968; Vredenberg and Cheney, 1971) are generally enriched in 34S upwards to the 834S values identified with associated evaporites.

1.3.3 Fossil fuels Thode and Monster (1965) found petroleum and associated H2S to be on the average some 16%0depleted in 34S as compared to related evaporites. However, 834S values for petroleum may be altered during certain oil transformation processes (Orr, 1974; Thode, 1981). In certain hightemperature transformations, hydrogen sulphide formed by the thermochemical reduction of evaporitic sulphate, with little or no isotope fractionation, is introduced into the petroleum, thus altering its 834Svalues. Similar hightemperature reduction of sulphate can account for the high 834S values of H2S in the sour gas wells of Alberta, Canada, in the range of those for the associated evaporites (Krouse, 1980). Although petroleum 834Svalues vary from -10 to + 40%0 (Figure 1.4), individual large pools and pools of genetically related deposits are exceedingly uniform in isotopic content (Thode and Rees, 1970). Sulphur isotope variations in petroleum have been reviewed by Krouse (1977) (see also Sections 4.5.3 and 4.5.4). Coal originates from plants in continental or swamp areas. In most coals, the sulphur occurs primarily as organic and pyrite sulphur. In freshwater coastal plain peats and many lignites, organic sulphur dominates; sulphate is important and pyrite is minor. Thus the transformationi of peat to coal appears to involve an increase in pyrite principally at the expense of organic sulphur. Recent studies of sulphur transformation in peat across the Everglades basin in Florida indicate that pyrite formation in organic-rich swamps depends on the use of organic oxysulphur compounds in dissimilatory respiration by sulphur reducing bacteria (Altschuler et at., 1983). Since this process would take place essentially in a closed system, the net sulphur isotope fractionation would in most cases be small. Thus in low sulphur

16

Stable Isotopes

coals,the 534S values would essentially reflectthat of the sulphatenutrient


in the swamp area. The total sulphur and pyrite become more prominent in peats of brackish or marine environments and in coal derived from such peats. However, most of the sulphur in high sulphur coals (> 1% sulphur) comes from pyrite inclusions and the 834S ranges of coal of different ages and origins vary greatly, as indicated in Figure 4.15 (see also Sections 4.5.1 and 4.5.2). 1.4 SULPHUR ISOTOPES IN THE ATMOSPHERE

The concentration and isotopic composition of sulphur in the atmosphere can be highly variable and highly source dependent (see also Chapter 5). Nielsen (1974) has summarized the major contributors to atmospheric sulphur and in addition has given estimates of their isotopic compositions. According to Nielsen the major natural contributors to the global atmospheric sulphur budget are as follows: (a) Volcanic exhalations. The present-day addition of volcanic sulphur is probably rather minor (estimates range from 2-5%), although of course most of the sulphur on the surface of the earth had as its origin the outgassing of the deep crust and mantle. As mentioned previously, volcanic sulphur exhalations probably centre around 834S = with a total spread of 15%0 (see also Section 4.6). (b) Hydrogen sulphide from anoxic ocean waters. Transfer of such hydrogen sulphide to the atmosphere can only take place under special circumstances, as, for example, when deeper water masses are stirred by heavy storms. Under normal conditions, any hydrogen sulphide released from anoxic bottom waters will be oxidized to sulphate by dissolved oxygen in the water column long before it can reach the ocean surface. The 834S value of any hydrogen sulphide released in this manner will be negative with respect to seawater sulphate, reflecting its biological origin. (c) Hydrogen sulphide from sea marshes and intertidal flats. Again because of the fast oxidation of hydrogen sulphide in the presence of oxygen, this has been considered a minor source of isotopically light sulphur. However, direct measurements by Hansen et at. (1978) indicate very significant emissions from shallow littoral sediments of the North Sea ranging from 18 g S m-2 yr-1 from sediments with 1% organic matter to 4'\0 g S m~2 yr~1 for sediments highly enriched in organic matter. These results are in agreement with those of other investigations obtained for sediments on the coasts of the Barentzov and Caspian seas and the sea of Azov. The 834Svalues obtained for the release of H2S by 'sulphate reducers' from similar littoral sediments of the Eastern Sivash (Chukhrov et at., 1975) have an average value of 4%0 as compared to seawater
~

Sulphur Isotopes in Nature and the Environment: An Overview

17

sulphate at 21%0.Estimates of global emissions of light gaseous reduced

sulphur compounds into the atmosphere range up to 34 Tg yC 1 or -6%


of total emissions (Granat et at., 1976; Rjaboshapko et at., 1978). (d) Sea spray. Sulphate is transferred from the ocean to the atmosphere by wave action at the ocean surface producing aerosols. This sulphate has a 834S value of + 21%0, reflecting its ocean water source, and can be transported by wind onto continental regions. (e) Biogenic volatile sulphur compounds in continental areas from decay of organic matter. In a given area, this sulphur must isotopically reflect its derivation from sulphate in rain water, rivers, lakes, soils, etc. Such emissions are absent in arid regions but may be high in tropical swamps. In general this sulphur is isotopically lighter than seawater sulphate. It is necessary to add one further category to those considered by Nielsen. (f) Dimethyl sulphide. According to Maroulis and Bandy (1977) and Graedel (1979), for example, dimethyl sulphide is probably a major source of atmospheric sulphur. This compound, which can be produced by marine algae, is a more likely source of reduced sulphur from the oceans than is hydrogen sulphide. Little is known about the isotopic composition of dimethyl sulphide. An estimate is given in Section 5.5.6. The anthropogenic fluxes of sulphur to the atmosphere result from the utilization in industry of sulphur compounds themselves and from the utilization of other materials that contain sulphur as an unwanted or unavoidable by-product. These sources may be summarized as follows: (a) Native sulphur and hydrogen sulphide. These compounds are produced as a concomitant of natural gas use or from large sulphur deposits in salt domes of the Gulf Coast area and other parts of the world, and have 834S values -150/00light relative to associated sulphates, reflecting their biological origin (see Section 4.4). Sulphur dioxide and traces of hydrogen sulphide from the processing of 'sour' gas wells can be a major source of atmospheric sulphur in certain locations (see Section 8.2). (b) Sulphide ores. The 834Svalues of sulphide ores are highly variable and atmospheric contributions will be highly source dependent. Specific ore bodies are usually homogeneous and have distinctive 834S values (see Section 4.3). Sulphur from sulphide ore is emitted to the atmosphere from smelters and in the emission can be either as sulphate formed directly in high-temperature processes or as sulphur dioxide which can be oxidized further to sulphate during its atmospheric residence. (c) Coal. As mentioned above, most of the sulphur in coal comes from

18

Stable Isotopes

pyrite inclusions and its values therefore vary widely. However, low sulphur coals < 1% will have 834Svalues, reflecting that of the organic matter from which the coal was formed. (d) Petroleum. Although the isotopic composition of sulphur in petroleum is variable (Thode and Rees, 1970), the sulphur in particular oil reservoirs and horizons tends to be rather homogeneous isotopically. In general 834S oils tend to be 15%0light relative to the sulphate in the oceans at the time of oil formation.
~

VOLCANIC

CONTINENT

NATURAL

BIOGENIC

\\""\\,,\\\\\\\\\\\\\\\\\\\\\\\\\\-

OCEAN

BIOGENIC

H2S

SPRAY

COMBUSTION

COAL Oil

MANUFACTURING

CHEMICAL SMEl TING

-30

-20

-10

0 834S (%0)

+10

+20

Figure 1.4 Synopsis of 834Sdistributions in major source materials of atmospheric sulfur (Nielsen, 1974)

Figure 1.4 gives a synopsis of 834S distribution patterns in the major source materials of atmospheric sulphur. Thus the 834Svalue of a particular atmospheric sample reflects the averaging of many inputs. To the extent that there is mixing and long-distance transport, characteristic 834S values should be expected for continental, maritime, and transitional regions, particularly in remote areas from dominant local sources.

Sulphur Isotopes in Nature and the Environment: An Overview

19

1.5

RAIN WATER

The principal source of sulphate in rain water over ocean is sea spray. The 834S values of the ocean atmosphere (Atlantic and Pacific) far offshore range from + 12 to + 18%0(Chukhrov et at., 1980). These low values relative to sea sulphate have been explained by the mixing of the atmosphere with biogenic sulphur carried over from the continents. There is also the possibility of a biogenic source of sulphur from the ocean itself, perhaps involving both hydrogen and sulphur and dimethyl sulphide. Rainwater sulphate over the continents is, on the other hand, isotopically much lighter, indicating a relatively high proportion of biogenic sulphur. 1.5.1 Ocean sulphate in rain water A number of attempts have been made to determine the fraction of sulphate in rain water (atmosphere) in various regions derived from the oceans., which will have a 834Svalue of + 21%0.The estimates of the so-called excess sulphur, or non-marine sulphate in the atmosphere, are usually based on the ratio of (SO/-)/(CI-) in the atmospheric samples as compared to the ocean water. If no fractionation of the ions occurs during the transfer from

ocean to the atmosphere then (SOi-)cxcess = (SOi-)total - 0.14 CI-,


where the ratio of (SOi-)/(CI-) in sea water is 0.14. However, there is fractionation of the ions during transfer and according to Korhz (1976) the contribution of marine sulphates to the global atmosphere should be calculated according to the empirical formula:
(SOi-)excess

= (SOi-)total

- 0.38CI-

The mechanism of ion fractionation is not known; however, the values for (S04Z-)/(CI-) ratios from 0.14 to ~0.40 seem to relate to carrier particle sizes in the atmosphere. For example, at 6 km inland, the ratio increases by a factor of 2 from that offshore, suggesting that the smaller particles which penetrate deeper in the continent have considerably higher (SOi-)/ (CI-) ratios (Yaalon and Lomas, 1970). Estimates of the fraction of sulphate in the continental atmosphere derived from ocean spray range from 8 to 35%, based on (SOi-)/(CI-) ratios in precipitation and river runoff samples (Korhz, 1976; Inland Waters Directorate Environment Canada, 1977-79). Early 834Smeasurements of sulphate in snow and rain samples from remote continental areas gave values between +4.0 and +6%0, or ~ 15%0 light relative to seawater sulphate (Ostlund, 1959), thus providing evidence for a major component of natural sulphur of biogenic origin in the global atmosphere. Rainwater samples collected from remote areas in the Canadian

shield and other similar areas of the world have 034S values ranging from

20

Stable Isotopes

+2.2 to +4.2%0(Kramer and Snyder, 1977;Chukhrov et al., 1980; Nriagu and Coker, 1978b). By comparison, rainfall (sulphur) over oceans far offshore have 034S values ranging from + 12 to + 18%0 (Chukhrov et at., 1980; Rafter and Mizutani, 1967). The 034S values lower than +21%0 are attributed to biogenic sulphur in air masses coming from the continents or from the oceans themselves, perhaps due to methyl sulphide emissions as well as HzS. 1.6 LAKES AND RIVERS

A wide range of sulphur isotope ratios arise in the environment regardless of man's activities. For example, rivers flowing through a sedimentary rock terrain such as the MacKenzie River in Canada's Northwest Territory exhibit a wide range of 034S values which can be related to the geological strata along its course (Hitchon and Krouse, 1972). Also freshwater lakes with relatively high sulphate contents (karst lakes) which occur widely where evaporite-bearing rocks lie near the surface, will have 034Svalues reflecting those of the source sulphates enriched in 34S. Furthermore, considerable fractionation of the sulphur isotopes may occur in these lakes due to the bacterial reduction of sulphate (which may occur either in the shallow sediments or in the anoxic water column). The extent of sulphur isotopic fractionation in these lakes increases with the sulphate concentratinn, rising to a 55%0spread between sulphide and residual sulphate similar to that in the Black Sea at high sulphate concentrations (Deevey et at., 1963; Matrosov et at., 1975; Vinogradov et at., 1962). However, in lakes with very low sulphate concentrations (0.5-4.0 mg of SOi- /litre) in remote areas of continental shield terrain, there is no evidence of sulphur isotopic fractionation due to sulphate reducers and the 034Svalues of the sulphate reflect that of the rain water in the absence of other natural sources. In the same way, sulphur in the lake sediments, the soil, and vegetation in these areas will have 034Svalues centered around that of the rain water for that area. 1.7 VEGETATION

Land plants and organic matter in sediments derived from land plants may play an important role in environmental studies. For example, studies of organic matter in lake core sediments as a function of depth provide an opportunity to follow historical changes in the 034Sbase level values for an area or drainage basin (Nriagu and Coker, 1983; Thode and Dickman, 1983). Plants may absorb and metabolize sulphate from the soil or SOz from the atmosphere through the leaves (see Section 7.2.5). The main source of plant sulphur in soils whose parent rock contains little or no sulphides or

Sulphur Isotopes in Nature and the Environment: An Overview

21

sulphates is atmospheric precipitation. Since there is little or no fractionation of the sulphur isotopes in plant metabolism, plants in these areas will have 334S values reflecting that in the rain water. For example, the 334S values of the total sulphur in plant samples from the remote Kazalehstan steppe area in the Soviet Union centre around +2.0%0, which is consistent with average 334S values of 3.2 and +3.0%0 for the sulphate in rain and soil samples respectively for the same area (Chukhrov et at., 1980). 1.8 APPLICAnONS

Stable isotopes of sulphur have been used to trace the movements of sulphur compounds in the atmosphere, rivers, lake water, and ground water. A series of studies originating from the Brookhaven National Laboratories have demonstrated the usefulness of sulphur isotopes in tracing the sulphur emitted from power plants and the oxidation to sulphate of sulphur dioxide present in emission plumes (Newman et at., 1975; Forrest and Newman, 1977a, 1977b). Papers by Forrest and Newman (1977a, 1977b), Kramer and Snyder (1977), Nriagu and Coker (1978a, 1978b), and Nriagu and Harvey (1978) give details of isotopic studies of atmospheric and lake samples downwind from the Sudbury smelter. Holt et at. (1972) showed that sulphur with a unique 334Svalue could be followed from deep wells through sewage processing and into the effluent. Case and Krouse (1980) have used sulphur isotope ratio studies in the elucidation of sources, mixing and dispersion of sulphur compounds in the atmosphere and hydrosphere. In these studies plots of 334Svalues versus sulphur content at various distances from a known source proved useful in determining the impact of the source over the area and the fate of sulphur compounds. The oxygen isotopes used widely as indicators in the study of geological processes have been used along with sulphur isotopes to identify sulphate sources asnd to elucidate mode of formation of sulphate in the atmosphere (Holt et at., 1981; see Chapter 2). In summary, sulphur isotope studies have already been used extensively in testing regional sulphur flux models, and in identifying and determining the impact of natural and anthropological sources of sulphur on the environment. These studies have been reviewed by a number of authors (Nriagu, 1978; Krouse, 1980; Ivanov, 1981) and will be the subject of later chapters in this book. REFERENCES Alexander, M. (1971). MicrobialEcology, Wiley, pp. 432-6. Altschuler, Z.S., Schnepfe, M.M., Silber, c.c., and Simon, F.G. (1983). Sulphur diagenesis in everglades, peat and origin of pyrite in coal. Science, 221, 221-7.

22

Stable Isotopes

Austin, S.R. (1970).Some patterns of sulphur isotope distribution in uranium deposits. Earth Sci. Bull., 3, 5. Baas-Becking, L.G.M. (1925). Studies on sulphur bacteria. Ann. Bot., 39, 613-50. Bigeleisen, J. (1947). The relative reaction velocities of isotopic molecules. J. Chern. Phys., 17,675. Bigeleisen, J., and Mayer, M.G. (1947). Calculation of equilibrium constants for isotopic exchange reactions. 1. Chern. Phys., 38, 2457-66. Bigeleisen, J., and Wolfsberg, M. (1958). Theoretical and experimental aspects of isotope effects in chemical kinetics. Adv. Chern. Phys., 1, 15. Campbell, F.A., Evans, T.L., and Krouse, RR. (1968). A reconnaissance study of some western Canadian lead-zinc deposits. Econ. Geol, 63, 349-59. Case, J.W., and Krouse, H.R. (1980). Variations in sulphur content and stable sulphur isotope composition of vegetation near a S02 source at Fox Creek, Alberta, Canada. Oecologia (Berl.), 44, 248-57. Chambers, L.A., and Trudinger, P.A. (1978). Microbiological fractionation of stable sulphur isotopes. A review and critique. Geomicrobiol. J., 1, 249-93. Chukhrov, F.V., Churikov, V.S., Ermilova, L.P., and Nosik, L.P. (1975). Variations of sulphur isotope ratios in some natural waters. Geokhimiya, 3, 343-56. Chukhrov, F.V., Ermilova, L.P., Churikov, V.S. and Nosik, L.P. (1980). The isotopic composition of plant sulphur. Org. Geochem., 2, 69-75. Davies, C.E., and Krouse, H.R. (1975). Sulphur isotope distribution in Paleozoic evaporites, Canadian Arctic Archipelago. Geological Survey of Canada Paper 75-1, part B, pp. 221-5. Deevey, E.S. Jr, Nakai, N. and Stuiver, M. (1963). Fractionation of sulphur and carbon isotopes in a meromictic lake. Science, 139, 407-8. Evans, M.G., and Polanyi, M. (1935). Some applications of the transition state method to the calculation of reaction velocities. Trans. Faraday Soc., 31, 875-94. Eyring, H. (1935). Activated complex in chemical reactions. J. Chern. Phys., 3, 107. Ford, R.W. (1957). Sulphur isotope effects in chemical and biological processes. Ph.D. thesis, McMaster University, Canada, pp. 59-63. Forrest, J., and Newman, L. (1977a). Further studies on the oxidation of sulphur dioxide in coal-fired power plant plumes. Atmos. Environ., 11, 465-74. Forrest, J., and Newman, L. (1977b). Oxidation of sulphur dioxide in the Sudbury smelter plum. Atmos. Environ., 11, 517-20. Fry, B., Gest, R, and Hayes, J.M. (1983). Sulfur isotopic composition of deep-sea hydrothermal vent animals. Nature, 306, 51-2. Graedel, T.E. (1979). Reduced sulphur emissions from the open oceans. Geophys. Res. Left., 6, 329-31. Granat, L., Rodhe, H., and Hallberg, R.O. (1976). The global sulphur cycle. In: SCOPE Report 7, Econ. Bull., 22, 89-134. Grinenko, L.N., Kononova, V.A., and Grinenko, V.A. (1970). Sulphur isotope composition of sulphides in carbonatites. Geokhimiya, 1, 66-75. Grinenko, V.A., and Thode, H.G. (1970). Sulphur isotope effects in volcanic gas mixtures. Can. 1. Earth Sci., 7,1402-9. Hansen, M.H., Ingvorsen, K., and Jorgensen, B.B. (1978). Mechanism of hydrogen sulphide release from coastal marine sediments to the atmosphere. Limnol. Oceanog., 23, 68-76. Harrison, A.G., and Thode, H.G. (1957). The kinetic isotope effect in the chemical reduction of sulphate. Trans. Faraday Soc., 53, 1-4. Harrison, A.G., and Thode, RG. (1958). Mechanism of the bacterial reduction of sulphate from isotope fractionation studies. Trans. Faraday Soc., 54, 84-92. Heald, E.F., Naughton, 1.1., and Barnes, LL. Jr (1963). The chemistry of volcanic gases. J. Geophys. Res., 68, 545-57.

Sulphur Isotopes in Nature and the Environment: An Overview

23

Hitchon, B., and Krouse, H.R. (1972). Hydrogeochemistry of the surface waters of the MacKenzie River drainage basin, Canada. III. Stable isotopes of oxygen, carbon and sulphur. Geochim. Cosmochim. Acta, 36, 1337-57. Hoefs, J. (1973). Stable Isotope Geochemistry, Springer-Verlag, Berlin, New York, 140 pp. Holland, H.D. (1973). Systematics of the isotope composition of sulfur in the oceans during the Phanerozoic and its implications for atmospheric oxygen. Geochim. Cosmochim. Acta, 37, 2605-16. Holser, W.T., and Kaplan, LR. (1966). Isotope geochemistry of sedimentary sulphates. Chem. Geol., 1, 93-135. Holt, B.D., Engelkemeir, A.G. and Venters, A. (1972). Variations of sulphur isotope ratios in samples of water and air near Chicago. Environmental Sci. and Tech., 6, 338-41. Holt, B.D., Cunningham, P.T., and Kumar, R. (1981). Oxygen isotopy of atmospheric sulphates. Environmental Sci. and Tech., 15, 804-8. Inland Water Directorate Environment Canada (1977-79) Detailed surface water quality data, Northwest Territories. Ishii, M. (1953). Fractionation of the sulphur isotopes in plant metabolism of sulphur. M.C. thesis, McMaster University, Hamilton, Ontario, pp. 1-50. Ivanov, M.V. (1981). The global biogeochemical sulfur cycle. In: Likens, G.E. (Ed.) Some Perspective of the Major Biogeochemical Cycles, SCOPE Report 17, Wiley, pp. 61-78. Ivanov, M.V., Gogotova, G.I., Matrosov, A.G. and Zyakun, A.M. (1976). Fractionation of sulphur isotopes by phototrophic sulphur bacterial. Microbiol. Eng. Trans., 45, 655-9. Jones, G.E., and Starkey, R.L. (1957). Fractionation of stable isotopes of sulphur by microorganisms and their role in the deposition of native sulphur. Appl. Microbiol., 5, 111-18. Kaplan, LR., and Rafter, T.A. (1958). Fractionation of stable isotopes of sulphur by Thiobacilli. Science, 127, 517-18. Kaplan, I.R., and Rittenberg, S.c. (1964). Microbial fractionation of sulphur isotopes. 1. Gen. Microbiol., 34, 195-212. Kaplan, LR., Rafter, T.A., and Hulston, J.R. (1960). Sulphur isotope variations in nature. Part 8, Some applications to biogeochemical problems. N.Z. 1. Sci., 3, 338-61. Kaplan, LR., Emery, K.O., and Rittenberg, S.c. (1963). The distribution and isotopic abundance of sulphur in recent marine sediments off Southern California. Geochim. Cosmochim. Acta, 27, 297-331. Kemp, A.L.W., and Thode, H.G. (1968). The mechanism of bacterial reduction of sulphate and of sulphite from isotope fractionation studies. Geochim. Cosmochim. Acta, 32, 71-91. Korhz, V.D. (1976). Chemical exchange between air and sea as a factor in controlling the composition of salts in river water. Dokl. Acad. Nauk. SSR, 230 (2), 432-5. Kramer, J., and Snyder, W.H. (1977). Precipitation scavenging 1974. U.S. ERDA, 1977, 127-36. Krouse, H.R. (1977). Sulphur isotope studies and their role in petroleum formation. J. Geochem. Exploration, 7, 189-211. Krouse, H.R. (1980). Sulphur isotopes in our environment. In: Fritz, P., and Fontes, J. Ch. (Eds.) Isotope Geochemistry, Vol. 1. The Terrestrial Environment, Elsevier, Amsterdam, pp. 435-71. Krouse, H.R., Legge, A., and Brown, H.M. (1984). Sulphur gas emissions in the boreal forest: the West Whitecourt Case Study V: Stable sulphur isotopes. Water, Air, Soil Pollution, 22, 321-47.

24

Stable Isotopes

McKinney, C.R.,McCrea, J.M" Epstein, S.,Allen,H.A" andUrey,H.C.(1950).


Improvement of mass spectrometers for the measurement of small differences in isotopic abundance ratios. Rev. Sci. Inst., 21, 724-30. MacNamara, J., and Thode, H.G. (1950). Comparison of the isotopic composition of terrestrial and meteoritic sulfur. The Physical Rev., 78, 307-8. Maroulis, P.J., and Bandy, A.R. (1977). Estimate of the contribution of biologically produced dimethyl sulphide to the global sulphur cycle. Science, 196, 647-8. Matrosov, A.G., Chebotarev, Y.N., Kudryavtseva, A.J., Zyakun, A.M., and Ivanov, M. (1975). Sulphur isotope composition in fresh water lakes containing H2S. Geochem. Inter., 12 (3),217-21. Mekhtiyeva, V.L. (1971). Isotopic composition of sulphur from plants and animals in water basins of various salinities. Geokhimiya, 6, 725-30. Mekhtiyeva, V.L., and Kondrateva, E.N. (1966). Fractionation of stable isotopes of sulphur by photosynthesizing purple bacteria Rhodopseudomonas sp. (English trans.), Dok/. BioI. Sci., 166,80--83. Mekhtiyeva,V.L., and Pankina, R.G. (1968). Isotopic composition of sulphur in aquatic plants and dissolved sulphates. Geochem. Intern., 5, 624. Mitchell, R.H., and Krouse, H.R. (1975). Sulphur isotope geochemistry of carbonatites. Geochim. Cosmochim. Acta, 39, 1505-13. Nakai, N., and Jensen, M.L. (1964). The kinetic isotope effect in the bacterial reduction and oxidation of sulphur. Geochim. Cosmochim. Acta, 28, 1893-912. Newman, L., Forrest, J., and Manowitz, B. (1975). The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulphur dioxide in the plume from an oil-fired power plant. Atmospheric Environment, 9, 959-68. Nielsen, H. (1974). Isotopic composition of the major contributors to atmospheric sulphur. Tellus, 26, 213-21. Nielsen, H. (1979). Sulphur isotopes. In: Jager, E., and Hunziher, J.e. (Eds.) Lectures in Isotope Geology, Springer-Verlag, Berlin, pp. 283-312. Nielsen, H., and Ricke, W. (1964). Schwefel-Isotopenverhaltnisse von Evaporiten aus Deutschland: ein Beitrag zu Kenntnis von 34 S im Meerwasser-sulfat. Geochim. Cosmochim. Acta, 28, 577-91. Nriagu, J.D. (1978). Deteriorative effects of sulphur pollution on materials. In: Nriagu, J.D. (Ed.) Sulphur in the Environment, Part II, Wiley, New York, pp. 1-59. Nriagu, J.D., and Coker, R.D. (1978a). Isotopic composition of sulphur in precipitation within the Great Lakes Basin. Tellus, 30, 367-75. Nriagu, J.D., and Coker, R.D. (1978b). Isotopic composition of sulfur in atmospheric precipitation around Sudbury, Ontario. Nature, 274, 883-5. Nriagu, J.D., and Coker, R.D. (1983). Sulphur in sediments chronicles past changes in lake acidification. Nature 393, 692-7. Nriagu, J.D., and Harvey, J.J. (1978). Isotopic variation as an index of sulfur pollution in lakes around Sudbury, Ontario. Nature, 273, 223-4. Dana, S., and Ishikawa, H. (1966). Sulphur isotope fractionation between sulphur and sulphuric acid in the hydrothermal solution of sulphur dioxide. Geochem. J., 1, 45-50. Ohmoto, H. (1972). Systematics of sulphur and carbon isotopes in hydrothermal ore deposits. Econ. Geo!., 67, 551-78. Orr, W.L. (1974). Changes in sulphur content and isotopic ratios of sulphur during petroleum maturation-study of Big Horn Basin paleozoic oils. AAPG Bull., 50, 2295-318. Ostlund, G. (1959). Isotopic composition of sulphur in precipitation and sea water. Tellus, 11, 478-80.

Sulphur Isotopes in Nature and the Environment: An Overview

25

Postgate, J .R. (1959). Sulphate reduction by bacteria. Ann. Rev. Microbiol., 13, 505-20. Puchelt, H., Hoefs, J., and Nielsen, H. (1971). Sulphur isotope investigations of the Aegean volcanoes. Acta of the First International Science Congress on the Volcano of Thera, Athens 15-23 1969, pp. 303-17. Rafter, J.A., and Mizutani, Y. (1967). Oxygen isotope composition of sulphate. Preliminary results on oxygen isotopic variations in sulphates and relationship to their environment and to their 334Svalues. N.S. 1. Sci., 10, 816-40. Rafter, T.A., Kaplan, I.R., and Hulston, J.R. (1960). Sulphur isotopic variations in nature, part 7. Sulphur isotopic measurements on sulphur and sulphates in New Zealand geothermal and volcanic areas. N.S. 1. Sci., 3, 209-18. Rees, CE. (1970). The sulphur isotope balance of the ocean: an improved model. Earth Planet. Sci. Left., 7, 366-70. Rees, CE. (1973). A steady state model for sulphur isotope fractionation in bacterial reduction processes. Geochim. Cosmochim. Acta, 37, 1141-62. Rees, CE. (1978). Sulphur isotope measurements using SOz and SF6. Geochim. Cosmochim. Acta, 42, 383-9. Rjaboshapko, A.G., Sholayskene, D.A., and Erdman, L.K. (1978). Background levels of SOz and NOz and global balance of sulphur. Proceedings of Institute of Applied Geophysics, Vol. 39, Hydrometeorology, pp. 20-33. Ryznar, G., Campbell, F.A., and Krouse, H.R. (1967). Sulphur isotopes and the origin of the Quemont ore body. Econ. Geol., 62, 664-78. Sakai, H. (1957). Fractionation of sulphur isotopes in nature. Geochim. Cosmochim. Acta, 12, 150-69. Sakai, H. (1971). Sulphur and oxygen isotopic study of barite concretions from banks in the Japan Sea off the Northeast Honshu Japan. Geochem. 1., 5, 79-93. Sangster, D.F. (1968). Relative sulphur isotope abundances of ancient seas and strata-bound sulphur deposits. The Geol. Assoc. of Canada Proc., 19, 79-91. Sasaki, A., and Krouse, H.R. (1969). Sulphur isotopes and the Pine Point lead-zinc mineralization. Econ. Geolog., 64, 718-30. Schwarcz, H.P. (1973). Sulphur isotope analyses of some Sudbury, Ontario, ores. Can. J. Earth Sci., 10, 1444-59. SCOPE 19 (1983). In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle and Influence on It of Human Activity, Wiley, Chichester, 470 pp. Shima, M., Gross, W.H., and Thode, H.G. (1963). Sulphur isotope abundances in basic sills differentiated granites and meteorites. J. Geophys. Res., 68 (9),2838-47. Thode, H.G. (1963). Sulphur isotope geochemistry. In: Shaw, D.M. (Ed.) Studies in Analytical Geochemistry, The Royal Society of Canada, Special Publications 6, pp. 25-41. Thode, H.G. (1981). Sulphur isotope ratios in petroleum research and exploration; The Williston Basin. AAPG Bull., 65, 1527-35. Thode, H.G., and Dickman, M. (1983). Sediment sulphur distribution and its relationship to lake acidification in four lakes north of Lake Superior (Unpublished). Thode, H.G., and Goodwin, A.M. (1983). Further sulphur and carbon isotope studies of Late Archean iron-formations of the Canadian Shield and the rise of sulphate reducing bacteria. Precambrian Res., 20, 337-56. Thode, H.G., and Monster, J. (1964). The sulphur isotope abundances in evaporites and in the ancient oceans. Geochemical Conference in Moscow, 'Chemistry of the Earth's Crust', March 1963. Published in the Vernadsky Memorial, Vo\. II, March 1964 (Translated: Jerusalem, 1967).

26

Stable Isotopes

Thode, H.G., and Monster, J. (1965). Sulphur isotope geochemistry of petroleum, evaporites and ancient seas. AAPG Mem., 4, 367-77. Thode, H.G. and Rees, C.E. (1970). Sulphur isotope geochemistry of Middle East Oil studies. Endeavour, XXIX, 24-8. Thode, H.G., Graham, R.L., and Ziegler, J.A. (1945). A mass spectrometer and the measurement of isotope exchange factors. Can. J. Res., B23, 40-7. Thode, H.G., MacNamara, J. and Collins, C.B. (1949). Natural variations in the isotopic content of sulphur and their significance. Can. J. Res., B27, 361-73. Thode, H.G., Kleerekoper, H., and McElcheran, D.E. (1951). Isotope fractionation in the bacterial reduction of sulphate. Research, 4, 581-2. Thode, H.G., Harrison, A.G., and Monster, J. (1960). Sulphur isotope fractionation in early diagenesis of recent sediments of Northeast Venezuela. AAPG Bull., 44, 1809-17. Thode, H.G., Dunford, H.B., and Shima, M. (1962). Sulphur isotope abundances in rocks of the Sudbury District and their geological significance. Econ. Geol., 57, 565-78. Thode, H.G., Cragg, C.B., Hulston, J.R., and Rees, C.E. (1971). Sulphur isotope exchange between sulphur dioxide and hydrogen sulphide. Geochim. Cosmochim. Acta, 35, 35--45. Trofimov, A. (1949). Isotopic composition of sulphur in meteorites and in terrestrial objects. Dokl. Akad. Nauk. SSSR, 66, 181-4. Tudge, A.P., and Thode, H.G. (1950). Thermodynamic properties of isotopic compounds in sulphur. Can. 1. Res., B28, 567-78. Urey, H.C. (1947). The thermodynamic properties of isotopic substances. J. Chem. Soc., 1947, 562-81. Urey, H.C., Brickwidde, F.G., and Murphy, G.M. (1931). An isotope of hydrogen of mass 2 and its concentration. Phys. Res., 39, 864. Urey, H.C., and Greiff, L.J. (1935). Isotope exchange equilibria. 1. Am. Chem. Soc., 57, 321. Vinogradov, A.P., Grinenko, V.A., and Ustinov, V.I. (1962). Isotopic composition of sulphur compounds in the Black Sea. Geokhimiya, 10, 973-97. Vredenberg, L.D., and Cheney, E.S. (1971). Sulphur and carbon isotopic investigations of petroleum. Windy River Basin, Wyoming. AAPG Bull., 55, 1954-75. Yaalon, D.H., and Lomas, J. (1970). Factors controlling the supply and the chemical composition of aerosols in a near shore and coastal environment. Agricultural Meteorology, 7 (6), 443-54.

CHAPTER 2

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle


B.D. HOLT AND R. KUMAR

2.1

INTRODUCTION

Oxygen isotope ratio measurements are uniquely applicable to the study of mechanisms of formation of natural or man-made sulphates in the environment because the 180/160 ratios of sulphates that are formed from given supplies of SOz, water, and oxidants can differ, depending upon the mechanism by which they are formed. Some of the more prominent mechanisms of formation of SO i- from SOz are aqueous phase air oxidation, aqueous phase HzOz oxidation, and high-temperature oxidation to S03 (in combustion sources) accompanied by hydration to HzS04. Once formed, the sulphates of various isotopic ratios are extremely stable with respect to isotopic exchange with water, with which they may be subsequently associated. Information on mechanisms of formation of sulphates can be very useful in the establishment of limiting criteria for the origins of environmental sulphates and in strategies of pollution control. For example, if the 0180 (deviation in parts per thousand (%0) of the 180/160 ratio of the sample from that of standard mean ocean water (SMOW)) of atmospheric sulphates at a receptor site of interest is more characteristic of primary sulphates (i.e. sulphates formed within combustion sources before emission into the atmosphere) than of secondary sulphates (i.e. sulphates formed from gaseous sulphur compounds, principally SOz, by atmospheric conversion processes), it may be suspected that the major source of sulphate pollution at that site is of local rather than distant origin. In such a case, the most effective strategy for control of sulphate pollution and its contribution to acid rain in that area might be more efficient removal of sulphuric acid mist at the local sources of emission. 2.2 BASES FOR THE USE OF OXYGEN ISOTOPY

The concept of determining mechanisms of sulphate formation from oxygen isotope ratio measurements is based on the following phenomena:
Stable Isotopes ill the Assessmellt of Natural alld AllIhropogellic Sulphur ill the Ellvirollmelll Edited by H.R. Krouse and V.A. Grinenko @ 1991 SCOPE. Published by John Wiley & Sons LId

28

Stable Isotopes

(a) The rate of isotopic exchange between sulphate and water is extremely low. (b) The rate of isotopic exchange between S02 and associated water is very high in either aqueous or non-liquid systems. (c) The difference between 0180 levels in environmental water and oxidants is usually large. (d) S03 that is formed by high-temperature oxidation of S02 hydrates rapidly to form H2SO4. 2.2.1 Sulphate-water isotope exchange Measurements of isotope exchange factors between water and dissolved sulphate, as functions of temperature and pH (Lloyd, 1967; Kusakabe and Robinson, 1977; Chiba and Sakai, 1985), indicate that under typical ambient conditions, the rate of isotope exchange is extremely low. For example, it is evident from kinetic data presented by Lloyd (1967) that, even in a highly acidic rain of pH ~4, the half-time of oxygen atom exchange between sulphate and water is of the order of 1000 years. A significant result of the extremely slow exchange reaction is that, once the sulphate is formed, its 0180 (which may reflect its mechanism of formation) is preserved. Further, the 0180 of the sulphate is not appreciably affected by the heat treatments in acidified solutions that are necessary in analytical procedures for the precipitation of BaS04 (Holt et at., 1978a). 2.2.2 SOrwater isotope exchange

2.2.2.1 Aqueous systems When S02 dissolves in water, it rapidly hydrolyses according to the chemical equilibrium. S02 + H2O;;::::::: HS03
-

+ H+

(1)

(Eigen et at., 1961; Beilke and Gravenhorst, 1978). Isotopic equilibrium apparently accompanies chemical equilibrium in this reaction, so that, in the atmosphere or hydrosphere, the HS03 isotopically equilibrates with liquid water (in large excess) and upon oxidation the influence of the original 0180 of the S02 on the 0180 of the sulphate product is effectively lost (Holt et at., 1981a). Correspondingly, because of the overwhelming isotopic influence of the water on the intermediate ion HS03 - in the oxidation process, the 0180 of three of the oxygens in the sulphate product are strongly controlled by the 0180 of the water oxygens.
~

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle

29

In a variety of preparations of sulphate by Fe3+ catalysed aqueous air oxidation with water in large excess, Holt et at. (1981a) observed that the 8180 of the sulphate varied with the 8180 of the water according to regression curves of the general form
8180
504

2-

= -;).8180 4

H2O

+ C1

(2)
are

The slope of

-~ suggests

that three of the four oxygens in the sulphate

isotopically controlled by the water. In other experiments in which the 8180 of the water was held constant, the 8180 of the sulphate remained constant as the 8180 of the SOz was varied. These results confirmed that the isotopic influence of the SOz was erased by hydrolysis in excess water before oxidation. The effect of the rapid isotopic equilibration between water and dissolved SOz was further demonstrated by the 8180s01- versus 8180H20 relationship obtained when the SOz was oxidized by HzOz in aqueous solution. Using reagent-grade HzOz of constant 8180, the 8180 of the water was varied, yielding a regression curve of the form
8180
S04

-38180
"5

H2O

+C

(3)

The

slope suggests that three of the five oxygens in the intermediate

adduct HS03 -. HzOz were isotopically controlled by the solvent water, two by the oxidant and essentially none by the SOz. Figure 2.1 gives a comparison of isotopic results from Fe 3+-catalysed aqueous air oxidation, charcoalcatalysed aqueous air oxidation, and HzOz aqueous oxidation in curves A, B, and C respectively (Holt et at., 1981a).

2.2.2.2 Systems with no tiquid water Experimental results have indicated rapid isotopic equilibration between SOz and water vapour in mixtures of SOz, air, and water vapour (no liquid water), confined to a 3-litre glass container (Holt et at., 1983). Figure 2.2 shows the results of equilibrating SOz of constant 8180 (-14%0) with water vapours of various 8180 values. Curve A represents experiments in which the equilibrated SOz was recovered by evacuating the SOrair-water vapour mixture through a dry-ice cold trap (-79C) to remove the water vapour and a liquid-nitrogen cold trap to remove the S02; curve B represents experiments in which the procedure was the same, except that the water vapour was removed by passing the mixture through magnesium perchlorate for chemical absorption at 22C.

30
3Z
A 24 B

Stable Isotopes

!!.. <I> 0 .<::

-0 "e
Q.

16

:; "' 0 '" Co

-8

-16

. -20

-10

0
8'8 0, water (%0)

10

20

Figure 2.1 Isotopic results of oxidation of SOz to sulphate. Curve A, Fe3+-catalysed air oxidation; curve B, charcoal-catalysed air oxidation; curve C, aqueous phase HzOz oxidation

Rapid isotopic equilibration between the S02 and the water in these systems is indicated by the effect of temperature, at the point of water vapour separation, On the OIXOSOZ versus OIXOHzOrelationship. In both sets of experiments, curves A and B, the SOrair-water vapour mixtures were aIlowed to stand at room temperature for about half an hour before recovery of the S02. The results, curve A, suggest striking evidence that the isotopic equilibration responded very rapidly to the change in temperature as the gas mixture flowed through the cold trap (-79C) in which the water vapour was removed; the difference of 19%0between the y intercepts of the two curves is in approximate agreement with the difference in the thermodynamic fractionation factors that are calculable from spectroscopic data for the gaseous species, S02 and H2O, at the two temperatures, -79 and 22C. In the reaction by which the isotopic equilibration apparently occurs,
~

SOig) + H2O(h) ~ S02 . H2O

(4)

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle 70

31

60

Slope

LOl~

50
0

~
CJ) "0

x
40 t ~19
'"

~o(\J

~ .0
.;
CD ~c:.o

I
/00

g 30
I
[::,

20

- 0.94 10 X A Cryogenic separation at -79 C B Chemical separation at 22 C -20 -10


8'8 water

0
vapour (%0)

10

20

Figure 2.2 Isotopic equilibrations between S02 and water vapour at two temperatures (Holt et al., 1983)

the structural form of S02 . H2O (e.g. gas phase or surface adsorbed) is unspecified. One important consequence of rapid isotopic equilibration between S02 and water vapour is that in SOz-air-water vapour mixtures, in which the water vapour is in large excess (as in the atmosphere), the 8180 of the S02 does not remain unchanged from that at its point of origin; rather, it is dynamically controlled by the 8180 of the associated water vapour. Another consequence of rapid isotopic equilibration is that oxidation of S02 in humidified air yields sulphate, three oxygens of which are isotopically

32

Stable Isotopes

controlled by the 0180 of the associated water vapour. Thus, a three-quarters control of sulphate oxygens by associated water is possible in both aqueous and non-liquid systems, although the mechanisms responsible for the two transformation processes are entirely different. 2.2.3 0180 values of S02 oxidants and of associated water A potential advantage to the use of 8180 measurements in the study of mechanisms of sulphate formation is the relatively large difference between the 8180 of oxygen in air and the 8180 of waters of the atmosphere and the hydrosphere. A mechanism by which the 8180 of the sulphate is dominated by the 8180 of the water (0.0%0 for sea water, down to -50%0 for atmospheric water) is expected to yield a sulphate that is generally lower in 8180 than one by which the 8180 of the sulphate is dominated by the 8180 of air (~23%0; Horibe et at., 1973). 2.2.4 Rapid hydration of SO} Sulphur trioxide is formed at high temperatures in S02-0rSO} systems in exhaust gases from smelters, power plants, combustion engines, etc., and is continuously emitted into the atmosphere. As the exhausted gases cool, the S03 rapidly reacts with combustion-produced water vapour, and, since the combustion-produced water vapour is relatively high in 8180 (because of its prior high-temperature isotopic equilibration with air oxygen), the hydrated sulphuric acid that is formed has a relatively high 8180. This relatively high 8180 of H2SO4 of high-temperature S03 origin thereby characterizes its mechanism of formation (Holt and Kumar, 1984). 2.2.5 Sulphide oxidation On the basis of experiments bubbling O2 through Na2S solutions, Lloyd (1967) suggested that the inorganic oxidation of sulphide to sulphate takes place in two steps:
S2- + H2O + O2 ~ S032S032+

(5) (6)

i O2

S042-

He concluded that there is no oxygen isotope selectivity during the incorporation of water oxygen. For reaction (6), he attributed a kinetic isotope effect of 8.7%0favouring the lighter isotope during incorporation of O2, Lloyd calculated that two-thirds of the oxygen comes from water and one-third from molecular oxygen. This ratio is influenced by oxygen exchange between S032- and water (Section 2.2.2.1).

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle

33

Taylor et al. (1984) cited three reactions involved in the oxidation of FeS2: FeSz + 14 Fe3+ + 8 HzO ~ 15 Fe2+ + 2 SOiFez+ + i Oz + H+ ~ Fe3+ + ! HzO FeSz + ~02 + H2O ~ Fe2+ + 2 SOi+ 2 H+ + 16 H+ (7) (8) (9)

In reaction (7), all sulphate oxygen is derived from water molecules, whereas in reaction (9), stoichiometrically, 87.5% of the sulphate oxygen is derived from molecular oxygen and 12.5% from water molecules. Taylor et at. (1984) plotted OHIOvalues for SOi- in acid mine waters against those of associated HzO and, on the basis of theoretical lines, estimated the relative contributions of the above reactions. These theoretical lines were revised by van Everdingen and Krouse (1985).

2.2.6 Bacterial SOi- reduction Bacteria preferentially metabolize 160 during the reduction of SOi- to HzS. This leaves the remaining SOi- enriched in 180 according to the relationship
0180 - 01800 = 1000 (1 - <x)InF

where F is the fraction reacted; 0180 and 01800 refer to remaining and initial sulphate respectively. Lloyd (1967) found that 1000 (1 - <x) = -4.6%0. Mizutani and Rafter (1969b) reported that the ratio of 0I80 to 034S of the remaining SOi- was approximately 1:4. This observation is interesting because of the one sulphur atom to four oxygen atoms in the SO/structure. However, the same authors (1973) found departures from this ratio when waters of different 0180 values were used in the reduction medium. The 0180 value of the remaining sulphate was a function of that of the water. Since it is known that the rate of exchange between SO/and HzO is extremely low under the conditions of their experiments, it was concluded that the isotopic composition of SOiwas influenced by backreaction of intermediates that had exchanged oxygen with water.

2.2.7 Sulphate crystallization During crystallization of CaS04.2H20 in natural evaporation pans, the precipitates were found to be enriched in 180 by 3.6 ::!:: 0.9%0compared to dissolved sulphate (Lloyd, 1967).It is not certain whether this represents equilibrium isotope fractionation or a kinetic effect.

34

Stable Isotopes

2.3 APPLICATIONS OF 180STUDIESTO ENVIRONMENTAL SULPHATES


Oxygen isotope ratio measurements have been applied to studies of mechanisms of sulphate formation in the atmosphere, particularly in relation to acid rain and to studies of formation mechanisms and isotope thermometry (Mizutani and Rafter, 1969a) in the hydrosphere and the lithosphere.

2.3.1

Atmosphere

2.3.1.1 Precipitation sutphates Isotopic analyses of sulphates in precipitation have led to at least three significant observations. Firstly, the 01110of the sulphate varies seasonally and in phase with corresponding isotopic variations of the precipitation water (Cortecci and Longinelli, 1970; Longinelli and Bartelloni, 1978; Holt et at., 1981b; see Figure 5.8). Holt et at. (1979a) noted that in samples collected at Argonne, Illinois, the amplitude of the seasonal variation of

0180 in precipitation sulphate was ~i of that of the precipitation water. In


view of results from laboratory experiments discussed in Section 2.2.2.1, this suggests that the isotopy of three of the four oxygens in precipitation sulphate is controlled by the precipitation water, and therefore that essentially complete isotopic equilibration occurs between the precipitation water and

dissolved HS03 -, prior to significantoxidation to sulphate.


A second observation is that the 0180 of precipitation sulphates is higher, in relation to the associated precipitation water, than that of sulphates prepared in the laboratory by a variety of methods to simulate secondary sulphate formation and lower than that of sulphates prepared in the laboratory by high-temperature S03 formation to simulate primary sulphate formation. (Primary and secondary sulphates are defined in Section 2.1). This is illustrated by experimental results in Figure 2.3 (Holt et at., 1982). The rain and snow data represent samples collected at Argonne, Illinois, during the period October 1976 to March 1978. Curves 1 to 3 represent 0180S0/- versus 0180HzOrelationships for sulphates prepared in the laboratory by aqueous phase reactons (Holt et at., 1981a), corresponding to secondary sulphate formation in the atmosphere; curves 4 to 7, sulphates prepared by non-liquid phase reactions (Holt et at., 1983), also corresponding to secondary sulphates; and curve 8, sulphate prepared by high-temperature reactions (Holt and Kumar, 1984), corresponding to primary sulphates in the atmosphere. Specifically, curve 1 was by Fe3+-catalysed aqueous air oxidation; curve 2, charcoal-catalysed aqueous air oxidation; curve 3, aqueous HzOz oxidation; curve 4, non-liquid air oxidation in the presence of an electric spark (simulating lightning in the atmosphere); curve 5, non-

Oxygen Isotope -Fractionation for Understanding the Sulphur Cycle


50

35

40-

PRIMARY

-8

30
'0

1
SNOW RAIN 0

6 4
7

t ", 10
20

-ro

.
1

. r.,.

r... .

0 Q>0

.000Q0-

0 0 <iJ 0

-10

2 5 3 -20 -10 0
10 20

-20 -30
8180, liquid water (%0)

Figure 2.3 Isotopic comparison of atmospheric sulphates to laboratory-prepared sulphates (see text for identification of curves)

liquid air oxidation in the presence of N02 (simulating NOx pollution in the atmosphere); curve 6, non-liquid air oxidation in the presence of gamma irradiation (simulating OH radicals in the atmosphere); curve 7, non-liquid oxidation after adsorption of S02 on charcoal in air (simulating S02 adsorption on soot particles in the atmosphere); curve 8, air oxidation of S02 to S03 at high temperature (catalytic surface at ~450-500 q, accompanied by hydration of the S03 to H2SO4. The experimental results in Figure 2.3 illustrate that the 0180 of the 'primary' sulphate (derived from high-temperature S03) is much higher and less dependent on the 0180 of the associated water than the group of 'secondary' sulphates. The results also illustrate that the 0180 data for precipitation sulphates fall in between the expected values for primary and secondary sulphates, suggesting that they comprise a mixture of the two types.

36

Stable Isotopes

A third observation aboutprecipitationsulphates is that, at receptorsites near sea coasts, the 3180may be significantlyaffected by sea spray (Mizutani and Rafter, 1969c;Cortecci and Longinelli, 1970).
2.3.1.2 Aerosol sulphates During the same period of collection of precipitation samples of Figure 2.4 (December 1976 to March 1978), Holt et al. (1981b) collected consecutive 7-day samples of aerosol sulphates and water vapour at the Argonne site. A comparison of the month-averaged 3180 values for both precipitation sulphates and aerosol sulphates is shown in Figure 2.4. As substantiated by isotopic results on other samples that had been collected intermittently during 1975 and 1976, the oxygen isotopies of precipitation sulphate and aerosol sulphate indicate that the respective mechanisms of formation are not identical throughout the year. Heterogeneous, aqueous phase oxidation appears to be the predominant mechanism of precipitation sulphates at all times. In contrast, although aerosol sulphates may also be formed by the same mechanism during late autumn, winter, and early spring, a competing mechanism that tends to lower the 3180 may be present during the summer months. Despite these apparent differences in the seasonal variation of 3180 in precipitation and aerosol sulphates, both types fall between the primary and 20
PRECIPITATION 0".8.~ , 'B~--=>
o~-e~ ~&
0 0

15

10
......

0 co
'Q

G~~6)

,\

,\ ,,
\, 6

,:

,, , ,, , ,

,
-,is!, ''Q

, '~

0
u ~
Q

Z <
~

~ ~ ~

~ <

1976

~ ~

<

~ <

z ~
~

~ ~
~

~ ~

1977

<

~ ~

~ u
0

~ 0

u ~
Q

Z <
~

~ ~
~

~ 1978

~ <

Figure 2.4 Comparison of 81HOin precipitation sulphates and aerosol sulphates at Argonne, Illinois, USA (Holt et al., 1981b)

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle

37

secondary sulphate curves in Figure 2.3. This suggests that both types may contain appreciable amounts (10-30%) of primary sulphate at Argonne, Illinois (Holt et at., 1982). In other studies results indicated that, depending on meteorological conditions, the 8180 of aerosol sulphates often varies inversely with barometric pressure (Holt et al., 1981b) and that it can vary diurnally and in conjunction with varying origins of sulphur emissions, as determined by backtrajectories of air-mass movements (Holt et at., 1978b). 2.3.1. 3 Sutphur dioxide As suggested in Section 2.2.2, the rapid isotopic equilibration that occurs between S02 and either liquid water or water vapour has important implications regarding the 8180 of S02 in the atmosphere. One is that atmospheric S02 is very probably in dynamic isotopic equilibrium with the water vapour with which it is associated; consequently, any change in 8180 of water vapour (e.g. by precipitation, evaporation of ground water, etc.) probably causes an immediate corresponding change in the 8180 of the S02 in the same air mass. For this reason, the 8180 of S02 in the atmosphere cannot be used as a fingerprint of its origin (combustion source; volcano; product of oxidation of H2S, S, or other sulphur compounds in the atmosphere; etc.). Another implication is that the 8180 of S02 in the atmosphere cannot be sampled and converted to CO2 for mass spectrometric analysis without concurrent association with water with which it can immediately undergo isotopic exchange (Holt et at., 1979b). 2.3.2 Hydrosphere (see also Chapter 6) Measurements of oxygen isotope ratios in water and in dissolved sulphates afford a unique method of studying important bacterial, chemical, and physical processes that occur in the oceans, lakes, streams, and ground waters (Longinelli and Craig, 1967; Rafter and Mizutani, 1967). 2.3.2.1 Oceans It has been observed that the 8180 of ocean sulphate (~9.5%0) apparently does not correspond to isotopic equilibrium between ocean sulphate and ocean water (8180 = 0.0%0)(Lloyd, 1967, 1968; Holser et at., 1979; Claypool et at., 1980). The complete explanation of this apparent lack of isotopic equilibrium is still debatable. Some of the possible causes which have been proposed are: (a) oxygen isotope fractionation in bacterial reduction of SOi- to lower-valence species of sulphur compounds, (b) oxygen isotope

38

Stable Isotopes

exchange between sulphur compounds of intermediate valence states and ocean water, (c) oxygen isotope fractionation in bacterial and/or chemical oxidation of reduced species to SO}-, (d) a steady-state ratio of sulphate inputs and outputs of the sea, and (e) oxygen isotope fractionation in the physical processes of crystallization of sulphate minerals (to form evaporites) and oxidative weathering of pyrites. 2.3.2.2 Streams and lakes The 8180 of sulphates in streams and lakes can be quite varied, depending on natural and anthropogenic inputs and on bacterial activity. For example, results obtained by Longinelli and Cortecci (1970) on the Serchio River in Italy showed that the 8180, 834S,and concentration of the sulphates generally increased downstream along the course of the river. A 834S/8180ratio of 4:1 has been found for S042- obtained from different depths in Lake Yanda, Antarctica (Rafter and Mizutani, 1967). In contrast, for a stratified ice-covered lake on Ellesmere Island in the Canadian Arctic, the 834S/8180 ratio was highly variable with an average value near unity (Jeffries et al., 1984). 2.3.2.3 Ground waters Oxygen -18 analyses of ground waters have been effectively used in determining whether the sulphates in the ground waters are from solution of evaporites or from aqueous air oxidation of sulphides. Longinelli (1968) found no relation among &180 of sulphate, 8180 of water, and water temperature in thermal springs in Tuscany, Italy. This was interpreted as evidence that the sulphates in the thermal waters came from dissolution of Upper Triassic evaporites. Shakur (1982) also used 8180 and 834S data to study sulphate geochemistry of freshwater wells, springs, and associated deposits (see Section 6.7.3). In one such study, very negative 8180 values of radioactive barite sinters proved that HS- was being oxidized in the water before precipitation as Ba(Ra)S04 (Cecile et al., 1984). 2.3.3 Lithosphere (see also Chapter 4) Shakur (1982) used 8180 data to identify secondary alteration of primary deposits of marine evaporites (gypsum and anhydrite). The significant difference in the 8 values of Upper Cambrian and Lower Devonian evaporites (see Section 4.2) was used to locate the position of a thrust-fault in Palaeozoic strata in the Norman Range, Northwest Territory, Canada (van Everdingen et al., 1982). This represents a direct application of stable isotope data in subsurface mapping.

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle

39

2.4

CONCLUSIONS

Oxygen isotope analysis of atmospheric sulphates and atmospheric water has potential use in the estimation of the fractions of primary and secondary sulphates at any given receptor site. A relatively large fraction of primary sulphate corresponds to local sources of SOx emissions and a relatively small fraction to distant sources. The fractions of the two types of sulphates multiplied by the total loadings of sulphate, as a function of time, correspond to the respective temporal variations in loadings at a receptor site. Measurements of 8180 of sulphates and of associated waters can be applied to the study of bacterial, chemical, and physical processes that regulate the steady-state balance of the sulphur-oxygen system in the oceans. Similarly, 8180 measurements can be made to determine whether the sulphates in certain ground waters originate from solution of evaporite sulphates or from aqueous air oxidation of sulphides. Such information is important to geological subsurface mapping. REFERENCES
Beilke, S., and Gravenhorst, G. (1978). Heterogeneous S02 oxidation in the droplet phase. Atmos. Environ., 12,231-9. Cecile, M.P., Goodfellow, W.D., Jones, L.D., Krouse, H.R., and Shakur, M.A. (1984). Origin of radioactive barite sinter, Flybye Springs, Northwest Territories, Canada. Can. 1. Earth Sci., 21, 383-95. Chiba, H., and Sakai, H. (1985). Oxygen isotope exchange between dissolved sulfate and water at hydrothermal temperatures. Ceochim. Cosmochim. Acta, 49, 993-1000. Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H., and Zak, I. (1980). The age curves of sulphur and oxygen isotopes in marine sulphate and their mutual interpretation. Chem. Ceo!., 28, 199-260. Cortecci, c., and Longinelli, A. (1970). Isotopic composition of sulphates in rainwater, Pisa, Italy. Earth Planet. Sci. Left., 8, 3~0. Eigen, M., Kusten, K., and Maas, A. (1961). Die Geschwindegkeit der Hydration von S02 in wassriger Lasung. Z. Phys. Chem., 30, 130. Holser, W.T., Kaplan, I.R., Sakai, H., and Zak, 1. (1979). Isotope geochemistry of oxygen in the sedimentary sulphate cycle. Chem. Ceo!., 25, 1-17. Holt, B.D., and Kumar, R. (1984). Oxygen-18 study of high-temperature air oxidation of S02' Atmos. Environ., 18, 2089-94. Holt, B.D., Cunningham, P.T., and Engelkemeir, A.G. (1978a). Application of oxygen -18 analysis to the study of atmospheric sulphate formation. In: Robinson, B.W. (Ed.) Stable Isotopes in the Earth Sciences, New Zealand DSIR Bulletin 220, pp. 105-9. Holt, B.D., Kumar, R., Cunningham, P.T., Bouchard, M., Engelkemeir, A., Johnson, S.A., and Nielsen, E.L. (1978b). Regional oxygen-18 variations in particulate sulphate and water vapor at three sampling sites about 100 km apart. Environ. Sci. Tech., 12, 1394-8. Holt, B.D., Cunningham, P.T., and Kumar, R. (1979a). Seasonal variations of oxygen-I8 in atmospheric sulphates. Int. J. Environ. Ana!. Chem., 6, 43-53.

40

Stable Isotopes

Holt, B.D., Kumar, R., and Engelkemeir, A.G. (1979b). Interference by isotopic exchange in the determination of &180 in environmental sulphur dioxide. In: Klein, E.R., and Klein, P.D. (Eds.) Stable Isotopes: Proceedings of the Third International Conference, Academic Press, New York, pp. 223-9. Holt, B.D., Kumar, R., and Cunningham, P.T. (1981a). Oxygen-18 study of the aqueous-phase oxidation of sulphur dioxide. Atmos. Environ., 15, 557-66. Holt, B.D., Cunningham, P.T., and Kumar, R. (1981b). Oxygen isotopy of atmospheric sulphates. Environ. Sci. Tech., 15, 804-8. Holt, B.D., Kumar, R., and Cunningham, P.T (1982). Primary sulphates in atmospheric sulphates: estimation by oxygen isotope ratio measurements. Science, 217,51-3. Holt, B.D., Cunningham, P.T, Engelkemeir, A.G., Graczyk, D.G., and Kumar, R. (1983). Oxygen-18 study of nonaqueous-phase oxidation of sulphur dioxide. Atmos. Environ., 17, 625-32. Horibe, Y., Shigehara, K., and Takakuwa, Y. (1973). Isotope separation factors of carbon dioxide-water system and isotopic composition of atmospheric oxygen. J. Geophys. Res., 78, 2625-9. Jeffries, M.O., Krouse, H.R., Shakur, M.A., and Harris, S.A. (1984). Isotope geochemistry of stratified Lake 'A', Ellesmere Island, N.W.T, Canada. Can. 1. Earth Sci., 21, 1008-17. Kusakabe, M., and Robinson, B.W. (1977). Oxygen and sulphur isotope equilibria in the BaSOcHS04 --H2O system from 110 to 350 DCand applications. Geochim. Cosmochim. Acta, 41, 1033-40. Lloyd, R.M. (1967). Oxygen-18 composition of oceanic sulphate. Science, 156, 1228-31. Lloyd, R. (1968). Oxygen isotope behavior in the sulfate-water system. J. Geophys. Res., 73, 6099-6110. Longineeli, A. (1968). Oxygen isotopic composition of sulfate ions in water from thermal springs. Earth Planet Sci. Leu., 4, 206--10. Longinelli, A., and Bartelloni, M. (1978). Atmospheric pollution in Venice, Italy, as indicated by isotopic analysis. Water, Air, Soil Pollut., 10, 335-41. Longinelli, A., and Cortecci, G. (1970). Isotope abundance of oxygen and sulphur in sulphate ions from river water. Earth Planet. Sci., 7, 376--80. Longinelli, A., and Craig, H. (1967). Oxygen-18 variations in sulphate ions in sea water and saline lakes. Science, 156, 56--9. Mizutani, Y., and Rafter, TA. (1969a). Oxygen isotopic composition of sulphates: part 3. Oxygen isotopic fractionation in the bisulphate ion-water system. NZ: J. Sci., 12, 54-9. Mizutani, Y., and Rafter, TA. (1969b). Oxygen isotopic composition of sulphates: part 4. Bacterial fractionation of oxygen isotopes in the reduction of sulphate and in the oxidation of sulphur, NZ. 1. Sci., 12, 60-7. Mizutani, Y., and Rafter, TA. (1969c). Oxygen isotopic composition of sulphates: part 5. Isotopic composition of sulphate in rainwater, Gracefield, New Zealand. NZ. J. Sci., 12,69-80. Mizutani, Y., and Rafter, T.A. (1973). Isotopic behavior of sulphate oxygen in the bacterial reduction of sulphate. Geochem. 1.., 6, 183-91. Rafter, TA., and Mizutani, Y. (1967). Oxygen isotopic composition of sulphates: part 2. NZ. J. Sci., 10, 816--40. Shakur, M.A. (1982). &34Sand &180 variations in terrestrial sulphates. A doctoral thesis, submitted to the Department of Physics, University of Calgary, Calgary, Alberta, Canada.

Oxygen Isotope Fractionation for Understanding the Sulphur Cycle

41

Taylor, B.E., Wheeler, M.C., and Nordstrom, D.K. (1984). Isotope composition of sulphate in acid mine drainage as measure of bacterial oxidation. Nature (Lond.), 308 (5959), 538-41. van Everdingen, R.O., and Krouse, H.R. (1985). The isotope composition of sulphate generated by bacterial and abiological oxidation. Nature, 315, 395-6. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1982). 34S and IRO abundances differentiate Upper Cambrian and Lower Devonian gypsum-bearing units, District of Mackenzie, NWT-an update. Can. J. Earth Sci., 19, 1246--54.

CHAPTER 3

The Isotopic Analysis of Sulphur and Oxygen


C.E REES AND B.D. HOLT

3.1

THE ISOTOPIC ANALYSIS OF SULPHUR*

3.1.1 Introduction This chapter is concerned with the ways in which sulphur samples are treated chemically, converted to a suitable gaseous form, and analysed mass spectrometrically. In order for sulphur isotope analyses to be useful it is necessary that sample handling introduces no isotope fractionation. In general this means that extractions, chemical reactions, etc., should be quantitative. The total range of 834Svalues encountered in nature is ~ 150%0. Modern mass spectrometers can analyse sulphur isotope ratios with a reproducibility of ~:i:0.05%0, whereas the overall reproducibility for sample preparation and analysis is usually closer to :i:0.2%0.Precision can be checked within a laboratory by replicate analysis of particular samples. In order to check accuracy, though, it is desirable to perform analyses on standard materials. There are two isotope reference materials for sulphur available from the International Atomic Energy Authority, Vienna (Gonfiantini, 1978), while in addition a set of 10 silver sulphide samples prepared at McMaster University has been widely circulated (Rees, 1978). 3.1.2 The chemical pretreatment of sulphur samples 3.1. 2.1 Introduction Chemical pretreatment of sulphur samples is necessary in order to convert the sulphur in the sample to a form (usually silver sulphide, Ag2S, or barium sulphate, BaS04) that is suitable for further processing to sulphur dioxide (or sulphur hexafluoride) for mass spectrometric analysis (see Sections 3.1.3 and 3.1.4). In simple cases there may be only one form of sulphur present
*

C.E.Rees.

Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko @ 1991 SCOPE, Published by John Wiley & Sons Ltd

44

Stable Isotopes

in the sampleor different forms of sulphur may be combined for the analysis of total sulphur. In other cases it may be desirable to separate different forms of sulphur so that their individual concentrations and isotopic compositions may be determined. Thus it is clearly not possible to set out a procedure that will be applicable under all circumstances. The chemical pretreatment strategy will vary according to what forms of sulphur are present, in what matrix (if any) they are contained, and the extent to which separation and individual analysis of particular components are required. This section does not give complete details of particular schemes of chemical pretreatment. Instead a number of such schemes are outlined and the sources are indicated where complete details may be found. 3.1. 2.2 The conversion of barium sulphate to silver sulphide This conversion is necessary if sulphur dioxide or sulphur hexafluoride is to be produced from silver sulphide. Rafter (1967) described the conversion of barium sulphate to barium sulphide (BaS) by reaction with graphite at 1050-1100C. The BaS so produced is converted to silver sulphide by reaction with silver nitrate (AgN03). A modification of this method can be used not only to produce BaS but also to convert the oxygen from the sulphate to carbon dioxide for oxygen isotopic analysis of sulphate. Thode et al. (1961) described the use of 'reducing mixture' to convert barium sulphate and other sulphates to hydrogen sulphide. The reducing mixture, originally introduced by Auger and Gabillon (1911), consists of 500 ml HI, 816 ml concentrated HCl, and 245 ml H3P02 (50%). Reduction is carried out in a 200-ml flask provided with a reflux condenser. The H2S swept out with a stream of nitrogen is washed with distilled water and finally absorbed in a solution containing cadmium acetate and distilled water. The cadmium sulphide produced is converted to silver sulphide by adding silver nitrate. 3.1.2.3 Sulphur in the atmosphere and in water Forrest and Newman (1973) have reported a method for sampling atmospheric oxides and for processing the samples for the determination of 834Svalues. Particulate sulphate is collected on glass filters while atmospheric sulphur dioxide is absorbed on alkali-impregnated filter paper. Filter packs are assembled from a commercial high-volume sampler modified to hold the particulate prefilter followed by one or more sulphur dioxide absorbers. The same technique has also been used to sample sulphur compounds in the plumes from the stacks of power plants and smelters (Newman et al., 1975; Forrest and Newman, 1977a, 1977b). Forrest et al. (1973) devised a high-

The Isotopic Analysis of Sulphur and Oxygen

45

temperature probe to collect separately sulphur dioxide and sulphur trioxide in power plant flue gases. Chemical pretreatment of such samples prior to the production of sulphur dioxide for mass spectrometric analysis includes the oxidation of sulphur dioxide to sulphate by treatment with hydrogen peroxide (HzOz), the precipitation of barium sulphate by the addition to sulphate in acid solution of barium chloride, and the conversion of barium sulphate to silver sulphide as described in Section 3.1.2.2 above. Sulphur is determined at the silver

sulphide stage by using silver containing a

110Ag

tracer and gamma counting

the resultant activity in the silver sulphide (Forrest and Newman, 1977c). Sulphate in water samples such as rain, river water, lake water, ocean water, etc., is converted to BaS04 by the addition of barium chloride. Sulphate may be preconcentrated from dilute solutions using ion exchange columns (e.g. Mizutani and Rafter, 1969). The BaS04 is further processed as required (e.g. Holt et al., 1972; Nriagu and Coker, 1978; Rees et al., 1978). 3.1.2.4 Sulphur in vegetation, soils, and sediments Sulphur in samples of vegetation can be extracted by Parr bomb conversion to sulphate (Case and Krouse, 1980) or by oxidizing fusion to sulphate (Mekhtiyeva et al., 1976), precipitation of barium sulphate following in each case. For samples with very low sulphur contents, preliminary ashing to reduce the bulk of the sample may be desirable (Johnson and Nishita, 1952). For soil samples it may be desirable to separate different forms of sulphur for individual analysis. Thus Lowe et al. (1971) separated out soluble sulphate fractions by leaching with distilled water. The polysaccharide sulphate fraction was extracted by leaching the residue of the first step with 1% NaCl and humic acid fractions were obtained by extraction with 0.1 N NaOH. Sulphur from all three fractions was converted to barium sulphate prior to conversion to silver sulphide. Freshly formed sediments contain appreciable amounts of pore water which may contain hydrogen sulphide (if the sedimentation conditions are anoxic) and will also contain dissolved sulphate. In addition, the sediment may contain acid-soluble sulphides, elemental sulphur, pyrite, and organic sulphur. Care must be taken when handling such sediments that hydrogen sulphide and iron monosulphides are not oxidized accidentally by exposure to the atmosphere. Various procedures have been used to extract different forms of sulphur from both fresh and ancient sediments, although none of them seems to be entirely satisfactory. Detailed procedures are given in, for example, Thode et at. (1960), Vinogradov et at. (1962), Kaplan et at. (1963), Smith et at. (1964), Nriagu (1975), and Nriagu and Coker (1976).

46

Stable Isotopes

A typicalprocedurewould be:
(a) Strip the sample of HzS by flushing with nitrogen. The HzS is absorbed in a cadmium acetate solution for further processing. (b) Add HCI to the residue of the previous step to liberate HzS from acidsoluble sulphide. The HzS is absorbed in a cadmium acetate solution. (c) Filter the residue of the previous step and add BaClz to the filtrate to precipitate sulphate as BaS04. (d) Extract elemental sulphur from the solid residue of the previous step by Soxhlet extraction with benzene. The extracted sulphur is oxidized to sulphate with a HNOrBrz mixture and precipitated as BaS04' (e) Treat the residue of the previous step with 6N HN03 to oxidize pyrite sulphur to sulphate. The material is filtered and the sulphate in the filtrate is precipitated as BaS04' (f) Convert the organic sulphur in the solid residue of the previous step to sulphate by fusion oxidation or treatment with HNOrBrz.

The difficulties which may arise include the accidental oxidation of hydrogen sulphide and acid-soluble sulphide before analysis,the overestimation of elemental sulphur by producing it during the extraction of the acidsoluble sulphide step (Berner, 1974),the partial extraction of organic sulphur together with pyrite, and the failure to extract all pyrite, some of which may then be included in the organic sulphur fraction. With regard to the separation of pyrite sulphur and organic sulphur, Monster (1978a) has critically compared the Li-AI-H4 method (Smith et al., 1964) and the nitric acid method, outlined above. Some separation procedures employ solvent extraction for organic sulphur. Monster (1978b) has pointed out the well-known fact that only a fractional amount of the total amount of organic material in recent sediments is extractable with organic solvents (see also Smith, 1952, 1954). Sasaki et al. (1979) have cautioned against the use of the reducing mixture reaction for the extraction of whole rock sulphur, pointing out that this reagent reacts with pyrite slowly so that quantitative extraction of sulphur is a barely accomplished event after several tens of hours processing. Further, because silicates are not decomposed by the reducing mixtures, fine-grained inclusions of sulphur-bearing material therein may have no chance to react with the solution. Tin (H)-strong phosphoric acid (Kiba reagent) was first described by Kiba et al. (1955) as a powerful reducing agent capable of taking sulphate to hydrogen sulphide. Sasaki et al. (1979) described the use of this reagent in sulphur isotope studies. They found that Kiba reagent converts both sulphate and pyrite to hydrogen sulphide but some other minerals including copper

The Isotopic Analysis of Sulphur and Oxygen

47

sulphides, arsenopyrite, and molybdenite do not react completely. Therefore caution must be used when determining whole rock sulphur. More recently Ueda and Sakai (1983) discussed the use of Kiba reagent in the extraction from rock samples of sulphate sulphur as sulphur dioxide and of sulphide sulphur as hydrogen sulphide. These gases may be separated by vacuum distillation for volumetric determinations of sulphate and sulphide concentrations and subsequent isotopic analysis. 3.1. 2.5 Sulphur in petroleum, coal, and sulphur ores Sulphur in petroleum is extracted by Parr bomb oxidation to sulphate followed by precipitation of barium sulphate. Coal contains sulphur in various chemical forms and oxidation statesincluding pyrite, other metallic sulphides, sulphates, and organic sulphur. In some cases it may be sufficient to measure the total sulphur concentration and its isotopic composition whereas in other cases it may be desirable to determine the isotopic compositions of different extracts. For total sulphur determinations either the Eschka (ASTM, 1971) or Parr bomb method may be used. Hicks et al. (1974) have recommended the Parr bomb method as being faster than the Eschka method, but personal experience in this laboratory suggests that the Parr bomb method is subject to variable yields. Various schemes exist for the sequential extraction of different forms of sulphur for isotopic analysis. A recent article by Westgate and Anderson (1982) refers to previous work and gives details of an extraction scheme which includes: (a) (b) (c) (d) (e) Massive pyrite removal Extraction of acid-soluble sulphides Extraction of sulphate sulphur Extraction of different disseminated pyrite Extraction of organic sulphur

Thus the extraction of sulphur from coal is similar to the extraction of sulphur from other sediment samples in terms of the need to make effective separations of different forms of sulphur which may differ drastically in their isotopic compositions. Two recent developments deserve mention. Krouse et al. (1987) described experiments involving the release of HzS from fossil fuels during linearly temperature programmed pyrolysis. This technique may have application as an analytical technique for sulphur compounds and as a means of understanding the organic geochemistry of sulphur in fossil fuels.

48

Stable Isotopes

Kelly et at. (1982)described the extractionof total sulphurfrom coalfor sulphur concentration determination by isotope dilution mass spectometry. This technique is likely to become the method against which other techniques will be tested. Sulphur ores can be treated with hydrochloric acid if they are acid-soluble forms or oxidized to sulphate using HNOrBrz, as in the case of pyrite. 3.1.3 The preparation of sulphur samples for mass spectrometry 3.1.3.1 Introduction The isotopic analysis of sulphur requires the production of a suitable gaseous compound of the element which can be introduced into the source region of a gas source mass spectrometer. The normal gas used for sulphur isotope analysis is sulphur dioxide and there are three distinct methods used for its production: (a) The high-temperature combustion of sulphide sulphur in a stream of oxygen or oxygen plus nitrogen (b) The high-temperature oxidation of sulphide sulphur using a solid oxygen donor such as vanadium pentoxide (V20s) or copper oxide (CuO or CU20) (c) The direct, high-temperature, conversion of sulphate sulphur to sulphur dioxide The first of those methods is now largely of historical interest and is not widely used. However, a brief description of the method is given below because it serves as a convenient introduction to the advantages of the latter two methods. A small number of laboratories use or have used the gas sulphur hexafluoride (SF6) for sulphur isotope analysis. The preparation of sulphur hexafluoride is less convenient than the preparation of sulphur dioxide but the sulphur hexafluoride method has some advantages over the sulphur dioxide method when very high accuracy and precision are required. Since fluorine has only one stable isotope, the sulphur hexafluoride is also to be recommended when measurements of the rare isotopes 33S and 36S are to be made. 3.1. 3.2 Combustion of sulphides in oxygen Descriptions of this method have been given by Thode et al. (1949), Rafter (1957), and Thode et al. (1961). The last cited authors, for example, used silver sulphide samples which were placed in quartz boats and burned in a

The Isotopic Analysis of Sulphur and Oxygen

49

stream of oxygen inside a fused quartz tube at a temperature of 1350 C, the temperature of the tube being maintained by an electric furnace. The oxygen used was purified by passage through ascarite (NaOH), anhydrone (MgCl04), and concentrated sulphuric acid. Hydrocarbons were elimimated by passing the oxygen through a zircon tube at 1350 C. After the removal of excess oxygen, together with water and carbon dioxide, the sulphur dioxide samples were collected in pyrex breakseal tubes which were sealed off and stored for mass spectrometric analysis. It is necessary to use high temperatures of combustion with this method in order to minimize the production of sulphur trioxide (S03)' This compound is produced according to the reaction

S02 + ! O2 ;= S03
At low temperatures the production of S03 is favoured. With increasing temperature the ratio of S02 to S03 increases (see, for example, Robinson and Kusakabe, 1975). Isotope exchange between S02 and S03 leads to the enrichment of 32S in S02; i.e. for the isotope exchange reaction
34S02 + 32S03 ;= 32S02 + 34S03

the isotope equilibrium constant is greater than unity (Tudge and Thode, 1950). Sakai and Yamamoto (1966) have calculated values for this equilibrium constant of 1.0046 and 1.0035 at temperatures of 1000 and 1200C respectively, so that both the yield of S03 and the isotope fractionation between S02 and S03 decrease with increasing temperature. The influence of S03 production on measured 1)34Svalues is easy to calculate. If combustion at 1200 C produces 95% S02 and 5% S03 which equilibrate isotopically then by the isotopic mass balance 0.95 1)34S(S02)+ 0.05 1)34S(S03)= 1)34S (original sulphur) and 1)34S(S03)- 1)34S(S02)= 3.5%0 so that 1)34S(S02)= 1)34S (original sulphur) - 0.175%0 This shift in the isotopic composition of S02 relative to the starting material will not introduce serious difficulties so long as all samples and standards are combusted identically, with constant yields of S03' Thode et al. (1961) minimized this isotopic shift by combusting at the highest practical temperature (13500q. Sakai and Yamamoto (1966), on the other hand, showed that combustion at lower temperatures, with a reduced partial pressure of oxygen, could achieve the same end.

50

Stable Isotopes

In their experiments, Sakai and Yamamoto (1966) combusted silver sulphide samples in streams of oxygen that contained different proportions of nitrogen. They showed that the production of S03 was reduced when the partial pressure of oxygen was reduced. The optimum operating conditions determined involved a 15:1 nitrogen to oxygen ratio at a combustion temperature of 1200 c. Under these conditions 98% yields of SOz were obtained and it was found that the lifetime of the combustion tube was considerably extended compared to the lifetime for combustions with pure oxygen. 3.1. 3.3 Oxidation of sulphides using a solid oxygen donor Vinogradov et al. (1956) produced sulphur dioxide for mass spectrometric analysis by reacting pyrite with lead oxide (PbO) at 850-900 C under vacuum. Subsequently Gavelin et al. (1960) reported the use of vanadium pentoxide (VzOs) as an oxidant. The VzOs method was refined by Ricke (1964) and is still in use in a number of laboratories. Grinenko (1962) first reported the use of cupric oxide (CuO) for oxidation of sulphides, as did Kaplan et al. (1970). Monster (1973) performed oxidations with both CuO and CUzO and recommended the use of CuO because he was able to obtain better yields at lower temperatures than with CUzO and because CuO is stable with respect to storage in atmospheric oxygen whereas CUzO is not. Fritz et al. (1974) also performed experiments with both oxidants and also concluded that CuO was to be preferred to CUzO. They noted that commercially available CUzO can contain up to 0.3% oil, which is used as a preservative against oxidation by atmospheric oxygen. Unless removed, this oil of course gives rise to the coproduction of carbon dioxide and water along with the sulphur dioxide. Robinson and Kusakabe (1975) showed that pure CUzO could be made from reagent grade CuO by heating it at 800C in the furnace tube of a vacuum line maintained at 10-4 bar. They advised against the preparation of large quantities of CUzO at one time because of the possibility of it oxidizing or absorbing gases with time. They also presented cogent reasons for the use of CUzO as oxidant rather than CuO, pointing out that when excess CuO is used for burning at high temperatures the P2 conditions may favour the production of S03 or metal sulphates. According to them, burnings with CUzO favour low P 2 conditions and S03 production can be ignored. The method employed by Robinson and Kusakabe (1975) is as follows. Sulphide samples are intimately mixed with an excess (twice the stoichiometric amount) of CUzO and ground to a grain size of about 150 f.Lmin a small agate mortar. The sample plus CUzO in a small ceramic boat that has been preheated to remove sulphur contamination, is placed inside a quartz tube,

The Isotopic Analysis of Sulphur and Oxygen

51

which is sealed at one end, together with a magnetic pusher, and the quartz tube is connected to a vacuum line and evacuated. Following evacuation and the attainment of a temperature of 800C in a small furnace surrounding the quartz tube the sample boat is pushed into the heated region. The chosen temperature of 800 C is low enough to enhance the life of the furnace windings and high enough to obtain 99-100% of the theoretical yield of SOz from silver sulphide (AgzS) within 10 minutes. Gas yields are measured using a Wallace and Tiernan absolute pressure gauge with an accuracy of :!: 0.4 % . Robinson and Kusakabe (1975) noted, as did Fritz et al. (1974), that difficulties are experienced burning minerals such as ZnS, PbS, etc., so that they are converted to AgzS. This procedure also removes silicates and other impurities from the minerals. Pyrite may be combusted directly but requires 15 minutes at 1000C. 3.1.3.4 Direct conversion of barium sulphate to sulphur dioxide Holt and Engelkemeir (1970) reported a method for the direct conversion of barium sulphate (BaS04) to SOz according to the net reaction BaS04 ---?> BaO + SOz + iOz According to their method the BaS04, covered with pulverized quartz powder in a fused quartz tube, was heated to the softening point of quartz (about 1400 q, the SOz produced being collected in a cold trap and the oxygen pumped away. This method offers the advantage that oxidized forms of sulphur need not be processed chemically to silver sulphide before preparation of sulphur dioxide for mass spectrometry. Bailey and Smith (1972) suggested a modified technique in which the vacuum system included an auxiliary furnace containing copper turnings heated to 800C. They argued that the method of Holt and Engelkemeir gave rise to the production of S03, because of the liberation of oxygen, with consequent displacement of the 034S of the sulphur dioxide from that of the original sample. The purpose of the copper heated to 800C was to reduce the partial pressure of oxygen in the system and to facilitate the conversion of any S03 to SOz. The temperature at which BaS04 is converted to SOz was reduced significantly in the method reported by Haur et al. (1973) by the addition to the reaction mixture of vanadium pentoxide (VzOs). These workers used a mixture of BaS04, VzOs and SiOz in the proportions by weight 1:3:2 and a reaction temperature of 1000-1050 C. Copper was used to bind free oxygen formed in the reaction by including copper wire in the reaction tube.

52

Stable Isotopes

Coleman and Moore (1978) usedcuprous oxide (CU20) rather than V205 and employed reaction mixtures of 15 mg BaS04 ground with 200 mg CU20 and 600 mg Si02 with a reaction temperature of 1120 dc. A separate furnace containing copper at 800 C was used for oxygen suppression and the
elimination of S03'
.

Halas and Wolacewicz (1981) and Halas et al. (1982) have described the use of sodium metaphosphate (NaP03) at 1000 C to produce S03 by the reaction BaS04 + NaP03 ~ NaBaP04 + S03 The S03 is converted to S02 on copper in a cooler part of the reaction chamber. The V205 method introduced by Haur et al. (1973) has been refined by Yanagisawa and Sakai (1983). These authors recommend the use of mixtures of BaS04, V205 and Si02 in the proportions 1:10:10 with a reaction temperature of 900C. As in other methods the partial pressure of oxygen is minimized by the use of copper wire placed a few centimetres above the mixture to assure negligible formation of S03' The high ratio of V205 + Si02 to BaS04 ensures that the oxygen isotopic composition of the S02 is completely controlled by these components and is independent of the original oxygen isotopic composition of the BaS04. Ueda and Krouse (1987) have extended this technique to other sulphates and sulphide minerals. 3.1. 3.5 The purification and storage of sulphur dioxide The preparation methods mentioned in Sections 3.1.3.2 and 3.1.3.4 above usually give rise to sulphur dioxide which is contaminated to a greater or lesser extent with water and carbon dioxide. The vacuum lines used for sample preparation must therefore include provision for the elimination of these contaminants before storage of sulphur dioxide prior to mass spectrometry. The normal purification procedure is to trap the H2O + S02 + CO2 mixture in a U-trap, using liquid nitrogen, and then to raise the temperature of the U-trap to -80C by replacing the liquid nitrogen with a dry ice-acetone slush so that H2O is retained while S02 and CO2 are vacuum transferred to a further trap at liquid nitrogen temperature. The temperature of this trap is then raised to -131C, the temperature of freezing n-pentane (Oano and Ishikawa, 1966), so that CO2 is pumped away to waste while S02 is retained. The arrangement of traps employed by Coleman and Moore (1978) appears to be particularly convenient. Finally, the temperature of the trap containing the S02 is raised further so that the S02 can be transferred to a storage vessel.

The Isotopic Analysis of Sulphur and Oxygen

53

Rather than using liquids with characteristic temperatures (dry iceacetone, freezing n-pentane) to control temperatures it may be more convenient to use a true variable temperature cold trap. Hayes et at. (1978) have described a trap of this type for the purification of carbon dioxide samples for mass spectrometric analysis, where the trap is cooled by liquid nitrogen but can also be heated electrically via windings around the body of the U. With this arrangement, any desired temperature or series of temperatures can be obtained by adjustment of the current to the heater windings. It is uncommon for sulphur dioxide samples to be taken directly from the preparation line to the inlet system of the mass spectrometer-normally the samples are stored for hours, days, or weeks prior to analysis. The simplest storage device is a pyrex tube which is fused at one end and has a stopcock at the other, together with a suitable fitting for attachment to the preparation line or mass spectrometer. The stopcocks can either be of the type that use vacuum grease or of the newer type with teflon barrels and elastomer 0ring seals. If long-term storage is required, there is the risk of leakage and the possible need for repurification of samples. In such cases it is more desirable to store samples in sealed vessels. Conventional breakseal tubes offer secure long-term storage but are difficult to make or expensive to purchase. The tube cracker device described by Desmarais and Hayes (1976) permits samples to be stored in simple lengths of 6 mm o.d. pyrex or quartz tubing fused at both ends, until required for mass spectrometry, when the tube is cracked under vacuum and the sample transferred to the spectrometer. 3.1. 3.6 The preparation of sutphur hexafluoride The use of sulphur hexafluoride (SF6) as a gas for sulphur isotope analyses was pioneered at the Carnegie Institution of Washington. Puchelt et at. (1971) have described the use of bromine trifluoride (BrF3) as a fluorinating agent. Hulston and Thode (1965), on the other hand, used elemental fluorine for this purpose, while Thode and Rees (1971) used bromine pentafluoride (BrFs) and Leskovsek et at. (1974) again used elemental fluorine. The SF6 method has now been used routinely at McMaster since 1970 with samples requiring measurement of 33S and 36S or very high precision with 34S and for samples that are too small for analysis as S02 (see, for example, Rees et at., 1978; Thode and Rees, 1979; McEwing et at., 1980, and references therein). Silver sulphide samples are reacted with a 20x molar excess of BrFs for 16 h at 300C in nickel reaction tubes. The SF6 produced is separated from unreacted BrFs and other reaction products by a series of vacuum distillations from traps cooled with dry ice-acetone baths to traps cooled with liquid nitrogen. Final SF6 clean-up is effected gas

54

Stable Isotopes

chromatographically using a coiled 6-ft column of ~ in o.d. copper tubing packed with a 5- A molecular sieve. Care must be taken at this stage because experience in this laboratory as well as the results of Moiseyev and Platzner (1976) indicate that isotope fractionation can occur in the gas chromatography unless complete recovery of the SF6 is effected.

3.1.4 The mass spectrometric analysis of sulphur The isotopic analysis of sulphur is almost invariably performed by gas source mass spectrometry, usually with sulphur dioxide and less commonly with sulphur hexafluoride as the analysis gas. The advent of quadrupole mass spectrometers with sample introduction from solution and ionization in an inductively coupled plasma may be important in the future. Similarly, some work is in progress, at the National Bureau of Standards of the United States, on the solid source mass spectrometry of very small samples of sulphur (Paulsen and Kelley, 1984). These latter two types of analysis are not considered in this article. The sulphur-containing gas is analysed in a conventional sector-type mass spectrometer equipped with two or more Faraday cups for ion detection. The sample of interest and a standard reference gas are admitted alternately so that isotopic differences can be determined under constant conditions and with consequent elimination of many instrumental biases (Nier, 1947; Nier et ai., 1947; McKinney et ai.., 1950; Wanless and Thode, 1953). It is increasingly common for isotopic analyses to be performed with commercially available mass spectrometers. Because of this, the technical details of mass spectrometer design and construction will not be considered here. Even with commercially supplied instruments it is necessary to consider the corrections to raw data that are required because of background peaks in the mass spectrum, peak tailing, and inlet leak offsets (Craig, 1957; Deines, 1970; Mook and Grootes, 1973; Beckinsale et ai., 1973). In the case of the mass spectrometric analysis of sulphur dioxide it is further necessary to take into account the interference in the SO+ or S02 + mass spectrum by ions containing 170 and 180. Thode et ai. (1949), Hulston and Shilton (1958), Ault and Kulp (1959), Hulston (1962), Holt and Engelkemeir (1970), Rees (1978), and Nord and Billstrom (1982) have developed corrections to be applied to raw mass spectrometric data which take into account this oxygen isotopic interference. Rees (1978) has also pointed out that sulphur dioxide, unlike other gases such as nitrogen or carbon dioxide, gives rise to a mass spectrometer memory effect caused by the slow flushing of unknown samples and the standard reference gas from the inlet line of the mass spectrometer. This memory effect causes isotopic differences between samples to be underestimated.

The Isotopic Analysis of Sulphur and Oxygen

55

There are not isotopic interferences to be considered in the case of the mass spectrometric analysis of SF6 because fluorine has only one stable isotope. Rees (1978) has presented details of the conversion of raw mass spectrometric data to 034Svalues for this type of analysis. 3.2 OXYGEN ISOTOPES*

3.2.1 Introduction
Oxygen isotopic analyses of environmental materials of the atmosphere, hydrosphere, lithosphere, and biosphere are carried out by the following steps: (a) collection of samples, (b) conversion of the oxygen in each sample to either CO2 or O2, and (c) mass spectrometric analysis of the CO2 or O2. 3.2.2 Sample collection The technique of collecting samples of sulphates and S02 in the atmosphere (Forrest and Newman, 1973) and sulphates in the hydrosphere, lithosphere, and biosphere for oxygen -18 analysis are generally identical to those for sulphur- 34 analysis. These techniques are described or referenced in Section 3.1.2. Samples of sulphates in precipitation water (rain or snow), surface water (streams, lakes, and oceans), or ground water (springs and wells) must be protected from bacterial alteration, either by refrigeration or by the addition of a bacteriocide (e.g. CuCI), if the samples are to be stored for more than a day or two before isotopic analysis. Environmental water samples that are to be analysed for 0180 of the H2O must also be protected from alteration by evaporation of diffusion through container walls, prior to analysis. Water vapour in the atmosphere is sampled by conducting a stream of air through a dry-ice cold trap for quantitative removal. The condensate is allowed to melt and the liquid water (5-10 ml) is sealed without delay in a small glass bottle or vial to avoid evaporation before analysis (Holt et al., 1978). 3.2.3 Sample preparation 3.2.3.1 Introduction Complete isotopic studies of mechanisms of sulphate formation may require the isotopic analysis not only of the sulphate but also of each of the reactants

* B.D. Holt.

56

Stable Isotopes

from which it was formed. For example, for the net reaction of catalysed aqueousair oxidation of SOz, SOz+ ! Oz + HzO ~ HzS04 (1)

the isotopic study may be based on the 0180 values of the SOz, Oz, HzO, and SOi-. To obtain the respective 0]80 values, each of these species is quantitatively converted to either COz or Oz for mass spectrometric analysis. The 1ZC160180/1ZC]60160 ratio, or the 160180/160160 ratio, of the sample is then compared to the corresponding ratio of a standard to obtain the 0180 values. Conversion of oxygen-bearing compounds to Oz is achieved by complete fluorination of each compound (Clayton and Meyeda, 1963). Conversion of many oxygen-bearing compounds to COz can be achieved by reaction with hot carbon. Each of the compounds in equation (1) can be converted to COz by reaction with hot graphite. Holt (1977) has described techniques by which a single graphite reduction furnace is used for this purpose.

3.2.3.2 Sulphate Using standard analytical procedures, a sulphate sample is quantitatively converted to pure BaS04 by the addition of BaClz to a hot aqueoussolution of the sulphate acidified by HCl. After a few hours of aging to enhance crystallite growth, the BaS04 is separated from the supernatant liquid, washed with deionized distilled water until free of chlorides, intimately mixed with pure graphite powder, and heated to convert the oxygen in the sulphate to a mixture of CO and COz by the reaction

(x + 2y) BaS04 + 4(x + y) C~ (x + 2y) BaS + 4x CO + 4y COz

(2)

where x = 0 and y = 0 correspond to the two extremes of 100% COz and 100% CO production respectively. The carbon reduction of BaS04 can be carried out, with conventional furnaces (Rafter, 1967), electrical resistance heating of the sample holder (Mizutani, 1971; Sakai and Krouse, 1971), and radio-frequency induction heating (Clayton and Epstein, 1958; Longinelli and Craig, 1967; Lloyd, 1968; Holt, 1977). Figure 3.1 shows the resistance heating reactor used by Shakur (1982). Most of the reactor was constructed of stainless steel. The base of the reactor has eight vertical electrical leads, each of which is connected to a stainless steel rectangular block. Four platinum boats, containing the BaSOcgraphite mixture, are clamped between pairs of block holders which are appropriately wired to heat the platinum by high-current, low-voltage power.

The Isotopic Analysis of Sulphur and Oxygen

57

~ TO DISCHARGE CONVERTER COPPER WIRES (40 A rating) Figure 3.1 Electrical resistance reactor for the conversion of BaS04 to CO2 for 81xO measurements (Shakur, 1982)

WATER

Figure 3.2 shows the graphite crucible assembly used by Holt (1977). By this procedure, a slurry of the BaS04 and graphite powder is filtered directly from the precipitation beaker into a porous graphite capsule. After washing and drying, the capsule containing the mixture of BaS04 and graphite powder is placed in the graphite crucible. The crucible is mounted in a water-cooled transparent quartz chamber and heated by a surrounding. induction-heating coil. By either system of heating, the BaSOcgraphite mixture is first vacuum heated at ~600 C for a few minutes to expel moisture and organic contaminants; then the temperature is raised to 1000-1150 DC.The CO2 and CO are expelled into a vacuum line where the CO2 is collected in a liquidnitrogen cold trap and the CO is subjected to a high-voltage (~1 kV) electrical discharge between two platinum electrodes. The discharge causes the CO to disproportionate into carbon and CO2, The newly formed CO2 is continuously removed from the gas phase by condensation in the liquidnitrogen cold trap. At the end of the heating period (~15-20 min), essentially all of the sample oxygen is collected in the cold trap as CO2; BaS remains in the sample holder (platinum boat or graphite capsule, depending on the method used). The CO2 is cryogenically transferred through a cold trap (temperature of melting pentane ~-132 DC) to a gas sample bulb (cooled to -196 DC) for subsequent transfer to a mass spectrometer.

58

Stable Isotopes

116mmr

I 25mm
6mm

-r--~'
Ffj

wj/

VENT HOLES IN LID CAPSULE CRUCIBLE

I)-GRAPHITE

25mm

--- GRAPHITE

I~
6mm

6.4 mm PLATINUM

TUBE

-"$

10/18

QUARTZ

JOINT

Figure 3.2 Radio-frequency heated graphite crucible for the conversion of BaS04 to CO2 (Holt, 1977)

3.2.3.3 Water By an older method for the determination of the 0180 of water, a measured quantity of the water is equilibrated with a measured quantity of CO2 known isotopic composition (Epstein and Mayeda, 1953). The equilibration is usually carried out by frequent agitation of the water and CO2 for about three days at a constant temperature; with continuous agitation, the equilibration time can be reduced to less than one day. If the quantity of H2O is much greater than that of the CO2, then comparison of the isotopic compositions of the equilibrated CO2 is equivalent to comparing the H2O samples. If the H2O sample is very small, the isotope separation factor for the equilibration temperature and a comparison of the 180/160 ratios in the CO2 before and after the equilibration are necessary to calculate the 0180 of the water. Majzoub (1966) developed a method for converting all of the oxygen in microquantities (~4 /-lL) of water to CO2 for mass spectrometric analysis. The method was modified by Holt (1977) for adaptation to the same graphite-reduction apparatus that was used for BaS04 samples. The procedure consists of hot-graphite reduction of the water to CO and H2, separation of the H2 from the CO by diffusion through a palladium membrane, and conversion of the CO to CO2 and carbon as described above. The advantages of the carbon-reduction method are that only very small samples are required and that the long equilibration period is avoided.

The Isotopic Analysis of Sulphur and Oxygen

59

Similar advantages can be realized by reaction of H2O with guanadine (Boyer et at., 1961). 3.2.3.4 502 and O2 Holt (1977) used a Toepler pump to circulate samples (~2 mL, STP) of either O2 or S02 through a closed loop for exposure to the inductively heated (~1200 c) graphite crucible. The sample gas was quantitatively converted to CO and subsequently to CO2 as described above. The procedure for S02 or O2 is not applicable to oxygen in air because the N2 in air combines with the O2 to form nitrogen oxides in the high-voltage discharge. Instead, the graphite crucible (Figure 3.2) is covered with platinum gauze (catalyst) and heated to not more than ~600 C (Horibe et at., 1973). Under these conditions, the air oxygen is converted directly to COb with negligible formation of CO. 3.2.4 The mass spectrometric analysis of oxygen The CO2 (or O2, if the sample is fluorinated) which is obtained from the carbon reduction of BaS04, H2O, S02, or O2 is analysed with a highprecision, double-beam, isotope-ratio, gas mass spectrometer as described in Section 3.1.4. REFERENCES
ASTM D-271 (1971). Sampling and analyses of coal and coke, Sections 21 to 24. Auger, V., and Gabillon, J. (1911). Nouveau procede de dosage de l'acide sulfurique et des sulfates. C.R. Acad. Sci., Paris., 424. Ault, W.U., and Kulp, J.L. (1959). Isotope geochemistry of sulphur. Geochim. Cosmochim. Acta, 16, 201-35. Bailey, S.A., and Smith, J.W. (1972). Improved method for the preparation of sulfur dioxide from barium sulfate for isotope ratio studies. Anal. Chem., 44, 1542-3. Beckinsale, R.D., Freeman, N.J., Jackson, M.C., Powell, R.E., and Young, W.A.P. (1973). A 30 cm radius 90 sector double collecting mass spectrometer with a capacitor integrating detector for high precision isotopic analysis of carbon dioxide. Int. J. Mass Spectrom. Ion Phys., 12,299-308. Berner, R.A. (1974). Iron sulfides in pleistocene deep Black Sea sediments and their paleooceanographic significance. In: Degens, E.T., and Ross, D.A. (Eds.) The Black Sea-Geology, Chemistry and Biology, Memoir 20, American Association of Petroleum Geologists, Tulsa, Oklahoma, pp. 524-31. Boyer, P.D., Graves, D.J., Suelter, C.H., and Dempsey, M.E. (1961). Simple procedure for conversion of oxygen of orthophosphate or water to carbon dioxide for oxygen-18 determination. Anal. Chem., 33,1906-9. Case, J.W., and Krouse, H.R. (1980). Variations in sulphur content and stable sulphur isotope composition of vegetation near a S02 source at Fox Creek, Alberta, Canada. Oecologia, 44, 248-57.

60

Stable Isotopes

Clayton,R,N., andEpstein, S, (1958), Therelationship between IHOj160 incoexisting


quartz, carbonate, and iron oxides from various geological deposits. 1. Ceol., 66, 352-71. Clayton, R.N., and Meyeda, T.K. (1963). The use of bromine tetrafluoride in the extraction of oxygen from oxides and silicates for isotopic analyses. Ceochim. Cosmochim. Acta, 27, 43-52. Coleman, M.L., and Moore, M.P. (1978). Direct reduction of sulfates to sulfur dioxide for isotopic analysis. Anal. Chem., 50, 1594-5. Craig, H. (1957). Isotopic standards for carbon and oxygen and correction factors for mass spectrometric analysis of carbon dioxide. Ceochim. Cosmochim. Acta, 12, 133-49. Deines, P. (1970). Mass spectrometer correction factors for the determination of small isotope composition variations of carbon and oxygen. Int. 1. Mass Spectrom. Ion Phys., 4, 283-95. Desmarais, D.J., and Hayes, J.M. (1976). Tube cracker for opening glass-sealed ampoules under vacuum. Anal. Chem., 48, 1651-3. Epstein, S., and Meyeda, T. (1953). Variation of OIHOcontent of waters from natural sources. Ceochim. Cosmochim. Acta, 4, 213-24. Forrest, J., and Newman, L. (1973). Sampling and analysis of atmospheric sulphur compounds for isotope ratio studies. Atmos. Environ., 7, 561-73. Forrest, J., and Newman, L. (1977a). Further studies on the oxidation of sulfur dioxide in coal-fired power plant plumes. Atmos. Environ., 11, 465-74. Forrest, J., and Newman, L. (1977b). Oxidation of sulfur dioxide in the Sudbury Smelter Plume. Atmos. Environ.., 11,517-20. Forrest, J., and Newman, L. (1977c). Silver-110 microgram sulfate analysis for the short time resolution of ambient levels of sulfur aerosol. Anal. Chem., 49, 1579-84. Forrest, J., Klein, J.H., and Newman, L. (1973). Sulphur isotope ratios of some power plant flue gases: a method for collecting sulphur oxide. J. Appl. Biotechnol., 23, 855-63. Fritz, P., Drimmie, R.J., and Nowicki, V.K. (1974). Preparation of sulfur dioxide for mass spectrometer analyses by combustion of sulfides with copper oxide. Anal. Chem., 46, 164-6. Gavelin, S., Parwel, A., and Ryhage, R. (1960). Sulfur isotope fractionation in sulfide mineralisation. Econ. Ceol., 55, 510-30. Gonfiantini, R. (1978). Standards for stable isotope measurements in natural compounds. Nature, 271, 534-6. Grinenko, V.A. (1962). The preparation of sulphur dioxide for isotopic analysis. Zh. Neorgan. Khim., 7, 2478-83. Halas, S., and Wolacewicz, W.P. (1981). Direct extraction of sulfur dioxide from sulfates for isotopic analyses. Anal. Chem., 53, 686-9. Halas, S., Shakur, A., and Krouse, H.R. (1982). A modified method of S02 extraction from sulphates for isotopic analysis using NaPO,. lsotopenpraxis, 18, 433-6. Haur, A., Hladikova, J., and Smejkal, V. (1973). Procedure of direct conversion of sulfates into S02 for mass spectrometric analysis of sulfur. Isotopenpraxis, 9, 329-31. Hayes, J.M., Desmarais, D.J., Peterson, D.W., Schoeller, D.A., and Taylor, S.P. (1978). High precision stable isotope ratios from microgram samples. Adv. Mass Spectrom., 7A, 475-80. Hicks, J.E., Fleenor, J.E., and Smith, H.R. (1974). The rapid determination of sulfur in coal. Analytica Chimica Acta, 68, 480-3.

The Isotopic Analysis of Sulphur and Oxygen

61

Holt, B.D. (1977). Preparation of carbon dioxide from sulphates, sulphur dioxide, air, and water for determination of oxygen isotope ratio. Anal. Chern., 49, 1664-7. Holt, B.D., and Engelkemeir, AG. (1970). Thermal decomposition of barium sulfate to sulfur dioxide for mass spectrometric analysis. Anal. Chern., 42, 1451-3. Holt, B.D., Engelkemeir, AG., and Venters, A. (1972). Variations of sulfur isotope ratios in samples of water and air near Chicago. Environ. Sci. Tech., 6, 338-41. Holt, B.D., Cunningham, P.T., and Engelkemeir, A.G. (1978). Application of oxygen-18 analysis to the study of atmospheric sulphate formation. In: Robinson, B.W. (Ed.) Stable Isotopes in the Earth Sciences, New Zealand DSIR Bulletin 220, pp. 105-9. Horibe, Y., Shigehara, K., and Takakuwa, Y. (1973). Isotope separation factors of carbon dioxide-water system and isotopic composition of atmospheric oxygen. 1. Geophys. Res., 78, 2625-9. Hulston, J.R. (1962). New Zealand sulfur standards in relation to meteoritic sulfur. In: Jensen, M.L. (Ed.) Biogeochemistry of Sulfur Isotopes, Proceedings of NSF Symposium at Yale University, 12-14 April 1962, pp. 36-41. Hulston, J.R., and Shilton, B.W. (1958). Sulphur isotopic variations in nature. Part 4-measurement of sulphur isotopic ratio by mass spectrometry. N. S. J. Sci., 1, 91-102. Hulston, J.R., and Thode, H.G. (1965). Variations in S33, S34, and S36contents of meteorites and their relation to chemical nuclear effects. J. Geophys. Res., 70, 3475-84. Johnson, C.M., and Nishita, H. (1952). Microestimation of sulfur. Anal. Chern., 24, 736-42. Kaplan, LR., Emery, K.O., and Rittenberg, S.c. (1963). The distribution and isotopic abundance of sulphur in recent marine sediments off Southern California. Geochirn. Cosrnochirn. Acta, 27, 297-331. Kaplan, LR., Smith, J.W., and Ruth, E. (1970). Carbon and sulphur concentration and isotopic composition in Apollo 11 lunar samples. Geochirn. Cosrnochirn. Acta, Supp!. 1 (2), 1317-29. Kelly, W.R., Paulsen, P.J., and Garner, E.L. (1982). Determination of sulfur in coal SRM's 2682, 2683, 2684, 2685, and 1635 by thermal ionization mass spectrometry. Report of analysis, U.S. Department of Commerce, National Bureau of Standards, Washington, 20 September 1982. Kiba, T., Takagi, T., Yoshimura, Y., and Kishi, L (1955). Tin (H)-strong phosphoric acid. A new reagent for the determination of sulfate by reduction of hydrogen sulfide. Bull. Chern. Soc. Japan, 28, 641-4. Krouse, H.R., Ritchie, R.G .G., and Roche, R.S. (1987). Sulphur isotope composition of HzS evolved during nonisothermal pyrolysis of sulphur containing materials. 1. Anal. Appl. Pyrol., 12, 19-29. Leskovsek, H., Marsel, J., and Kosta, L. (1974). The application of SF6 in the geochemical research. Isotopenpraxis, 10, 375-7. Lloyd, R.M. (1968). Oxygen isotope behavior in the sulphate water system. J. Geophys. Res., 73, 6099-110. Longinelli, A, and Craig, H. (1967). Oxygen-18 variations in sulphate ions in sea water and saline lakes. Science, 156, 56-9. Lowe, L.E., Sasaki, A., and Krouse, H.R. (1971). Variations of sulfur- 34: sulfur- 32 ratios in soil fractions in Western Canada. Can. J. Soil Sci., 51, 129-31. McEwing, C.E., Rees, C.E., and Thode, H.G. (1980). Sulphur isotope effects in the dissociation and evaporation of troilite: a possible mechanism for the :14S enrichment of lunar soils. Geochirn. Cosrnochirn. Acta, 44, 565-71.

62

Stable Isotopes

McKinney, C.R., McCrea, J.M., Epstein, S., Allen, H.A., and Urey, H.C. (1950). Improvements in mass spectrometers for the measurement of small differences in isotope abundance ratios. Rev. Sci. Inst., 21, 724-30. Majzoub, M. (1966). Une methode d'analyse isotopique de l'oxygene sur des microquantites d'eau determination des coefficients de partage a l'equilibre de I'oxygene 18 entre H2O et CO2, D20 et CO2, 1. Chern. Phys., 63, 563-8. Mekhtiyeva, V.L., Gavrilov, E.Y., and Pankina, R.G. (1976). Sulfur isotopic composition in land plants. Geokhimiya, 11, 1755-9. Mizutani, Y. (1971). An improvement in the carbon-reduction method for the oxygen isotopic analysis of sulphates. Geochem. 1., 5, 69-77. Mizutani, Y., and Rafter, T.A. (1969). Isotopic composition of sulphate in rainwater, Gracefield, New Zealand. N.z. J. Sci., 12, 69-80. Moiseyev, N., and Platzner, 1. (1976). Isotope effect in gas-solid chromatography of SF6. 1. Chromatog. Sci., 14, 143-8. Monster, J. (1973). Preparation of S02 samples for mass-spectrometry using copper oxide as an oxidant. McMaster University Internal Report 97. Monster, J. (1978a). Comparison of the LAH and the 6N HN03 method for separating pyrite and organic sulphur in sediments. McMaster University Internal Report 134. Monster, J. (1978b). The isotopic composition of coexisting pyrite and organic sulphur in sediments. McMaster University Internal Report 139. Mook, W.G., and Grootes, P.M. (1973). The measuring procedure and corrections for the high precision mass-spectrometric analysis of isotopic abundance ratios, especially referring to carbon, oxygen, and nitrogen. Int. 1. Mass Spec. Ion Physics, 12, 273-98. Newman, L., Forrest, J., and Manowitz, B. (1975). The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulfur dioxide in the plume from an oil-fired power plant; and The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulfur dioxide in the plume from a coal-fired power plant. Atmos. Environ., 9, 959-68 and 969-77. Nier, A.O. (1947). A mass spectrometer for isotope and gas analysis. Rev. Sci. Inst., 18, 398-411. Nier, A.O., Ney, E.P., and Inghram, M.G. (1947). A null method for the comparison of two ion currents in a mass spectrometer. Rev. Sci. Inst., 18, 294-7. Nord, A.G., and Billstrom, K. (1982). A system for stable isotope analyses of geological samples. Geologisha Foreningens i Stockholm Forhandlingar, 104 (2), 113-20. Stockholm ISSN 0016-786X. Nriagu, J.O. (1975). Sulfur isotopic variations in relation to sulfur pollution of Lake Erie. In: Isotope Ratios as Pollutant Source and Behavior Indicators, IAEA, pp. 77-93. Nriagu, J.O., and Coker, R.D. (1976). Emission of sulfur from Lake Ontario sediments. Limnol. Oceanog., 21, 485-9. Nriagu, J .0., and Coker, R.D. (1978). Isotopic composition of sulfur in precipitation within the Great Lakes Basin. Tellus, 30, 365-75. Oano, S., and Ishikawa, H. (1966). Sulfur isotopic fractionation between sulfur and sulfuric acid in the hydrothermal solution of sulfur dioxide. Geochem. J., 1, 45-50. Paulsen, P.J., and Kelley, W.R. (1984). Determination of sulfur as arsenic monosulfide ion by isotope dilution thermal ionization mass spectrometry. Anal. Chern., 56, 708-13. Puchelt, H., Sabels, B.R., and Hoering, T.e. (1971). Preparation of sulfur

The Isotopic Analysis of Sulphur and Oxygen

63

hexafluoride for isotope geochemical analysis. Geochim. Cosmochim. Acta, 35, 625-8. Rafter, T.A. (1957). Sulphur isotopic variations in nature. Part I-the preparation of sulphur dioxide for mass spectrometer examination. N.Z. J. Sci. Technol., 838, 849-57. Rafter, T.A. (1967). Oxygen isotopic composition of sulphates: part 1. N.z. J. Sci., 10, 493-510. Rees, C.E. (1978). Sulphur isotope measurements using S02 and SF6. Geochim. Cosmochim. Acta, 42, 383-9. Rees, C.E., Jenkins, W.J., and Monster, J. (1978). The sulphur isotopic composition of ocean water sulphate. Geochim. Cosmochim. Acta, 42, 377-81. Ricke, W. (1964). Praparation von Schwefeldioxid zur massenspektrometrischen Bestimmung des Schwefel-Isotopen- Verhaltnisses C4SP2S)in natiirlichen Schwefelverbindungen. Zeitschrift fur analytische Chemie, 199, 401-13. Robinson, B.S., and Kusakabe, M. (1975). Quantitative preparation of sulfur dioxide, for 34Sp2S analyses, from sulfides by combustion with cuprous oxide. Anal. Chem., 47, 1179-81. Sakai, H., and Yamamoto, M. (1966). Fractionation of sulfur isotopes in the preparation of sulfur dioxide. An improved technique for the precision analysis of stable sulfur isotopes. Geochem. J., 1, 35-42. Sasaki, A., Arilawa, Y. and Folinsbee, R.E. (1979). Kiba reagent method of sulfur extraction applied to isotopic work. Bull. Geol. Surv. Japan, 30, 241-245. Shakur, M.A. (1982). (\34Sand (\180 variations in terrestrial sulphates. A doctoral thesis, submitted to the Department of Physics, University of Calgary, Calgary, Alberta, Canada. Smith, J.W., Young, N.B., and Lawlor, D.L. (1964). Direct determination of sulfur forms in Green River oil shale. Anal. Chem., 36, 618-22. Smith, P.V. (1952). Preliminary note on origin of petroleum. Am. Assoc. Petrol. Geologists Bull., 36, 411-13. Smith, P.B. (1954). Studies in origin of petroleum: occurrence of hydrocarbons in recent sediments. Am. Assoc. Petrol. Geologists Bull., 38, 377-404. Thode, H.G., and Rees, C.E. (1971). Measurement of sulphur concentrations and the isotope ratios 33Sp2S, 34Sp2S and 36Sp2Sin Apollo 12 samples. Earth Planet. Sci. Left., 12, 434-8. Thode, H.G., and Rees, C.E. (1979). Sulphur isotopes in lunar and meteorite samples. Proceedings of Tenth Lunar Planetary Science Conference, pp. 1619-36. Thode, H.G., MacNamara, J., and Collins, C.B. (1949). Natural variations in the isotopic content of sulphur and their significance. Can. J. Res., 827, 361-73. Thode, H.G., Harrison, A.G., and Monster, J. (1960). Sulphur isotope fractionation in early diagenesis of recent sediments of Northeast Venezuela. Am. Assoc. Petrol. Geologists Bull., 44, 1809-17. Thode, H.G., Monster, J., and Dunford, H.B. (1961). Sulphur isotope geochemistry. Geochim. Cosmochim. Acta, 25, 159-74. Tudge, A.P., and Thode, H.G. (1950). Thermodynamic properties of isotopic compounds of sulphur. Can. J. Res., 828, 567-78. Ueda, A., and Krouse, H.R. (1987). Direct conversion of sulphide and sulphate minerals to S02 for isotope analyses. Geochem. J., 20, 209-12. Ueda, A., and Sakai, H. (1983). Simultaneous determinations of the concentration and isotope ratio of sulfate- and sulfide-sulphur and carbonate-carbon in geological samples. Geochem. J., 17, 185-296. Vinogradov, A.P., Chupakhin, M.S., Grinenko, V.A., and Trofimov, A.V. (1956).

64

Stable Isotopes

Isotopic composition of sulfur in relation to the problem of the age of pyrites of sedimentary genesis. Geochimiya, 1956, 96--105. Vinogradov, A.P., Grinenko, V.A., and Ustinov, V.I. (1962). Isotopic composition of sulfur compounds in the Black Sea. Geokhimiya, 10, 974-97. Wanless, R.K., and Thode, H.G. (1953). A new mass spectrometer for high precision isotope ratio determinations. 1. Sci. Inst., 30, 395-8. Westgate, L.M., and Anderson, T.F. (1982). Extraction of various forms of sulfur from coal and shale for stable sulfur isotope analysis. Anal. Chem., 54, 2136-9. Yanagisawa, F., and Sakai, H. (1983). Preparation of SOz for sulfur isotope ratio measurements by the thermal decomposition of BaSOcVzOs-SiOz mixtures. Anal. Chem., 55, 985-7.

CHAPTER 4

Lithospheric Sources of Sulphur


H. NIELSEN, J. PILOT, L.N. GRINENKO, V.A. GRINENKO, A. YU. LEIN, J.W. SMITH, AND R.G. PANKINA

4.1

SULPHUR IN THE COSMOS AND THE CRUST-MANTLE SYSTEM*

All the sulphur contributed to our environment by natural processes and human activity originates from reservoirs in the earth's crust. In many cases, the 034S can serve as a 'fingerprint? to identify sources of sulphur and trace its fate in the environment. On a cosmic scale, sulphur belongs to the 10 most abundant elements. The rank order to atomic abundances in our solar system is: H, He, 0, C, N, Ne, Mg, Si, S, Fe, AI, Ca, Ni, Na, Cr, etc. (Goles, 1969). It is widely believed that our planets and the planetary parent body(ies) of meteorites originate from a common reservoir of cosmic matter of homogeneous chemical and isotopic composition. Thus, the study of meteorites provides information about the mean sulphur concentration and 034S value of the earth as a whole. Typical S concentrations for iron and stone meteorites lie between 3 and 50 g kg-l (Keil, 1969), and therefore an estimate of 10 g kg-1 for the mean sulphur content of the earth seems realistic. The mean 034S value seems easier to calculate. Troilite (FeS) of iron meteorites is isotopically uniform and samples from Canon Diablo and Shikote Alin are used as the 'primary' standard defining the zero point of the 034Sscale (see Chapter 1). Thus, we can postulate that the earth as a whole has a mean 034Svalue of 0%0. With the differentiation of the earth into core, mantle, and crust, the bulk of the sulphur accumulated in the inner shells and the upper crust became systematically depleted in this element. This arose because sulphur was predominantly bound as iron sulphide droplets which had a higher density and proved to be immiscible in the silicate melt. A preliminary estimate of the S concentration in this depleted crust can be obtained from analyses of present-day igneous rocks which are also depleted in sulphur. This gives a range from about 300 mg kg-l (mean value for granites) to 800 mg kg-l (mean value for mafic rocks) (see Table 16- B-2a in Nielsen, 1978, and Schneider, 1978). * H. Nielsen.
Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko <9 1991 SCOPE, Published by John Wiley & Sons Ltd

66

Stable Isotopes

Because of the high temperature at which this differentiation took place, isotopic fractionation should have been very small and mantle and lower crust sulphur should have preserved the primaeval 834S value of zero. In reality, however, samples of mantle rocks (ultramafic inclusions in kimberlite pipes, materials from Ivrea structures, etc.) are slightly enriched in the heavy isotopes. This is believed to result from crustal contamination. The sulphur also brought up by present-day volcanic emissions is non-juvenile to a large extent. The 834S values close to zero have been found in uncontaminated ultramafic rocks from Czechoslovakia and, on average, the 834S value for continental basalts is + 1.4%0(SCOPE 19, 1983). Immediately after the solidification of the primitive crust, volcanic eruptions and fumarolic activity began to convey immense quantities of H2O, CO2, S species, and other 'volatiles' to the surface. This in turn led to extensive weathering and inaugurated the 'sedimentary history' of our planet. Under the varying redox and pH conditions of the newly established 'exogeneous cycle', sulphur acquired different valence states, and formed gaseous compounds, sulphates, sulphides, and a manifold of S organic compounds. This led to an extremely inhomogeneous distribution in the sedimentary envelope. The sulphur concentrations vary from a few milligrams per kilogram in freshwater limestones to about 200 g kg-l in anhydrite strata and to even more than 500 g kg-l in local pyrite accumulations. From the isotopic viewpoint, the most exciting facet was the emergence of larger isotope fractionation during dissimilatory bacterial S042- reduction. As the biological sulphur cycle became established, large isotopic variations were promoted in sedimentary rocks. This large variability renders it extremely difficult to determine the 'true' mean values for S content and 834S of crustal sulphur (Figure 4.1). All the hitherto published values are speculative and are a matter of controversial debate. As an example, Holser and Kaplan (1966) estimate the mean S content and 834Sof the sediments as +4.9 ::I:: 1.3 g kg-1 and +7.1 ::I:: 4.5%0, respectively. Another compilation by Ronov, Wedepohl, and others cited in Nielsen (1978) suggests concentrations between 3.6 and 5.6 g kg-l and 834S values between -9.0 and +0.7%0. The extreme values arise from different estimates of the total mass of evaporite sulphate. The slightly positive 834Svalues in crustal sulphur are not only predicted by sophisticated budget evaluations (cf. SCOPE 19, 1983) but can easily be deduced by visual inspection of S isotope histograms based on representative and sufficiently numerous samples (Figure 4.1).
Figure 4.1 The 834S histograms of samples investigated at the G6ttingen Stable Isotope Laboratory during the last two decades. Reduced sulphur species are drawn above and the oxidized species below the baseline. Numbers in brackets are the number of samples used in each histogram. Values lower than -40%0 or larger than +40%0 are shown at the ends of the scale

Lithospheric Sources of Sulphur


-40

67
0
~

-30

-20

-10

+10 ....

-120

+30 .

+40 .

All reduced

sur ohur

(nn)

,All oxldlzed sulphur


I I I I I I r I I I I

(5126)

"Total sulphur"

(2057>

~rmjhi1T11T1TI
HzS and elemental su'_ph',,-

~-

IIIIIIIIIIIIIIII-m-mn-n-n m..m
I r

D~
DIssolved sulphate Pyrite

(441)

<1791>

CalcIum. magnesium sulphat..


Metastable sedimentary Fe sur ph ;d.

(2076)

(134)

~~
Surphat. In silicate
(147)

Pb-Cu- Zn su'-ph;cte s

(3347)

Baryte.

coelestlte (1053)

Organic sulphur Phosphorite Sulphur

(10') (39)

, -40

T -3U

-20

-10 8"5, %0

T +10

T +20

+30

T +40

68

Stable Isotopes

4.Z SULPHUR AND OXYGENISOTOPECOMPOSITION OF MARINE EVAPORITE AND BARITESTRATA*


4.2.1 Introduction Evaporites are relicts of former oceans or non-marine water bodies which usually contain abundant anhydrite, gypsum, and a variety of other minerals such as polyhalite, glaserite, kieserite, kainite, langbeinite, leonite, halite, sylvite, and carnallite. Gypsum dust is added to the environment during mining of evaporite deposits and manufacturing of building products. In addition, acidic pollutants can enhance dissolution of shallow natural evaporites. Sulphur and oxygen isotope data can serve to study these activities. Evaporite strata tend to form in basins on the border of continents when prevailing conditions favour evaporation. Basins may also form during early stages of rifting where ocean water of a different provenance spills into a sub-sea-level graben (Burke, 1975). Whereas present-day oceanic sulphate has a (')34S value near + 21%0,marine evaporite sulphate varies between + 10 and + 35%0.It is generally believed that these reflect actual variations in the isotopic composition of S042- in the ancient oceans. Many scientists have studied the (')34S-agecurve as constructed from evaporite data (Thode and Monster, 1964; Nielsen, 1965, 1972, 1973; Holser and Kaplan, 1966; Yeremenko and Pankina, 1971; Pilot et at., 1972; Claypool et at., 1980). In such studies, care must be taken to exclude samples influenced by isotopically selective processes such as sulphate reduction and contamination by SOi- from continental waters. The (')34S-agecurve throughout the Phanerozoic has been studied so well that it can be used to differentiate gypsum-bearing units for subsurface mapping (van Everdingen et at., 1982). In contrast, the secular isotopic variations are not so well charted in the Precambrian due to the scarcity of evaporite deposits. Many attempts have been made to understand the mechanisms for changing the (')34S values of marine SOi- (Nielsen, 1965; Holser and Kaplan, 1966; Rees, 1970; Holland, 1973; Schidlowski et at., 1977; Claypool et at., 1980; Hoefs, 1981). Some models include the concept that the sulphur cycle is linked with the oxygen and/or carbon cycles (Schidlowski and Junge, 1981). With fewer data, a (')ISO-age curve has also been constructed for the Phanerozoic evaporites (Solomon et at., 1971; Sakai, 1972; Pilot et at., 1972; Claypool et at., 1980; Viezer et at., 1980).

* J. Pilot.

Lithospheric Sources of Sulphur

69

4.2.2 Compilation of 8348 and 8180 data for marine evaporites 4.2.2.1 General remarks on the data Tables 4.1 and 4.2 summarize marine evaporite 034S and 0180 data from the following references: Nielsen and Ricke, 1964; Thode and Monster, 1965; Holser and Kaplan, 1966; Muller et al., 1966; RosIer et al., 1968; Yeremenko and Pankina, 1971; Perry et al., 1971; Sakai, H., 1972; Pilot et al., 1972; Klaus and Pak, 1974; Pak, 1974; ~ambow and Nielsen, 1976, cited in Claypool et al., 1980; Hinze and Nielsen, 1976, cited in Claypool et al., 1980, p. 212; Claypool et al., 1980; Cortecci et al., 1981; Shakur, 1982. Various criteria must be used to select the 034Sand 0180 values that best reflect the isotopic composition of contemporaneous oceanic SO/-. Obviously the age and the horizon of the samples must be well documented. Isotope fractionation prior to and during precipitation must also be considered. Samples containing dispersed sulphides were probably influenced by bacterial reduction and the 034S value of unreacted SOi- may have increased substantially prior to precipitation. Because of this phenomenon, Thode and Monster (1965) preferred to use minimum 034Svalues for a given age. However, there are also processes that deplete SO/- in 34S.Precipitated CaS04 is 1.65 :!: 0.12 and 3.6 :!: 0.9%0enriched in 34Sand 180, respectively, compared to SOi- (Thode and Monster, 1965; Lloyd, 1967; Holser et al., 1979). In addition, secondary sulphate may form by oxidation of 34S-depleted sulphides. Further, continental weathering and deep crustal sources may contribute. Hence, many authors prefer to use the mean value of a data set. The suitability of the selection criteria is confirmed if the 034Sdata from different basins and/or different regions of the same age (comparable stratigraphic horizons) coincide. Table 4.1 presents data from samples for which horizons are well known. In most cases, the horizons have the authors' designation and not the international nomenclature. The time scale of Odin and Kennedy (1982) is used. The 0180 data are normalized to seawater SO/- taken as +8.6%0, in accordance with Claypool et al. (1980). It is noted that many authors prefer a value closer to +9.5%0. For Table 4.2, the data measured at UCLA are corrected by +0.5%0, those from Cortecci et al. (1981) and Shakur (1982) by -0.9%0 and from Pilot et at. (1972) and Pilot and Harzer (1969) by -0.5%0. Data are arbitrarily classified by region: North America, Middle America, Asia, Europe, Australia, and Africa. Samples from South America were only available from northern parts and are included in Middle America.

-.J

Table 4.1 A 1)J4Sdata summary for marine evaporites: (number), minimum average, maximum North America Tertiary Pliocene Miocene Oligocene Eocene Palaeocene (21), 18.4 20.6, 24.3 (10), 19.0 19.7, 20.3 (10), 10.9 17.8, 25.5 (2), 17.1 21.3, 25.5 (8), 10.9 16.9, 19.9 (2), 19.0 19.3, 19.6
-

Middle America

Europe

Africa

Asia

Australia

Arctic

Total

(I) 20.5 (9), 20.9 21.8, 23.9

(2), 17.4 17.9, 18.3

(3), 17.4 18.8, 20.5 (9), 20.9 21.8, 23.9 (2), 17.1 21.3, 25.5 (30), 10.9 19.5, 24.3 (14), 19.0 19.7, 20.4

(1) 17.8 (2), 19.5 19.9, 20.4


---

Total

(32), 18.4 20.3, 24.3

(11), 19.0 21.3, 23.9

(3), 17.4 17.9, 18.3

(2), 19.5 19.9, 20.4

(58), 10.9 19.9, 25.5

Cretaceous Senonian Turonian Cenomanian Albian Aptian Neocamian (4), 15.8 16.1, 16.5 (24), 12.8 14.9, 18.1 (17), 13.4 16.0, 21.8

(1) 22.7 (2), 16.8 17.7, 18.5 (20), 12.8 14.7, 18.1 (17), 13.4 16.0, 21.8 (7), 13.9 15.1, 18.6

(1) 22.7 (2), 16.8 17.7, 18.5 (27), 12.8 14.8, 18.6 (17), 13.4 16.0, 21.8 (4), 15.8 16.1, 16.5 (51), 12.8 15.6, 21.8

Total

(10), 13.9 16.4, 18.6

-..J ......

-..) N

Table 4.1 Continued North America Jurassic Kimmeridgian Oxfordian Upper Jurassic Undefined Jurassic Lower Jurassic (7), 15.4 16.6, 17.5 (1), 16.0 (11), 15.1 16.0, 17.0 (1) 15.6 Middle America Europe Africa Asia Australia Arctic Total

(1) 16.7

(1) 16.7 (7), 15.4 16.6, 17.5 (35), 15.4 17.1, 21.0 (36), 12.9 15.5, 17.1 (1) 8.1

(34), 15.4 17.1, 21.0 (18), 12.9 15.1, 17.1 (1) 8.1
-----

(1) 18.6

(5), 13.6 15.1, 16.7

Total

(19), 15.1 16.2, 17.5

(1) 15.6

(53), 8.1 16.3, 21.0

(2), 16.7 17.7, 18.6

(5), 13.6 15.1, 16.7

(80), 8.1 16.1, 21.0

Tria.5sic Upper (Keuper) Middle (Muschelkalk) Lower (Buntsandstein) Upper Buntsandstein (Rot) Upper Buntsandstein II Middle Buntsandstein Lower Buntsandstein

(9), 13.4 14.2, 15.7

(55), 10.9 15.3, 18.3 (23), 17.6 19.6, 21.8

(9), 15.7 16.7, 17.8

(73), 15.3, (23), 19.6,

13.4 17.8 17.6 21.8

(2), 28.2 28.3, 28.3

(17),25.1 26.3, 28.7 (15), 18.6 20.6, 23.4 (7), 10.7 12.8, 14.8 (4),9.5 10.2, 11.7 ---

(19), 25.1 26.5, 28.7 (15), 18.6 20.6, 23.4 (7), 10.7 12.8, 14.8 (4),9.5 10.2, 11.7 (141), 9.5 17.8, 28.7

Total

(11), 13.4 16.8, 28.3

(121), 9.5 18.0, 28.7

(9), 15.7 16.7, 17.8

J VJ

--J
Table 4.1 Continued Africa Asia Australia Arctic Total

.j:>.

North America Permian Upper Permian (Zechstein, Ochoan) Middle Permian (Guadalupian) Lower Permian

Middle America

Europe

(18), 9.0 11.0, 13.8 (27), 10.5 11.9, 13.4 (26), 12.2 13.7, 16.0 (71), 9.0 11.6, 13.8

(195),7.9 10.7, 14.1

(213), 7.9 10.7, 14.1 (27), 11.9, (75), 12.8, 10.5 13.4 9.8 16.2

(45), 9.8 12.8, 16.2

(4), 12.9 13.3, 13.9


----

Total

(240), 7.9 11.1, 16.2

(4), 12.9 13.3, 13.9

(315), 7.9 11.2, 16.2

Carboniferous Pennsylvanian Mississippian

(5), 13.3 16.3, 18.6 (188), 13.7 17.3, 20.6 (193), 13.3 17.3, 20.6

(16), 13.4 15.2, 20.8

(21), 13.3 15.5, 20.8 (188), 13.7 17.3, 20.6

-----

Total

(16), 13.4 15.2, 20.8

(209), 13.3 17.1, 20.8

Devonian Upper Devonian Souris River alone Middle Devonian Lower Devonian

(79), 22.3 27.1, 34.0 (29), 24.0 28.4, 31.6 (9), 18.5 19.4, 21.8 (23), 15.6 17.0, 17.9

(26), 19.7 24.0, 28.3

(1) 19.1

(2), 21.8 22.3, 22.7

(108), 19.1 26.2, 34.0 (29), 24.0 28.4, 31.6 (21), 18.4, (29), 18.4, 15.7 21.8 15.6 24.6

(12), 15.7 17.6, 19.3 (6), 18.6 20.3, 24.6

Total

(111), 15.6 24.6, 34.6 (34), 23.5 26.5, 28.8 (17), 15.1 27.6, 35.8

(26), 19.7 24, 28.3

(1) 19.1

(20), 15.7 18.9, 24.6

(158), 15.6 23.7, 34.0 (34), 23.5 26.5, 28.8

Silurian

Ordovician

(2), 25.5 25.9, 26.3

(19), 15.1 27.4, 35.8

Cambrian Middle Cambrian Lower Cambrian Total

(3), 27.6 29.3, 31.1

(37), 25.2 28.8, 33.3 (25), 29.0 31.8, 35.5 ---

(4), 26.8 29.5, 32.0

(2), 22.0 29.5, 30.0

(46), 25.2 28.9, 33.3 (25), 29.0 31.8, 35.5

--(4), 26.8 29.5, 32.0 (2), 29.0 29.5, 30.0 (71), 25.2 29.9, 35.5 --.) U1

(3), 27.6 29.3, 31.1

(62), 25.2 30.0, 35.5

--.J 0\

Table North America Middle America

4.1 Continued

Europe

Africa

Asia

Australia

Arctic

Total

Proterozoic Lower Proterozoic Without Asian

(3), 14.6 15.1, 15.9

(8), 15.8 18.6, 21.0

(4), 33.0 33.7, 35.0

(13), 14.6 18.0, 24.3

(28), 10.1, (24), 17.8,

14.6 35.7 14.6 24.3

. . (number), minimum Table 4.2 o,xO data summary for manne evapontes: . average, maxImum
North America Tertiary Pliocene Miocene Oligocene 11.9, 11.65 (2), 12.17 11.9, 11.7 (2), 12.2 (5), 10.8 12.4, 12.9 (2), 12.1 13.1, 14.1 (5), 10.8 12.4, 12.9 Middle America Europe Africa Asia Australia Arctic Total

(2), 12.1 13.1, 14.1

(2), 12.1 13.1, 14.1 (5), 10.8 12.4, 12.9 (2), 12.17 11.9, 11.65 (9), 10.8 12.4, 14.1

Total

Cretaceous Turonian

(2), 15.4 15.4, 15.4

(2), 15.4 15.4, 15.4

-.J -.J

-.) 00

Table

4.2 Continued

North America Jurassic Oxfordian Upper Jurassic Undefined Jurassic Lower Jurassic (1) 10.0

Middle America

Europe

Africa

Asia

Australia

Arctic

Total

(2), 12.8 13.5, 14.3 (12), 11.9 14.1, 16.6

(2), 12.8 13.5, 14.3 (12), 11.9 14.1, 16.6 (1) 10.0

(1) 15.9

(1) 15.9

Total

(3), 10.0 12.3, 14.3

(13), 11.9 14.2, 16.6

(16), 10.0 13.8, 16.6

Triassic Upper Triassic (Keuper) Middle Triassic (Muschelkalk) Upper Buntsandstein (Rot) Upper Buntsandstein II Middle Buntsandstein Lower Buntsandstein Total

(2), 11.9 12.2, 12.5

(27), 10.6 14.6, 18.4 (4), 14.4 14.8, 15.8 (8), 12.9 15.3, 16.0 (11), 13.1 15.2, 16.3 (8), 10.3 12.9, 16.0 (5), 10.8 13.7, 15.3

(5), 9.6 12.7, 13.8

(34), 9.6 14.2, 14.6 (4), 14.4 14.8, 15.8 (8), 13.9 15.3, 16.0 (11), 13.1 15.2, 16.3 (8), 10.3 12.9, 16.0 (5), 10.8 13.7, 15.3

--(2), 11.9 12.2, 12.5 (63), 10.3 14.5, 18.4

-(5), 9.6 12.7, 13.7 (70), 9.6 14.4, 18.4

Permian Upper Permian (Zechstein, Ochoan) Middle Permian (Guadalupian) Lower Permian

(32), 8.1 11.8, 15.2 (2), 9.6, (24), 13.2, 9.3 9.9 11.5 14.7

(12), 10.6 12.6, 14.5 (44), 8.1 12.0, 15.2

(32), 11.8, (2), 9.6, (36), 13.0,

8.1 15.2 9.3 9.9 10.6 14.7

Total

(26), 9.3 12.9, 14.7

(70), 8.1 12.4, 15.2

-..) \0

00 0

Table North America Middle America Europe

4.2 Continued Africa Asia Australia Arctic


-------

Total

Carboniferous Pennsylvanian Mississippian Devonian Upper Devonian Souris River alone Middle Devonian Lower Devonian

(37), 12.3 13.3, 15.4

(37), 12.3 13.3, 15.4

(9), 12.9 15.1, 17.3 (6), 14.0 15.7, 17.3 (4), 14.3 14.8, 15.5 (1) 15.0 --(20), 12.9 15.2, 17.3

(4), 15.1 15.8, 16.3

(1) 13.7

(2), 17.1 17.5, 17.8

(3), 12.5 13.4, 14.6

(16), 12.9 15.5, 17.8 (6), 14.0 15.7, 17.3 (7), 12.5 14.2, 15.5 (1) 15.0

--

----

--

Total

(4), 15.1 15.8, 16.3

(1) 13.7

(5), 12.5 15.0, 17.8

(30), 12.5 15.2, 17.8

Silurian

(1) 10.6 (1) 15.9

(1) 10.6 (1) 15.9

Ordovician

Cambrian Middle Cambrian Lower Cambrian (10), 10.7 14.5, 17.1 (10), 10.7 14.5, 17.1

(1) 13.1

(1) 13.1 (10), 10.7 14.5, 17.1

Total

(1) 13.1

(11), 10.7 14.3, 17.1

Proterozoic Lower Proterozoic

(6), 14.2 16.7, 18.9

(1) 15.9

(7), 14.2 16.6, 18.9

00 .......

82

Stable Isotopes

4.2.2.2Description of the6345- and 6180-age curves


From the (')34Sdata, it is possible to construct an age curve for the Phanerozoic and part of the Precambrian (Figure 4.2). Yeremenko and Pankina (1971) stated that the curve for samples of the Soviet Union coincide with the curve for other parts of the world. Claypool et ai. (1980) stressed the reality of a universal age curve. In the Precambrian, (')34S values were + 18%0(the number of data is small and the age is not so precise as in the Younger Phanerozoic). There is an increase to more than +30%0in the Cambrian. Through the Ordovician and Silurian ages, there is a small decrease to +26.5%0. In the Upper Silurian, the (')34Svalues decrease sharply and reach a minimum at 18.4%0 in the Lower and Middle Devonian. This minimum is accentuated in the Lower Devonian in North America and the Middle Devonian in Australia. In the Upper Devonian, there is a sharp rise to a mean value of +26.2%0 in North America and +27.1%0 in Europe. The mean (')34Svalue for the Souris River formation, Canada, was found to be +28.4%0 and the highest value, +34%0. The few recorded values from Australia and Asia are lower. Throughout the Carboniferous and Permian, there is a continuous decrease in (')34S.The minimum of all marine evaporites near +10.7%0 is found in the Upper Permian (Zechstein, Ochoan, Tatarian). In the Lower and Middle Lower Triassic, there is a sharp increase in the mean (')34S value in the Upper Lower Triassic (Rot) to +26.3%0 in Europe and +28.3%0 in North America (Monkeoppi formation). The highest reported value is +28.7%0. Then, (')34Svalues decreased to + 19.6%0 in European Middle Triassic and to + 16%0in Upper Triassic (North America, Europe), thereafter remaining relatively constant with a small increase in the Upper Jurassic. It is noted that the Rot data significantly affect the calculations for the Triassic in Table 4.1. Their omission decreased the average (')34S by 2%0. In the Cretaceous, there is a minimum in the Apt/Albian and thereafter a continual increase to +21.8%0 in the Miocene. Samples of European Oligocene from the Rhine Valley were not included in the data summary because of possible admixture of Upper Permian salt. The (')180 data are plotted in Figure 4.3, and are compared with the (')180-age curve of Claypool et ai. (1980). The (')180-age curve starts in the late Precambrian at + 16.5%0, decreases in the Cambrian to + 13.5%0,and remains there through the Ordovician and Silurian to Middle Devonian. There is a small increase in the Upper Devonian to + 15.5%0,after which there is a steady decrease to a minimum in the Middle Permian. There is a sharp increase in the Upper Permian and Lower Triassic to +14.4%0. Thereafter, there appears to be a slow decrease through the Jurassic and Cretaceous with perhaps an increase through the Miocene to Pleistocene. However, data from Tertiary samples are sparse. In summary, the variation in (')180, :t 3.5%0, is much smaller than that found for (')34S(more than :t 10%0). One contributing factor is the smaller

Lithospheric Sources of Sulphur


slocene Miocene

83
0

T K J

Oligocene
Eocene

P'8iOOCene

Senonian Turonian anlan Albian Aptian Neocomlen U -

my
G>

100

M L

200
Rb"

R
---:U

U 'M -L"

.~=~

p C
-

.M
L
U
L

300

D
0

u
J'!'1. L

400

/'

500

~:::':"'!

p
+10 +20 +30 8348 (%0)
Figure 4.2 The o34S-age curve for marine evaporites (data from Table 4.1; curve from Claypool et al., 1980)

84

Stable Isotopes

Plio-Pleistocene

Oligocene Eocene Poloocone Turonian

Miocene

my
<:>

K
-

c.;;;nlan
Albian
Noocomlen

100

U -

J -M
L

200

R
-U

U 'M
'L" Ri:It'

P
C

M
L

300

0 0

U M L
<:>

400

a
\

500

p
+10

, ,
...

+20 8180 (%0)

Figure 4.3 The olRO-age curve for marine evaporites (data from Table 4.2; curve from Claypool et ai" 1980)

Lithospheric Sources of Sulphur

85

oxygen isotope fractionation during bacterial SOl- reduction. More data are required for parts of the curves and 8180 data are particularly lacking for deposits before the Devonian and after the Triassic. 4.2.2.3 Minima and abrupt rises in the 834S-age curve Considerable discussion has taken place in the literature on the rate of change of 834S in oceanic SOl- with time (Nielsen, 1965; Holser and Kaplan, 1966; Holser, 1977). Rates lower than 0.3%0 per Ma have been proposed (Nielsen, 1974) and disputed (Holser, 1977). Differences among continents may be due to local events. However, Holser (1977) described five common features about the 834Svariations among continents which may be summarized as follows: (a) The mean 834S-age curve established by worldwide sampling rIses sharply at certain times (Figure 4.4). (b) After a rise, the 834Svalue falls almost as rapidly so the event appears as a peak on the age curve. (c) Each event seems to have taken place within one stratigraphic unit, sometimes in much less than 5 Ma. (d) Each event is followed by a period when the mean oceanic 834S value is higher than the preevent level. (e) The mean curve for 8180 in marine sulfates at equivalent times appears either to display no discernible rise (Yudomski and Souris event) or a relatively small one (Rot event).

MIDDLE

UPPER

LOWER

~
I

CAMBRIAN

Yudomski

...J

T ~50Ma
w

LOWER

SOURIS ---"" ....


...

-1

UPPER

I-

LOWER

+10 +20 8345 I %0

I SILURIAN I I I ----1+20 +30 +20 8348, %0

....

PROTEROZOIC

+30 +40 8345, %0

Figure 4.4 The 534S variations in marine sulphate during the Rot, Souris, and Yudomski

86

Stable Isotopes

The concordance of these 834S peaks in different basins and continents reaffirms that most evaporite data are representative of oceanic S042- over time. Discordant 834Svalues from different basins and continents probably indicate mixing with continental waters and/or postdepositional alteration. 4.2.2.4 Non-marine evaporites Non-marine evaporites are usually not older than Tertiary. An exception is the Lower Permian Rotleigend salt deposits where the 834S value is from +5 to +9%0 (Pilot et al., 1972; Holser, 1979). The 834Svalues were found to range from +10 to +20%0 for Oligocene evaporites in the Rhine Valley. Lower values, e.g. + 11%0for the Buggingen deposits, are strong evidence for a contribution from leaching of Permian Zechstein evaporates by deepseated circulating brines (Nielsen, 1967). Birnbaum and Coleman (1979) traced the source of sulphur for a non-marine evaporite in the Tertiary Ebro basin. 4.2.3 Data from marine barites There are other natural sulphate minerals that might provide data from the oceanic isotope time curve but it is necessary to establish sample selection criteria. In particular, barite is interesting because it can precipitate under non-evaporitic conditions. Cecile et al. (1983) studied the isotope composition of sulphur and oxygen in samples of various barite occurrences (nodules, thin strata, thick beds, massive, etc.) in sediments from the Middle Devonian to Middle Mississippian age. This investigation and earlier data (Sakai, 1971; Rye et al., 1978; Morrow et al., 1978; Schroll and Pak, 1980) show that 834Sand 8180 values for barites are either higher than, or correspond to, those of evaporites of the same age. Therefore, the lowest 834S and 8180 values must correspond to unaltered SO/- in basins where they were generated. The best coincidence of isotopic data for barites and evaporites was observed for barites sampled from massive sediments as well as thin strata occurring in sediments with low carbon content. Barite concretions and thin strata occurring in organic-rich rocks are characterized by 834S and 8180 values as high as + 90 and + 32%0respectively, and must reflect sulphate reduction prior to sedimentation. 4.2.4 Reservoir of evaporite sulphur Based on estimates of evaporite volumes of different ages, Migdisov et al. (1983) calculated the mass of sulphur they contain. Using the masses and average 834S values of different age evaporites, they determined an overall

Lithospheric Sources of Sulphur

87

average 034S. The results are summarized in Table 4.3. The global mass of sulphate sulphur in evaporites is computed to be 108Tg S with a mean 034S of + 19.5%0. 4.3 ISOTOPIC COMPOSITION OF SULPHUR IN SULPHIDE ORES*

Sulphide ores are a major source of anthropogenic sulphur. About onethird of the world's manufactured sulphuric acid is produced from sulphide ores. In 1980, this corresponded to about 16-18 million tonnes of sulphur based on the estimated total sulphur production of 50-55 million tonnes (Nriagu, 1978). Moreover, a substantial amount of sulphur is released to the atmosphere during smelting of sulphide ores. According to Ryaboshapko (1983), this process contributes 20-25% of the global emissions of sulphur oxides which reached 98 Tg S y-l in the late 1980s. Therefore, the total production of sulphur from sulphide ores by that time was 20-25 Tg S,
Table 4.3 Volume, mass, and isotopic composition of sulphate sulphur in continental sediments (SCOPE 19, 1983) Volume (106 km') Mass of 13'4Sof sulphate rock (10HTg) sulphate (%0) 2.97 1.47 2.97 3.86 3.27 3.43 1.56 3.47 0.01 0.46 5.50 1.38 30.35
----

Stratigraphic interval Neogene Paleogene Cretaceous Jurassic Triassic Permian Carboniferous Devonian Silurian Ordovician Cambrian Upper Proterozoic

Gypsum and anhydrite 0.129 0.064 0.129 0.168 0.142 0.149 0.068 0.151 0.006 0.020 0.239 0.060 1.325

Total evaporite 0.410 0.250 0.570 0.750 0.550 1.430 0.240 0.311 0.032 0.045 1.525 0.360 6.473 Totals

+22.4 +18.1 +16.6 +16.6 +18.8 +10.8 +16.7 +20.1 +26.2 +28.5 +29.3 +16.3 +19.5 Average

* L.N. Grinenko and V.A Grinenko.

88

Stable Isotopes

which constituted an appreciable part of the total sulphur (169 Tg S y-l) extracted from the lithosphere (Ivanov, 1983). In order to use isotopic analyses of sulphur for monitoring anthropogenic sulphur fluxes to the atmosphere and hydrosphere, it is necessary to consider 834Svariations in different sulphide ores. The main sulphide ores are pyrite, chalcopyrite, polymetallic (dominantly Pb-Zn), and copper-nickel deposits. The use of sulphur isotope data to establish the systematics of sulphide ore formation (sulphur source, temperature, and physicochemical parameters) has been reported in several hundred publications (cf. Barnes, 1979). These data were used to construct histograms of the distribution of sulphur isotopes in various deposits in order to find their average 834S (Grinenko and Grinenko, 1974). There are sulphur isotope data for more than 70 pyrite-chalcopyrite deposits in various regions of the Soviet Union, Japan, Canada, Australia, Africa, and Central Europe. These are summarized in Figure 4.5 along with data from pyrite-polymetallic deposits. Many deposits have 834S values ranging from -2 to +5%0. Exceptions are highly metamorphic Precambrian deposits in Finland, Sweden, Karelia, the United States, Canada, and India where the 34Sdepletion is greater. In deposits of the Urals, Altai, Caucasus, and Japan, the average 834Svalues range from +0.5 to +4%0. The average 834S value for sulphides of such deposits is +3.5%0, quite close to the meteorite value. Several pyrite-polymetallic deposits such as Meggen, Rammelsberg (Central Europe), Shimokava (Japan), Casteliano, and Santa Lucia (Cuba) exhibit wide variations of 834S values within individual ore bodies and average values far removed from the meteorite composition. Deposits such as Meggen, Rammelsberg, and Shimokova have 834S values up to + 18%0 whereas deposits in Cuba are depleted in 34S (down to - 23%0). High depletions in 34S (average, -18%0) are observed also in stratiform deposits of Nairne, Australia.

-20

-10
8345(%0)

Figure 4.5 Distribution of 834Svalues in pyrite-chalcopyrite and pyrite-polymetallic deposits. One average value is entered for each deposit

Lithospheric Sources of Sulphur

89

Ores of almost 50 copper-molybdenum deposits of the United States, Peru, Mongolia, Kazakhstan, East Transbaikal, and Canada (Figure 4.6) are characterized by a relatively small range (5%0) in sulphur isotopic variation within an individual deposit. The average 834S values in ores of some deposits are close to that of meteorite sulphur, whereas for all deposits, the mean value ranges from -5 to +3%0 (average +0.3%0). Of the world's significant copper-nickel ore deposits, many in the Soviet Union have been extensively studied with sulphur isotopes. The 834Svalues of most of these deposits are close to zero (Figure 4.7). Deposits in the north-west part of the Siberian platform are an exception with 834Svalues for sulphides in the range from +8 to + 10%0.Within an individual deposit, the isotopic composition of sulphides varies by less than 10%0.The average 834S values for many deposits, regardless of continent, age, morphology, structure, ore content, etc., fall between 0 and +3%0. The mean 834Svalue of sulphide sulphur in copper-nickel deposits is +3.5%0 (Figure 4.7). Of particular interest are stratal-copper deposits in sandstone found in the United States and the Soviet Union. As seen in Figure 4.8, the isotopic composition of sulphur in such deposits varies over a wide range, from -18 to + 18%0. An individual deposit is also characterized by a large variation of 834S values in sulphides. The average 834S value is + 1.5%0for this type (Figure 4.8). The literature contains about 90 publications on the isotopic composition of sulphur in lead-zinc deposits, most of which belong to the stratiform type

-10

0 8345(%0)

+10

Figure 4.6 Distribution of 834Svalues in copper-molybdenum

deposits

-10

0 8345(%0)

+10

Figure 4.7 Distribution of 834Svalues in copper-nickel deposits

90

Stable Isotopes

0
-20

D
-10

n rl

0
8345(%0)

JL,
+10 +20

Figure 4.8 Distribution

of 834S values in stratal-copper

sandstones

-20

r--t

-10

~
0 834

+10

&
+20

CJ

S(%o)

Figure 4.9 Distribution of 834Svalues in lead-zinc deposits

(Figure 4.9). Some large deposits such as Broken Hill (Australia), Cartagena (Spain), Greenhow and Skyreholme (England) are isotopically uniform with mean 834S values close to the meteorite value. However, most lead-zinc deposits are characterized by large sulphur isotope variations with 834S values varying from -14 to +22%0, the average being +5.6%0 (Figure 4.9). In summary, the mean 834S value of sulphur emissions produced by processing copper-molybdenum ores should be close to the meteorite value. The same is true for most pyrite and copper-nickel sulphide ores, although some exceptions are enriched in 34Sor more rarely 32S. Sulphur from many copper and lead-zinc ores differs greatly in isotopic composition from that of meteorites. The isotopic signatures of different ores can be applied to regional monitoring of industrial sulphur acquisition by the environment during ore processing. The largest 'isotopic leverage' (difference between natural background 834S values and those of anthropogenic sulphur) is observed during processing of lead-zinc and copper ores from stratiform deposits. In many mixed-type deposits, i.e. pyrite-polymetallic, the isotopic leverage can also be large. Data for all types of deposits summarized in Figure 4.10 provide an estimate of the average isotope composition of sulphur compounds released during sulphide ore processing. Most of the deposits are slightly enriched in 34S, and the average 834S value may be taken as +3.4%0.

Lithospheric Sources of Sulphur

91

-20

-10

0 8345(%0)

+10

+20

+30

Figure 4.10 Distribution

of &34Svalues in main types of sulphide

deposits

4.4

NATIVE ELEMENTAL SULPHUR DEPOSITS*

The world annual production of native sulphur is about 17 million tonnes or nearly one-third of the total sulphur consumed by the chemical industry (Nriagu, 1978) and 10% of the sulphur produced annually from the lithosphere in the late 1970s (Ivanov, 1983). The main uses of native sulphur are for sulphuric acid production, chemical manufacture, and fertilizers. Therefore the bulk of produced native sulphur enters the soil as SO/- and is transported into continental basins. Major production of native sulphur occurs in the United States, Poland, the Soviet Union, Mexico, Iraq, China, Italy, and South America (Table 4.4). Most industrial sulphur deposits occur in sedimentary rocks. These socalled exogenic deposits constitute 90% of the world reservoir of native sulphur. Endogenous or volcanic sulphur makes up the remaining 10% (Vlasov, 1971). Sulphur isotope data for major and some minor elemental sulphur deposits are summarized in Table 4.5. Negative 034Svalues are found in small native sulphur deposits such as Novodimitrovskaya, USSR, Boraitotto, Italy, and Moss Bluff, USA. In some industrial sulphur deposits, the native sulphur
* A. Yu, Lien.

92

Stable Isotopes

Table 4.4 Annual world production of native sulphur and its isotope composition (data from Ruckmick et al., 1979) Sedimentary Syngenetic 034S TgS (%0) 1.90 5.00 2.57 0.75 0.20 0.04 0.01 0.01 10.48 62.79 6.7 12.9 13.6 15.9 9.9 11.9 Epigenetic 034S TgS (%0) 4.30 1.70 6.00 35.95 7.7 7.7 7.7 Volcanic 034S Tg S (0/00) 0.01 0.03 0.05 0.09 0.02 0.01 0.21 1.26 1.0 1.0 1.0 1.0 ? 1.0

Country USA Poland USSR Mexico Iraq China South America Italy Turkey Others Total Percent

Total Tg S 6.21 5.00 2.60 1.70 0.75 0.25 0.09 0.06 0.01 0.02 16.69 100.0

is isotopically similar to associated evaporite sulphate. This is attributed to intensive bacterial sulphate reduction with small isotopic selectivity. An estimate of the average isotopic composition of produced native sulphur can be based on large occurrences. These include deposits in cap rocks of salt domes in the Gulf of Mexico, the Delaware field in west Texas, USA (Ault and Kulp, 1959), deposits in the Polish and Soviet SubCarpathians, the Mishrak deposit in Iraq, and stratal sulphur deposits in Sicily. The average 334Svalues of native sulphur in two large deposits, Spindleton and Boling, Gulf of Mexico, USA, are +3.1 and +5.7%0 respectively, whereas in Challenger Knoll it is + 14.2%0. The average 334Svalue for stratal and stratiform deposits of the Delaware basin is +6.7%0. It is estimated that sulphur production in this region will substantially increase in the near future (Bodenlos and Nelson, 1979; Ruckmick et al., 1979). Tarnobzheg, the largest deposit in Poland, contains 20% elemental sulphur with an average 334Svalue of + 12.9%0. In the Soviet Sub-Carpathians, the average 334Svalue for industrial sulphur is + 13.6%0.

Lithospheric Sources of Sulphur

93

Table 4.5 Comparison of average 334Svalues of native sulphur and sedimentary sulphates in exogenic deposits 334S (%0) Deposit Large industrial deposits USSR Shor-Su, deposit 'L-K' deposit 'K' Gaurdak Middle Volga Sub-Carpathian Iraq, Mishrak Sicily USA, Gulf of Mexico, Spindleton Boling Challenger Knoll Delaware basin (USA) Sulphate Sulphur References

18.4 18.1 14.0 13.6 19.8 19.4 +20.7 17.0 16.2 14.8 10.7

+17.4 +8.8 +4.5 +12.6 +13.2 +15.9 +10 +5.7 +3.1 +14.2 +6.7

Vinogradov et aI. (1964) Vinogradov et aI. (1964)

(Figure 4.11) Feely and Feely and Davis and (1979) Davis and (1970)" Kulp (1957) Kulp (1957) Kirkland Kirkland

Small deposits USSR Novodmitrovskoe Romnenskoe Krasnovodskoe Italy, Borraitotto, Sicily USA, Moss Bluff, Gulf Coast
a

11.0 22.6 12.3 21.5 17.3

-4.3 -6.5 -19.0 -0.9 -6.4

Dessau et al. (1962) Feely and Kulp (1957)

Data of author.

The world's largest sulphur deposit, Mishrak in Iraq, has the heaviest isotopic composition of sulphur, with an average 334Sof +15.9%0. The 334S values of sulphur ores in four rather large deposits of Sicily (Baccarato, Cozzodisi, Trabia, Ciavolotta) vary from +4 to + 14%0,with an average of + 10%0(Figure 4.11). Native sulphur in volcanogenic deposits corresponding to about 1% of the world production have 534S values within a narrow range near + 1%0 (Lein, 1968; Vinogradov, 1970).

94

Stable Isotopes

GYPSUM.
So.

PRIMARY

SECONDARY

.L
L.
I

BACCARATO
COZZODISI BORRAITOTTO TRABIA CIAVOlOTT A

...

. OJ.rnDEb CD

.a:fh

..... .DC .. . . [J
-20 -10 0 +10 +20

+30

834$, %0
Figure 4.11 Histogram of 834S values for sulphate and native sulphur deposits of Sicily (Dessau et al., 1962)

Three peaks in the distribution of exogenic native sulphur deposits are evident in the stratigraphic scale corresponding to arid continental periods and evaporite formation (Figure 4.12). The first peak occurred in the Upper Permian, and includes stratiform deposits of native sulphur in west Texas, USA, and middle Volga, USSR. The second peak in the Jurassic includes the large sulphur deposits in cap rocks of salt domes in and around the Gulf of Mexico and the Gaurdak deposit in the Turkmenian SSR. The third peak
en OJ t-

" Sulphate Rocks

o
~

. S'
I

en2

10

=> en LL 0 en en :::< 0

0 a: UJ ~ t J: ~11Cn5

en 0> Ia

l' 'f
\ I I I I

~
-8

Figure 4.12 Stratigraphic distribution of the world's reservoirs of sulphate rocks and native sulphur

Lithospheric Sources of Sulphur

95

in the Miocene includes stratiform sulphur deposits in Sicily, Soviet and Polish Sub-Carpathians, Crimea, and Iraq. These peaks correspond to more than 90% of the total exogenic sulphur. Three main factors determine the isotopic composition of native sulphur from exogenic deposits: (a) Age and isotopic composition of initial evaporites (b) Isotope fractionation during sulphate reduction (c) Extent of secondary changes depending primarily on the depth of occurrence The average isotopic composition of native sulphur produced globally, estimated from production data for different countries and the average 034S value of their largest deposits (Tables 4.4 and 4.5), is + 10%0. 4.5 SULPHUR ISOTOPE COMPOSITION OF FOSSIL FUELS

4.5.1 Peat and lignite* Peat represents the first stage of maturation of dead plant material. Therefore, its sulphur must be considered in the sequence of the transformations from living plants to lignite and coal. Sulphur is an essential constituent of the living cell. Plants can take up SOi- or S02 directly from the environment. The S organic compounds resulting from assimilatory SOi- reduction are depleted in 34Sby 1 to 5%0. Thus the 'primary' organically bonded sulphur in peat will have slightly lower 034Svalues than the source sulphate. Peat-producing bog areas beneath rain slopes of mountain chains have practically no contact with groundwater sulphate so that the mosses and other plants must obtain sulphur from the atmosphere. Assuming a 034S value of + 3 to +6%0 for this atmospheric sulphur (Chapter 5), the expected range for peat is about 0 to +4%0. Later, some of the buried sulphur is remobilized to volatile sulphur organics (dimethyl sulphide, etc.) with a slight isotope fractionation, and the range of 034S values in the remaining peat becomes broader. Peat samples from rain slopes in the Harz Mountains, FRG, vary in 034Svalue from -3 to + 10%0.The SIC ratio in all samples is lower than 1:100, which is similar to that of land plants. Dissolved sulphate in groundwater will penetrate bogs in local depressions. This is a favourable habitat for SOi- reducing organisms, particularly at deeper burial levels. The reduced sulphur reacts with the decaying organic matter and iron to form S organic compounds and iron sulphide concretions.
* H. Nielsen.

96

Stable Isotopes

During mining, these concretions are readily weathered under the more aerobic conditions to form sulphuric acid. The range of 034S values for sulphide in the concretions depends on the conditions of SO/- reduction. Low concentrations of SOi- will be reduced more or less quantitatively and the 034Svalue of the sulphide will be nearly the same as the SO i- . At higher concentrations, the remaining SOi- will follow the Rayleigh equation if the system is effectively closed (Chapter 1). In that case, the cores of the concretions are more depleted in 34Sthan the rims. The range of 034S values in a given concretion depends upon its formation time with respect to the extent of SO i- reduction. If S04Zreduction approaches 100%, the average 034S values of the concretions should be the same as the initial SOi-. An example is the lignite deposit in the impact crater of the Norlinger Ries in southern FRG where the 034S values for sulphide range from -27 to +25%0. Evaporite strata are lacking around the Ries structure and therefore the S04Z- content of the ground

water must have been low. The mean value of 034S= -3.5

:::+:: 15%0 suggests

that this groundwater SOi- originated from the leaching of sedimentary rocks. In northern Central Europe, the ground waters are usually much richer in SOi- due to the wide distribution of Zechstein salt domes. Such ground waters form an almost 'infinite' sulphate reservoir so that reduction proceeds under quasi-open system conditions. Therefore, the sulphide is uniformly light and the S content is high. The sulphide 034S values in lignite seams from the mining district south of Helmstedt, eastern FRG, range from -24 to -10%0. The mean 034S = -15.2 :::+:: 6.2%0 agrees with the assumed provenance of these sulphides from reduction of Zechstein solutions with 034S = + 11%0. In some cases, elemental sulphur may be formed in abundance in the upper 0.5 m of peat deposits. Sulphur isotope data revealed that one such occurrence arose by sulphide oxidation (Krouse, 1980; see Section 7.1.3). 4.5.2 Coal* The extent of sulphur contribution to the environment from the mining and utilization of coal is well known. For example, in 1969 the combustion of coal was responsible for 58.2% of atmospheric pollutant SOz in the United States (US Department of Health, Education and Welfare, 1969). Although SOz is the main form in which sulphur is released into the atmosphere by coal utilization, sulphur may also be introduced into the environment as sulphate during the mining of coal or from the weathering of coal stockpiles

* J. W. Smith.

Lithospheric Sources of Sulphur

97

1.0

. Marginal

,\O~
0.8
..... 0 "D <f. 0.6 \-:J
.r::. Q. -:J If)

.
/

"0
\)'
. <ec,

. 0

0'

~e< ~o

marine

"0

0.4

;
0.2

/'

~
/
/'

0 /'

~
/'./'

/' .
/'

/'

L .fT

/'

ower delta
plain

.
d
elta plain

/'

.
.

Fluvial-upper

Braided alluvial fan

. Well characterized
0 Undifferentiated fluvio -deltaic

-2.0

0.0

+2.0

+4.0

+6.0

+8.0

834 Sorgonic(%o)(canyon diablo troilite) Figure 4.13 Effect of deposition environment on the content and isotopic composition of sulphur in coal

and waste dumps. Subsequent bacterial action may then convert this sulphate to S-containing gases. Earlier workers were most interested in the source(s) of sulphur in coal and concentrated their studies on atypical coals with unusually high organic or total sulphur contents (Kavcic, 1958; Rafter, 1962; Vinogradov and Kizilshtein, 1969). Coals of the former type are comparatively rare, and coals in the latter category are now unlikely to be utilized without beneficiation to reduce the content of sulphur-containing minerals. In addition, within-seam variations in the isotopic composition of coal seam components are now firmly established. These within-seam variations in the distribution and isotopic composition of organically combined sulphur and of finely disseminated and massive

Table 4.6 Relationship between. isotopic composition and concentration of sulphur in coal and environments of deposition and maturation (from data of Smith and Batts, 1974; Price and Shieh 1979; and Smith et al., 1982) Environment of deposition and maturation
Freshwater conditions during deposition and maturation

Sulphur content

Forms of sulphur

834S (%0)

\0 00

<1%

Mainly organic with minor disseminated pyrite Organic and pyrite sulphur contents of same order

Organic + 4 :t 3

High sulphate concentrations, normally marine, during plant growth and early deposition Freshwater environment during plant growth followed by marine invasion in early deposition

Commonly

~ 2%

Organic and pyrite sulphur with similar values, e.g. -16% reflecting single sulphate source Organic + 4 :t 3 Pyrite along margins + 20 reflecting rapid complete reduction of available marine sulphate Organic + 4 :t 3 Pyrite along margins increasing from +5 to +28 with increasing depth of penetration of sulphate reflecting kinetically controlled sulphate reduction Organic and pyritic sulphur still reflect environments of deposition and maturation; sulphate usually + 13 to +30

In unaffected portions of seam < 1%. High pyrite sulphur contents along seam margins

Mainly organic with minor disseminated pyrite in unaltered coal. Massive pyrite along seam margins

Freshwater environment during plant growth followed by later marine invasion during maturation

In unaffected portions of seam < 1%. High pyrite sulphur contents along seam margins

Mainly organic with minor disseminated pyrite in unaltered coal. Massive pyrite along seam margins

Postiithitication invasion of seam by sulphate in solution

Usually ~ 1%

Main feature high additional sulphate tilling veins

Lithospheric Sources of Sulphur

99

pyrite (Smith and Batts, 1974; Price and Shieh, 1979) have provided fundamental information that allows five environments for coal deposition and maturation to be recognized (Table 4.6). It is clear from the above that isotopic data for total sulphur cannot be used to assess the isotopic composition of emissions to the atmosphere. Further, the few data from a given deposit may not be representative of the coal actually recovered. However, Table 4.6 serves as a guide to the isotopic composition of coals after projected beneficiation. For example, in the Australian and Illinois coals, which have been described in detail, only massive pyrite is likely to be separated by commercial beneficiation processes. and it is possible to calculate the effect of the removal of this pyrite on the isotopic composition and sulphur content of the washed coal. In low sulphur coals containing 1% total sulphur and 0.2% pyritic sulphur from Australian fluvio-deltaic sequences, variations in organic and total sulphur contents reflect directly associated variations in brackishness (available SOi-) (l.W. Hunt and l.W. Smith, unpublished data, 1984). Figure 4.13 demonstrates that even in coals of undoubted freshwater origin and low pyrite content, significant variations in sulphur isotope composition exist.

YUGOSLAVIA

H Istria

INDIAH

HighS

SOUTH AFRICA f ANTARCTICA f Danbas ~ <1%S

USSR

British Columbia j i Albert a


CANADA"" I

Alberta I (foothills) 0 rganic Massive

I (plains) 1 pyrite I NEW ZEALAND I


+10

ILLINOIS!

North American (not specified) I

Black coals AUSTRALIA -30 Total


1

1<1%S

Brawn coals I
+20 +30

---L -20

-10 631,Sorganic(%O)

Figure 4.14 The isotopic composition of sulphur in coal (data from references cited; *data from H.R. Krouse, personal communication)

100

Stable Isotopes

The published isotopic data on sulphur in coal are displayed in Figure 4.14. No attempt has been made to weight this information with regard to number of samples examined, size of deposit, etc. The only objective of the figure is to demonstrate the variability in measured 034S values. In conclusion, with the exception of some studies described in Chapter 8, data on the isotopic composition of sulphur in coal were not collected for the purpose of evaluating sulphur emissions to the atmosphere, and on examination are clearly unsuited to that purpose. Acceptable data will require an extensive global assessment of the isotopic composition of coals as utilized.

4.5.3 Oil*
4.5.3.1 Introduction Oxidized sulphur compounds formed during combustion and refining of oil may enter the atmosphere, continental water basins, and the world ocean systems. Oil and oil products per se may introduce sulphur into sediments of oceans and continental reservoirs as the result of tanker spills and offshore drilling practices. Data on the isotopic composition of sulphur in oil were first reported by Thode et at. (1958) for reservoirs of different ages in Canada and the United States. Later, Thode and Monster (1970) studied oils in the Middle East (Arabia, Kuwait, Northern and Southern Iraq, Iran). These studies revealed that in large reservoirs and oil provinces of the same age the sulphur isotopic composition of oil is markedly uniform, even where the concentration of sulphur varies widely. In contrast, the 034Svalues vary considerably among oilfields (from -5.2 to +28.2%0). Similar investigations in the Soviet Union initiated by Eremenko (1960), Pankina (1978) and Eremenko and Pankina (1962, 1963) dealt with oils occurring over a wide stratigraphic range. These studies, which included entire columns from many oil- and gas-bearing provinces, revealed a much greater spread in 034Svalues (about 50%0). 4.5.3.2 Summary of 0345 data for oil Sulphur isotope data for oils can be summarized according to the age of the reservoir strata. Oil is difficult to date and in many locations such as the Middle East (Thode and Rees, 1970) and Timano-Pechora, USSR, rocks of different ages may contain the same oil pool. The most ancient oil-bearing deposits are found in Proterozoic Riphean and Vendian complexes. There are different productive zones in the East Siberian platform. Minor showings are observed also in the central (Moscow syncline) and eastern (Upper Kama depression) parts of the Russian
* R. G. Pankina.

Lithospheric Sources of Sulphur

101

platform. Oil of the Vendian and Riphean periods was found to have an average S content and 334S value of 1% and + 15%0 respectively (Table 4.7). Oils in Cambrian deposits of the East Siberian platform (Irkutsk amphitheatre) and Sub-Baltic syneclise are the most enriched in 34Sof the entire stratigraphic column (+ 19 to +25%0, Table 4.8). The 334S values of these two regions differ by only 4%0even though the locations are thousands of kilometres apart and the depositional environments are substantially different. Ordovician deposits in the Gusev (Baltic syncline) and Kibarty (TimanoPechora Province) areas have average 334S values of +6.8 and 7.3%0 respectively and low sulphur contents (0.28-0.31%). Similar data were
Table 4.7 Sulphur isotopic composition of oils from Proterozoic reservoir rocks Sulphur Content (%) 1.3 0.68 0.7 0.6 0.98 0.88

Age Vendian Vendian Vendian Vendian Vendian Riphean

Area Upper Kama depression Upper Kama depression Irkutsk amphitheatre Irkutsk amphitheatre East Siberian platform East Siberian platform

Field, well Vereshchaginskoe, 60 Sokolovskoe, 52 Markovskoe, 28-R Krivolukskoe, 3-Sp Diyavolskoe Kuyumbinskoe, 9

S34S(%0) +13.8 +14.0 +15.2 +15.2 +17.8 +14.2

Table 4.8 Sulphur isotopic composition of oils from Cambrian reservoir rocks Sulphur Content (%)

Area Irkutsk amphitheatre Irkutsk amphitheatre Irkutsk amphitheatre Baltic syneclise

Field, well

Reservoir strata

S34S (%0)

Markovskoe, 1 Markovskoe, 61 Osinskoe, 1 Plunge

Osinsky horizon Osinsky horizon Osinsky horizon Tiskretskaya suite

1.01 0.7 0.4 0.58

+22.3 +22.5 +25.3 +19.3

102

Stable Isotopes

obtained for deposits in Ontario, Canada, and west Texas, North Dakota, and Montana, USA (Thode et ai., 1958; Thode, 1981). Minor oil occurrences in Silurian deposits of the Kuldig area, Sub-Baltic (syneclise), were found to have an average 834Svalue of + 3.4%0.In Ontario, west Texas, North Dakota, and Montana, the mean 834Svalues were closer to +9%0. The discrepancy may relate to the fact that in the Timano-Pechora gas- and oil-bearing province, Silurian and Devonian reservoir rocks contain the same oil pools. Economically viable reservoirs of oil occur in Devonian deposits of the Soviet Union, United States and Canada. The average 834S values among different fields span a wide range (0 to + 16%0, Table 4.9). In western Canada and the United States most of the data fall between + 10 and + 15%0. It is interesting to note that the excursion in the evaporite 834Stime curve during this time (Figure 4.2) is about the same as the range found for the oil. Oils produced from Upper Carboniferous strata from many regions of the Soviet Union are characterized by a rather homogeneous 834S value near +5%0 in spite of the significant range in sulphur content (0.4-3.2%) (Table 4.10). In contrast, many data in the Pennsylvanian of Wyoming, USA, are near -5%0 (Vredenburgh and Cheney, 1971) and a range from -6 to +9%0 has been found in the Mississippian in Wyoming and western Canada (Thode et ai., 1958; Vredenburgh and Cheney, 1971; Orr, 1974). Oil in Permian deposits of the Soviet Union tends to have average 834S values near -2.5%0 whereas the mean for Wind River Basin, Wyoming, is about 1%0lighter (Vredenburgh and Cheney, 1971). Orr (1974) found the mean 834Svalue for the Big Horn Basin, Wyoming, to be near + 1%0(Table 4.11). Triassic commercial oil pools in the East Sub-Caucasus have low sulphur contents (0.14%). The average 834S value for the Urozhainenskoe deposit is +6.2%0. Oil in the Karagandinskoe deposit in the Sub-Caspian has a 834S value of +5.2%0. Triassic oils in the Miziiskaya platform (Bulgaria) are also enriched in 34S (+ 3.1%0). Triassic deposits in northern Iraq have a mean 834Svalue around 2%0whereas negative values around - 5%0were found for the Wind River Basin in Wyoming (Table 4.12). Commercial oil pools in Jurassic reservoirs of large areas of the Soviet Union and the Wind River Basin (Table 4.13) are variably depleted in 34S. Cretaceous oils vary in S content from less than 0.1% in the Wind River Basin to greater than 5% in the Athabasca oil sands of Alberta, Canada. The 834S values vary from -7 to + 12%0.For some areas, the 834S values are markedly uniform throughout the Cretaceous at locations up to 200 km apart (e.g. -5.5::':: 2%0 in northern Iran). In contrast, in the Lower Cretaceous of Alberta, Canada, heavy oils with greater than 2% S and light

Lithospheric Sources of Sulphur

103

Table 4.9 Sulphur isotopic composition of oils from Devonian reservoir rocks Sulphur Area Field Age Content (%) 834S (%0)

USSR Rechitskoe Pripyat trough ZhigulevskoSol Ravine Pugharevsky ZhigulevskoMukhanovskoe Pugharevsky MikhailovskoZhigulevskoKokhanovskoe Pugharevsky ZhigulevskoPokrovskoe Pugharevsky SernovodskoRadaevskoe Abdulinsky Romashkinskoe Tatarsky roof Tatarsky roof Shkapovskoe Leonidovskoe Tatarsky roof Permsko-Bashkir sky roof Kushkulskoe Rechitskoe Pripyat trough Ostashkovichskoe Pripyat trough Tishkovskoe Pripyat trough Canada (Thode et al., 1958) Alberta Leduc Alberta Excelsior Alberta Stettler Alberta Stettler Alberta Leduc Alberta Golden Spike USA (Thode, 1981) Montana 18T 30NR 48E

Dz D3 D3 D3 D3 D3 D3 D3 D3 D3 D3 D3 D3 Dz Dz D3 D3 D3 D3 D

0.3 0.72 0.72 0.66 1.08 2.8 1.6 1.81 1.55 1.6 0.58 1.02 0.7 0.3 0.66 1.45 1.78 0.26 0.19 -

+7.4 +6.4 +7.6 +10.2 +8.5 +4.4 +2.4 +0.1 +2.7 +5.8 +12.5 +14.0 +15.9 +12.0 +15.3 +13.1 +10.9 +12.5 +15.0 +10.1

oils with less than 1% Shave 034S values near +6 and + 11%0respectively (Thode et ai., 1958). The S content and 034S value in the Gazli deposit of Middle Asia are 0.1% and +3.2%0 respectively. In the Armavir deposit, West Caucasus, the average 034S value is +8.2%0 whereas the S content is 0.16%. In West Siberia, the 034Svalue and S content vary from +0.5 to +6.6%0 (+ 3.2 on average) and from 0.7 to 1.93% respectively. As seen in Table 4.14, there is no apparentrelationship between the S content and 034Svalues.

104

Stable Isotopes

Table 4.10 Sulphur isotopic composition of oils from Carboniferous reservoir rocks Sulphur Area USSR Peschano- Umetskoe Zhigulevskochevsky roof Zhigulevskochevsky roof Zhigulevskochevsky roof Zhigulevskochevsky roof Zhigulevskochevsky roof SernovodskoAbdullinskaya depression Bashkirsky roof USA Oklahoma" Oklahoma" Kansas" Wyoming Wyoming Wyoming Wyoming Wyoming Wyoming Wyoming Wyoming Canada" Saskatchewan Saskatchewan Saskatchewan Alberta
U

Field

Age (formation) Pennsylvanian (Bobrikovsky) (Bobrikovsky) (Bobrikovsky) (Bobrikovsky) (Bobrikovsky) (Bobrikovsky)

Content (%)

(\34S (%0)

0.4 2.6 1.38 1.59 1.25 2.2

+3.8 +4.7 +5.5 +3.8 +5.6 +4.9

Karpovo-Syboskoe Sol ravine Krasnoyarskoe Mukhanovskoe Tarkhanskoe

Radaevskoe Arlanskoe Circle Ridgeb Black Mountain" Fourbear" Manderson" Forget Black Mountain" Coleville-Smiley Turner Valley

(Bobrikovsky) (Bobrikovsky) (Springer) (Misener) (Cherokee) (Tensleep) (Tensleep) (Tensleep) (Tensleep) (Phosphoria) (Phosphoria) Mississippian (Madison") (Madison") (Charles) (Mission Canyon) (Banff Sand) (Madison)

3.2 3.02 0.17 0.06 0.24 2.39 2.6 3.0 3.1 1.8 1.8 1.9 3.2 0.54 1.8 2.6 0.35

+3.2 +5.8 -0.1 -5.6 -3.9 -5.1 -5.1 -3.8 +5.0 +5.6 +5.0 -3.9 -3.0 +1.8 +5.0 +4.9 +8.8

Thode et al., 1958. Vredenburgh and Cheney, 1971.


1974.

c Orr,

Lithospheric Sources of Sulphur

105

Table 4.11 Sulphur isotopic composition of oils from Permian reservoir rocks Sulphur Field or area USSR Mukhanovskoe Vostochno-Chernovskoe Novo-Klyuchevskoe Kohkanovskoe USA Wind River Basin, Wyoming" Big Horn Basin, Wyomingb West Texas"
a h

Content (%)

8}4S (%0)

1.82 1.82 1.82 2.5 2 2.5

-2.3 -2.9 -1.6 -2.0 -3.3 +1.1 +1.0

Vredenburgh and Cheney, 1971. Orr, 1974.


and Monster, 1965.

C Thode

Table 4.12 Isotopic composition of sulphur oils from Triassic deposits Sulphur Field or area USSR Urozhoinenskoe, East Sub-Caucasus Karagandinskoe, Sub-Caspian depression Bulgaria Miziiskaya platform Iraq (Thode and Monster, 1970) Butmah Field Alan Field USA (Vredenburgh and Cheney, 1971) Crow Mountain, Wind River Basin, Wyoming Nugget, Wind River Basin, Wyoming 1.9 2.2 Content (%) 8}4S (%0)

0.14

+6.2 +5.2 +3.1 +1.9 +2.8 -6.0 -4.0

Cenozoic oils have a 40%0 spread in 034S values, greater than for any other age (Table 4.15). The most positive 034Svalue (+28%0) is found along with low S contentni the Uinta Basin, Utah, USA. The most negative 634S values (average, -11%0) are found in oils in the Tadzhik depression, USSR.

106

Stable Isotopes

Table4.13 LIsotopiccomposition of sulphurin oils from Jurassic reservoirrocks


Sulphur
----

Field or area USSR Shurchi, Amydarjinskaya syneclise Akdzhar, Kaganskoe upwelling Dzharkak, Kaganskoe upwelling Karaul-Bazar, Kaganskoe upwelling Shurtepe, Kaganskoe upwelling Yulduzkak, Kaganskoe upwelling Urtobulok, Chardzhousky step Zamnokulskoe, Sub-Caucasus Kraxhanbas, Sub-Caucasus Martyshi, Sub-Caucasus Mulymjnskoe, West Siberia Mortymjinskoe, West Siberia USA (Vredenburgh and Cheney, 1971) Wind River Basin, Wyoming

Content (%)

834S (%0)

0.87 0.90 0.74 0.47 0.96 0.76

-7.5 -7.9 -6.4 -7.1 -7.7 -6.8 -8.8 -5.6 +3.7 +26.3 -2.5 -2.2 -4.6

2.2

Table 4.14 Sulphur isotopic composition of Lower Cretaceous oils in Western Siberia Sulphur Field, well Fyodorovskoe, 79 Fyodorovskoe, 79 Fyodorovskoe, 79 Chipalskoe, 54 Chipalskoe, 58 Chipalskoe, 58 Pokachevskoe, 35 Pokachevskoe, 23 Malobalykskoe, 7 Malobalykskoe, 7 Aganskoe, 12 Aganskoe, 3 Ust-Balykskoe, 63 Average Content (%) 1.17 1.93 0.93 1.02 1.22 1.0 0.7 0.8 1.37 1.06 0.90 0.86 1.35 834S (%0) +2.0 +4.4 +5.9 +4.7 +5.0 +6.6 +1.9 +4.1 +5.5 +6.5 +0.5 +2.0 +2.8 +3.2

Lithospheric Sources of Sulphur Table 4.15 Sulphur isotopic composition of Cenozoic oils Sulphur
--

107

Area USSR Tadzhik depression Fergana depression Sub-Caucasus Sub-Carpathians Iraq (Thode and Monster, 1970) Kirkuk Bai Masson Field Jambur Field USA (Thode et al., 1958) Green River, Utah

Age (formation)

Content (%)

334S (%0)

(Alay) (Bukhara) (Kyma) Eocene

0.60 0.50 0.80 0.42

-11.0 -4.5 -1.2 +3.8 -5.5 -7.1 6.5

Tertiary Tertiary Tertiary Eocene 0.11

+28.2

Variations in the 034Svalues of sulphur in oil are shown according to the age of the reservoir rocks in Figure 4.15. The evaporite 034S-time curve is also drawn. Two observations are noteworthy. In Precambrian and Cambrian host rocks, the 034S values of the oil are homogeneous and close to the evaporite value. The spread in 034S values seem to be greater in younger reservoirs. Perhaps this reflects variations in shallower deposition environments and migration. Older oil may migrate upwards into younger rocks. Although migration of younger oil into older rocks has been observed, it is less likely to occur. 4.5.3.3 Estimate of the mean 0345 value of oil The sulphur content and isotopic composition of oils along with their distribution in Phanerozoic geological systems permit a rough estimate of the mean 034Svalue of oil sulphur that can potentially enter the environment. In spite of significant variations, 034S values for the bulk of samples examined fall within the range -5 to + 10%0 (Figure 4.15). Oils from Cretaceous, Devonian, Carboniferous, and Jurassic reservoirs constitute most of the world resources. The sulphur content of these oils is usually about 1%. The sulphur isotopic composition of oils characterized by a high S content (2.5%) greatly differs from the bulk of oils. Such oil pools are usually small in a given stratum. Since their sulphur isotopic compositions are divergent,

108
106 YEARS BP

Stable Isotopes

ERA

PERIOD

-10

QUATERNARY CENOZOIC TERTIARY


100 - MESOZOIC JURASSIC 200 'RIASSIC PERMIAN CRETACEOUS

--

+40

300
400

PENNSYLVANIAN

I-

MISSISSIPPIAN PALAEOZOIC DEVONIAN SILURIAN ORDOVICIAN

of oil with age of reservoir rocks

500

t-

CAMBRIAN

600 PRECAMBRIAN 700

Figure

4.15

Variation

of 834S values

one can consider the possibility that their deviations mutually balance, resulting in an average sulphur isotopic composition near zero. Therefore, it would appear that the isotopic composition of oil sulphur that can potentially contaminate the environment varies from -5 to + 10%0. Data for oil utilization in Australia are given in Section 8.8.2. 4.5.3.4 Use of sulphur isotope analyses to determine petroleum contamination in sediments Venkatesan et al. (1982) examined hydrocarbon distributions in pristine and polluted marine environments. In both reducing and oxidizing pristine sediments, the 034Svalues for extractable organic sulphur were lower than -15%0 as the consequence of bacterial SOireduction. In contrast, petroleum-derived S in compounds resistant to weathering was found to be more enriched in 34S: 0%0for Prudhoe Bay, Alaska; -5 to +5%0 for Gulf Coast oils; and +8 to + 15%0for natural seeps in southern California. At a number of locations, positive 034Svalues were found in sediments for which organic geochemical analyses indicated crude oil contamination.

Lithospheric Sources of Sulphur

109

4.5.4 HzS in hydrocarbon deposits* Hydrocarbon deposits contain different concentrations of HzS as a free gas or dissolved in formation waters. During recovery and refining, HzS and products of its oxidation may contaminate the environment, especially the atmosphere (see Section 8.2). Thode et al. (1958) were the first to study the isotopic composition of HzS in gas deposits. They analysed HzS associated with oil in Devonian reservoir rocks in Alberta, Canada. Although the HzS content varied significantly, 034Svalues were markedly uniform and approximated those of oils from the same wells (Table 4.16). The authors concluded that the HzS was formed during maturation of the oil with negligible isotope fractionation. The isotopic composition of HzS from gas- and oilfields was studied in many areas of the Soviet Union including the Volga-Ural region and the Tadzhik, Fergana, and Amu-Dariya depressions (Eremenko and Pankina, 1962, 1963; Pankina and Mekhtiyeva, 1964; Pankina, 1978). In these deposits, production zones are found at depths of less than 1000 m and temperatures less than 60C. The HzS content is low and the 034S values have a total spread of 40%0.Diversity of 034Svalues was found in individual strata: -14 to +8%0 in Lower Carboniferous, -22 to +10%0 in Middle Carboniferous, -3 to -1%0 in Lower Permian, -12 to +11%0 in Upper Permian, -16 to -1%0 in Jurassic, and + 11 to + 17%0in Palaeocene deposits. In contrast to the data of Table 4.16, the isotopic composition of HzS in these shallow occurrences often differed from that of the associated oil. HzS of younger deposits was relatively enriched in 34S whereas in older gases HzS was either depleted in 34S with respect to the oil or occasionally had the same isotope composition. Isotopic data and microbiological examination revealed that HzS at these low temperatures was formed by the biological
Table 4.16 Isotopic composition of HzS and associated oil in Devonian reservoir rocks, Alberta, Canada (Thode et al., 1958) 034S (%0) Formation D-2 D-3 Location
--

HzS (%) 0.016 1.6 3.1

HzS +15.5 +12.6 +13.7 +13.7 +12.6 +13.1 +13.9

Oil +11.9 +12.6 +11.0 +10.8 +11.0 +11.7 +13.2

Leduc Leduc Leduc Stettler Stettler Big Valley Bashaw

* R. G. Pankina.

110
0

Stable Isotopes

1500
)WINO
RIVERol

\
E 1000 Q. Q.
~ J.!ot'

. .
0

~
o~ \~ ALBERTAol

en '--' 500

\,
0

0\
0 \ 00
\00 0
0

\
0 -10 0 +10 8345 +20 +30

(%0)

Figure 4.16 Sulphate concentration versus 334Svalues of H2S in formation waters of the Bobrikov horizon, Volga-Urals (Pankina and Mekhtiyeva, 1964), Wind River Basin, Wyoming (Vredenburgh and Cheney, 1971), and Devonian strata of Alberta, Canada (after Krouse, 1977). The lines are drawn to show the trends. Theoretically the data should fit a logarithm function if the kinetic isotope effects were constant throughout the deposit

reduction of S04Zin the formation waters. Significant variations in the isotopic composition of HzS arise from the diversity of concentrations and isotopic composition of S042- as well as the isotopic selectivity and the extent of reduction (Pankina, 1978). As reduction proceeds, the HzS becomes more enriched in 34S (Figure 1.2) and the SO/- concentration decreases. If the system remained closed, the accumulated H2S at 100% reduction would have the same 834Svalue as the initial SO/-. Trends consistent with this concept were found not only in the Volga-Urals (Pankina and Mekhtiyeva, 1964) but also in the Wind River Basin of Wyoming (Vredenburg and Cheney, 1971) and the Devonian of Alberta (Figure 4.16). The large gas deposit at Lacq, France, is found at 150C in Lower Cretaceous-Jurassic strata at a depth of 3300-4300 m. The HzS content is 15.3% and the 834S value of refined sulphur is + 14.5%0 (Grinenko and

Lithospheric Sources of Sulphur

111

Vdovykin, 1966). These authors concluded that the H2S formed by complete reduction of Jurassic sulphates. The Orenburg gas-condensate deposit has been extensively studied (Gavrilov et at., 1973; Pankina, 1978). It is situated on the northern side of the Caspian depression and is associated with Lower Permian-Upper Carboniferous systems. The gas bed is 520 m thick in the central part and the thickness of oil seams is 20 m. The stratal temperature is rather low, from +27 to + 32C. The H2S content changes from 1.3%0 in the upper dome and western part to +4.5% in the east pericline. The (334Svalues increase from +2.0 to +5.8%0 in the same direction. The H2S in this deposit was either formed by SOi- reduction in formation waters of more ancient Carboniferous systems (Pankina, 1978) or in the same strata downdip (Gavrilov et at., 1973). Vredenburgh and Cheney (1971) studied H2S in gases associated with oils down to a depth of 4000 m in the Phosphoria (Permian), Tensleep (Pennsylvanian), and Madison (Mississippian) formations of the Wind River Basin, Wyoming, USA. The (334S values of H2S varied from +0.2 to + 13.2%0. Above 2700 m, the differences between the isotopic compositions of H2S and SOi- were rather significant. The authors concluded that this H2S was generated by bacterial SO/- reduction. Below this depth, the isotopic composition of H2S and S042- were similar. This deeper H2S was attributed to thermochemical processes. Bely and Vinogradov (1972) studied H2S in some deposits of the AmuDaria syncline associated with Upper Jurassic systems. They attributed the isotopically heavy H2S ((334Svalues from +13.4 to +18.2%0) to SO/reduction by organic matter. Orr (1974) considered H2S to have several sources: (a) Microbial reduction of SO i-. This is the main source at temperatures below 60C. (b) Thermal decomposition of organic sulphur compounds in oils, bitumens, and kerogen. The rate of this process increases with temperature and H2S generation may be accompanied by a small isotope selectivity. The quantity and isotopic composition of H2S thus formed depend on the reduced sulphur which enters the organic matter during deposition and diagenesis. (c) Non-microbial SO/- reduction at temperatures above 80C. This may be accompanied by a slight isotopic fractionation. Two other less likely sources are reaction of pre-existing elemental sulphur with organic matter and volcanic emanation (Orr, 1977). Maximov et al. (1975) differentiated two groups of gases in deposits of the Jurassic system of the Amu-Dariya syncline: (a) gases with low

112

Stable Isotopes

concentrations of HzS (0.01-1%) depleted in 34S(-16.0 to -0.7%0) and (b) gases with high concentrations of HzS (1.0-6.5%) enriched in 34S (+ 12.5 to +25.0%0, typical of evaporites). Deposits with high concentrations of HzS are situated at depths from 2500 to 3200 m and temperatures from 95 to 110C. This HzS was attributed to chemical reduction of sulphates. Data for subsequently discovered HzS deposits in the Amu-Dariya syncline (Belenitskaya et ai., 1981) were consistent with the trends reported by Maximov et ai. (1975). In the south of the Caspian depression, the Astrakhan gas-condensate deposit is productive in the Middle carboniferous at a depth interval of 4035-4837 m and temperatures above 100C. The HzS content is as high as 25% and the (334Svalues vary from + 11 to + 13%0. HzS with similar isotopic composition occurs in the Tengiz (+ 12.3%0)and Zhanazhol (+ 11.7%0) deposits situated in the south-east Astrakhan region and on the eastern margin of the Caspian depression respectively (Pankina et ai., 1983). In all of these deposits, HzS is evidently formed by chemical reduction of sulphates. Krouse (1977) summarized data for HzS in Devonian strata of Alberta, Canada. The (334S values for shallow depths and temperatures below 80C varied from + 10 to +20.0%0, whereas in deeper horizons and higher temperatures it ranged from +20 to +25%0 (Figure 4.17). He concluded that HzS in the shallower environments was biogenic, whereas above 80C it formed by thermochemical reduction of sulphate. In deep 'sour gas' (HzSrich) occurrences in Alberta, the HzS content may exceed 90% (Figure 8.2). On the basis of carbon isotope data, Krouse et ai. (1988) concluded that in

+25

- +20
0

,g ~
..

..
H2S

en
+15

"eo

+10

DEVONIAN,ALBERTA

Figure 4.17 Variation of 834Svalues of HzS with reservoir temperature in Devonian strata of Alberta, Canada (after Krouse, 1977)

Lithospheric

Sources of Sulphur

113

some reservoirs in western Canada, HzS was generated through reduction of sulphate by light hydrocarbon gases. The 034S values of HzS dissolved in oil and formation water are similar in a given stratum (Figure 4.18). Free HzS and HzS dissolved in underlying formation water usually have similar isotopic compositions. For instance, 034S values of HzS and water were respectively: +4.0 and +4.7%0 in the Orenburg deposit, +12.1 and +12.6%0 in the Astrakhan deposit, and +14.6 and +14.1%0 in the UrtabulakDengizkul deposit. Occasionally, wide variations are found for the dissolved species (Table 4.17). Figure 4.19 shows that the isotopic composition of HzS in strata of the same age in the same area is highly dependent upon HzS concentration. Negative and positive 034Svalues were found to be associated with low and

CHANGIRMASH
w Z w

0 w <I -.J 0:

<->

PALVENTISH

J
:::>

PALVENTISH

w lL Z a::0 wCD a.a::

0 a::

a...: :::>u

11 CIZRANS7

II
0..

KALlNOVSKOYE

H2S

-10

+10

+20

834S (%0)
Figure 4.18 Relationship between 834Svalues of H2S dissolved in oil and water in
different regions of the Soviet Union

AMU-DARY A SYNECLISE
Fields Bulchara of terrace
~

(J)
Fields Chardzhou of terrace 2I 54
H2S

.........j::.

'#.
(f)
.-4

I'6

H.S

. -5
0

-15

-10

+5

+10

RUSSIAN PLATE, CASPIAN DEPRESSION


Fields of middle area Voiga

C2-P, 'i
~

Orenburg field

(f) .. I

2L504 f

:~ ~~2~
-5 0
+ 20 8345,(%0)

Fields

of middle

Astrakhan
9...,- C 2

swell t

Voiga area

"*

I
25
20 16 10

SO24

I
I +6 I +10 I +16 I +20 834S,(%0)
I.

H2S
1 -16 I' -10
11111111111111111111111111

J4
0

-6

UTIIIIII 1

~2

-13

1 Concentration of H25 < 2 % 2 Concentration of H25 >2% 3 Limits of variation of 8345 in sulphates
Figure 4.19 Relation of the isotopic composition of H2S to concentrations in gas accumulations in the Amu-Darya syneclise and the Russian Plate (Pankina and Mekhtiyeva, 1981)

...... ...... Ut

116

Stable Isotopes

Table4.17 Isotopiccomposition of H~Sdissolved in water(Pankina,1978)


Number of determinations Variations of (')345values (%0) Average (')345 value (%0)
----

Age of deposit

Region

Paleogene Jurassic Permian Carboniferous

Middle Asia Middle Asia and Volga-Urals Middle Asia and Volga-Urals Middle Asia and Volga-Urals

11 12 2 15

-11.3 to +18.2 -17.0 to +4.7 -9.8 to -1.9 -8.9 to +11.8

+1.8 -1.4 -5.5 +3.0

high concentrations respectively. In the case of the latter, mean values of +13, +5, and +18%0 can be assigned to Carboniferous, Permian, and Jurassic strata. This arises because H2S generated by thermochemical reduction of evaporites is consistent with the oceanic (')34S-timecurve (Figure 4.2). Estimation of the worldwide average (')34Svalue for H2S in oil- and gasfields must consider the following: (a) Higher concentrations of H2S tend to have (')34Svalues slightly higher than associated evaporites (Figure 4.19). (b) The average (')34Svalue for evaporites has been estimated at + 19%0 (Section 4.2). (c) H2S is not uniformly associated with evaporites of all ages. (d) Minor occurrences of H2S usually have negative (')34S values. It is difficult to quantitatively incorporate all of these factors. A qualitative estimate for the mean value is + 15%0. 4.6 FLUX OF VOLCANOGENIC SULPHUR TO THE ATMOSPHERE AND ISOTOPIC COMPOSITION OF TOTAL SULPHUR*

4.6.1 Volcanogenic sulphur flux to the atmosphere The sulphur flux to the atmosphere from the lithosphere during surface volcanism was estimated at 28 Tg S yr-I (Lein, 1983). This value is higher by one order of magnitude than those published previously (Kellogg et at., 1972; Friend, 1973; Cadle, 1975; Granat et at., 1976). The estimate of Lein
* A. Yu. Lien.

Lithospheric Sources of Sulphur

117

(1983) considered not only eruptions but also gas emissions during fumarolic and solfataric stages. This approach, described by Ivanov (1981), was made possible by direct determinations of sulphur concentrations in fumarole and solfatara gases and their fluxes to atmosphere (Stoiber and Jepsen, 1973; Okita and Shimozuru, 1975; Berlyand, 1975). As the result of improved instrumentation, many direct determinations of SOz, HzS, COS, and S03 fluxes to the atmosphere from volcanoes have been recently reported. A review of these data (Tables 4.18 and 4.19) leads to two basic conclusions: (a) Volcanic activity during the past two decades has appreciably increased. Since 1961, many large and disastrous eruptions were recorded. Up to 13.4 X 105 tonnes of SOz were transported to the stratosphere during the eruption of El-Chichon in Mexico (Evans and Kerr, 1983). (b) Direct determinations of the sulphur flux from more than 40 volcanoes with fumarole-solfatara activity (compared to 11 reported previously) confirmed the previous estimate of the average SOz emission from fumaroles of an individual volcano, i.e. 245 tonnes SOz day-l (Lein, 1983). The daily average emission per volcano was estimated at 275 tonnes by Berresheim and Jaeschke (1983). These investigators concluded, as did Lein (1983), that the sulphur flux from unstable gas jets of volcanoes is at least
Table 4.18 Volcanogenic sulphur flux during eruptions S02 (tonnes d-l) 3740 1000--5000 1600 2000 10 000 55 000 1500 1280 86 400 86 400 60 000 1500 1 340 000 5888 10 000 References
..------

Volcano Etna Etna Etna Cerro Negro Fuego Fuego Santiaguito Kilauea St Augustine St Augustine St Helens Pacaya EI-Chichon Tolbachik Ardukoba

Haulet et al. (1977) Malinconico (1979) Jaeschke et al. (1982) Taylor and Stoiber (1973) Rose et al. (1973) Crafford (1975); Rose et al. (1982) Stoiber and Jepsen (1973) Naughton et al. (1975) Stith et al. (1978) Stith et al. (1978) Evans and Kerr (1983); Casadevall et al. (1981); Hobbs et al. (1981) Cadle et al. (1979) Evans and Kerr (1983) Abramovsky et al. (1977) Allard (1980)

118

Stable Isotopes

Table 4.19 Sulphur flux to the atmosphere from fumaroles


Volcano Japan Mihara Osima Asama Asama Guatemala Santiaguito Fuego Fuego Fuego Pacaya Pacaya Nicaragua Telika Momotombo San-Kristobal Masaya USA Kilauea Kilauea Mauna-Vlu Sulphur Banks St Augustine St Augustine Martin Mountain State Washington Mt St Helens ' Mt St Helens Mt St Helens Italy Etna Etna Etna Etna Etna Karymsky SOz (tonnes d-l) References

345 345 142 787 420 40 423 393 260 300 20 50 360 180 280 200 30 7 432 86 3 600-3600 10-50 1000-1900 900 3325 1130 1000 142 955-1600 173

Berresheim and Jaeschke (1983) Okita and Shimozuru (1975) Okita and Shimozuru (1975) Okita and Shimozura (1975) Stoiber and Jepsen Stoiber and Jepsen Crafford (1975) Rose et al. (1982) Stoiber and Jepsen Stoiber and Bratton (1973) (1973) (1973) (1978)

Stoiber and Jepsen (1973) Stoiber and Jepsen (1973) Stoiber and Jepsen (1973)

Naughton et al. (1975) Harding and Miller (1982) Stoiber and Malone (1975) Stoiber and Malone (1975) Stith et al. (1978) Stith et al. (1978) Casadevall et al. (1981) Hobbs et al. (1981) Casadevall et al. (1981) Hobbs et al. (1981) Hobbs et al. (1981) Zettwoog and Haulet (1978) Zettwoog and Haulet (1978) Malinconico (1979) Jaeschke et al. (1982) Jaeschke et al. (1982) Berlyand (1975)

one order of magnitude higher at the fumarolic stage than during eruptions. Proceeding from this conclusion, these authors obtained a total annual flux to the atmosphere of 11.8 Tg S including 7.6 Tg S as S02 (15.2 Tg S02 yr-I). Of this amount, 14.2 Tg S02 are released annually from fumaroles and only 1 Tg S02 during eruptions.

Lithospheric Sources of Sulphur

119

The value obtained by Berresheim and Jaeschke (1983) is almost three times lower than the current estimate. This can be explained by the different numbers of active volcanoes assumed in the calculations. We considered 578 active volcanoes (Vlodavets, 1966) with average daily emissions of 245 tonnes S02 (or 0.092 million tonnes annually) from the fumarolic area of each volcano. Berresheim and Jaeschke considered 520 volcanoes but used only 100, with an average emission of 275 tonnes d-I per volcano. This number seems low since they classified 365 volcanoes as very active. If, from a total of 578 volcanoes, one subtracts 55 that erupt annually and 100 with low daily fumarolic emissions (10-50 tonnes S02)' 423 remain with individual average emissions of 275 tonnes S02 d-I. Their global emission would be 42.8 Tg S02 or 21.4 Tg S ycl, which is close to our estimation. Therefore, from recent estimates, the flux of volcanic sulphur to the atmosphere varies from 11.8 to 28 Tg S ycl. It should be emphasized that: (a) Not all sulphur compounds such as H2S, COS, and S03 are included since their measurement is difficult (Hobbs et at., 1981). (b) The sulphur flux during large eruptions may be underestimated by an order of magnitude. (c) Most data refer to flux to the stratosphere rather than to lower layers of the atmosphere. Therefore, the higher estimate of 28 Tg S ycl is favoured.

4.6.2 Isotopic composition of volcanogenic sulphur Sulphur occurs in different valence states in volcanogenic emissions: H2S and its derivatives, So, S02> and H2SO4 (acid and its salts). Sulphur enters the atmosphere usually in the form of gaseous compounds (S02 and to a lesser extent H2S, S03, COS) as well as solid particles which may contain sulphur minerals and absorbed sulphur compounds. The isotopic composition of individual sulphur compounds depends on the stage of volcanogenic activity, type of eruption, temperature, kinetic istope effects and exchange during redox reactions, valence, and contamination of the magmatic source by crustal sulphur. In summary: (a) At high temperatures and pressure, isotopic equilibrium between some sulphur compounds is established rather quickly. The S02 and H2S are enriched and depleted in 34S respectively, compared to total sulphur. (b) The differences in isotopic compositions of the various molecules (1l834S values) decreaseas the temperature increases. (c) The greater solubility of S02 compared to H2S favours the enrichment of 32S in the gas phase and 34S in the water phase.

120

Stable Isotopes

4.6.2.1 Isotopic composition of gases emitted to the atmosphere during eruption The chemical and isotopic compositions of evolved gases were measured during a rift eruption (7-14 November 1978) of basalt magma near Ardukoba volcano, Azal rift, Ethiopia (Allard, 1980). Four samples taken immediately above the magma in the eruptive rift at 1070 C were found to have consistent 034Svalues near -3%0. Two samples from lava flows near 800C had 034Svalues near 0%0whereas a third sample from the same environment was -4.8%0. The 034S value of the total sulphur in gases (mainly SOz), evolved from basalt lavas in two eruptions of the Kiluaea volcano (1971 and 1974), varied from +0.3 to +5.4%0, averaging +0.9%0. These values approximate those of total sulphur entrapped in the basalt (Sakai et at., 1982). Degassing is peculiar to subaereal basalt eruptions (Moore and Fabbi, 1971; Sakai et at. 1982). Assuming that the primary basalt magma contained 1100 ppm sulphur, then a sulphur concentration of 150 ppm after degassing means that nearly 90% of the total sulphur was removed from the magma as SOz. The sulphur flux to the atmosphere from the eruption of Ardukova is calculated from the portion of sulphur released (950 ppm) and the amount of basalt discharge (43 x 106tonnes). This corresponds to 1.1 x 104tonnes SOz d-I. This is higher than the average value of SOz flux to the atmosphere from Mount Etna (ct. Haulet et at., 1977; Malinconico, 1979). Therefore, at rift discharges of basalt magmas, nearly 6 x 103 tonnes of sulphur may enter the atmosphere per day, the 034S value being close to that of meteoritic sulphur. 4.6.2.2 Isotopic composition of ash emitted to the atmosphere during eruptions Volcanic eruptions are accompanied as a rule by the discharge of ash to the atmosphere. The volume and the height of the rise of this material depend on the type and scale of the eruption (Table 4.20). The sulphur isotopic compositions of mineral particles in ash emitted with gases and thermal waters during the eruption of Ontake (Japan) on 8 October 1979 were studied by Kusakabe et at. (1982). Ontake is a Quaternary andesite stratovolcano in the central parot of Honshu Island, similar to andesite volcanoes of Kamchatka and Kuril Islands. It was a weak eruption (t = 110-185 c) lasting less than a day. The volcanic dust sampled soon after the eruption contained pyrite, anhydrite, and elemental sulphur. The 034S values of pyrite and native sulphur varied from -1.8 to -6.4%0, the average value for 18 samples being -4.8%0. The 034S values of anhydrite

Lithospheric Sources of Sulphur

121

Table 4.20 Classification of volcanic activity (Berresheim and Jaeschke, 1983) Rise (km) 1 1 Max. 2 2-7 5-10 10-20 20-40 40 S02 (tonnes d-I) 103 103 300-103 104 7 x 104 1-2 x 105 1-2 x 105 Frequency of activity per year"

Type A -1 A- 2 A- 3 A -4 A -5 A -6 A -7 A -8 Iceland Hawaiian Strambolian Vulcanian Vexuvian Plinian Pelean Krakatao

Explosive index" 0-5 0-5 5-20 20-40 40-60 60-80 80-99 99-100

Compositionh a, ab a ab ab, b ab, b b, bc b, bc b, bc

10.46 760.00 9.40 4.13 0.53 0.12

U Explosive index = mass of solid material/total mass of solid and liquid material (Rittmann. 1981). b Composition of products: a---dominance of lava; b--prevalence of fresh solid material; cdominance of ancient volcanic rocks. (' Average number of erupting volcanoes 55 per year; data for the period from 1961 to 1979.

fell within the range +8.3 to + 14%0, the average being + 10.8%0 (seven samples) . Soluble sulphate on the surface and total S extracted with Kiba reagent from ash emitted by the explosion of Mt St Helens, Washington, USA, on 18 May 1980 were found to have 334S values near +9%0 (H.R. Krouse, unpublished data). This suggests that the surface sulphate resulted from high-temperature oxidation of sulphur-containing minerals in the ash. The above data, based on fresh samples, suggest that the 334Svalues of the total sulphur in volcanic ash discharges tends to be positive. Positive 334Svalues could arise if the sulphur is magmatic (near 0%0)and the heavier isotopes are favoured in the condensed phases. However, the sulphur may have other than a magmatic origin. In the case of Mt St Helens, the northern flank was blown away to the extent of 4 km3 and a decrease in elevation of 600 m. Therefore during an explosive eruption, a variety of geological strata with different 334Svalues might be pulverized. 4.6.2.3 The sulphur isotope composition of gaseous emissions from fumaroles and solfataras It is well known that the concentration and isotopic composition of sulphur in fumaroles (t = 200-1000 c) and solfataras (t = 100-200C) are rather

variable and depend primarily on the state of the volcanic system in which they are found.

122

Stable Isotopes

Table 4.21 contains the phenomenological classification of non-eruptive volcano activities and their duration. Evidently, an important contribution of sulphur to the atmospheric budget might be expected only from the last two categories including fumarolic and solfatara activity over the time scales of months to centuries. The isotopic composition of high-temperature fumarolic gases approximates the meteorite value (Nakai and Jensen, 1967; Grinenko and Grinenko, 1974). Recently, data were reported for long-term observations of the chemical and isotopic compositions of high-temperature gases emitted from the hottest fumarole, A-1,in Showashinzan volcano on Hokkaido (Mizutani and Sugiura, 1982). The 834SI01 was +2%0 whereas the isotopic composition of SOz varied within the range +0.3 to +5.6%0 (Table 4.22). The above authors suggested that temporal changes were caused by increased infiltration of surface waters and atmospheric gases, as well as shifts in chemical reactions with decreased temperature and pressure. In a complex geochemical study of fumarolic gases from volcanoes of the Satsuma Iwo-Jima group (Matsubaya et ai., 1975), 834Svalues for HzS were found to vary within the range +8.9 to +11.2%0 (Table 4.22). This was explained by fractionation of isotopes during chemical reactions in gas jets. The above data show that 834Svalues for all sulphur compounds studied in high-temperature gases (600-850 c) fall within the range -13 to + 12%0 with an average of +5.6%0. It is generally accepted that sulphur compounds in moderate- and lowtemperature fumarole and solfataras gases have a more diverse isotopic

Table 4.21 Classification of volcanic activity between eruptions (Berresheim and Jaeschke, 1983) SOz (tonnes d-J) Number of volcanoes

Stage B-1 Pre eruptive B-2 Intraeruptive

Duration Days-weeks Hours-days

Type Rhythmical gas evolutions before eruption Gas evolutions between paroxysmal eruptions Fumarole activity Fumarole activity and solfatar

600-10'

55

B-3 Posteruptive B -4 Extraeruptive

Weeks-years Years-centuries

500-10' 50-500 10-50

365 100

Lithospheric Sources of Sulphur Table 4.22 Isotopic composition of sulphur compounds in gases of hightemperature fumaroles 034S (%0) Volcano Showashinzan, A -1 Date of sampling (ReferenceO) (1) (2) 8.1960 8.1961 8.1962 10.1973 9.1974 8.1977 11.1977 10.1978 Showashinzan, B-5 Showashinzan, C-2 Showashinzan, A -6 Nasudake, M-1 Satsuma, Iwo-lima (3) (3) (3) (3) (3) (3) (3) (3) (3) (3) (3) (3) (4) (4) (4) (4) (5) (5) (5) (3) Temperature of gas Cc) 800 661 559 483 722 703 635 600 588 642 617 554 446 605 695 489 835 751 732 751 700 480 500
--

123

H2S

S02 -4.8 -1.8 +0.4 -1.3 +4.4 +0.3 +1.8 +5.2 +5.0 +5.6 +4.8 +3.9 +9.3 +7.1 +5.5 +4.4 +11.6 +11.7 +12.2 +11.8 +12.2 +8.3 +3.7 +10.6

Sulphur

-8.0 -1.7 -11.0 +4.4 +5.1 -0.7 +9.9 +11.2 +8.9 +9.3 + 1.4 -0.9 -12.7 +7.3

+4.4 +0.3 +1.4 +1.9 +4.6 +6.8 +5.5 +1.9 +12.1 +12.1 +12.2 +11.5

Bezymyanny Avachinsky Mutnovsky Kuyu

8.1974 8.1974 9.1974 9.1974 1967 Ku-5

+6.9

" (1) Sakai (1957). (2) Sakai and Nagasawa (1958). Mizutani and Sugiura (1982). (4) Matsubaya el al. (1975). (5) Vinogradov (1980).

composition than in high-temperature cases (compare Figure 4.20 and Table 4.22). Enrichments in 34S, up to +20%0, are observed for sulphate sulphur from some solfataras and volcanic thermal springs. The reasons for such enrichmen.ts are unclear but might be explained by chemical reactions (Sakai, 1957; Rafter et al., 1958; Grinenko and Lein, 1967; Taylor and Stoiber, 1973; Grinenko and Grinenko, 1974) and in certain cases (Mendeleev volcano, Kunashir Island; Cape Reykjanfs, Iceland) by the involvement of isotopically heavy seawater sulphates (Vinogradov, 1964, 1966, 1970; Hubbertenet at., 1975; Sakai and Matsubaya, 1977).

124

Stable Isotopes

SHIRANE SHIVELUCH CRATER OF UZON EBEKO REVUSHCHAYA

GULOVNINAI
KAMBALNY GEYSER VAL MUTINOVSKY.
A V ACHINSKY.

KUYU
NASUDAKE SHOWASHIN: NEA KAMENI
I

ISI H,S
IQJ

SO

II S02
0 S042-

. TOTAL

1-10

+10
8348(%0)

+20

Figure 4.20 Sulphur isotope composition of different compounds in medium (20O-400C; data above line) and low (50--200C; data below line) temperature fumaroles and soIfataras

4.6.3 The isotopic composition of total volcanic sulphur entering the atmosphere Numerous attempts were made to calculate average 034Svalues for volcanic sulphur. Using data from six fumaroles in the White Island volcano, Rafter et at. (1958) obtained an average of +2.5%0. A similar value, +2.2%0, was obtained by Borisov (1970) who applied statistical mathematics to 267 isotopic analyses of various sulphur compounds. When these data are combined with those for gases emitted from basaltic magma (Allard, 1980; Sakai et at., 1982), it would appear the total volcanic sulphur to the atmosphere during eruptions has a 034Svalue near zero.

Lithospheric Sources of Sulphur Table 4.23 Sulphur isotope balance for volcanic emissions Total S (106 tonnes yr-1) 1.45 26.6 ? 28

125

Activity Eruptions Fumarole activity Ash Total

()34S(%0) 0 +5 +10 +5

The quantity of sulphur entering the atmosphere annually during eruptions is 1.45 Tg S (Lein, 1983). To this should be added 26.6 Tg S yc1 emitted from fumaroles and solfataras with an isotopic composition equal to + 5%0 (Table 4.23). The quantity of ash sulphur reaching the troposphere is difficult to estimate. During more intensive eruptions (e.g. Plinian, Pelean, Krakatoa), huge incandescent clouds of ash were discharged. One can affirm only that the basic forms were oxidized sulphur compounds. The amount of ash released from EI-Chichon, Mexico, in March-April 1982 was comparable to the Krakatoa eruption in 1884. On the basis of few measurements, surface secondary sulphate minerals were found to have a 034Sof about + 10%0. In summary, the total volcanic sulphur entering the atmosphere is enriched in 34S compared to the meteorite value, with an average of 034S value of about + 5%0. REFERENCES
Abramovsky, B.P, Alimov, V.H., Ionov, V.A., Nazarov, I.M., Patrakeev, c.1., Fedatov, C.A., and Chirkov, V.A. (1977). The ejection of gases and aerosols to the atmosphere during the eruption of Tolbachik Volcano, Kamchatka (in Russian). Dokl. AN-SSSR, 237 (6), 1479-82. Allard, P. (1980). Caracteristiques geochimiques des volatils emis par l'eruption volcani que de novembre 1978 dans Ie rift d'Asal. Bull. Soc. Geol. France, 22 (6), 825-9. Ault, W.U., and Kulp, J.L. (1959). Isotopic geochemistry of sulphur. Geochim. Cosmochim. Acta, 16 (4), 201-35. Barnes, H.L. (Ed.) (1979). Geochemistry of Hydrothermal Ore Deposits, 2nd edn, Wiley-Interscience, New York, 798 pp. Belenitskaya, G.A., Golubchina, M.V., Gurevich, M.S., and Mishina, T.A. (1981). Isotopic composition of sulphur and genesis of HzS in natural gases of AmuDaria basin (in Russian). Litologia i poleznye Iskopaemye, 2, 118-38. Bely, V.M., and Vinogradov, V.I. (1972). Isotopic composition of sulphur and genesis of high concentrated HzS of gas and oil bearing fields (in Russian). Geolog. nefti i gaza, 2, 37-41. Berlyand, M.E. (1975). RecentProblemsin Atmospheric Diffusion and Pollution (in Russian). Gydrometeoisdat, Leningrad, 448 pp.

126

Stable Isotopes

Berresheim, H., and Jaeschke, W. (1983).The contributionof volcanoes to the global atmospheric sulphur budget. J. Ceophys. Res., 88 (66), 3732-40. Birnbaum, S.J., and Coleman, M. (1979). Source of sulphur in the Ebro Basin (Northern Spain) Tertiary nonmarine evaporite deposits as evidenced by sulphur isotopes. Chem. Ceol., 25, 163-8. Bodenlos, A.J., and Nelson, c.P. (1979). Sulfur. Econ. Ceol., 74 (2), 459-61. Bomsov, O.G. (1970). Source of native volcanogenic sulphur. Ceokhimiya, 963, 332-43. Burke, K. (1975). Atlantic evaporites formed by evaporation of water spilled from Pacific, Tethyan and Southern Ocean. Ceology, 3, 613--16. Cadle, R.D. (1975). Volcanic emissions of halides and sulphur compounds to the troposphere and stratosphere. 1. Ceophys. Ress., 80 (12), 1650--2. Cadle, R.D., Lazrus, A.L., Huebert, B.J., Heidt, L.E., Rose, W.J., Woods, D.C., Chuan, R.L., Stoiber, R.E., Smith, D.E., and Zielinski, R.A. (1979). Atmospheric implications of studies of Central American volcanic eruption clouds. J. Ceophys. Res., 84 (C11), 6961-8. Casadevall, T.J., Johnston, D.A., Harris, D.M., Rose, W.J., Malinconico, L.L., Stoiber, R.E., Bronhorst, R.J., Williams, S.N., Woodruff, L., and Thompson, J.M. (1981). S02 emission rates at Mount St Helens from March 29 through December 1980. Ceol. Surv. Prof. Pap., 1250, 193-200. Cecile, M.P., Shakur, M.A., and Krouse, H.R. (1983). The isotopic composition of western Canadian barites and the possible derivation of oceanic sulphate 834S and 8180 age curves. Can. J. Earth Sci., 20, 1528--35. Claypool, G.E., Holser, W.T., Kaplan, 1.R., Sakai, H., and Zak, 1. (1980). The age curves of sulfur and oxygen isotopes in marine sulfate and their mutual interpretation. Chem. Ceol., 28, 199-260. Cortecci, G., Reyes, E., Berti, G., and Casati, P. (1981). Sulphur and oxygen isotopes in Italian marine sulphates of Permian and Triassic ages. Chem. Ceol., 34, 65-79. Crafford, T.C. (1975). S02 emission of the 1974 eruption of Volcano Fuego (Guatemala). Bull. Volcanol., 39(4), 536--56. Davis, J.B., and Kirkland, D.W. (1970). Native sulfur deposition in the Castile formation, Culberson County, Texas. Econ. Ceol., 65, 107-21. Davis, J.B., and Kirkland, D.W. (1979). Bioepigenetic sulfur deposits. Econ. Ceol., 74 (2), 462-8. Dessau, G., Jensen, M.L., and Nakai, N. (1962). Geology and isotopic studies of Sicilian sulfur deposits. Econ. Ceol., 57, 410--38. Eremenko, N.A. (1960). Variation of the sulphur isotope composition in U.S.S.R. crude oils according to stratigraphic sequence (in Russian). Ceol. Nefti i. Caza, 4 (11), 9-10. Eremenko, N.A., and Pankina, R.G. (1962). Sulphur isotopes in gas and oil of Volgo-Ural deposits and other regions of the Soviet Union (in Russian). Ceol. nefti i gaza, 9, 43-8. Eremenko, N.A., and Pankina, R.G. (1963). Variations in the isotopic composition of sulphur in gas and oil with the age of reservoir rocks. Chemistry of the earth crust (in Russian). Nauka, 1, 424-31. Evans, W.F.Y. and Kerr, J.B. (1983). Estimates of the amount of sulphur dioxide injected into the stratosphere by the explosive volcano, Mt. St. Helens. Ceophys. Res. Left., 10 (11), 1049-51. Feely, H.W., and Kulp, J.L. (1957). Origin of Gulf Coast salt dome sulfur deposits. Am. Assoc. Petrol. Ceologists Bull., 41, 1802-53.

Lithospheric Sources of Sulphur

127

Friend, J.P. (1973). The global sulphur cycle. In: Rossol, S.J. (Ed.) Chemistry of the Lower Atmosphere, Plenum Press, New York, pp. 177-201. Gavrilov, E. Ya., Grinenko, V.A., Zhurov, Yu. A., et al. (1973). Peculiarities of the distribution of argon, sulphur and carbon isotopes in gases of Orenburg gas condensed deposit (in Russian). Geolog. nefti i gaza, 8, 26-31. Goles, G.G. (1969). Cosmic abundances. In: Wedepohl, H. (Ed.) Handbook of Geochemistry, Vol. I, Springer, Berlin, Heidelberg, New York, pp. 116-33. Granat, L., Rodhe, H., and Hallberg, R.O. (1976). The global sulphur cycle. Ecol. Bull. (SCOPE Report 7), 22, 89-134. Grinenko, V.A., and Grinenko, L.N. (1974). Geochemistry of Sulphur Isotopes (in Russian). Nauka, Moscow, 272 pp. Grinenko, V.A., and Lein, A. Yu. (1967). Variations of the isotopic composition of various sulphur compounds in crater lake ore deposits, Telaga Bodas, Indonesia (in Russian). Geokhimiya, 3, 290-5. Grinenko, V.A., and Vdovykin, G.P. (1966). Isotopic composition of sulphur of gas and oil bearing deposit Lucq (South-West France) (in Russian). Geokhimiya, 3, 351-3. Harding, D., and Miller, J.M. (1982). The influence on rain chemistry of the Hawaiian volcano Kilauea. 1. Geophys. Res., 87 (C2), 1225-30. Haulet, R., Zettwoog, P., and Sabroux, J.e. (1977). Sulphur dioxide discharge from Mount Etna. Nature, 268 (5622),715-17. Hobbs, P.V., Radke, L.F., Eltgroth, M.W., and Hegg, D.A. (1981). Airborne studies of the emissions from the volcanic eruptions of Mount St. Helens. Science, 211 (4484),816-18. Hoefs, J. (1981). Isotopic composition of the ocean-atmospheric system in the geologic past. In: Evolution of the Earth, Vol. 5, Geodynamics Series, pp. 110-19. Holland, H.D. (1973). Systematics of the isotopic composition of sulphur in the oceans during the Phanerozoic and its implications for atmospheric oxygen. Geochim. Cosmochim. Acta, 37, 2605-16. Holser, W.T. (1977). Catastrophic chemical events in the history of the ocean. Nature, 267 (5610), 403-8. Holser, W.T. (1979). Rotielgend evaporites, Lower Permian of Northwestern Europe. Geochemical confirmation of the non-marine origin (in German). Erd61 und Kohle-Erdgas-Petrochemie vereinigt mit brennstoffchemie, 32 (4), 159-62. Holser, W.T., and Kaplan, 1.R. (1966). Isotope geochemistry of sedimentary sulphates. Chem. Geol., 1 (2), 93-135. Holser, W.T., Kaplan, 1.R., Sakai, H., and Zak, 1. (1979). Isotope geochemistry of oxygen in the sedimentary sulphate cycle. Chem. Geol., 25,1-17. Hubberten, H.W., Nielsen, H., and Puchelt, H. (1975). The enrichment of 34S in the solfataras of the Nea Kameni volcano, Santorini Archipelago, Greece. Chem. Geol., 16 (3), 197-205. Ivanov, M.V. (1981). The global biogeochemical sulphur cycle. In: Likens, G.E. (Eds.) Some Perspectives of the Major Biochemical Cycles, Wiley, Chichester, pp. 61-78. Ivanov, M.V. (1983). Major fluxes of the global biogeochemical sulphur cycle. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE 19, Wiley, pp. 449-63. Jaeschke, W., Berresheim, H., and Georgii, H.W. (1982). Sulphur emissions from Mt. Etna. J. Geophys. Res., 87 (C9), 7253-61. Kavcic, R. (1958). Study of the concentration of sulfur isotopes in high sulfur coals (in Russian). Vestn. Slov. Kem. Drus., 5, 7-11.

128

Stable Isotopes

Keil, K. (1969). Meteorite composition. In: Wedepohl, H. (ed.) Handbook of Geochemistry, Vol. I, Springer, Berlin, Heidelberg, New York, pp. 116-33. Kellogg, W.W., Cadle, R.D., Allen, R.R., Lazrus, A.L., and Martell, E.A. (1972). The sulphur cycle. Science, 175 (4022),587-96. Klaus, W., and Pak, E. (1974). Neue Beitriige zur Datierung von Evaporiten des Ober-Perm (in German). Carinthia, 164 (84), 79-85. Krouse, H.R. (1977). Sulphur isotope studies and their role in petroleum exploration. J. Geochem. Explor., 7, 189-211. Krouse, H.R. (1980). Sulphur isotopes in our environment. In: Fritz, P., and Fontes, J. Ch. (Eds.) Handbook Environmental Isotope Geochemistry. The Terrestrial Environment, Vol. 1, Elsevier, Amsterdam, pp. 435-71. Krouse, H.R., Viau, C.A., Eliuk, L.S., Veda, A., and Halas, S. (1988). Chemical and isotopic evidence of thermochemical sulphate reduction by light hydrocarbon gases in deep carbonate reservoirs. Nature, 333 (6172), 415-19. Kusakabe, M., Mizutani, Y., and Kometani, M. (1982). A preliminary stable isotope study of volcanic ashes discharged by the 1979 eruption of On take Volcano, Nagano, Japan. Bull. Volcanol., 45 (3),203-9. Lein, A. Yu. (1968). Mineral composition of sulfur ores in some volcanic deposits of Kamchatka (in Russian). Tr. Geol. Inst. Kazan, 20, 80-90. Lein, A. Yu. (1983). The sulphur cycle in the lithosphere: Part II, cycling. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE 19, Wiley, New York, pp. 95-129. Lloyd, R.M. (1967). Oxygen-18 composition of oceanic sulphate. Science, 156, 1228-31. Malinconico, L.L. (1979). Fluctuations in SOz emission during recent eruptions of Etna. Nature, 278 (5699), 43-5. Matsubaya, 0., Veda, A., Kusakabe, M., Matsuhisa, Y., Sakai, H., and Sasaki, A. (1975). An isotopic study of the volcanoes and the hot springs in Satsuma Iwo Jima and some areas in Kyushu. Bull. Geol. Surv. Japan, 26 (8), 375-92. Maximov, S.P., Pankina, R.G., Smakhtina, A.M., et al. (1975). Genesis of HzS in gases of Amu-Daria syneclise (by isotopic composition of sulphur) (in Russian). Proc. VNIGNI, 174, 122-35. Migdisov, A.A., Ronov, A.B., and Grinenko, V.A. (1983). The sulphur cycle in the lithosphere. Part 1, reservoirs. In: Ivanov, M.V. and Freney, J.F. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE 19, New York, pp. 25-95. Mizutani, Y., and Sugiura, T. (1982). Variations in chemical and isotopic compositions of fumarolic gases from Showashinzan Volcano, Hokkaido, Japan. Geochem. 1., 16 (2), 63-71. Moore, J.G., and Fabbi, B.P. (1971). An estimate of the juvenile sulphur content of basalt. Cont. Mineral. Petrol., 33 (2), 118-27. Morrow, D.W., Krouse, H.R., Ghent, E.D., Taylor, G.c., and Dawson, K.R. (1978). A hypothesis concerning the origin of barite in Devonian carbonate rocks of northeastern British Columbia. Can. J. Earth Sci., 15, 1391-406. Muller, G., Nielsen, H., and Ricke, W. (1966). (a) Schwefel-Isotopen-aVerhiiltnisse in formations-Wassern undo Evaporiten Nord-und Siiddeutschlands (in German). Chemical Geol., 1, 211-20. Nakai, N., and Jensen, M.L. (1967). Sources of atmospheric sulphur compounds. Geochem. J., 1, 199-210. Naughton, H.J., Lewis, V., and Thomas, D. (1975). Fume compositions found at various stages of activity at Kilauea Volcano Hawaii. J. Geophys. Res., 80 (21), 2963-6.

Lithospheric Sources of Sulphur

129

Nielsen, H. (1965). Schwefelisotope im marinen Kreislauf und das (')34S der friiheren Meere (in German). Geol. Rund., 55, 160-72. Nielsen, H. (1967). Sulphur isotopes in the Rhine graben evaporite sulphates (in German). The Rhinegraben Progress Report, pp. 27-31. Nielsen, H. (1972). Sulphur isotopes and formation of evaporite deposits. In: RichterBernberg, H. (Ed.) Geology of Saline Deposits, Proceedings of Hann. Symposium, pp. 91-102. Nielsen, H. (1973). Model estimates of sulphur isotope balance in ancient oceans (in Russian). In: International Geochemical Congress, 1st Proceedings, Sedimentary Processes, Vol. 4, No.1, VINITI, Moscow, pp. 127-40. Nielsen, H. (1974). Isotope composition of the major contributors to atmospheric sulfur. Tellus, 26, 1-2, 213-21. Nielsen, H. (1978). Sulphur isotopes in nature. In: Wedepohl, H. (Ed.) Handbook of Geochemistry, Vol. II, Springer, Berlin, Heidelberg, New York, Ch. 16-B, pp. 1-40. Nielsen, H., and Ricke, W. (1964). Schwefel-Isotopenverhaltnisse von Evaporiten aus Deutschland; ein Beitrag zur Kenntnis von (')34Sin Meerwasser-Sulfat. Geochim. et Cosmochim. Acta, 28 (5), 577-91. Nriagu, J.O. (1978). Determinative effects of sulfur pollution on materials. In: Nriagu, J.O. (Ed.) Sulfur in the Environment, Part 1, Wiley, New York, pp. 1-59. Odin, G.S., and Kennedy, W.J. (1982). Geochimie et geochronologie isotopiques. C.R. Acad. Sci. Paris, 294, 383-6. Okita, T., and Shimozuru, D. (1975). Remote sensing measurements of mass flow of sulphur dioxide gas from volcanoes. Bull. Volcanol. Soc. Japan, 19 (3), 151-7. Orr, W.L. (1974). Changes in sulphur content and isotopic ratios of sulfur during petroleum maturation-study of Big Horn Basin Paleozoic oils. Bull. Am. Assoc. Petrol. Geologists, 58, 2295-318. Orr, W.L. (1977). Geologic and geochemical controls on the distribution of hydrogen sulfide in natural gas. In: Campos, R., and Goni, J. (Eds.) Advances in Organic Geochemistry, 1975, Enadisma, Madrid, Spain, pp. 571-97. Pak, E. (1974). Schwefelisotopenuntersuchungen am Institut fiir Radiumforschung und Kernphysik. I. Anz. Osterr. Akad. Wiss., Math. Naturwiss. KL, 1974, 166-74. Pankina, R.G. (1978). Geochemistry of Sulphur Isotopes of Oils and Organic Matter (in Russian). Nauka, Moscow, p. 247. Pankina, R.G., and Mekhtiyeva, V.I. (1964). Isotopic composition of H2S of associated gases from Bobrikov horizon in Volgo-Ural region (in russian). Geokhimiya, 9, 866-71. Pankina, R.G., and Mekhtiyeva, V.L. (1981). Origin of H2S and CO2 in hydrocarbon accumulations as indicated by isotope data (in Russian). Geologica Nefti i Gaza, 12, 44-8. English translation in International Geology Rev., 25, No.1, 63-8 (1983). Pankina, R.G., Mekhtiyeva, V.L., and Maximov, S.P. (1983). Formation of H2S and CO2 in oils of Astrakhan deposit (by isotopic data) (in Russian). Geol. nefti i gaza, 4, 45-51. Perry, E.C. Jr, Monster, J., and Reiner, T. (1971). Sulfur isotopes in Swazilandsystem barites and the evolution of the Earth's atmosphere. Science, 171 (3975), 1015-16. Pilot, J., and Harzer, D. (1969). Sauerstoff- und Schwefelisotopenuntersuchungen an Sulfaten. Isot. Titles, 1, 160-78. Pilot, J., Rosier, H.J., and Muller P. (1972). Zur geochemischen Entwicklung des Meerwassers und mariner Sedimente im Phanerozoikummitte\s Untersuchungen

130

Stable Isotopes

von S-, 0- und C-Isotopen. NeueBergbautechnik, Z (3)\ 161-8, Price, F.T., and Shieh, Y.N. (1979). The distribution and isotopic composition of sulfur in coals from the Illinois Basin. Econ. Geol., 74, 1445-61. Rafter, T.A. (1962). Sulfur isotope measurements on New Zealand, Australian and Pacific Island specimens. In: Jensen, M.L. (Ed.) Biogeochemistry of Sulfur Isotopes, National Science Foundation Symposium, pp. 42-60. Rafter, T.A., Wilson, S.H., and Shilton, B.W. (1958). Sulphur isotopic variations in nature. Part 6. Sulphur isotopic measurements on the discharge from fumaroles on White Island. N.Z. J. Sci., 1, 154-71. Rees, e.E. (1970). The sulphur isotope balance of the ocean: an improved model. Earth Planet. Sci. Lett., 4, 366-70. Rittmann, A. (1981). Vulkane und ihr Totigkeit (Volcanoes and Their Activities) 3rd edn, Ferdinand Enke, Stuttgart, FRG, pp. 399. Rose, W.J., Bonis, S., Stoiber, R.E., Keller, M., and Bickford T. (1973). Studies of volcanic ash from two recent Central American eruptions. Bull. Volcanol., 37 (3), 338-64. Rose, W.J., Stoiber, R.E., and Malinconico, L.L. (1982). Eruptive gas compositions and fluxes of explosive volcanoes: budget of Sand Cl emitted from Fuego volcano, Guatemala. In: Thorpe, R.S. (Ed.) Andesites: Organic R.S. Andesites and Related Rocks, Wiley, Chichester, pp. 669-76. Rosier, H., Pilot, J., Harzer, P., and Kruger, P. (1968). Isotopengeochemische Untersuchungen (O.S.e.) an Salinar- und Sapropelsedimenten Mitteleuropas. 23rd International Geological Congress, Vol. 6, pp. 89-100. Ruckmick, J.e., Wimberly, B.H., and Edwards A.F. (1979). Classification and genesis of biogenic sulfur deposits. Econ. Geol., 74 (2),469-74. Ryaboshapko, A.G. (1983). The atmospheric sulphur cycle. In: Ivanov, M.V. and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE 19, Wiley, New York, pp. 203-96. Rye, R.O., Shawe, D.R., and Poole, F.G. (1978). Stable isotope studies of bedded barite at East Northumberland Canyon in Toquina Range, central Nevada. 1. Res. U.S. Geol. Surv., 6, 221-9. Sakai, H. (1957). Fractionation of sulphur isotopes in nature. Geochim. Cosmochim. Acta, 12, 150-69. Sakai, H. (1971). Sulfur and oxygen isotopic study of barite concretions from banks in the Japan Sea off the Northeast Honshu, Japan. Geochem. J., 5, 79-83. Sakai, H. (1972). Oxygen isotopic ratios of some evaporites from Precambrian to recent ages. Earth Planet. Sci. lett., 15 (2), 201-5. Sakai, H., and Matsubaya, O. (1977). Stable isotopic studies of Japanese geothermal systems. Geothermics, 5 (1-4), 97-125. Sakai, H., and Nagasawa, H. (1958). Fractionation of sulphur isotopes in volcanic gases. Geochim. Cosmochim. Acta, 15 (1-2), 32-9. Sakai, H., Casadevall, T.J., and Moore, J.G. (1982). Chemistry and isotope ratios of sulphur in basalts and volcanic gases at Kilauea Volcano, Hawaii. Geochim. Cosmochim. Acta, 46 (5), 729-38. Schidlowski, M., and Junge, e.E. (1981). Coupling among the terrestrial sulphur, carbon and oxygen cycles: numerical modeling based on revised phanerozoic isotopes record. Geochim. Cosmochim. Acta, 45, 581-94. Schidlowski, M., Junge, e.E., and Pietrak, H. (1977). Sulphur isotope variations in marine sulphate evaporites and the phanerozoic oxygen budget. J. Geophys. Res., 82, 2557-65. Schneider, A. (1978). Sulphur abundance in common igneous rocks. In: Wedepohl,

Lithospheric Sources of Sulphur

131

H. (Ed.) Handbook of Geochemistry, Vol. II, Springer, Berlin, Heidelberg, New York, Ch. 16-C. Schroll, E., and Pak, EW. (1980). Schwefelisotopenzusammensetzung von Baryten aus den Ost- und Siidalpen. Tschermaks Mineralogische und Petrographische Mitteilungen, 27, 79-91. SCOPE 19 (1983). Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, Wiley, Chichester, 470 pp. Shakur, M.A. (1982). i)34Sand i)IHOvariations in terrestrial sulphates. Ph.D. thesis, The University of Calgary, Alberta, Canada. Smith, J.W., and Batts, B.D. (1974). The distribution and isotopic composition of sulfur in coal. Geochim. Cosmochim. Acta, 38, 121-33. Smith, J.W., Gould, K.W., and Rigby, D. (1982). The stable isotope geochemistry of Australian coals. 1. Org. Geochem., 3, 111-31. Solomon, M., Rafter, T., and Dunham, K. (1971). Sulfur and oxygen isotope studies in northern Pennines in relation to ore genesis. lnst. Min. Metall. Trans., 80B, 259-75. Stith, J.L., Hobbs, P.V., and Radke, L.F. (1978). Airborne particle and gas measurements in the emissions from six volcanoes. J. Geophys. Res., 83 (C8), 4009-17. Stoiber, R.E., and Bratton, G. (1978). Airborne correlation spectrometer measurements of S02 in eruption clouds from Guatemalan volcanoes. EOS, 59, 1222. Stoiber, R.E., and Jepsen, A. (1973). Sulfur dioxide contributions to the atmosphere by volcanoes. Science, 182 (4112), 577-8. Stoiber, R.E., and Malone, G.B. (1975). S02 emission at the crater of the Kilauea, at Mauna Ulu and at Sulphur Banks, Hawaii. Trans. Am. Geophys. Union, 56 (6), 461. Taylor, P.S., and Stoiber, R.E. (1973). Soluble material on ash from active Central American volcanoes. Geol. Soc. Am. Bull., 84 (3), 1031-42. Thode, H.G. (1981). Sulfur isotope ratios in petroleum research and exploration: the Williston basin. AAPG Bull., 65 (9), 1527-35. Thode, H.G., and Monster, J. (1964). The sulphur isotope abundances in evaporites and in ancient oceans (Geochemical Conference in Moscow, March 1963) (in Russian). In: Vinogradov, A.P. (Ed.) Vernadsky Memorial, Vol. I, Khimija Zemnoi Kori, pp. 589-600. Thode, H.G., and Monster, J. (1965). Sulphur-isotope geochemistry of petroleum, evaporites and ancient seas. In: Young, A., and Galley, J.E. (Eds.) Fluids in the Subsurface Environments, American Association of Petroleum Geologists, Memoir 4, pp. 367-77. Thode, H.G., and Monster, J. (1970). Sulphur isotope abundance and genetic relations of oil accumulations in Middle East basins. Am. Assoc. Petrol. Geologists Bull., 54, 627-37. Thode, H.G., and Rees, C.E. (1970). Sulfur isotope geochemistry and Middle East oil studies. Endeavour, 29, 24-8. Thode, H.G., Monster, J., and Dunford, H.B. (1958). Sulphur isotope abundances in petroleum and associated materials. Am. Assoc. Petrol. Geologists, 42 (11), 2619-41. United States Department of Health, Education and Welfare (1969). Air quality criteria for sulfur oxides. National Air Pollution Control Administration Publication AP-50, Washington, D.C. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1982). i)34Sand i)IRO abundances differentiate Upper Cambrian and Lower Devonian gypsum-bearing

132

Stable Isotopes

units. District of Mackenzie, N.W.T.-an update. Can. J. Earth Sci., 19. 1246-54. Venkatesan, M.I., Kaplan, I.R., Makiewig, P., Ho, W.K., and Sweeny, R.E. (1982). Determination of petroleum contamination in marine sediments by organic geochemical and stable sulphur isotope analyses. In: Ernst, W.G., and Morin, J.G. (Eds.) The Environment of the Deep Sea, Vol. 2, Prentice-Hall. Viezer, J., Holser,W.T., and Wilgers, C.K. (1980). Correlation of DC/12Cand 34S/ 32Ssecular variations. Geochim. Cosmochim. Acta, 44, 579,-87. Vinogradov, V.I. (1964). Isotopic composition and the origin of volcanic sulphur (in Russian). Geol. Rud. Meet., 3 (3). Vinogradov, V.I. (1966). On the question of the origin of volcanic sulphur. In: Reports on the Geochemistry of Endogenic and Hypergenic Processes (in Russian). Nauka, Moscow. Vinogradov, V.I. (1970). Isotopic composition of sulphur in the thermally active Kamchatka and Kurile islands and its genetic importance. In: Reports on the Geochemistry of Mercury, Molybdenum and Sulphur in Hydrothermal Processes (in Russian), Nauka, Moscow, pp. 258-71. Vinogradov, V.I. (1980). Isotopic composition of elements and the problem of mantle degassification and the formation of the gas-water cover of the Earth (in Russian). In: Kropotkin, P.N. et al. (Eds.) Degazatsiya Zemli i geotektonkia, Vol. 1, Nauka, Moscow, pp. 23-30. Vinogradov, V.I., and Kizilshtein, L. (1969). Isotopic composition of sulfides in coals of the Donetz Basin (in Russian). Litol. Polezn. Iskop, 5, 149-51. Vinogradov, A.P., Grinenko, V.A., and Ustinov, V.I. (1964). Isotopic composition of the sulphur and carbon in ore of the Shor-su deposit of Uzbekistan (in Russian). Geokhimiya, 11, 1075-85. Vlasov, G.M. (1971). Volcanic Sulfur Deposits and Some Problems of Hydrothermal Ore Formation (in Russian), Nauka, Moscow, p. 360. Vlodavets, 0.1. (1966). Volcanic activity on the Earth in historical time (in Russian). In: Recent Volcanism, Vol. 1, Nauka, Moscow, pp. 7-17. Vredenburgh, L.D., and Cheney, E.S. (1971). Sulphur and carbon isotopic investigation of petroleum. Wind River Basin, Wyoming. Am. Assoc. Petrol. Geologists, 55 (11), 1954-75. Yeremenko, N.A., and Pankina, R.G. (1971). Variation of 334Sin the sulphates of modern ancient marine basins of the Soviet Union (in Russian). Geokhimiya, 1, 81-91. Zettwoog, P., and Haulet, R. (1978). Experimental results on the S02 transfer in the Mediterranean obtained with remote sensing devices. Atmos. Environ., 12 (1-3), 795-6.

CHAPTER 5

Sulphur Isotope Variations in the Atmosphere


L. NEWMAN, H.R. KROUSE, AND V.A. GRINENKO

5.1

INTRODUCTION

The measurement of the isotope ratios of sulphur and oxygen can in principal be used to assess sulphur inputs into, transformation within, and removal from, the atmosphere. Major inputs arise from both anthropogenic and biogenic activities (Nakai and Jensen, 1967). Transformations arise from oxidation, neutralization, and other chemical reactions. Advection causes dilution and the main removal processes are dry deposition (governed by gravitation and diffusion) and rain. The admixture of sources can be discerned from their isotopic signatures whereas transformations and removal can be followed from the isotopic fractionation that might occur. In this chapter, the atmospheric sulphur cycle and the associated chemistry are summarized. Also presented is information on natural isotopic variations and fundamental concepts relating to the use of isotopic data to delineate anthropogenic S in the atmosphere. Examples of successful applications of these concepts are given. These are supplemented by case studies in Sections 8.2, 8.4, and 8.5. Finally, consideration is given to the potential of using isotopically enriched sulphur to study transport and transformation of atmospheric S compounds. 5.2 SULPHUR SPECIES IN THE ATMOSPHERE

5.2.1 Atmospheric S compounds and their concentrations Sulphur in the atmosphere occurs in the gaseous (HzS, CH3SCH3 (DMS), CH3SSCH3 (DMDS), CH3SH, cas, CSz, SOz), liquid (HZS03, SOl-, and SOl-), and solid (sulphates, SO) phases. A detailed discussion of these species worldwide is given in Chapter 4 of SCOPE 19 (1983) and will not be repeated here. A summary of pertinent data for some species is given in Table 5.1.
Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko <9 1991 SCOPE, Published by John Wiley & Sons LId

134

Stable Isotopes

Table 5.1 Concentration, residence time, and mixing height of different forms of sulphur in the atmosphere (SCOPE 19, 1983)
Species DMS H2S S02 SO/CS2 Concentration (fLgS m-3) 0.04 0.05 0.2 0.5 0.7 Residence time (d) < 1 1 3 4 70 Mixing height (km) 1.5 :I::0.5 1.5 :I::0.5 6.5:1:: 0.5 2.5:1:: 0.5 6.5 :I::1

cas

0.2

500

6.5 :I::1

5.2.2 The atmospheric chemistry of sulphur The chemistry of sulphur in the atmosphere is complex. The reduced S compounds are subject to direct photochemical oxidation and attack by other photochemically produced species including ozone and the OH radical. Cox and Sheppard (1980) have summarized reaction rates of S02, H2S, and organic sulphides with OH radicals in the troposphere. A simplified summary of oxidative reactions follows. 5.2.2.1 H2S oxidation Most global balance calculations propose significant emission of H2S to the atmosphere. There are a variety of possible reactions for the destruction of H2S. Sprung (1974) gives rates for a variety of homogeneous gas phase reactions. Hydroxyl radical is the most important destroyer of H2S followed, at a rate three orders of magnitude lower, by ozone. The reaction of H2S with the OH radical leads to the formation of HS radical and water; OH + H2S ~ HS + H2O The HS radical may react with O2 to form S02, or may proceed directly to sulphuric acid (Thiemens, 1977): HS + 02~HS02 HS02 + 02~HS04 HS04 ~ OH + SO) SO) + H2O ~ H2SO4 It was subsequently shown that the HS radical goes quantitatively to S02 by one of two paths:

Sulphur Isotope Variations in the Atmosphere

135

I.

HS + O2--7 SO + OH SO + O2--7 S02 + 0 II. HS + O2 --7HS02 HS02 --7HOSO HOSO + O2--7H02 + S02

In either case, all of the oxygen in the resultant S02 comes from atmospheric O2. Hales et at. (1974) examined the heterogeneous oxidation of H2S by ozone and concluded that heterogeneous processes on the surface of their reactors were unimportant. Hence, they concluded that heterogeneous oxidation is probably unimportant in the atmosphere. The rates of H2S oxidation by O2 in solution at pH values below 6 are also negligible (Chen and Morris, 1972). 5.2.2.2 So dust In view of the rapid fallout, oxidation of sulphur dust in the atmosphere is negligible. Upon reaching the ground, it is converted to sulphate by oxidation and hydrolysis reactions which may be mediated by sunlight and sulphur oxidizing bacteria such as Thiobacillus sp.

5.2.2.3 502 chemistry Since most of the anthropogenic emissions are in the form of S02, and since biogenic sulphides are oxidized to S02, the chemistry of S02 in the atmosphere is of major importance. Sulphur dioxide may be removed from the atmosphere directly by dry deposition (e.g. Garland, 1978; Platt, 1978; Mayer and Ulrich, 1978). Garland (1978) estimates that about half of the S02 emitted to the atmosphere is removed by dry deposition. The remaining half of the S02 in the atmosphere is removed by wet deposition, consisting of oxidation, hydration, and condensation. These processes may occur in a variety of sequences. Oxidation may not occur in some droplets before removal from the atmosphere by gravitational settling. Newman (1980, 1981) reviewed the literature on the atmospheric oxidation of sulphur dioxide in relation to power plant and smelter plume studies. The average rate of oxidation is generally less than 1% h -I in clean air, but in polluted air the rate can easily double (Forrest and Newman, 1977a, 1977b). A diurnal variation in the rate of oxidation is observed that is near zero at night and approximately 3% h-l during mid-day. There is no basis for a choice between the homogeneous or the heterogeneous pathway as the dominant oxidation mechanism.

136

Stable Isotopes

Eggleton and Cox (1978) reviewed the homogeneous oxidation of SO;; in the atmosphere. Reactions with free radicals, especially OH, are most important in both polluted and unpolluted air: OH + S02 -7 HOS02 Calvert et at. (1978) concluded that the most likely reaction of the HOS02 radical is with O2:
HOS02 + O2 -7 HOS02OO

They also noted that the disproportionation reaction forming H02 and S03 is endothermic and therefore non-competitive with the preceding reaction. The resulting HOS02OO radical is extremely reactive. It should eventually gain a H and lose an 0 atom to become H2SO4, or lose both an 0 and a H atom to become S03 which hydrates to H2SO4' The H2SO4 can then be neutralized by ambient ammonia to form (NH4hS04' Davis et at. (1978) estimated the rates of a variety of reactions involving HS03 and HSOs radicals. They believe that the dominant processes are the oxidation of HS03 to HSOs by atmospheric O2 with a slightly slower rate for the hydration of the HS03 radical. The resulting HSOs radical is hydrated (probably with several water molecules) at a rate several orders of magnitude greater than any competing reaction. It is not clear at what point the reactions can be considered to change from homogeneous gas phase reactions to heterogeneous solution reactions. The heterogeneous oxidation of S02 in solution has been reviewed by Beilke and Gravenhorst (1978) and by Hegg and Hobbs (1978). The primary step in these reactions is the solution of S02 in water and its equilibration in the sulphurous acid system to form bisulphite and sulphite:
S02 + H2O:;;:::::H2S03 :;;::::: H+ + HS03 - :;;::::: 2H+

+ SO/-

The dissociation reactions are very fast (order of 106 S-l). In solution,

S032- is present in negligibleamounts at pH values less than 6 and HS03 dominates in rain and cloud droplets. The oxidation of SlY to SYIin droplets may proceed by three mechanisms: oxidation by O2 without metal catalysts; oxidation by O2 in the presence of transition metal catalysts; and oxidation by strong oxidants, particularly hydrogen peroxide or ozone. Beilke and Gravenhorst (1978) concluded that the uncatalysed reaction with O2 is unimportant in droplets. The catalysed reaction may be important where the transition metal component in the solution is extremely high, but oxidation reactions by hydrogen peroxide and ozone are probably the dominant mechanisms. Hegg and Hobbs (1978)

Sulphur Isotope Variations in the Atmosphere

137

suggest that some mixed salt catalysts (equimolar Mn2+ and Fe3+) may promote oxidation of SOz in urban atmospheres. Newman (1981) suggested that oxidation within cloud droplets is an important means through which air obtains its sulphate and acidity. Recent evidence substantiates this hypothesis (Daum et al., 1982). 5.3 ANALYTICAL TECHNIQUES

5.3.1 Sampling of atmospheric S compounds The first requisite for sampling sulphur compounds from the atmosphere for isotopic measurements is the preservation of their isotopic composition. Consequently all steps in the procedure should be quantitative to assure that isotopic fractionation does not occur. Concentrations in the atmosphere are usually low (ppb or less). The sampling rate has to be designed to obtain enough material in a time period commensurate with the temporal resolution dictated by the experiment (Forrest and Newman, 1973). About 1 mg of material is required in order to perform a precise isotopic ratio measurement on the mass spectrometer. Further, it is generally desirable to sample and measure the isotopic ratio for each of the sulphur substances present in the atmosphere. A requisite during sampling is the necessity of preventing transformation of one sulphur compound into another, e.g. the conversion of SOz to sulphate when it is desired to measure the isotopic ratio of the sulphate particles in the atmosphere. The flue gas of power plants can be sampled through a quartz wool filter, followed by a heated probe to prevent condensation in the line (Forrest et al., 1973). The S03 is then condensed and collected in a coil maintained at a temperature above the dew point of water. Finally the SOz is collected and oxidized to sulphate in an impinger containing alkali and hydrogen peroxide. Huygen (1963) successfully collected SOz by forcing air through cellulose filters impregnated with potassium hydroxide plus glycerol or triethanolamine. This procedure was further developed by Forrest and Newman (1973) who collected ambient samples with a filter pack arrangement consisting of a quartz prefilter to trap aerosol sulphate followed by KOH-triethanolamineimpregnated fast-flow cellulose filter paper to absorb SOz. They achieved air flow rates of 2 m3 min-1 with 8 x 10 in filters in high-volume samplers. The filter paper can also be impregnated with silver and mercury salts to collect HzS (Hitchcock et al., 1978). Another technique to sample SOz for sulphur isotope analyses employs adsorption on a molecular sieve (Holt, 1975). Several procedures for collecting organic sulphur compounds, primarily for final determination by gas chromatography, have been described.

138

Stable Isotopes

Hydrogen sulphide was absorbed by mercuric-chloride-treated paper (Hockheiser and Elfers, 1970), lead-acetate-soaked membrane filters (Okita et al., 1971), and silver-nitrate-impregnated cellulose filters (Natusch et al., 1972). Nakai et al. (Section 8.5.2.1) describe a circulating pumping and chemical trapping system for sampling hydrogen sulphide released by tidal flats. Okita (1970) used glass fibre filters impregnated with mercuric cyanide to adsorb mercaptans and with mercuric chloride to collect mercaptans and DMS. Black et al. (1978) used molecular sieve 5A packed in tubes to adsorb S02 and H2S. Cryogenic trapping at liquid nitrogen temperatures enabled Sandalla and Penkett (1977) to concentrate COS and CS2. Several organic compounds, including H2S, methyl mercaptan, DMS, and DMDS, were adsorbed on gold-coated beads in a quartz tube by McClenny et al. (1979). Although numerous procedures have been described for absorbing specific gases in liquids or solutions by bubblers or impingers, flow rates are generally too low for successfully accumulating sufficient quantities for isotopic analysis. Particulate sulphur is readily removed from an air stream by filters. By this technique, sulphur isotopic ratios of atmospheric aerosol sulphate have been measured by Forrest and Newman (1973) and by Saltzmann et al. (1983). Collected particles may be separated by chemical means. Leahy et al. (1975) used benzaldehyde to specifically dissolve and separate H2SO4 from other sulphates. Sulphur isotope analyses can also be carried out on lead-peroxide-treated paper placed on the inner surface of an open cylinder and exposed vertically to the atmosphere, typically for a one-month period. Krouse (unpublished data, 1976; see Section 8.2) found isotope data from these cylinders to be comparable to those obtained by averaging samples collected over shorter time periods with the filter pack of Forrest and Newman (1973). 5.3.2 Further characterization of atmospheric compounds for sulphur isotope analyses 5.3.2.1 Wind direction If concentrations and isotopic compositions of atmospheric S compounds can be characterized according to wind direction, more information might be obtained on sources and mixing phenomena. Isotopic variations in S02 with wind direction were documented near Whitecourt, Alberta, during September 1975 (Krouse et at., 1984). A highvolume sampler was operated on a 16-m high scaffold to study the effects of S02 emissions on mid-crown and upper-crown foliage of a mature conifer stand. On 18 September, the wind was north-west and the S02 had 534S values near +24%0, corresponding to a sour gas plant operation (Section

Sulphur Isotope Variations in the Atmosphere

139

8.2) in that direction. During the next day, the wind shifted easterly and the ~34Svalue decreased to + 14%0,testifying to SOz arriving from sources other than the dominant industry in the area. During the next two days, the wind swung southerly and then westerly and the ~34Svalues increased, reflecting increasing contributions from the industry. The above event can be depicted on a polar diagram where ~34Sis plotted as the radial coordinate and the angular coordinate corresponds to wind direction (Figure 5.1). It would appear that if the wind direction sector angle could be reduced further during sample collection, such a diagram could pinpoint sources of sulphur emissions very effectively. It must be emphasized that the samples were collected using one sampler with concurrent meteorological data, and interpretations were made later. This procedure is inefficient since the wind direction might change many times during the acquisition of a sample. The construction of Figure 5.1 was possible because the particular major synoptic event was not subject to short-term fluctuations.

5
Figure 5.1 Polar plot depicting changes in (')J4Svalues for atmospheric S02 with wind direction near Whitecourt Alberta, 18-21 September 1975(data from Krouse et al., 1984)

140

Stable Isotopes

On the basis of the above study, arrays of high-volume samplers, each of which responded to a preselected wind direction, were constructed. An array with four high-volume samplers was tested in 1976 on the same scaffold in the Whitecourt area (Krouse et al., 1984). Since the array was programmed by a simple electrical contact system within the head of the wind vane, some difficulties were experienced at low wind velocities because of internal friction. Nevertheless, the observations were similar to those recorded the previous year during the major synoptic event. The four-unit sampler array was also used for ground-level sampling in the Teepee Creek area of Alberta (Krouse and Case, 1981). Although there was a minor source of industrial emissions in the area, it was found that the highest S02 concentrations were not downwind of the plant. Further, a large component of the S02 flux appeared to pass through the study area. A mobile nine-unit high-volume sampler array was then built and tested (Krouse and Case, 1984). Eight samplers were individually triggered according to preselected wind directions whereas the ninth was direction ally independent and operated at wind speeds below a pre-set cutoff. An elapsed time counter on each sampler recorded its accumulated time of operation. Since air flow through each sampler was calibrated and controlled, the amount of air that had passed through the sample could be determined and used in concentration calculations. Filters would not be removed from a sampler that had operated for an insufficient time to provide enough sample for analysis. A criticism of the mobile array is that winds aloft are seldom in the same direction as those at ground level. Therefore, to avoid false source identification, the elevational dependence of the wind behaviour must be known. The argument can be reversed by stating that the isotope determinations per se help to differentiate sources and reduce interpretation errors. It would also seem that concentration and isotope measurements of atmospheric S at 'nose level' are justified from the viewpoint of human health. 5.3.2.2 Particle size The size of particles relates to their source, accretion, and transport in the atmosphere. Particles combining lowest density and smallest size arriving at an observation point can potentially come from further distances. Particles may be collected according to aerodynamic size using stacked metal plates with either staggered holes or slots (slotted cascade impactor) (Liu et al., 1980). Commercial units have typically five aerodynamic size classes: >7, 3.3 to 7.0,2.0 to 3.3,1.1 to 2.0, and <1.0 !J.m.These are considered to correspond to particles found on the human skin surface, and in the trachea and primary bronchi, secondary bronchi, terminal bronchi, and alveoli respectively. Aerodynamic sizing with a five-stage slotted cascade impactor was incorpor-

----------------

Sulphur Isotope Variations in the Atmosphere

141

ated in the mobile array described above. Data for a study in the midst of three sour gas processing gas plants are given in Table 5.2. It is seen that the smallest particles reaching the non-slotted filter paper are very much depleted in 34Sas compared to larger ones and the simultaneously collected S02. This suggests that the smallest particles bear little relationship to the nearby industrial operation. 5.3.2.3 Photographic examination of particles If more parameters are examined for particles, more information can be gained about their sources. Simple visual and microscopic techniques are very effective. The diverse nature of particulates is shown in the scanning electron microscope (SEM) photograph of Figure 5.2. which depicts abiological dust particles as well as biological materials such as pollen grains and broken plant debris. The latter should have an isotopic composition representing a mixture of sulphur from the atmosphere and soil (Section 7.2.4). Variations in 834S values of the total particles collected at different sampling heights can be explained by different proportions of abiological and biological matter (Krouse et al., 1984).

142

Stable Isotopes

Table 5.2 The OJ4Svalues for S02 and six aerodynamic class sizes of airborne particulate matter near Crossfield, Alberta, March-April, 1983 (Krouse and Case, 1984)
Sampler" Size (/Lm) >7.0 3.3-7.0 2.0-3.3 1.1-2.0 <1.0 Non-slotted S02 1 +18.8 + 13.2 +15.3 +13.1 +12.3 -2.0 + 23.4 2 +14.0 +12.0 +13.6 +9.5 +13.5 -3.4 +22.4 3 +16.6 +6.5 +14.5 +15.5 +18.9 +3.5 +15.2 4 +19.4 +17.1 +16.0 +15.6 +16.6 +6.1 +16.3

" Numbers refer to different wind directions.

5.3.2.4 Endogenous versus adsorbed sulphur Since endogenous sulphur and sulphur compounds adsorbed on a particle's surface may come from different sources, it is desirable to separate these components. One approach would be to wash the particulates and compare the isotopic composition of the filtrate and the residue. This is appropriate if water-soluble sulphur compounds are adsorbed to insoluble organic particulates. To date, such investigations have apparently not been attempted. 5.3.3 The use of sulphur isotope data from vegetation and soils Epiphytic lichens and many mosses acquire the bulk of their sulphur from the atmosphere (Krouse, 1977; see Sections 7.2.3. and 8.2). Therefore the sulphur isotope composition of these species represent a long-term average for atmospheric S. In addition to upward SO/- transport from the soil, rooted plants acquire some atmospheric S by gaseous transport through their stomata. Therefore needles and leaves reflect trends in the isotopic composition of atmospheric S gases if fluctuations in the 834Svalues of soil are minimal. These trends are also preserved in litter formed by fallen needles and leaves in upper soil horizons. The surface soil may in fact provide a better record of atmospheric trends than living foliage since sulphur from wet fall and dry deposition may be incorporated in the litter. The use of vegetation and soil isotope data as indicators of sources of atmospheric sulphur is documented in Chapter 8.

Sulphur Isotope Variations in the Atmosphere

143

5.4

SULPHUR ISOTOPE ABUNDANCE VARIATIONS IN ATMOSPHERIC COMPOUNDS

5.4.1 Fractionation of sulphur isotopes during transformations of atmospheric sulphur compounds A critical question is whether the isotope distribution is preserved during chemical transformations in the atmosphere (e.g. S02 to SOl-). If so, (334S and (3180values can be used as a label to identify the source of the precursor compounds. Newman and his colleagues assumed that isotope fractionation occurs in the formation of particulate S042- from precursor S02 power plant plumes, and that this shift can be employed to estimate the percent conversion of the S02 in the plume (Newman et al., 1971; Newman et al., 1975a, 1975b; Forrest and Newman, 1978). They attribute the fractionation to the following isotope exchange reaction in liquid surrounding atmospheric particles:
K

32S02 (g)

34S0/-

(I) ~

34S02 (g)

+ 32S0/-

(I)

Subsequent catalytic oxidation of the HS03 to SOl- should yield a (334S value for the sulphate that is established by the equilibrium constant K. On the basis of isotope fractionation between S02 and HS03 dissolved in water associated with Dowex exchange resins, Eriksen (1972) estimated K to be 1.02. Newman and his colleagues did not observe the anticipated shift of +20%0 in the sulphate relative to the S02 which is implied by this scheme. However, this could be due to other factors, including the presence of sulphate derived from other sources. Large shifts in the (334S value of S02 are also absent in plumes of sour gas processing plants (Figure 5.3). Forrest et al. (1973) found that sulphate formed in the combustion process or in stacks at high temperatures was only 1.5%0enriched in 34Scompared to the S02' Nriagu and Coker (1978b) found this enrichment to be 2.6%0 in a smelter stack at Sudbury, Canada. The isotopic composition of S02 and sulphate generated by oxidation of biogenic H2S in natural springs shows considerable variation which may in part be related to sampling problems (van Everdingen et al., 1982, 1985). The average values for H2SO4 fallout appear to be similar to the precursor H2S, whereas the ambient S02 is about 5%0depleted in 34S. In summary, sulphur isotope fractionation occurs during oxidation of anthropogenic S02 and biogenic H2S. It is smaller than that predicted by equilibrium exchange between S02 and SO~- In most cases, this fractionation will not preclude identification of sources because of the wide variations in
(334Svalues of emissions.

144

Stable Isotopes

AMOCO CROSSFIELD PL.UMESAMPL.ES


+32
11 October 1975
~

14 October 1975 1975

I 18 December
0 ;:,g 0
(f) '<t I') GO

+30

+26

2 DISTANCE

4 DOWNWIND

6 8 FROM STACK

10 (km)

Figure 5.3 Variations in 834S values for S02 in plume as a function of distance downwind of a sour gas processing plant Crossfield, Alberta, Canada

5.4.2 Natural sources Sulphur compounds in the atmosphere can originate naturally (volcanic, sea spray, aeolian weathering, biogenic) or anthropogenically (combustion and refining of fossil fuels, gypsum processing, ore smelting). This section summarizes sulphur isotope data pertaining to natural emissions whereas the next section (5.4.3) examines emissions from industrial processing and combustion. 5.4.2.1 Sea spray and marine aerosols It is well known that the 834S value of seawater S042- is about + 21%0 (Rees, 1978). However, lower values are found for marine aerosols. The superficial layer of about 30 fLmthickness at the air-sea interface has been found to be highly enriched in organic matter and associated elements such as I, P, and certain heavy metals (McIntyre and Winchester, 1969; McIntyre, 1970; Chesselet et al., 1972; Duce et al., 1975; Morelli, 1977).

It has been estimated that the sea produces between 103and 104Tg yC I
of atmospheric sea salt particles with radii less than 20 fLm(Eriksson, 1959, 1960; Blanchard, 1963; Petrenchuk, 1980). Most sea salt particles in this size range are produced by whitecap bubbles (Duce, 1981). When a bubble

laDle J.j

ISOTOpICcomposlIIon

or sUlpnur 10 me OceanIC atmospnere 834S (%0) V, i': "6'" ;:,i': Mean Remarks Reference

...
Q

Region

Compound

Range

t;:;<

Central Atlantic Ocean New Amsterdam Island, Indian Ocean (3747'S, 7732'E) Gulf of Guinea San Francisco Bay, California, USA Waimanolo Bay, Hawaii, USA Chesapeake Bay, Virginia, USA

SO/SO/-

2 to + 15

+11 +0.9

Aerosols a few m ASL Gravenhorst (1978) Nguyen and Cortecci, (unpublished data)

.g "> .... Q'


'"

SO/SO/-10.1 to +8.3

+14.7 -1.3 Wind from sea stratus droplets and particulates Two samples of 6 hours duration on separate days High sulfate concentration Low sulfate concentration High sulfate concentration Low sulfate concentration

Nguyen and Cortecci, (unpublished data) Ludwig (1976)

S.
">

Gaseous S Particulate S02

+5.9 to +6.0 + 15.0 to + 16.2 -0.1 to +1.6 -1.0 to +0.8 +0.1 to +1.6 -9.4 to +0.1

+6.0 +15.6 0.9 -0.1 +0.7 -1.9 +4.2 -3.8 +4.3


-2.1

Hitchcock (1976a, 1976b) Hitchcock and Black (1984)

' Q {J ;:,"> ... ">

Atl2ntic Coast Guisseny, Brittany,


France Mikawa Bay, Japan

SO/S02
S042HzS

+3.3 to +7.3 -4.1 to +0.5 +1.5 to +7.3


-21.9 to +12.1

Exposed algal beds during low tide


HzS released sediments from

Newman and Forrest (Section 8.4) Nguyen and Cortecci (unpublished data)
Nakai et ai. (Section

>-'
V1

.j:>.

8.5.2.2)

146

Stable Isotopes

bursts at the sea-air interface, atmospheric particles are produced both from the microlayer and from the subsurface sea water. Bonsang et at. (1980) have found that the SO/-/Na+ ratio of marine aerosols is always higher than that of sea water (0.25) and can reach 1.25. Moreover, they have shown that the marine aerosols with radii less than 1.1 fLm are enriched in SO/- from oxidation of organic sulphides released by biological activity at the sea surface. This secondary SO/- may be depleted in 34S depending on the Drigin of the organic sulphides (Section 5.4.2.2). Data from a number of studies summarized in Table 5.3 show that the 834S value of SO/- in marine aerosols ranges from -10 to + 16%0. In two studies where gaseous S was also measured, it was on the average 8-10%0depleted in 34Scompared to SO/-. Therefore the isotope data provide evidence that the SO/- in marine aerosols has lower 834S values because of oxidation of biogenic sulphide. 5.4.2.2 Biogenic emissions During dissimilatory SO/- reduction by strict anaerobes such as Desutphovibrio desutphuricans and Desutphotomacutum, copious dissolved sulphide species are produced in either anoxic waters or sediments. Investigators have found that the 834S values of sedimentary sulphide vary from 0 to -70%0 with respect to coexisting sulphate (e.g. Vinogradov et at., 1962; Kaplan et at., 1963; Ivanov, 1978; Weyer et at., 1979; Chambers and Trudinger, 1979). Intense emissions of H2S to the atmosphere occur in oceanic littorals, lagoons, salt marshes, rice-fields, springs, and wet tropical forest soils (Delmas and Servant, 1983; Delmas et at., 1978; see Sections 6.3,6.7, and 8.5.3). DMS concentrations in the atmosphere in some coastal areas where living algae are exposed at low tide can be much higher than those of H2S (Paugam et at., 1977). The isotope fractionation associated with the production of gaseous organic sulphides is more difficult to assess. If they arise from the decomposition of organic S compounds, the isotopic selectivity should be minimal since a given mass of organic debris should all eventually decay. Since the isotopic selectivity during SO/- assimilation is small (Section 6.2.2), the organic sulphide should be isotopically similar to the original SO/-. Another possibility is that gaseous organic sulphides are generated during S042reduction. This alternative is not well understood. In the open ocean, phytoplankton and zooplankton living principally in the euphotic layer assimilate seawater SO/- and generate sulphides such as DMS, DMDS, CH3SH, and H2S (Lovelock et at., 1972; Nguyen et at., 1978; Barnard et at., 1982; Maroulis et at., 1977; Andreae and Raemdonck, 1983). These sulphides are released to the atmosphere and oxidized to S02 and ultimately SO/-. Nguyen et at. (1974, 1983) observed open ocean atmospheric concentrations of 0.1 and 1 fLg/m3 for S02 and sulphate

Sulphur Isotope Variations in the Atmosphere

147

respectively in the sub-Antarctic and Antarctic. In laboratory experiments, Nguyen and Cortecci (unpublished data) found that SOz generated above algae under restricted air flow had a 034S value of -14.9%0, which was considerably lower than the value of -3.8%0 observed in situ. Data for gaseous S compounds in oceanic atmospheres are summarized in Table 5.3. The HzS released, in Mikawa Bay, Japan, had the largest range of 034S values (- 22 to + 12%0)with the minimum value recorded in the coldest month. The oceanic influence should decrease in going from near shore to continental environments. Saltzman et al. (1983) examined atmospheric SOz and SOiat the Hubbard Brook Experimental Forest Station, New Hampshire, USA. The 034Svalues of SOz and aerosol SO i- ranged from -1.1 to +2.3%0 (11 samples) and from +0.8 to +3.5%0 (14 samples) respectively. It is noted that these data are not significantly different from those at Chesapeake Bay (Table 5.3). However, one should not conclude that marine aerosol is the sole source or even the major component at Hubbard Brook in view of the industrial emissions in the northeastern USA which have a similar average 034S value (Section 8.4). Grey and Jensen (1972) found that 034S values from biogenic HzS at the margins of Great Salt Lake ranged from 0 to + 10%0.Since the maximum value is close to that of S04Z- in the lake, it might correspond to HzS generated during plant decay. Biogenic HzS from cold and warm springs appears to arrse almost entirely from SOi- reduction because of its significant depletion in 34S.For example, springs near Paige Mountain, Northwest Territories, Canada, have S04Zwith a 034S value near + 16%0.In contrast, the intense biogenic emissions contain HzS and SOz with 034S values of - 32 and -38%0 respectively. Aerosol HzS04 is similarly depleted in 34S (Krouse and van Everdingen, 1983). The isotopically selective generation of gaseous reduced S compounds is not unique to bacteria. Living vascular plants under severe S stress emit HzS which is depleted in 34S in comparison to the foliar S (Section 7.2.6). 5.4.2.3 Volcanic activity Flux and isotope data for different volcanic emanations have been compiled in Section 4.6. Whereas the isotope composition of individual gases and particulates may vary widely at a given site, the weighted average for total sulphur has a much smaller spread. Lein (Section 4.6.1) estimates that 28 x 106tonnes S are emitted annually from volcanoes with an average 034S value of +4.7%0. Volcanic activity influences the sulphur isotope composition of aerosols in the stratosphere. Extreme variations after the eruption of Mt Agung were reported by Castleman et al. (1974). At heights near 19 km in the Southern

148

Stable Isotopes

Hemisphere, the 834Svalue of sulphate increased to + 16%0during 1963 and then decreased to -24%0 by 1965. The variation was much less at the same altitude in the Northern Hemisphere and at 15 km height in the Southern Hemisphere. During several years without major eruptions, the 334Svalues

were found to be quite uniform (+2.6 ::!:0.3%0).


5.4.3 Anthropogenic sources 5.4.3.1 Combustion of coal The 334Svalues of coal range from -30 to +30%0(Section 4.5.2). Since much of this spread is due to coexistent pyrite, the isotopic composition of emissions from coal burning depend upon pretreatment for mineral sulphide and sulphate removal. The 334S values of flue gas from coal combustion can range from -0.5 to +20%0 with -1 to +3%0 being most common (Nielsen, 1974). Newman et al. (1975b) measured 334Svalues in the range +1.3 to +3.6%0 (average, +2.8%0) for S02 emitted from a coal-fired power plant at Keystone, Pennsylvania, USA. Buzek et al. (Section 8.9) found 334Svalues in the flue gas of the Chwaletice coal-fired power plant near Prague to range from -1.4 to +0.1%0 with an average of -0.7%0. Combustion of coal can also produce fly ash, the 334Svalues of which represent S-containing minerals in the coal with perhaps contributions from oxidation of organic sulphur compounds (see also Sections 8.3, 8.4, and 8.9). 5.4.3.2 Combustion and refining of oil and gas The 334S values of oils and H2S in gas-oil associations can range from - 5 to +30%0. A large number of data was compiled by Pankina (Sections 4.5.3 and 4.5.4). She concluded that the majority of 834Svalues for oil fall between 0 and + 10%0. The range for H2S is wider, with an estimated average of + 10%0. Nielsen (1974) found 834S values in the range +4.8 to +5.4%0 in flue gas of oil-fired power plants. Newman and Forrest (Section 8.4) found that fuel oils used in power plants of the north-eastern USA ranged in 334S value from -5 to + 14%0(see also Section 8.2). 5.4.3.3 Sulphide ores Roasting of sulphide ores is a major source of S02 in the atmosphere. The 834S values of sulphide ores are highly variable (- 25 to + 25%0) but the most common values are in the vicinity of 0 to +5%0. The average calculated by Grinenko and Grinenko was +3.4%0 (Section 4.3).

Sulphur Isotope Variations in the Atmosphere

149

5.4.3.4 Gypsum mining and processing Gypsum mining and wall board manufacturing introduces sulphate particulates into the atmosphere with positive 834S values in the range found for evaporites (Section 4.2). 5.4.4 Summary Data for 834Svalues of precipitation, gases, and particulates are summarized in Tables 5.4 and 5.5 and Figure 5.4. On average, the marine atmosphere is between 5 and 10%0 enriched in 34S as compared to the continental atmosphere. The wide range of sulphur isotope ratios arise in response to different natural and anthropogenic sulphur emissions. On the global scale, the statistics are inadequate to conclude whether the average isotopic composition of the total industrial area differs significantly from that of the total rural area. However, on the local scale, large differences in 834Svalues are found over short distances near dominant industrial and natural emitters. It is in these situations that stable isotope studies prove most successful (Section 5.5 and Chapter 8).

[ [
MARINE

- SEA SPRAY
VOLCANIC GASES H2S, DMS: BACTERIAL

[
[

H2S, DMS: PLANT DECAY

H2S, DMS: PLANT DECAY H2S, DMS: BACTERIAL COAL

CONTINENTAL

502

[
[
[
1 -20
I I I

OIL, GAS ORE S02

URBAN AIR
MARINE AEROSOLS
L-...

S042-

-40

0 834S (%0)

+20

+40

Figure5.4 Variationsof o34Svaluesfor differentsourcesof atmosphericsulphur


compounds

150

Stable Isotopes

Table 5.4 Isotopic composition ofSO/- in precipitation


(')34S (%0)

Region Italy Pisa Venice Israel Mediterranean Coast Central Galilee Jordan Rift Valley Japan Tokyo (1971-79) Nagoya (1971-79) Industrial regions Kurume (rural) Non-industrial regions Czechoslovakia Prague Beginning of rains Middle of rains End of rains Poland Lublin Sweden

Range

Mean

Reference

-2.5 to -0.2 to

+7.1 +8.3

+1.1 +3.0

Cortecci and Longinelli (1970) Longinelli and Bartelloni (1978) Wakshal and Nielsen (1982)

+5.3 to +14.3 +4.5 to +5.9 +5.2 to +5.4 -1.3 to - 1.1 to +5.5 +5.1

+8.9 +5.3 +5.3 + 1.7(1) + 1.9(1) +6.0(1) +12.8(1) + 14.0(2)

Nakai et al. Section 5.4.3 (1) Jensen and Nakai (1961) (2)

+3.2 to +7.3 +11.7 to +15.6 + 12.3 to + 15.6

+4.1 to +2.2 to +0.4 to +2 to

+8.4 +8.9 +9.7 +8 +8.2 +8.6 +7.6

+5.8 +5.4 +6.9 +3.7 +5.9 +5.9 +5.8 +10.6 +6.6

Busek, Cerny, and Sramek (Section 8.9)

Trembaczowski, and Halas (contributed data Ostlund (1959) Chukhrov et al. (1977)

+3.2 to

USSR Sakhalin and Vladivostok region +4.2 to Magadan region Siberia YakutiaU + 3.9 to

+ 3.0 to + 21.6 +4.9 to +8.5

Sulphur Isotope Variations in the Atmosphere

151

834S (%0) Region USSR Kazakhstan Steppes MountainsU Kirgizia mountains Tadzhikistan mountainsU Caucasus MountainsU Near Elbrusu Kola Peninsula Moscow and Novgorod regions Urals Rostov region + 3.8 to + 1.6 to + 5.1 to +5.9 +5.9 +5.7 Range Mean Reference

+2.1 to + 3.6 to

+3.7 +4.6

+3.2 +4.1 +6.9 +7.2 +0.7 +10.0 +4.0 +4.0 +5.4 +7.1

Chukhrov et al. (1977)

+5.6 to +8.6 +3.8 to +12.9 +0.7 to +2.4

+2.8 to +11.3

Gavrishin and Rabinovich (1971) Rabinovich (1971)

USA Utah, Salt Lake City Urban Rural New Zealand Gracefield Canada Sudbury, Ontario, within 90 km of smelters Great Lakes Basin Urban Remote and Rural Arctic Severnaya Zemlya, USSW

-1.5 to +8.0 to 0

+5.3 +0.2

+2.2 +9.0 +10.9

Grey and Jensen (1972)

to + 19.4

Mizutani and Rafter (1969)

+2.0 to +3 to

+6.8 +9.4b

+4.7 +6

Nriagu and,

Coker (1978a)

+2 to +6.5b

+4

+9.2 to +11.9

+10.7

M. Astratov et al. (1986)

152
Table 5.4 Continued 634S (%0) Region Antarctica George-von-Neumeyer Atlantic Ocean Pacific Ocean Coco Island, Rivera Range Mean

Stable Isotopes

Reference

Station"

+9

to + 22

+17 +13.7

+12.1 to +15.0

H. Nielsen (contributed data) Chukhrov et al. (1978) Chukhrov et al. (1978)

+9.4 to +16.3

+13.1

Measurementsin accumulated snow and ice.


Sulphate should have originated from precipitation.

b 634S values higher in winter. , River draining igneous rocks.

5.5

DELINEATION OF ANTHROPOGENIC AND NATURAL FLUXES OF ATMOSPHERIC S COMPOUNDS

5.5.1 Use of 534S and concentration data to identify sources of atmospheric gases Plots of 834S values versus various functions of concentration can prove effective for identifying sources and monitoring the fate of atmospheric sulphur pollutants. A number of hypothetical source combinations were examined by Krouse (1980). The conceptual and mathematical analyses assumed complete mixing without any isotope fractionation. In principle, the conclusions are valid for gases, ions, and fine particulates in fluids. It is noted that the approach fails in rivers where the flow is lamellar but has been successfully applied in some cases to soil and vegetation (Chapter 7). In the atmosphere, a frequently encountered situation is a source 'A' which is relatively constant in its emission rate and 834S value for a few hours or days, but upon which is superimposed a source 'B' varying in emission rate at a different fixed 834S value. The former might be the 'background' and the latter an industrial stack. The lowest concentration corresponds to contributions from source A only. With increasing concentration, the contribution from B increases so that the 8 value approaches that of B at very high concentrations (Figure 5.5a). It can be readily shown in this case that a plot of 834S versus the reciprocal of concentration is

Sulphur Isotope Variations in the Atmosphere

153

8s

I ",

8s

88A

I I .

~---

A
0
(0) CA

~
0
C -..
(b)

~
"'.
-1

SOURCE A CONSTANT SOURCE B VARIABLE


CA

C-1 -...

5.5 534S values versus (a) concentration and (b) concentration-I for mixtures Figure of sulphur from two sources. Source A with 5A has a constant emission rate. Source B with 58 varies in its emission rate

linear, the y intercept corresponding to the 034Svalue of the variable emitter B (Krouse, 1980; Figure 5.5b). This model applies over short periods of time since generally the background conditions do not remain constant due to factors such as changing wind direction and fluctuations in biological emissions. The case of a constant background and two variable sources is shown in Figure 5.6. In this case, data obtained by plotting 0 values versus the inverse of concentration fall within a triangular area. Data falling near either edge of the triangle correspond to a mixture of the background and one of the variable sources. Suppose that there is a third variable source. Its line on the 0 value versus concentration -[ plot will either fall within or outside the triangle of Figure 5.6b. In the latter case, a larger triangle results. Therefore it is difficult to delineate as to whether there are two or more variable emitters. Nevertheless, the plotting of 0 versus concentration-1 is a logical exercise that indicates the complexity of the situation. If the data do not fall within a triangle, then there was not a source emitting at a constant rate and 0 value during the observation period. If there is a basic uniform background 034Svalue for the global atmosphere or a large area thereof, then the model Figure 5.6 should apply. The value 'A' corresponding to the lowest concentration is the background. As additional sources contribute, the concentration increases as well as the range in 034S values. There should be a lower and upper limit to the atmospheric 834Svalue. These are at least -40%0 (biogenic emissions from a spring; van Everdingen et at., 1982) and + 30%0 (S02 emissions from a

...... U1 .j:>.

Table 5.5 Isotopic composition of selected atmospheric sulphur gases and particulates 834S(%0)
Type of atmosphere Stratosphere

Region 0-43S 18.3-19.8 km

Compound SO/-

Range +2.3 to +2.9 +20 to declining -24 0 to +10.0 -9 to -6 +3.6 to +9.0 +0.6 to +5.7 +7.5 to + 14.0

Mean +2.6

Remarks

References

Continental, biogenic

Great Salt Lake, Utah, USA Paige Mountain, NWT, Canada Prague, Czechoslovakia Calgary, Canada

H2S H2S
S042-

Urban

S02 SOz

+5.8 +2.3 +10.9

No (major) volcanic Castleman et al. eruption following (1974) the eruption of Mt Agung Lake margin Grey and Jensen (1972) Krouse and van Cold springs Everdingen (1983) Buzek, Cerny, and Sramek (Section 8.9) H.R. Krouse Lead peroxide exposure cylinders (contributed 1971 data)

Edmonton, Canada Red Deer, Canada New Haven, Conn., USA

S02 S02 SO/SOz

+ 10.5 to + 10.6 +14.0 to 20.9 +1.1 to +6.9 -2.1 to +8.7

+10.6 +19.0 +4.2 +2.6

1969-70 1968-70

Newman and Forrest (Section 8.4)

is' 2: "" t;;o ,g 1;;

VJ

Long

Island,

N.Y.,

SO/S02 S042S02 S02 S02

USA New York City, USA Power plants Oil Coal Coal Coal Northport, USA Keystone, USA Labadie, USA Akansk -Achinsk, USSR Chvaletice, Czechoslovakia Balzac, Alberta, Canada Whitecourt, Alberta, Canada

+ 1.9 +2.0 +3.2 -1.9 +4.1 +1.3 -5.0 +3.4

to to to to to to to to

+9.8 +4.7 +13.7 +3.9 +5.4 +3.7 +3.6 +5.0

+5.3 +2.1 +8.5 +1.5 +4.8 +2.9 -2.7 +4.1

1970-75 1970-75 1971-72 1968-72 Fluegas " " " " Forrest and Newman (1977a) (see Section 8.4) Grinenko and Grinenko (see Section 8.3) Busek et at. (see Section 8.9) Krouse and Case (1982) Krouse et at. (1984)

::: 'B"

c,..,

;:;::: ....
C)

t:;-

,g '"
....

. c:;. ;: "" s. '"


:J:..

Coal Sour gas processing plants, (see Section 8.2)

S02 S02 S02 S02 SO/SO/SO/SO/S02 S02

+1.4 to +0.1 +6 to +21 + 12 to +25 +5 to +9 +12 to +30 +6 to +11 +4.2 to +7.6 +0.9 to +2.6 + 1.1 to + 1.8

-0.7 +15 +20 +7 +21 +9 +3.7 +5.7 +1.3 +1.4

Flue gas

C) i3 ;:;'"

Smelter

Sudbury, Ontario, Canada

Downwind Upwind Downwind Upwind Three-stack samples Plume Stack and plume Downwind

Nriagu and Coker (1978b) Forrest and Newman (1977b)

>-' U1 U1

156

Stable Isotopes

~c

t
8

88 8A

.
A

II

I "

~
C~

~c, ... ...

8Bk'
8AI-

...~. .::...

SOURCE A CONSTANT SOURCES 8 AND C VARIABLE


0

0
(a)

CA

c-1 A

C-1 ---..
(b)

Figure 5.6 The 0345 values versus (a) concentration and (b) concentration-1 source A with OA and a constant emission rate and two sources Band different 0 values and varying in their emission rates

for one C with

sour gas plant; Krouse, 1980). A fundamental question is whether there are background ;)34S values which pertain to large areas of the earth. This question is addressed in Section 5.5.6. If the area surrounding a dominant industrial emitter is examined randomly on many occasions, the data tend to fit a pattern such as that shown in Figure 5.7. The spread in ;)34Svalues at low concentrations reflects many minor sources. These sources are not evident at higher concentrations which are dominated by the industrial source. A given point is primarily a function of wind direction which determines the relative contributions from the various sources. Biological activity may impose a seasonal dependence upon some sources.

A
8348

:>
Figure 5.7 The 0345 values versus concentration for a dominant source with OA and numerous minor sources with various 0345 values and emission rates. The minor sources are only evident at lower concentrations

Sulphur Isotope Variations in the Atmosphere

157 in

5.5.2 Use of 834S and concentration data to identify sources of SO/precipitation

Extension of the hypothetical models of Section 5.5.1 to precipitation and water bodies must recognize that variable evaporation and dilution may alter the [SO/-] drastically. Whereas the [S04Z-] alone may not be useful, the problem may be overcome by referring it to the concentrations of other ions. If there is a seawater SOi- component, the [Cl-] may be used (Mizutani and Rafter, 1969; Cortecci and Longinelli, 1970; Wakshal and Nielsen, 1982). The approach of these authors is illustrated in Section 5.5.4. The data of Mizutani and Rafter were alternately interpreted in terms of three sources by Krouse (1980). In that case, the approach of Figure 5.6 was used with normalized units corresponding to the amount of SO/- in a sample containing a fixed amount of Cl-. 5.5.3 Use of 8180 and concentration data to identify sources of SOiprecipitation in

If sulphate from two or more sources were dissolved in precipitation, then the mixing models discussed in Section 5.5.2 should apply. Whereas the sulphur isotope composition does not change markedly during oxidation of HzS and SOz to SO/-, the oxygen isotope composition varies drastically depending upon the mechanism and the relative amounts of oxygen incorporated from Oz and HzO molecules (Holt and Kumar, Chapter 2). If all the SO/- in precipitation resulted from in situ oxidation, then the 0180 value of SO/- should correlate well with that of the water. Such appears to be the case in Figure 5.8 where the following linear relationship was found:
01!;Osulphate = 16.0 :':: 0.9 + (0.4 :':: 0.1) 0180water

However, most data in the literature are not so well behaved. In the case of Gracefield, New Zealand, plotting of 0180 of SO/- versus that of HzO does not reveal a simple relationship (Mizutani and Rafter, 1969). For example, in this data set, the sample which had SO/- similar in both 034S and 0180 to the marine values, had a 0180 value of -15%0 for the H2O. Clearly the oxygen isotope composition of the bulk of the SO/- had been determined prior to its incorporation in the rain. In the same data set the average 0180 value for water was -5%0 but the 034S and 0180 values of SO/- were lower. If one compares average values for Gracefield (Mizutani and Rafter, 1969), Venice (Longinelli and Bartelloni, 1978), and Lublin (Figure 5.8), one finds the 0180 differences between SO/- and water are 15, 20, and

158

Stable Isotopes

. SNOW
I

+20

RAIN I

+10""" .. ..
'0

..

".. . ...

.. .
.,.

5042-

oJ! 0

. . . . .. .. -..0: .
..

co
/JO

-10

.. '.. .. .. ....
r
I

. .
I I

. . . . . . ..... .. .." . H2O . .


....
I I I I I

-201.....
II

III

IV

VI

VII VIII

IX

XI

XII

1979
Figure 5.8 The 0180 values for SOl- and H2O in precipitation in Lublin, Poland (data provided by A. Trembaczowski and S. Halas)

25%0corresponding to average 0130 values of precipitation of -5, -7, and -14%0, respectively. This implies that the ratio of previously formed SOi- to that generated by oxidation within droplets is quite different at the three locations. It is interesting to note that the average values in the Venice study fit the linear relationship found for Lublin whereas this is not the case for the New Zealand data. It is also noted that the average 0180 values for SOi- at the three locations differ by less than 3%0. In summary, mixing of previously formed SOi- with SOi- formed entirely by oxidation within precipitation droplets can be modelled and interpreted. However, in natural situations, mixing and oxidation parameters may vary to complicate the interpretation. For example, consider the emission of SOz from one hot stack. Sulphate formed near the stack would probably incorporate more Oz with quite positive olgO values (Section 2.2.4), whereas further downwind at cooler temperatures oxygen more depleted in 180 might be incorporated from H2O droplets. 5.5.4 Isotopic delineation of sources of sulphur emissions in local situations Isotopic differentiation of sources of atmospheric sulphur compounds have been most successful over distances of a few tens of kilometres. In an early study in the vicinity of a large copper smelter near Salt Lake City, Utah, USA (Dequasi and Grey, 1970; Grey and Jensen, 1972), the industrial

Sulphur Isotope Variations in the Atmosphere

159

operation effectively ceased during an extended workers' strike. It was found that the 834Sfor both precipitation and atmospheric sulphur averaged + 2%0 during smelting as compared to +6%0 during the strike. Values near +9%0 were found in a remote 'control' area. It was further concluded that during the shutdown period, about 90% of the atmospheric sulphur was biogenic whereas 10% came from other anthropogenic sources. Mizutani and Rafter (1969) determined the isotopic composition of both Sand 0 in rainwater SOi- over a five-month period in Gracefield, New Zealand. They plotted 834S and 8]80 values against a parameter A which measures the percentage of marine SO/- in a given sample on the basis of [CI-] and [SOi-]' where [CI-]
measu~ X

A = 7.1

100

X [S04Z-]

measured

On this basis (see Section 5.5.6), it was concluded that the intercept at A = 0 (no marine SOi-) corresponded to a source with 834S =0. These authors felt that they could not differentiate between biogenic and industrial contributions to this component. It is interesting to note that a three-source model with 834S values of +20 (ocean SOi-), +3, and -4%0 includes all the data points of Mizutani and Rafter (Krouse, 1980). By coincidence, Nakai et at. (see Section 8.5.5) selected values of +3.4 and -4%0 for natural HzS and anthropogenic sulphur respectively, to calculate the atmospheric sulphur balance in Japan. Studies around sour gas plant operations in Alberta have proved informative since the 834S values of the emissions (+ 15 to + 30%0) are substantially higher than those of environmental receptors (- 30 to 0%0) (Section 8.2). In a fashion similar to the copper refinery study cited above, the startup of a sour gas plant may be monitored by measuring the 834S values of ground-level SOz and particulates (Figure 5.9). The change in the isotopic composition of the SOz is better behaved than that of the particulates. Plotting of 834Sversus [SOz]-] yields straight lines for each sampling station which extrapolate to +29%0 on the y axis (Krouse, 1980). It is noted that this is consistent with data obtained by high-volume sampling of the plume with a helicopter (Figure 5.3). Examples of trends using wind directionally controlled high-volume sampling in the vicinity of sour gas plants are given in Section 8.2. The areal distribution of 834S in vegetation and soil receptors in response to atmospheric sulphur near these plants is also discussed. 5.5.5 Construction of regional balances of atmospheric sulphur As shown in the following examples and case studies in Chapter 8, stable isotopes can be used to evaluate regional balances of atmospheric sulphur.

160
+30

Stable Isotopes

0 cF-

+20

0 en
qr<>
G()

C\I

en + 10

PLANT DOWN

FULL OPERATION

16

17 JUN E 1975

18

Figure 5.9 Isotopic monitoring of a sour gas plant startup with ground-level highvolume sampling of atmospheric SOz, Crossfield, Alberta (from Krouse, 1980). The three sampling stations are indicated by different symbols

It is interesting to compare data from two different areas in the Mediterranean region: Pisa and Venice, Italy (Cortecci and Longinelli, 1970; Longinelli and Bartelloni, 1978), and northern Israel (Wakshal and Nielsen, 1982). From [CI-] and [SO/-] and isotopic measurements on S04Z-, Cortecci and Longinelli (1970) calculated that the seawater spray contribution in 64 rainwater samples at Pisa was 35% for 1 sample, 10-30% for 9 samples, and less than 10% for 54 samples. The main source of sulphate according to these authors was atmospheric oxidation of SOz from industrial emissions. The possible contribution of biogenic HzS was not evaluated. The range in 034S values (- 2.5 to + 7.1%0)was very similar to that reported for rainfall in Venice (-0.2 to +8.3%0) by Longinelli and Bartelloni (1978). The average value for the latter (+3.0%0) was slightly higher than that of the former (+1.1%0). In the Venice study, the 0180 values of SO/- were low between D~cember 1974 and January 1975, became more positive during the summer, and decreased again during the autumn and winter of 1975. Since this trend was consistent with that of the isotopic composition of the rain water, much of the SO/- was attributed to oxidation of local emissions. Sulphate crusts on

Sulphur Isotope Variations in the Atmosphere

161

monuments were also examined isotopically and the data used to argue that they formed as a consequence of industrial emissions. This is consistent with the fact that the crusts had not been observed prior to industrial activities. In northern Israel, 034Svalues for SOi- in rain water, ranging from +4.5 to + 14.3%0, were more positive than those of Pisa and Venice. In most samples, the marine SOi- contribution was calculated to be less than 40%. The SOi- was found to be more enriched in 34Son the Coastal Plain (+ 7.5 to +14.3%0; average, +9.5%0). The mean decreased to +5.3%0 upon approaching the summit of central Upper Galilea. Using the A parameter of Mizutani and Rafter (1969), the extrapolated 034Svalue for zero marine SOi- contribution was found to be between + 1 and +2%0dependent upon the sampling location. This value is not radically different from the means and the minimum 034S values found at Pisa and Venice. Nriagu and Coker (1978a) investigated sulphur isotope and [SOi-] variations over distances of the order of 1000 km in the Great Lakes of North America over two years from 1973 to 1975. Average values for urban, rural, and remote stations were found to show a strong seasonal dependence with the urban 034S values about 2%0heavier (Figure 5.10). The maximum 034S values in January-February were about 4%0heavier than the minima

+10

.1./."".'
0

~ ~
r<"J

(J) v

+ 5 I-

I. .

,'

,., .,' ,.'., ..- --II,

...

,/~I.

./'
I

'.~,

't. '"

ro

'.' "'''',

\ \ \ ./ I/1.\.~ \"'~-. I "


;6,

\.
".

I!' ,. .'

.--8

I. I'

I ,'~"

'...

'. "'.
'e'.
, I

.- -:...'.',' ,.. II
.

:I

:','

' :/

.........

Figure 5.10 Seasonal variations for average 034Svalues for precipitation in the Great Lakes Basin (after Nriagu and Coker, 1978a)

162

Stable Isotopes

occurring in July-August. In contrast, the [SOl-] fluctuated considerably and it is difficult to discern any seasonal dependence. They concluded that the seasonal trends reflected the dominance of anthropogenic inputs with greater biogenic emissions in the summer months. The biogenic component was judged to account for 10-30% of the total SOi-. The study of Nriagu and Coker (1978a) for sites generally north of the Great Lakes is complemented by those of Newman and Forrest over a larger area south and east of the Great Lakes (Section 8.4). It is noted that the range in 834S values for particulate sulphur of the latter (+ 1 to + 13%0)is slightly larger than that of SOi- in the former (+1 to +9%0). However, very few data of Newman and Forrest are above +9%0 (Figure 8.15). The similarity is not surprising, considering the large number of anthropogenic sources and the storm patterns over this large region. For these reasons, precipitation in the remote areas studied by Nriagu and Coker (1978a) contain significant anthropogenic sulphur. Nakai et at. (Section 8.5) have estimated average 834S values for three main sources of SOiin rain water in Japan. They calculated the contribution of these three sources at polluted sites by the following equations:
034SS042-

= a834Sa +

b{j34Sb

C834Sc

(1) (2)

a + b + c = 1.0

~ = 0.58 a

(3)

where a is the fraction of sea spray SOi- with 834Sa = +20.3%0, b is the fraction of SOifrom oxidation of natural H2S emissions with
834Sb

= +3.4%0,and c is the fraction of SOi- from oxidation of anthropog-

enic S02 with 834Sc = -4.0%0. They concluded that the contribution of anthropogenic S02, which was 40-50% of the total atmospheric sulphur in industrial and urban sites of Japan in 1960, increased to 75% in 1973 and then decreased to about 60%. In summary, for a number of regions of the world, rain water contains minor marine SOi- and one or more components with 834Svalues not far removed from 0%0. In many cases, these have been attributed to industrial emissions. However, this small range in 834S values suggests a common source such as marine sulphide emissions, particularly in coastal regions. At high latitudes, the seasonal dependence of 834Smay delineate anthropogenic and natural emissions as in the study of Nriagu and Coker (1978a). Nakai et at. (Section 8.5) not only isotopically differentiated anthropogenic from natural emissions in Japan but divided the latter into marine and continental.

Sulphur Isotope Variations in the Atmosphere

163

5.5.6 Estimate of the isotopic composition of biogenic sulphur in the atmosphere An average 334Svalue for biogenic sulphur released to the atmosphere from oceans and continents is required for modelling the atmospheric sulphur budget. At the present time, relevant data are sparse. Nonetheless, some general ideas may be formulated from previous investigations. In earlier studies, biogenic sulphur was identified with HzS, and it was assumed that the isotopic composition of other reduced compounds (DMS, CSz, COS, etc.) did not differ from that of HzS. The present analysis recognizes that reduced biogenic sulphur compounds originate from essentially two different sources. Firstly, HzS and possibly other reduced sulphur compounds are released during bacterial SOl- reduction. The isotope composition of this HzS depends upon the rate of SOl- reduction and redox conditions (Nakai and Jensen, 1964; Kemp and Thode, 1968; Rees, 1978) in the environment. Its 334S values vary from - 28 to -15%0 in shallow marine regions, whereas in estuaries and coastal tidal zones the range is - 22 to + 12%0(Grinenko and Grinenko, 1974). The average 334Sfor HzS and its derivatives formed under such conditions is probably close to -10%0. In freshwater lakes with low [SOl-], the evolved HzS has nearly the same 334Svalue as the SOl- (-20 to +20%0; average +8%0; SCOPE 19, 1983, Chapter 5). Secondly, HzS and other reduced sulphur compounds are evolved during decay of sea and land biota. Complete decomposition must correspond to no isotope fractionation. Since organic S of marine organisms has 334Svalues close to that of seawater SOl- (Kaplan and Rittenberg, 1964; Mekhtiyeva and Pankina, 1968), the 334Svalue of reduced sulphur compounds from this decay should be near + 18%0.The same was found for terrestrial flora and fauna on Heron Island, Great Barrier Reef, Australia (Krouse and Herbert, 1988). Since the assimilated S for freshwater lake biota is familiar in isotope composition to the SOl-, reduced S compounds from their decay should average +8%0. Although the most frequent 334Svalues for SOl- in rivers is about +6%0, the mean weighted flux to the ocean is estimated to have a 334S value of +8%0 (SCOPE 19, 1983, Chapter 5). Since the [SOl-] is low, reduction is controlled by SOl- uptake and hence will not display a large isotope selectivity. Further, under the typical aerobic conditions of flowing rivers, the amount of HzS escaping to the atmosphere should be small. When water levels drop, there may be some decay of aqueous biota which would release a variety of sulphide gases to the atmosphere. Their mean 334Svalue should also be +8%0. However, one must bear in mind that bacterial SOlreduction also occurs in continental saline water bodies. The total surface area of such water bodies is three times that of the oceanic shelf. Therefore,

164

Stable Isotopes

it may be assumedthat their sediments accumulate triple the amount of reduced sulphur (9 Tg S y-l). Since the 334S value for SO/- in these reservoirs is, on average, close to + 15%0and the rate of bacterial reduction is somewhat higher compared to that on the shelf, the mean 334Svalue for biogenic sulphide may be taken as -10%0 (SCOPE 19, 1983, Chapter 6). The amount of reduced sulphur buried in freshwater sediments is 16 Tg S y-l. If it is assumed that bacterial HzS emitted to the continental atmosphere from fresh and saline basins is proportional to sulphide buried in their sediments, then this HzS has a mean value near + 1%0 (Figure 5.11), as calculated from the mass isotopic balance equation. 9 Tg S (-10%0) + 16 Tg S (+8%0) = 25 Tg 334Smean.If the value +6%0 is used for freshwater sulphate, the mean 334S value is closer to zero. A similar mass balance equation also gives a mean 334S value near 0%0 for total sulphide emissions from the oceans. Terrestrial plants contain sulphur with an isotopic composition similar to that of soil sulphur if ambient SOz levels are low (Chapter 7). According to Chukhrov et at. (1977, 1978), the 334S values for soils and plants over the Soviet Union average + 3%0. This appears to be a reasonable average value for all soil and plant data in the literature (Sections 7.1.3 and 7.2.1).

.
<!>

-cos
RANGE

S0421 +11%.

5 0 10 Z LONG o~ a.E 00,:,:. w~ -CS2 Q:w


Q:::E

+3%.1 COS 1+18%.

0o~ w!:!:: 104


0-1 z
~::E 00-:;)

CS2

..
RANGE

CONTINENT

OCEAN---

-::E MEDIUM o- x 1-<1: 1-502 ~::E 103

- sol-

. soli
s02
1

+2%.

S02

~g z <I:
Q:

I-

S +3"1.. 2 SHORT H2S 01 - DMS RANGE+8 -10%. ~ ,.. I DMS DMS FRESH WATER SALINE 102

-H

H2S T+18%. H2S 1-10%. DMS

BACTERIAL solREDUCTION

PLANT DECAY soi-

BACTERIAL REDUCTION

Figure 5.11 Estimated 8345values for atmospheric biogenic sulphides as a function of the maximum transport distance

Sulphur Isotope Variations in the Atmosphere

165

Therefore it may be assumed that OMS, HzS, CSz, COS, etc., from biological decay have &34S values that average +3%0. At the time of writing, there were no data on the amount of HzS released to the continental atmosphere from bacterial reduction and plant decay. However, even if these fluxes differ from each other by a factor of 3, the mean &34Svalue may still be taken as + 1%0. When considering the atmospheric sulphur cycle, one should bear in mind that various reduced sulphur compounds are oxidized to SOz at different rates (SCOPE 19, 1983, Chapter 4). The average residence time is the least

for OMS

1 day). It is about 1 day for H2S and SOz, and 70 and 500

days for CSz and COS respectively (SCOPE 19, 1983, Chapter 5). Consequently, these compounds are transported over different mean distances from their place of origin (Figure 5.11). The fluxes of OMS and H2S from seas and continents should mix only within a few hundred kilometres of the coast line. The compounds that are more resistant to oxidation, such as CSz and COS, can mix over a longer range. If it is assumed that the fluxes of these sulphur compounds are proportional to the organic matter produced by photosynthesis (roughly the same for oceans and continents), then the mean &34S value for such sulphur is +11%0. Note that the contributions from plants in freshwater and saline reservoirs were omitted in this mass balance calculation. However, their extent is relatively minor and, further, their mean &34Svalue is also close to + 11%0 (for simplicity, this flux is omitted in Figure 5.11). The mean isotopic composition of total biogenic sulphide from the oceans and continents is difficult to determine since the ratio of emissions from SOi- reduction and plant decay is unknown. This problem may be solved by analysis of atmospheric sulphur isotope data published for coastal environments, as summarized in Table 5.3. If it is assumed that sampling was performed during the period of sea-to-shore winds, biogenic sulphur was then the main source of S02 and H2S. The only possible mode of S02 production should be oxidation of reduced sulphur compounds in the atmosphere. For this reason, gaseous compounds are ideal for characterizing the mean isotopic composition of biogenic sulphur from oceans. The subsequent process of S02 oxidation to sulphate should not be accompanied by significant sulphur isotope fractionation (Section 5.2.3). The data of Table 5.3 suggest that the &34Svalue for gaseous forms of sulphur varies from -4 to +6%0 with a mean value of 0%0.These data are too few and it is noted that the spread of &34Svalues is much larger in sulphate aerosols. Another approach may be used to estimate the mean isotopic composition of the biogenic sulphur flux. Sulphur isotope data have been published for rainwater SOi- in coastal zones of New Zealand and Israel (Mizutani and Rafter, 1969; Wakshal and Nielsen, 1982). It is possible to estimate the contribution of marine sulphate using [SOi-]/[CI-] ratios. It is noted that

166

Stable Isotopes

most points for both data sets are located near the line passing through a 1)34Svalue equal to +20%0 (100% marine SO/-) and approximately 0%0 (0% marine SO/-) (Figure 5.12). In Figure 5.12, a line is drawn from the 1)34Svalue of +20%0 (100% marine SO/-) to -10%0 (0% marine 8042-). The latter value is that estimated for biogenic sulphide produced by SO/- reduction. A nearhorizontal line at 1)34S just under + 20%0would correspond to an admixture of marine SO/- and biogenic sulphide derived entirely from the decay of marine plants. However, the data fall well below this hypothetical line, implying that sulphide production by this mode is smaller than by S042reduction. It is seen that, with one exception, all data fall below the line from +20%0 (100% marine SO/-) to +5%0 (0% marine SO/-). If it is assumed that the anthropogenic influence is negligible, isotope balance calculations show that the line going through 0%0corresponds to the biogenic sulphide generated from SO/- reduction, being twice that arising from decay of marine plants. Table 5.4 presents 1)34Svalues for SO/- in atmospheric precipitation over remote regions of the globe. Sulphur isotope data may be used to attempt estimates of the percentage of biogenic sulphur in the atmosphere over the open ocean. It may be assumed that the 2+20 MARINE S04 MARINE PLANTS

~~

.",
+10

",," ..

""'-

~ 0
...

'"

en <t

" ",. . '0"'-",0.. "0 .


~

eo 0

"'" .
0

~
"'", '"
0

NEW ZEALAND

ISRAEL
50

-10

L 100

A
Figure 5.12 Relationship between 334Sand A value of rain near coastal areas (data from Rafter and Mizutani, 1967, and Wakshal and Nielsen, 1982

Sulphur Isotope Variations in the Atmosphere

167

contributions of anthropogenic and volcanic sulphur in precipitation over the open ocean and oceanic islands are rather small. The rainwater SOimust be composed almost entirely of sea spray S042- and oxidized biogenic sulphur. The ;)34Svalues in precipitation over oceans vary within the range of +9 to + 16%0,with the average value of + 13%0(Table 5.4). If we accept the ;)34Svalue of 0%0for the total biogenic sulphur released from oceans, the mass isotopic balance shows that biogenic sulphur accounts for one-third and sea spray for two-thirds of the total sulphur. Of particular interest are the results of snow and ice sampling in Arctic and alpine regions. Most samples have rather high ;)34Svalues (Table 5.4). Comparative analyses of sulphur, sodium, and chloride concentrations in such samples indicate that only 10% of the sulphate is of marine origin (SCOPE 19, 1983, Chapter 4). Consequently, high ;)34S values in these samples are due to sources other than marine SOi-. It is precisely in these samples that one should find biogenic sulphur from CS2 and COS, since these compounds may be transported large distances. Such sulphur should be particularly detectable in central regions of the Antarctic, Greenland, and the Arctic. At present, the only data available are from snow and ice on Severnaya Zemlya Island (Table 5.4). The average ;)34S value was +10.7%0 and the variation between 1911 and the present time was only 1%0. This suggests that these samples were not influenced by industrial sulphur. If 10% of the sulphur is attributed to seawater SOi-, a ;)34S value of +8.8%0 is obtained for the remaining sulphur (90%). Considering the uncertainties, this value compares reasonably well with the long-range

estimate of + 11%0 based on CS2 and COS in Figure 5.11. Alternatively, if the value of + 11%0 is assumed for biogenic sulphide, one must conclude
that this component in the Severnaya Zemlya ice exceeds 70%. If the

difference between the values +8.8 and + 11%0 is real, this implies that
roughly 20% of the sulphur should be attributed to other sources, such as volcanic sulphur and bacterial SOi- reduction. Using the above arguments with other data in Table 5.4, one concludes that CS2 and COS are proportionally lower in high-mountain regions and do not exceed 50% of the total sulphur in snow and ice. It is noteworthy that in cold alpine regions of the Caucasus located between the Black and Caspian Seas, ;)34Svalues for SOi- in precipitation are close to -1%0 (Table 5.4). These are probably indicative of a major contribution from marine biogenic gases, but volcanic activity is another possible contributor.

168

Stable Isotopes

5.6 POTENTIAL OF USING ENRICHED SULPHUR ISOTOPES FOR STUDYING TRANSPORT, TRANSFORMATIONS, AND REMOVAL OF ATMOSPHERIC SULPHUR 5.6.1 Introduction The most desirable tracer for a substance is the substance itself. Consequently, it is attractive to use stable isotopes of sulphur to follow transport and transformation of SOz in the atmosphere. The technical feasibility of stable tracers must address not only the long-range transport question but also whether they could be employed in directed embedded experiments such as examining transformations in a given cloud. 5.6.2 Background fluctuations The technical feasibility can be addressed by considering questions raised in the following example. Assume that we want to follow the mixing and transformation of emitted SOz with an isotope ratio different from the background. What are the limitations for detecting this SOz as it co-mingles with the background SOz? The ultimate permitted dilution corresponds to the 834S value of the mixture SOz falling within the random variability of the 834S value of the background or alternately within the precision of the measurements. The question can be expressed by the equation E(!E) + B(l - !E) = B :t f1B where E = isotopic ratio of the emitted sulphur f E = fractional amount contributed to the background such that the change in the measured isotopic ratio of a sample either falls within the natural fluctuations of the analytical precision
B
f1B

isotopic ratio of the background

random fluctuations in the background isotopic ratio and the analytical error

The equation can be rewritten in terms of 834Svalues and rearranged to yield f E834

S B

A numerical example follows: if the emitted SOz has a 834Svalue that is 100%0higher than the background (e.g. going from a 34Sj3ZS of 0.045 to 0.0495)

Sulphur Isotope Variations in the Atmosphere

169

and the random fluctuation ofthe background is l%o,fE = 0.01. Consequently, the emitted S02 would contribute a change in the isotopic ratio that is no greater than the random fluctuation of the background when its fraction of the measured background concentration becomes 0.01. This value, or more appropriately about three times it, would be the ultimate dilution that would still permit recognition of the isotopic label. Other examples are shown in Table 5.6. Tagging the S02 with an oxygen isotope is not viable because the oxygen atoms readily exchange with those in ambient H2O. However, SOi- might be labelled with both Sand 0 isotopes for tracing its movement since the oxygens are not labile. 5.6.3 Dispersion and dilution This aspect of the problem can be considered using arguments from a longrange transport experiment. Plumes from point sources do not follow straight lines as they sweep across large areas in response to wind. However, along any individual path, the plume expands in a fairly consistent way. Depending on meteorological conditions, the average lateral dispersion generally ranges between 5 and 15 degrees. Dilution at the source obviously depends on the wind speed u. If Q is the emission rate at a point source, then the dilution associated with an assumed 'top hat' plume concentration cross-section is given by

Q = xuhD tan e
where x is the downwind concentration of the emitted substance, h the height of the mixed layer, D is the distance travelled, and consequently D
Table 5.6 Systematics of isotopic enrichment studies of atmospheric S Fraction contributed to the background by the source to make the measured isotopic ratio of a sample equal to the: The amount which the del value of the injected sulphur is different from background
(0/00)

Potential longrange and spatial fluctations of 10%0 in the background ratio

Spatially homogeneous nearfield fluctuations of 10/00 in the background ratio

Limitation based on the analytical accuracy of 0.1%0 in ratio

20 100 1000

0.5 0.1 0.01

0.05 0.01 0.001

0.005 0.001 0.0001

170

Stable Isotopes

tan e is the width of the plume at the downwinddistance.For typical conditions experienced during a long-range transport experiment (u = 10 m S-1 h = 1 km, D = 1000km, and e = 10), x
Q = 6 X 10-10 Sm-3

Newman et al. (1975a, 1975b) have found that the isotopic ratio of 34S/ 32S in the atmosphere is generally about 0.045 corresponding to 34S comprising 4.3% of the total. Random variations of this ratio have been observed to be as much as 10 del units (1%) over periods of hours (Newman, 1969; Krouse et al., 1984). For a typical atmospheric concentration of 10 ppb for S02 (13 f1g S m-3), a potential variation of 34S in the atmosphere of about 6 ng 34Sm-3 is anticipated. In order to perform a successful experiment it would seem reasonable that it is necessary to exceed this random variation by aLfactor of 10, so that the source should deliver about 60 ng 34Sm-3. Therefore the point source for a 1000-km experiment would have to emit
Q = 100 gm 34S S-1

It must be recognized that the above is an optimistic argument since transformations and dry and wet deposition have been neglected. At a distance of 500 km, measurements have shown that these processes can cause up to 90% removal (SCOPE 19, 1983, Chapter 4). The necessary emission rate can be substantially reduced if one is able to select atmospheric conditions when the random fluctuations in del are less than 10%0, perhaps as low as 1%0. In addition, W.R. Kelly (personal communication, 1984) of The National Institute of Standards and Technology suggested that by measuring both 34Sp2Sand 34S/32S ratios, one could factor out the random atmospheric variations. Such measurements can be carried out mass spectrometrically using SF6 (Chapter 3). 5.6.4 Economic evaluation Releases that were found useful in a long range transport experiment were typically of 3 hours duration (R.N. Dietz, Brookhaven National Laboratory personal communication, 1983), which corresponds to 103 kg 34S02 using the above emission rate. Such quantities are not available and it would be prohibitively expensive to produce enough material using existing facilities. For short range experiments, natural sources with unique isotope compositions can be employed (Section 8.4; Newman, 1984). A variety of isotope separation processes including liquid thermal diffusion, f@Scentrifugation, laser separation, and distillation techniques have been

Sulphur Isotope Variations in the Atmosphere

171

examined in the hope of reducing the cost by a factor of 103 in order to have an affordable experiment. This goal seems highly improbable for the immediate future (Newman, 1984). Consideration should also be given to the use of 36S since its natural abundance is only 0.017% (compared to 4.2% for 34S). Experiments could be performed by releasing fewer grams of material. However, the production costs are much higher because of its low abundance, unless one could employ laser isotope separation.
REFERENCES Andreae, M.O., and Raemdonck, H. (1983). Dimethylsulphide in the surface ocean and the marine atmosphere: a global view. Science, 221, 744-7. Astnatov, M., Zaveryaeva, I., and Ryaboshapko, A. (1986). Variations of sulfate contents in Arctic atmospheric precipitation during the last 90 years. Meteorology and Hydrology, 1, 43-6. Barnard, W.R., Andreae, M.O., Watkins, W.E., Bingemer, H., and Georgii, H.W. (1982). The flux of dimethylsulfide from the oceans to the atmosphere. J. Geophys. Res., 87, 8787-96. Beilke, S., and Gravenhorst, G. (1978). Heterogeneous SOz-oxidation in the droplet phase. Atmos. Environ., 12, 231-40. Black, M.S., Horbst, R.P., and Hitchcock, D.R. (1978). Solid absorbent preconcentration and gas chromatographic analysis of sulphur gases. Anal. Chern., 50, 848-51. Blanchard, D.C. (1963). Electrification of the atmosphere by particles from bubbles in the sea. In: Progress in Oceanography, Vol. 1, Oxford, Pergamon, Oxford, England, pp. 71-202. Bonsang, B., Nguyen, B.C., Gaudry, A., and Lambert, G. (1980). Sulfate enrichment in marine aerosols owing to biogenic gaseous sulfur compounds. J. Geophys. Res., 85, 7410-16. Calvert, J.G., Su, F., Bottenheim, J.W., and Strausz, O.P. (1978). Mechanism of homogeneous oxidation of sulphur dioxide in the troposphere. Atmos. Environ., 12, 197-226. Castleman, A.W. Jr, Munkelwitz, H.R., and Manowitz, B. (1974). Isotopic studies of the sulfur component of the stratospheric aerosol layer. Tel/us, 26, 222-34. Chambers, L.A., and Trudinger, P.A. (1979). Microbiological fractionation of stable sulfur isotopes: A review and critique. Geomicrobiol. J., 1, 249-93. Chen, K., and Morris, J. (1972). Kinetics of aqueous oxidation of sulphide by Oz. Environ. Sci. Tech., 6, 529-37. Chesselet, R., Morelli, J., and Buat Menard, P. (1972). Variations in ionic ratios between reference sea-water and marine aerosols. J. Geophys. Res., 77, 5116--31. Chukhrov, F.V., Churikov, V.S., Ermilova, L.P., and Nosik, L.P. (1977). Isotope composition of atmospheric sulfur and its possible evolution in the history of the Earth (in Russian). Izv. Akard Nauk SSSR, Ser. Geol., 7, 5-13. Chukhrov, F.V., Yermilova, L.P., Churikov, V.S., and Nosik, L.P. (1978). Sulfurisotope phytogeochemistry. Geokhimiya, 7, 1015-31. Cortecci, G., and Longinelli, A. (1970). Isotopic composition of sulphate in rain water, Pisa, Italy. Earth Planet. Sci. Left., 8, 36--40. Cox, R.A., and Sheppard, D. (1980). Reactions of OH radicals with gaseous sulphur compounds. Nature, 284, 330-1.

172

Stable Isotopes

Daum, P.H., Schwartz, S.E., and Newman, L. (1983). Studies on gas and aqueous phase composition on stratiform clouds. In: Pruppacher, M.R., Semoniz, R.G., and Slinn, W.G.N. (Ed.) Precipitation Scavenging, Dry Deposition, and Resuspension, Elsevier, N.Y., pp. 31-52. Davis, D., Ravishankara, A, and Fischer, S. (1978). SOz oxidation via the hydroxyl radical: atmospheric fate of HSOx radicals. Geophysical Research Letters, 6, 113-116. Delmas, R., and Servant, J. (1983). Atmospheric balance of sulphur above an equatorial forest. Tellus, 35B, 110-20. Delmas, R., Baudet, J., and Servant, J. (1978). Mise en evidence des sources naturelles de sulfate en milieu tropical humide. Tellus, 30, 158-68. Dequasi, H.L., and Grey, D.C. (1970). Stable isotopes applied to pollution studies. Am. Lab., December 1970, 19-27. Duce, R.A, (1981). Biogeochemical cycles and the air/sea exchange of aerosols. Working paper of SCOPE Workshop on some perspectives of the major biogeochemical cycles, Orsundsbro, Sweden. Duce, R.A., Hoffman, G.L., and Zoller, W.H. (1975). Atmospheric trace metals at remote northern and southern hemisphere sites: pollution or natural? Science, 187, 59-61. Eggleton, A., and Cox, R. (1978). Homogeneous oxidation of sulphur compounds in the atmosphere. Atmos. Environ., 12, 227-30. Eriksen, T.E. (1972). Sulfur isotope effects. I and II, Acta. Chern. Scand., 26, 573-84. Eriksson, E. (1959). The yearly circulation of chloride and sulphur in nature; meteorological, geochemical and pedological implications. Part 1. Tellus, 11, 375-403. Eriksson, E. (1960). The yearly circulation of chloride and sulfur in nature; meteorological, geochemical and pedological implications. Part II. Tellus, 12, 63-109. Forrest, J., and Newman, L. (1973). Sampling and analysis of atmospheric sulphur compounds for isotope ratio studies. Atmos. Environ., 7, 561-73. Forrest, J., and Newman, L. (1977a). Further studies on the oxidation of sulfur dioxide in coal fired power plant plumes. Atmos. Environ., 11, 465-74. Forrest, J., and Newman, L. (1977b). Oxidation of sulfur dioxide in the Sudbury smelter plume. Atmos. Environ., 11, 517-20. Forrest, J., and Newman, L. (1978). Oxidation of sulfur dioxide in power plant plumes. Am. Inst. Chem. Eng., 74, 48-53. Forrest, J., Klein, J.H., and Newman, L. (1973). Sulphur isotope ratios of some power plant flue gases: a method for collecting the sulphur-oxide. J. Appl. Chem. Biotech., 23, 855-63. Garland, J. (1978). Dry and wet removal of sulphur from the atmosphere. Atmos. Environ., 12, 349-62. Gavrishin, AL, and Rabinovich, A.L. (1971). On the possibility of using the sulphur isotopic composition of sulphate in natural water as an indicator of chalcedon deposits in the middle Urals (in Russian). Geokhimiya, 7, 873-5. Gravenhorst, G. (1978). Maritime sulfate over the North Atlantic. Atmos. Environ., 12 (1-3, Sulfur Atmosphere), 707-13. Grey, D.C., and Jensen, M.L. (1972). Bacteriogenic sulphur in air pollution. Science, 177, 1099-100. Grinenko, V.A, and Grinenko, L.N. (1974). Geochemistry of Sulphur Isotopes, Nauka, Moscow, 272 pp.

Sulphur Isotope Variations in the Atmosphere

173

Hales, J., Wilkes, J., and York, J. (1974). Some recent measurements of HzS oxidation rates and their implications to atmospheric chemistry. Tel/us, 26, 277-83. Hegg, D., and Hobbs, P. (1978). Oxidation of sulphur dioxide in aqueous systems with particular reference to the atmosphere. Atmos. Environ., 12, 241-54. Hitchcock, D.R. (1976a). Microbiological contributions to the atmospheric load of particulate sulfate. In: Nriagu, J.O. (Ed.) Environmental Biogeochemistry, Vol. I, Ann Arbor Science Publishers, Ann Arbor, Mich., pp. 351-67. Hitchcock, D.R. (1976b). Atmospheric sulfates from biological sources. 1. Air. Pol/ut. Control, 26 (3), 210-15. Hitchcock, D.R., and Black, M.S. (1984). 34S;3zS evidence of biogenic sulfur oxides in a salt marsh atmosphere. Atmos. Environ., 18, 1-17. Hitchcock, D.R., Black, M.S., and Herbst, R.P. (1978). Sulphur isotope studies of biogenic sulphur emissions at Wallops Island, Virginia. Document P-2058 prepared for National Science Foundation. Hockheiser, S., and Elfers, L.A. (1970). Automatic sequential sampling of atmospheric HzS by chemisorption on mercuric chloride-treated paper. Environ. Sci. Tech., 4, 672-6. Holt, B.D. (1975). Determination of stable sulphur isotope ratios in the environment. Prog. Nuclear Energy, Anal. Chem., 12, 11-26. Huygen, C. (1963). The sampling of sulphur dioxide in air with impregnated filter paper. Analyt. Chem. Acta, 28, 349. Ivanov, M.V. (1978). Influence of micro-organisms and microenvironment on the global sulphur cycle. In: Krumbein, W.E. (Ed.) Environmental Biogeochemistry and Geomicrobiology, Vol. 1, Ann Arbor Science Publishers, Mich., pp. 47-61. Jensen, M.L., and Nakai, N. (1961). Sources and isotopic composition of atmospheric sulfur. Science, 134 (3496), 2102-4. Kaplan, LR., and Rittenberg, S.c. (1964). Microbiological fractionation of sulphur isotopes. J. Gen. Microbiol., 34, 195-210. Kaplan, LR., Emery, K.O., and Rittenberg, S.e. (1963). The distribution and isotopic abundance of sulphur in recent marine sediments of southern California. Geochim. Cosmochim. Acta, 27, 297-331. Kemp, A.L., and Thode, H.G. (1968). The mechanisms of the bacterial reduction of sulphate and of sulphite from isotope fractionation studies. Geochim. Cosmochim. Acta, 32, 71-91. Krouse, H.R. (1977). Sulphur isotope abundances elucidate uptake of atmospheric sulphur emissions by vegetation. Nature, 265, 45-6. Krouse, H.R. (1980). Sulphur isotopes in our environment. In: Fritz, P., and Fontes, J.e. (Eds.) Isotope Geochemistry, Vol. 1, The Terrestrial Environment, Elsevier, Amsterdam, pp. 435-71. Krouse, H.R., and Case, J.W. (1981). Sulphur isotope ratios in water, air, soil and vegetation near Teepee Creek Gas Plant, Alberta. Water, Air, Soil Pol/ut., 15, 11-28. Krouse, H.R., and Case, J.W. (1982). Sulphur isotope tracing of acid emissions into the environment of Alberta. In: Proceedings of Symposium on Acid Forming Emissions in Alberta and Their Ecological Effects, Edmonton 1982, Alberta Department of Environment Canadian Petroleum Association, Oil Sands Environmental Study Group, pp. 207-32. Krouse, H.R., and Case, J.W. (1984). Design and development and field testing of a mobile nine unit high volume air sampler array. Final Report to Alberta Environment, 45 pp. Krouse, H.R., and Herbert, M.K. (1988). Sulphur and carbon isotope studies of

174

Stable Isotopes

food webs. In: Kennedy, B.V., and LeMoince, G.M. (Eds) Diet and Subsistence, Proceedings of 19th Annual Conference of CHACMOOL, Arch. Association of the University of Calgary, pp. 315-22. Krouse, H.R., and van Everdingen, R.O. (1983). 834Svariations in vegetation and soil exposed to intense biogenic sulphur emissions near Paige Mountain, N.W.T., Canada. Water, Air, Soil Pollut., 23, 61-7. Krouse, H.R., Legge, A.H., and Brown, RB. (1984). Sulphur gas emissions in the boreal forest: the West Whitecourt case study V: stable sulphur isotopes. Water, Air, Soil Pollut., 22, 321-47. Leahy, D.F., Siegel, P., Klotz, P., and Newman, L. (1975). The separation and characterization of sulphate aerosol. Atmos. Environ., 9, 219-29. Liu, B.Y., Raabe, O.G., Smith, W.B., Spencer, H.W. III, and Kuykendal, W.B. (1980). Advances in particle sampling and measurement. Environ. Sci. Tech., 14, 392-7. Longinelli, A., and Bartelloni, M. (1978). Atmospheric pollution in Venice, Italy, as indicated by isotopic analyses. Water, Air, Soil Pollut., 10, 335-41. Lovelock, J.E., Maggs, R.J., and Rasmussen, R.A. (1972). Atmospheric dimethyl sulfide and the natural sulfur cycle. Nature, 237, 452-3. Ludwig, F.L. (1976). Sulfur isotope ratios and the origins of the aerosols and cloud droplets in California stratus. Tellus, 28, 427-33. McClenny, W.A., Shaw, R.W., Baumgardner, R.E., Paur, R., Coleman, A., Braman, R.S., and Ammons, J.M. (1979). Evaluation of techniques for measuring biogenic airborne sulphur compounds. U.S. Environmental Protection Agency, EP A-600/2-79-004. McIntyre, F. (1970). Geophysical fractionation during mass transfer from sea to air by breaking bubbles. Tellus, 22, 451-61. McIntyre, F., and Winchester, J.W. (1969). Phosphate ion enrichment in drops from breaking bubbles. 1. Phys. Chem., 73, 2163-9. Maroulis, P.J., Torres, A.L., and Bandy, A.R. (1977). Atmospheric concentrations of carbonyl sulfide in the southwestern and eastern United States. Geophys. Res. Leu., 4, 510--12. Mayer, R., and Ulrich, B. (1978). Input of atmospheric sulphur by dry and wet deposition to two central European forest ecoysystems. Atmos. Environ., 12, 375-8. Mekhtiyeva, V.L., and Pankina, R.G. (1968). Isotopic composition of sulphur in aquatic plants and dissolved sulphate in water bodies. Geokhimiya, 6, 739-42. Mizutani, Y., and Rafter, T.A. (1969). Oxygen isotopic composition of sulfates-part 5. Isotopic composition of sulfate in rain water, Gracefield, New Zealand. N.Z. 1. Sci., 12, 69-80. Morelli, J. (1977). Contribution a l'etude du cycle du potassium marin. These d'Etat, Universite de Paris VI. Nakai, N., and Jensen, M.L. (1964). The kinetic isotope effect in the bacterial reduction and oxidation of sulfur. Geochim. Cosmochim. Acta, 28, 1893-912. Nakai, N., and Jensen, M.L. (1967). Sources of atmospheric sulfur compounds. Geochem. J., 1, 199-210. Natusch, D.R.S., Klonis, H.B., Axelrod, RD., Teck, R.T., and Lodge, J.P. Jr (1972). Sensitive method for measurement of atmospheric hydrogen sulphide. Anal. Chem., 44, 2067-70. Newman, L. (1969). The atmospheric diagnostics program at Brookhaven National Laboratory. In: Tucker, W.D. (Ed.) Second Status Report, BNL 50206 (T-553). Newman, L. (1980). Atmospheric oxidation of sulfur dioxide. In: Shriner, D.S., Richmond, C.R., and Lindberg, S.E. (Eds.) Proceedings of Second Life Science

Sulphur Isotope Variations in the Atmosphere

175

Symposium on Atmospheric Sulfur Deposition Environmental Impact and Health Effects, Ann Arbor Science, Ann Arbor, Mich., pp. 129-43. Newman, L. (1981). Atmospheric oxidation of sulphur dioxide: a review as viewed from power plant and smelter plume studies. Atmos. Environ., 15, 2231-9. Newman, L., Smith, M., Forrest, J., Tucker, W., and Manowitz, B. (1971). A tracer method using stable isotopes of sulfur. In: Proceedings of American Nuclear Society Topical Meeting on Nuclear Methods in Environmental Research, Columbia, Mo. Newman, L., Forrest, J., and Manowitz, B. (1975a). The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulfur dioxide in the plume from an oil-fired power plant. Atmos. Environ., 9, 959-68. Newman, L., Forrest, J., and Manowitz, B. (1975b). The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulfur dioxide in the plume from a coal-fired power plant. Atmos. Environ.., 9, 969-77. Nguyen, B.C., Bonsang, B., and Lambert, G. (1974). The atmospheric concentration of sulphur dioxide and sulphate aerosols over antarctic, subantarctic areas and oceans. Tellus, 26, 241-9. Nguyen, B.C., Gaudry, A., Bonsang, B., and Lambert, G. (1978). Reevaluation of the role of dimethyl sulfide in the sulfur budget. Nature, 275, 637-9. Nguyen, B.c., Bonsang, B., and Gaudry, A. (1983). The role of the ocean in the global atmospheric sulfur cycle. Jour. Geophys. Res., 88, No. CIS, 10.903-10.914. Nielsen, H. (1974). Isotopic composition of the major contributors to atmospheric sulphur. Tellus, 26, 213-21. Nriagu, J.O., and Coker, R.D. (1978a). Isotopic composition of sulfur in precipitation within the Great Lakes basin. Tellus, 30, 365-74. Nriagu, J., and Coker, R.D. (1978b). Isotopic composition of sulfur in atmospheric precipitation around Sudbury, Ontario. Nature, 274, 883-5. Okita, T. (1970). Filter method for the determination of trace quantities of amines, mercaptans and organic sulphides in the atmosphere. Atmos. Environ., 4, 93-102. Okita, T., Lodge, J.P. Jr, and Axelrod, H.D. (1971). Filter method for the measurement of atmospheric hydrogen sulphide. Environ. Sci. Tech., 5, 532-4. Ostlund, G. (1959). Isotopic composition of sulphur in precipitation and seawater. Tellus, 11, 478-80. Paugam, J.Y., Nguyen, B.C., Bonsang, B., and Fongang, S. (1977). Production de noyaux aitken et de composes sourfres dans l'air au-dessus d'une zone littorale. Chemosphere, 6, 333-9. Petrenchuk, O.P. (1980). On the budget of sea salts and sulfur in the atmosphere. J. Geophys. Res., 85, 7439-44. Platt, U. (1978). Dry deposition of S02. Atmos. Environ., 12, 363-8. Rabinovich, A.L. (1971). Sulphur isotopic composition of sulphate ions in some surface waters and factors of their formation. Abstract of Ph.D. thesis, Novocherkask Hydrochemical Institute, 20pp. Rafter, J.A., and Mizutani, Y. (1967). Oxygen isotope composition of sulphate. Preliminary results on oxygen isotopic variations in sulphates and relationship to their environment and to their 34Svalues. N.z. J. Sci., 10,816-40. Rees, C.E. (1978). Sulphur isotope measurements using S02 and SF6. Geochim. Cosmochim. Acta, 42, 383-9. Saltzman, E.S., Brass, G.W., and Price, D.A. (1983). The mechanism of sulfate aerosol formation chemical and sulfur isotope evidence. Geophys. Res. Left., 10, 513-16. Sandalla, F.J., and Penkett, S.A. (1977). Measurements of carbonyl sulphide and carbon disulphide in the atmosphere. Atmos. Environ., 11, 179-99.

176

Stable Isotopes

SCOPE 19 (1983). Ivanov, M.V., and Freney, J,R. (Eds.) TheGlobal Biogeochemical Sulphur Cycle, Chichester, 470 pp. Sprung, J. (1974). Tropospheric oxidation of H2S. In: Metcalf, R., Lloyd, A., and Pitts, J. (Eds.) Advances in Environmental Science and Technology, Wiley, New York, 400 pp. Thiemens, M.H. (1977). The atmospheric oxidation of H2S and the oxygen isotopic ratios of tropospheric sulphates. Ph.D. dissertation, Florida State University, 203 pp. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1982). Isotope geochemistry of dissolved, precipitated, airborne, and fallout sulfur species associated with springs near Paige Mountain, Norman Range, N.W.T. Can. J. Earth Sci., 19, 1395-407. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1985). Role of corrosion by H2SO4 fallout in cave development in a travertine deposit-evidence from sulfur and oxygen isotopes. Chern. Geol., 49, 205-11. Vinogradov, A.P., Grinenko, V.A., and Ustinov, V.I. (1962). Isotopic composition of sulfur compounds in the Black Sea. Geochemistry, 10, 973-97. Wakshal, E., and Nielsen, H. (1982). Variations of 034S (SOl-), 0180 (H2O) and Cl-/SO/- ratio in rainwater over northern Israel, from the Mediterranean Coast to Jordan Rift Valley and Golan Heights. Earth Planet. Sci. Leu., 61, 272-82. Weyer, K.U., Krouse, H.R., and Howard, W.e. (1979). Investigations of regional geohydrology of south of Great Slave Lake, N.W.T., Canada, utilizing natural sulphur and hydrogen isotope variations. In: Isotope Hydrology 1978, Vol. I, Proceedings of IAEA, Vienna, pp. 251-64.

CHAPTER 6

Hydrosphere
1.0. NRIAGU, C.E. REES, V.L. MEKHTIYEVA, A. Yu. LEIN, P. FRITZ, R.l. DRIMMIE, R.G. PANKINA, R.W. ROBINSON,AND H.R. KROUSE

6.1

INTRODUCTION*

Although isotope hydrology is a well-established field, the sulphur isotope systematics of natural waters have not been extensively studied. The isotopic composition of sulphur in sea water is remarkably constant, and that of sulphur in fresh waters so variable that it is difficult to gain fundamental insights of key processes and pathways. With the exception of the pioneering work of Deevey and his collaborators in the early 1960s (see Section 6.4), little was published on the isotope geochemistry of sulphur compounds in unpolluted freshwater ecosystems until the early 1970s. With the recent development of 'environmental isotope' techniques, a number of detailed investigations were undertaken to (a) identify sources of sulphur; (b) determine the pollution component; and (c) trace the fate of sulphur in aquatic ecosystems. Simultaneously, significant advances were made in the use of sulphur and oxygen isotopes as complimentary tools in other areas of applied geochemistry. Of particular interest is the use of isotopic techniques to (a) 'date' ground water; (b) identify the recharge area and origin of sulphur; (c) infer groundwater circulation patterns, and (d) evaluate the migration rate, reactions, and extent of equilibration of the sulphur compounds (see Section 6.7). More recently, interest in sulphur isotope measurements has been heightened by the association of sulphur pollution with acid rain. Acid rain is threatening fisheries' resources and water quality of many lakes and rivers in Europe and North America. Variations in the isotope abundances of sulphur and oxygen in sulphate and sulphur oxides are ideally suited for acid rain studies. Isotopic compositions can be used for fingerprinting the sources of sulphur pollution as well as monitoring the long-range and transboundary movement of the pollutant sulphur. This chapter includes illustrative case studies pertaining to stable sulphur/oxygen isotopes in relation to lake and river acidification.
* J.O. Nriagu.

Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko <9 1991 SCOPE, Published by John Wiley & Sons Ltd

178

Stable Isotopes

Foremost, this chapter should serve to emphasize the fact that sulphur/ oxygen isotopy provides the basis for a powerful but undeveloped tool in the aquatic sciences. Only recently have sulphur isotopes been applied to foodweb studies. Furthermore, the available data are sparse and focus mainly on local areas in Europe and North America. Little has been done on sulphur isotope systematics in the ecosystems of developing countries. With the current data base, it would be premature to attempt even a rudimentary model of sulphur isotope balance on a regional or global scale. An isotopic balance model should constrain current estimates of sources (natural versus anthropogenic) and elucidate sulphur transfer processes and the rates of sulphur turnover in natural waters. In other words, it should provide much needed checks for various models of the global sulphur cycle. 6.2 6.2.1 Dissolved sulphate* Sulphur is present in the ocean in the form of dissolved sulphates, predominantly NaS04 -, MgS04, CaS04 -, KS04 -, and SOi- (Garrels and Thompson, 1962; Kester and Pytkowicz, 1969), with an average concentration of 0.904 g S kg-l (Horn, 1969). Taking the mass of water in the world ocean to be 1.43 x 1012Tg gives the total mass of sulphate sulphur as 1.277 x 109Tg S (Horn, 1969). The flux of sulphur to and from the ocean is ~300 Tg S yr-I (see below) so that the residence time of sulphur in the ocean is 1.277 X 109 Tg S -7-300 Tg S yC 1 or ~4 X 106 yr. In contrast to the long residence time of sulphate in the ocean, the
~

OCEANS

relatively short oceanic mixingtime of

1000years ensures that the sulphur

isotopic composition of ocean sulphate is essentially uniform except in regions where local influences such as the mouths of rivers are important. Measurements of the 034S values of suites of ocean water samples have been made by Ault and Kulp (1959), Thode et at. (1961), Sasaki (1972), and Rees et at. (1978). The results of these studies are summarized in Table 6.1. The mean values obtained in the first three studies agree very well whereas the decrease in the spreads of 034S values obtained in successive studies is indicative of the improvement with time of analytical precision. The mean value of +20.99%0 obtained by Rees et at. (1978) is markedly different from the mean values of around +20.0%0 obtained in the earlier studies. This difference is due to the fact that this latter study was made using sulphur hexafluoride (SF6) as the gas for isotope analysis whereas the previous studies employed sulphur dioxide (S02). Rees (1978, Section 2.1)

* C.E Rees.

Hydrosphere

179

Table 6.1 Summary of 834Sdeterminations for present-day ocean water sulphate Number Range of of samples 834Svalues N (%0) +18.9 to +20.7 +19.3 to 20.8 +19.62 to 20.32 +20.74 to +21.12 Spread of Mean 834S 834S (%0) (%0)
(J" In
(Jb I

Study

(%0)

(%0)

Ault and Kulp (1959) Thode et ai. (1961) Sasaki (1972) Rees et ai. (1978)
a

25 16 20 25

1.8 1.5 0.70 0.38

+20.0 +20.1 +20.03 + 20.99

0.1 0.10 0.05 0.02

0.54 0.38 0.21 0.08

Standard deviation of the mean.


Standard deviation of an individual determination 0"; = O"m YN.

points out that results obtained using SF6 are more precise and, more importantly, more accurate than those obtained using S02' It is of interest to speculate on whether the sulphur isotope composition of present-day ocean sulphate is constant or if its 034S value is increasing or decreasing with time. The 034S value of ocean water sulphate is controlled by the magnitudes of the various fluxes to and from the oceanic sulphate reservoir and by the isotopic compositions of these fluxes. Examination of the 034S values of evaporitic sulphates which were derived from the dissolved sulphate in ancient oceans has shown that the isotopic composition of oceanic sulphate has not been constant over geological time (Thode and Monster, 1964, 1965; Nielsen, 1965; Holser and Kaplan, 1966; Schidlowski et al., 1977; Claypool et al., 1980). The implication of these variations in terms of variations of the relative rates of evaporite and sedimentary sulphide formation, as well as of the global atmospheric oxygen budget, has been discussed by the authors cited above and by Rees (1970) and Holland (1973). The topic is covered in detail by Pilot (Section 4.2). With regard to the present-day ocean, it is important to note that its sulphate content is possibly not in balance. Table 6.2 contains some flux data taken from Table 7.1 of Ivanov (1983). Some of the individual fluxes listed by Ivanov have been combined for the purpose of the calculations developed later in this section. The total flux of sulphur to the ocean (P 10 + P20) is estimated to be ~480 Tg S yr-l while the total flux of sulphur from the ocean (PIs + PI9 + P21 + P22) is estimated to be ~Tg S yr-I.

Thus the estimated input exceeds the estimated ouput by


or ~60%.

180Tg S yr-I

Table 6.2 Calculations

of the rate of change

of the present-day

O,4S value of ocean water sulphate (Tg S yr -I) 6 7


------

>--'

00 0

Flux values Fluxes (Ivanov, 1983) This work O,4S value (%0)
Og

10

11

Continental atmosphere to Pg oceanic atmosphere (Px) and volcanic emissions (P17) PIO River (PIO) , underground runoff (PI I)' shore abrasion (P13) Biogenic emission to Pig oceanic atmosphere (PIx) PI9 Ocean spray sulphate to oceanic atmosphere (PI9) P211 Atmospheric sulphur compounds to ocean (P211) Removal of reduced P21

= +5

110

160

010 = +5

220

270

0/\

- ex
0/\

20 140 260

70 190 310 19() 3W

OAT 0/\

- f3
0/\

110 30
-

160 80
----------

sulphur in sediments (P21) Pn Removal of sulphate in sediments (Pn)

ex(%0) f3 (%0) 034Sof atmospheric sulphate (%0) 034Socean sulphate (%0per million years)

15 50
-_._----

20

15 40

15 5()

+13.4 +12.1 + 13.4 + 12.2 + 14.6 + 13.4 + 13.4 + 13.4 + 13.0 + 13.4 + 14.6 +0.2 0 -0.4 +0.6 +0.5 -0.1 +2.2 +0.2 +0.2 -0.6 +0.2

Hydrosphere

181

It is not clear whether this estimated difference between input and output is real or if it is an artefact of the uncertainties in the individual flux estimates. For example, previous estimates of the biogenic emission from the ocean have varied widely-170 Tg S yr-I (Eriksson, 1960),30 Tg S yr-I (Robinson and Robbins, 1970), 48 Tg S yr-I (Friend, 1973), 27 Tg S yr-l (Granat et al., 1976). Recently Nguyen et al. (1978) have estimated a marine flux of dimethyl sulphide (DMS) of at least 89 Tg S yr-l, possibly in addition to other biological contributions. Because there are uncertainties in the various flux estimates, calculations are performed below to determine the rate of change with time of the present-day 034S value of oceanic sulphate for both the flux estimates presented by Ivanov (1983) and for other possible values. Let A represent the present-day value of the sulphate sulphur content of the ocean and let Ii> Fj (i = 1,2,...; j = 1,2, . . .) represent the various present-day fluxes of sulphur into (1) and from (F) the ocean. The rate of change of A with time is dA = 2.Ii - 2. Fk
dt i J For 32S and 34S write
32 d32A

dt

= 2. 32]i - 2.
i j

Fj

and

d34A
dt

= 2. 34]i - 2.
i j

34Fj

This last equation can be expressed in terms of del values by writing


34A
34] . I

= 32A Ro(1 + OA) = 321 R ( 1 + 0, )


I 0 i

and
34

FJ. =

32FR
J

( 1 + OF) J

where Ro represents the 34Sf32Svalue of the standard reference material (Canyon Diablo troilite) and where A, OJ, OF represent the fractional del values of A, J, and F relative to the standard reference material. Making these substitutions in the equation for d34A/dt gives

182

Stable Isotopes

dp2A Ro(l

+ OA)]= 2: 321iRo(1 i

+ 01) -

2: ]

320

R,(1 + OFj)

Cancelling Ro throughout and making the assumption that the behaviour of the element, A, and that of the abundant isotope, 32A, are essentially identical gives d[A(1 + OA)] = 2: li(1 + 01) dt i

2:0(1 + 1

OF)

so that, expanding the differential,

(1 + OA)dt + A dt = L.. li(1 + 01) - L.. 0(1 + OF) [ ]


Using

dA

dOA '"

'"

dA

dt

= 2:
i

Ii

- 2:
]

Fj

and rearranging the previous equation gives

dOA '"

dt

= L.. li(o/i - OA)- L.. 0 (OFj- OA) [ ]

'"

In Table 6.2, 034S assignments for the various sulphur fluxes are given. The assignment of +5%0 for PI',and PlO may not be exact but will serve for calculation purposes. The 034Svalue of the flux of atmospheric sulphate to the ocean, P2o, is assumed to be equal to the 034S value of atmospheric sulphate and is designated OATin the table. It will be assumed that OATis determined by the various fluxes to the oceanic atmosphere:
UAT-

"

osPs + OAPI9+(OA
PS+PI9+P1S

ex)PIs

This is equivalent to assuming that none of the fluxes from the reservoir of oceanic atmospheric sulphate involves isotope fractionation relative to that reservOir. The equation for doA/dt can now be written as

dt

dOA

PIO(OIO

- OA) + P20(OAT - OA)

Hydrosphere

183

- PIS(OA
-

IX

- OA) - PI9(OA - OA)

-P2I(OA - 13 - 8A) - P22(8A - 8A) PIO(81O - 8A) + P2o(8AT - 8A) + IX Pu-: + 13 P21

In this expression, IXrepresents the difference in 834Svalues of ocean water sulphate and the biogenic sulphur flux to the atmosphere, while 13represents the difference in the 834Svalues of ocean water sulphate and sulphide being deposited in the ocean sediments. Column 1 of Table 6.2 shows the flux values quoted by Ivanov (1983) rounded off to the nearest 10 Tg S yc 1. These fluxes, their assigned del values, and values for IX(15%0)and 13(50%0)were inserted in the equations developed above. The calculated values of 834Sfor the atmosphere and the rate of change of 834Sof ocean sulphate are + 13.4 and +0.2%0 per million years respectively. The extent to which these values (tabulated at the bottom of column 1) are realistic depends on the accuracies of the assigned fluxes and del values. Columns 2 to 10 show the effects of altering the various parameters individually including IX and 13.Each flux is increased in turn by 50 Tg S yr-I. Thus, for example, column 4 shows that if all the parameters are correct except for the estimate for PIs (the biogenic emission of sulphur to the atmosphere should be 70 Tg S yr-I rather than 20 Tg S yc1), then the rate of change of 834Sof ocean sulphate should be +0.6 rather than +0.2%0 per million years. Column 11 differs from column 1 in that PI9 and P20 have been altered together. Altering P19 alone changes the estimated rate of change of 834S from +0.2 to +0.5%0 per million years (column 5). Alteration of P19 and P20 together gives no net change (compare column 1 with column 11). The values in Table 6.2 show that it is not possible to state unequivocally whether the 834Svalue of ocean sulphate is currently increasing, decreasing, or constant. A clear answer requires estimates of the various important fluxes to a much higher degree of certainty than is presently possible. 6.2.2 Sulphur of aquatic organisms* The hydrosphere is a major habitat which actively influences the distribution of chemical elements, particularly sulphur. The total S in animal and plant tissue varies from less than 1 up to 9% by dry weight. On a dry weight basis, the S content of aquatic plants and animal tissues varies from 0.5 to 5 and 0.1 to 3.3% respectively (Vinogradov, 1953; Kaplan et al., 1963; Mekhtiyeva and Pankina, 1968; Mekhtiyeva et al., 1976; Mekhtiyeva, 1971). (Unless stated otherwise, percent S will be given on a dry weight basis.)
* V.L. Mekhtiyeva.

184

Stable Isotopes

In order to understand their roles in the hydrospheric sulphur cycle, it is important to know the mass and isotopic composition of sulphur forms in marine organisms and changes in these parameters after their death. The available information is sparse. Mollusc shells, composed mainly of CaC03 (up to 95%), contain sulphate and iron sulphur compounds. The average sulphur content is 1% but certain species contain up to 4%. Vinogradov (1953) concluded that sulphur also occurs in conchiolin, the external epidermic layer of shell composed of scleroprotein. Cephalopoda are particularly rich in sulphur (Vinogradov, 1953). The S content of shells of molluscs from water bodies of different salinities is in the range of 0.01 to 0.13% (Mekhtiyeva, 1974). Kaplan et al. (1963) found 0.26% S in shells of Chlamis latiaurata. There is no direct correlation between water salinity and sulphate content in mollusc shells due to metabolic factors which are species and age dependent. However, the sulphate content in shells generally increases with increasing [SOi-].

. SO~2
0

~ PLANTS . ANIMALS

MOLLUSCS ZOSTERA MANGROVE

0
+10
3) 834S

~ Q
I:: co Q) U "'1::", oiJI::. '" co.E ~ ~~ 00 U i! <;' .~ '0 ~
:;:t::; "uo o..g:

0 .~

.!!1 ;:(

c:

m. oiJ u r5i ~ 0 '" co~ I:: ~ Q)go .!!1 0 Cf)!f, "0 Q) 0


. .~E .s:::,,:5:>'-

~ ~ co] Q).=; Cf) ~ ~


a; CD

0 "

> 0
0

CO:2: N Q) Cf)~ Q):; _0 :E~ :5:s

~ ~ ~ ~ "..c:
.
z

'"

~ ~~

Q)" ~Cii .Cf)~~.E Lu


00.8 ~

co::::.::;

Q)
Q) co J

00
Q) ;>. 0 ~ 00 I::

a:

cr.i Q)

15 C{j ;:: 'a.~ OON ~ I::~

CO"::J 0'" a..


Q) ~ co J

OJ, F"

..s::: -;>. u .c :) ::J 0 a:

Q) ;>. ~ 00 > 0

Figure 6.1

Hydrosphere

185

The isotopic composition of sulphate in shells of molluscs is practically identical to that of dissolved sulphate in their habitat (Kaplan et al., 1963; Mekhtiyeva, 1974; Figure 6.1). After an organism's death, the amount of sulphur and 334Svalues in skeletons do not change. This fact may be used for palaeohydrochemical reconstructions of sediment accumulation in ancient basins (Mekhtiyeva, 1974). Trace sulphide and sulphate can be extracted from geological specimens using Kiba reagent under vacuum conditions (Ueda and Sakai, 1983). Using this technique, a Puaua shell from New Zealand was found to have 334S values for sulphate (490 ppm S) and sulphide (30 ppm S) of +21.4 and +7.7%0 respectively (Krouse and Ueda, 1987). The error for the latter is relatively large because of its low concentration. The chitinic tubes of polychaeta are depleted in 34S by about 13%0 compared to dissolved sulphate. This is apparently due to the presence of pyrite inclusions (Kaplan et al., 1963). Sulphur is generally supplied to plants in the form of sulphate which is reduced in the course of metabolism to thiol groups reacting with serine. As a result, cysteine is produced which is the precursor of the whole series of sulphur-containing amino acids. Natural thiols and disulphides are readily oxidized to produce numerous derivatives of sulphur oxide compounds. In addition to sulphate, plants may assimilate other sulphur compounds, including hydrogen sulphide (Thomas et al., 1943; Ellis, 1969; Faller, 1972; Brunold and Erisman, 1974, 1975; Fry et al., 1982; see also Chapter 7). The isotopic composition of sulphur in aquatic plants is slightly lighter than that of dissolved sulphate (Figure 6.1). The depletion in 34Sfor total sulphur in

40
Pacific Ocean Indian Ocean'

--0
'0'2-

White Sea

. Black Sea ;: I- 20
CJ)

Z :J

. Asov Sea
0 0
Roublyovskoye Reservoir 18 I I I

SULPHUR CONTENT (%)


Figure 6.2

186

Stable Isotopes

aquatic plants is on the average of 5%0for freshwater samples, 1.8%0for the Black Sea, 2.3 %0for the White Sea, 6.5 %0for the Azov Sea, and 0.8%0 for the California Gulf. However, one cannot use the average values since species age and season influence metabolic processes and sulphur isotope composition. The isotope fractionation of sulphur in plant tissue is determined by two factors: diffusion through the cell membrane and biochemical sulphate reduction within the cell. The isotopic effect during SOi- reduction to H2S in plant cells is essentially determined by the relatively slow initial S-O bond rupture (SOi- to SOj-) whereas the SOj- to H2S reduction steps are extremely fast. Further transformations of sulphur without a valency change are not accompanied by significant selectivity. In most cases, assimilatory sulphate reduction does not lead to significant isotope fractionation. For aquatic organisms, the k32/k34 value is usually within the range of 1.000-1.003. Plants in mangrove marshes are unusual in that their total sulphur may be 20%0lighter than aqueous sulphate. This is attributed to H2S assimilation by plant roots in the mud (Fry et al., 1982). Another exception is marine grass Zostera with maximum depletion of 12.7%0in 34S as compared to dissolved sulphate. Comparison of total sulphur content and isotopic data for aquatic plants leads to the following conclusions. The sulphur content falls within the range of 0.2-5%, with the lower values occurring in freshwater species. The sulphur concentration in aquatic plants depends upon their morphology and environmental conditions. Plants with well-developed base tissues (water florals, benthic weeds, macrophytes) contain less sulphur than delicate algae such as Cladophora, Enteromorpha, Polysiphonia subilifera, Codium vermila, etc. (Mekhtiyeva, 1975). In some plants, the sulphur content depends directly upon the sulphate concentration in water. Enteromorpha is representative of plants with high fractionation factors and low total sulphur contents; Zostera is quite the opposite with lower fractionation factors and higher sulphur contents (Table 6.3). In Table 6.3, there is an indication of an inverse relationship between the S content and the shift in 834Svalue of aquatic plant tissue with respect to SOi-. This is consistent with diffusional fractionation duringSOi- transport through the cell membrane. Higher animals are not capable of reducing sulphate to sulphide and receive reduced sulphur from their food. However, for lower forms in the food chain, the diffusional transport of sulphate through membranes is not excluded. The sulphur content of tissues of aquatic animals is somewhat lower than that of plants. Of all the forms studied, zooplankton from the Indian Ocean proved to be richest in sulphur (3.3%), whereas Actinia from the Black Sea were the poorest (0.1%). For a variety of similar biotypes, it was found that organisms in basins with higher salinity generally have

Hydrosphere Table 6.3 Concentration

187
and isotopic composition of sulphur in Zostera and Enteromorpha from various basins Caspian Black White Great Salma area

Sea

Azov

Location Zostera (percent S) 034S (%0) Enteromorpha (percent S) 034S (%oY

Near Yuryevka

Near Yalta

Near Derbent

Kazakh Bay

Near Gelendijik

2.07 -1.9 0.29 -12.7

-0.8 0.38 -10.2

-0.1

2.57 -1.0

4.99 0 0.81 -11.6

-6.2

" Compared to dissolvedsulphate.

higher S contents (Mekhtiyeva et at., 1976; Mekhtiyeva, 1975; Mekhtiyeva, 1971; Mekhtiyeva and Pankina, 1968). Limited data for marine animals show that tissues are depleted in 34Swith respect to sulphate by up to 7%0.Some interesting exceptions are observed. Shrimp from mangrove regions contain sulphur with 834S = +6%0. The isotope composition of sulphur in animals in hydrothermal ecosystems is close to that of coexisting sulphide minerals with 1)34S values ranging from + 1 to +4%0 (Fry et at., 1983). Therefore, isotopically light sulphide can obviously be involved in trophic levels of plants and microorganisms. Fractionation of sulphur isotopes during metabolism in animals is apparently small. Even in tissues of predators such as salmon, seal, and polar bears, which are high in the trophic chain, the isotopic composition of sulphur is close to SOi- assimilated by the lowest numbers (see Section 7.3). For evaluating the contribution of aquatic organisms to the sulphur pool in sediments, one should know the sulphur isotopic composition of insoluble organic derivatives in living tissues. Mineral and soluble organic compounds are washed away after an organism's death, whereas insoluble forms are buried in sediments. Kaplan et at. (1963) divided sulphur in marine organisms into organic and mineral forms. Compared to total sulphur, the organic fraction was depleted in 34S by 1-2%0, whereas the mineral fraction was heavier by 3-4%0. Mekhtiyeva et at. (1976) distinguished three forms of sulphur: (a) soluble organic, represented mainly by glutathione, cystine, cysteine, and coenzyme A, (b) insoluble organic, composed principally of proteins, and (c) mineral sulphur, mainly sulphate.

188

Stable Isotopes

In both plants and animals, insoluble organic sulphur compoundsdominate. The second important form is soluble organic sulphur compounds in animals, but sulphate in plants. In unpolluted environments, the 834S of the three sulphur forms are similar for a given specimen with maximum differences of 4.8%0 for plants and 5.7%0 for animals. In all organisms, the soluble organic sulphur (compared to total sulphur) is isotopically heavier, whereas the insoluble fraction is lighter. Sulphate in plants is enriched in 34S,whereas in animals it has a variable composition compared to total sulphur. The comparative 34S enrichment in sulphate in plants is consistent with it being the residual of partial metabolism by the cell. Reduced sulphur is concentrated in proteins and thereafter it is transformed into soluble organic compounds which are dominantly oxidized forms. Sulphate in marine animals has a dual origin. It is formed by oxidation of reduced forms and is also supplied from sea water. This conclusion is corroborated by the direct dependence between sulphur contents in animal tissues and [SOi-] in the environment. Osmotic permeability of cell membranes differs in aquatic organisms. The organisms endowed with high permeability contain high concentrations of sulphate (aspidium, starfish, and sea comb). Animals closely related in classification, e.g. all fish species, and inhabiting similar environments contain nearly equal amounts of various forms of sulphur. Isotopic differences between sulphur in aquatic organisms and the environment are essentially determined by mechanisms of sulphate supply to cells. The gross production of phytoplankton in oceans is assessed as 2 x 104Tg Corg yr-l or 3.6 x 104Tg of organic matter (dry basis) (Romankevich, 1977). Volkov and Rozanov (1983) assume the average sulphur content in phyto- and zooplankton to be 1%, which suggests that 360 Tg S are contributed annually to the oceanic sulphur. Based on Mekhtiyeva (1971), the average sulphur content in oceanic plankton is considerably higher (2.5-3.3%). This suggests that the annual production of planktonic sulphur in the oceans is about 103Tg. Neither estimate is well founded since there are no sulphur analyses from representative samples of oceanic phytoplankton. Data are available on the content of sulphur in mixed samples of phyto- and zooplankton, and on the variable sulphur content in algae-macrophytes. The higher estimate appears to be more probable as the sulphur content in small plant forms with salt-permeable envelopes is usually higher than in macrophytes. According to Romankevich (1977), the annual production of phytobenthos in oceans is 300 Tg of dry matter. For an average sulphur content of 2%, this corresponds to 6 Tg of total sulphur or 4.2 Tg of insoluble organic sulphur per annum. The above estimates suggest that the isotopic composition of total sulphur in aquatic organisms is quite close to that of water sulphate. For planktonic forms, the fractionation factor is 1.0034 (average of three measurements).

Hydrosphere

189

Insoluble organic sulphur, which accounts for about 70% of the total sulphur in aquatic organisms, is slightly lighter isotopically (up to 2%0)compared to total sulphur. Decomposition of sulphur organic compounds from dead plankton occurring in the water column may lead to a heavier isotopic composition residual sulphur (by 1-2%0). Thus organic sulphur compounds supplied to the oceans with particulate matter should have ~34Svalues close to +16%0. For benthic macrophytes, the fractionation factor is lower and in most cases is close to 1.001. The bulk of the benthic biomass is formed by the most distributed oceanic algae Chlorophyta, Phaeophyta, and Rhodophyta, whereas contributions from Zostera and mangroves are insignificant. An average ~34S value of + 18.5%0 is estimated for the biomass of benthic vegetation, whereas that of their organic sulphur derivatives may be assumed to be + 17%0. Dead plants are buried in the proximity of their habitat. Therefore, no significant changes in their mass and isotopic composition occur during the period from their death to burial. 6.3 MODERN OCEAN SEDIMENTS*

When preparing SCOPE 19, data reported up to 1980 were used to calculate the isotopic balance of sulphur buried in oceanic sediments. New information on the content of various sulphur forms in sediments, sulphur isotopic composition, and sulphate reduction (Lew, 1981; Lein et al., 1981; Chambers, 1982; Skyring et al., 1983) warranted an updated review. Moreover, the previous SCOPE report did not address fully the geochemical activity of sulphate reducing bacteria in shallow oceanic waters. Interesting ecosystems such as sediments of marshes, tidal flats, and their North European analogues, lidos, were not considered. Intensive sulphate reduction has been studied in these ecosystems in the United Kingdom (MacLeod, 1973; Nedwell and Abram, 1978, 1979; Banat et al., 1981; Nedwell and Banat, 1981), Atlantic USA (Calvert and Ford, 1973; Atkinson and Hall, 1976; Skyring et al., 1979; Shink, 1979; Howarth and Teal, 1979, 1980; Howarth et al., 1983; Stround and Paynter, 1980; King and Wiebe, 1980; Luther et al., 1982; Bouleque et al., 1982; Peterson et al., 1983; Howarth and Giblin, 1983), Pacific USA (Kaplan et al., 1963), and Australia (Chambers, 1982; Skyring et al., 1983). 6.3.1 Intensity of sulphate reduction Recent investigations showed that microbial sulphate reduction processes are abundant in bottom sediments of continental and marginal seas, various

* A. Yu. Lein.

190

Stable Isotopes

geomorphological zones of the ocean to depths of 3000-5000 m, and to limited depths in deep-water trenches. The maximum depth at which sulphate reduction can occur in oceanic sediments has not been established but the process has been documented at a depth of 5-6 m from the silt surface and down to 13 m in Baltic Sea sediments (Lein, 1983). The intensity of microbial sulphide generation varies from less than 1 J.Lg S kg-1 d-1 in deep-water oceanic sediments to tens of mg S kg-l d-I in sediments of the continental Baltic Sea. However, the highest intensities of sulphide production are found in marshes and tidal flats, including lowland plains along seashores, flooded during high tides or storms. In marshes, soils rich in humus are formed on silt and sand-silt drifts with high Corg content. Such sediments are typical of lowland shores of Great Britain, the Federal Republic of Germany, the Netherlands, and sections of the Atlantic shore of the United States. The total area of such high-tide and low-tide zones is estimated at 6 x 104 km2 (Atlas of the Oceans, 1977). Despite the large number of investigations, the isotopic composition of sulphur has been studied only in sediments of Newport marsh, California, USA (Kaplan et al., 1963) and the Mambray Creek marsh of South Australia (Chambers, 1982). In marsh sediments, the concentration of sulphate increases with depth. For example, in Newport marsh the sulphate sulphur content in the surface horizon is 0.073% as compared with 0.091 % at 30 cm below the surface (Kaplan et al., 1963). Surface water and water squeezed out of green algae had sulphate sulphur concentrations of 0.094 and 0.098% respectively. Therefore, the most active sulphate reduction occurs in the upper (10-20 cm) of marsh sediments. Lower in the column, in the sand underlying the surface silt, reduced sulphur is oxidized due to flushing with Oz-saturated H2O. Stratified sulphate reduction is characteristic of most marshes. This is also confirmed by changes in Eh values from - 300 mV in upper horizons to + 100 mV deep in the sediment (Table 6.4). The Corg and Spyritccontents decrease similarly from 1.7 and 0.92% (dry weight basis) in the upper horizon to 0.4 and 0.3% at a depth of 30 cm (Table 6.4). The f>34S values of pyrite change with depth from -20.0%0 (0-5 cm horizon) to -27.7%0 (60-65 cm horizon). Just as in normal sea sediments, free H2S and acidsoluble sulphides are enriched in 34S by 4-5%0 compared to pyrite sulphur (Table 6.4). A comprehensive study of the sulphur cycle was carried out from November 1977 to November 1978 in a lO-cm stratum of Mambray Creek marsh sediments (Chambers, 1982). The f>34S values of pyrite varied from -9 to -27%0 (average -17.7%0, 80 determinations at three stations) (Figure 6.3). Rates of sulphate reduction, measured with 35S-labelled sulphate, ranged from 1.7 to 41 mg S kg-1 d-I or 100-2400 g S m-2 yr-l in sediments of

~
""C3

{J :::'" .... '"

Table 6.4 Content and isotopic composition of sulphur compounds in Newport marsh sediments and pore water (Kaplan et al., 1963) 534S (%0) Eh (mY) -300 -50 +100 SO/- in pore water (%S)" 0.073 0.077 0.091 Pyrite dry sediment (%S) 0.926 0.283 0.190 FeS acid soluble -15.0 -18.8 -19.9 Corg in sediment (%) 1.7 0.4 0.2

Horizon (cm) 0-5 30-35 60-65

SO/+ 17.0 +14.3 +11.6

H2S free -17.1 -

Pyrite -20.0 -27.1 -27.7

" Percentage relative to interstitial water content. Other entries on dry weight basis.

...... \0 ......

192

Stable Isotopes

-30

-20

-10

8348 (%0)

Figure 6.3 Values of the 334S sulphide sulphur from the Mambray Creek deposits (Chambers, 1982)

various marshes (Table 6.5). The annual production of hydrogen sulphide and its derivatives obtained by integrating over high-tide and low-tide zones ranges from 6 x 106 to 144 X 106 tonnes of sulphur. Most of the reduced sulphur formed is oxidized to sulphate, but part of the HzS, which is difficult to estimate, evolves to the atmosphere. Evidently only a small fraction of the reduced sulphur is buried in marsh sediments in the form of pyrite with an average 034S value of about -17.7%0. Table 6.6 summarizes data on the sulphur isotope composition of reduced sulphur compounds and annual intensities of sulphate reduction in sediments of various zones of the ocean. The data available are insufficient to calculate the reduced sulphur flux in marsh sediments and bottom sediments of shallow water gulfs, lagoons,
Table 6.5 Intensity of sulphate reduction and average isotopic composition of pyrite sulphur in marsh sediments Rate of sulphate reduction (g S m-2 yr-I)

Sampling site Newport, USA Mambray Creek, Australia New England, USA Essex, England

334S (%0)

Reference Kaplan et al. (1963) Skyring et al. (1983); Chambers (1982) Howarth and Teal (1979) Nedwell and Abram (1978)

200-400 2400 99-140

-23.6 -17.7

Hydrosphere

193

Table 6.6 Intensity of sulphate reduction, average isotopic composition, production of reduced sulphur, and organic carbon consumption in oceanic sediments Area of sediments (106 km2) 11 (2)b 16 (3) (15) (61) 257 360 (81) 55.6 37.2 28.2 5.5 1.0 Production of reduced sulphur (Tg yr-l) -" - a 79.2 53.1 201.0 158.6

Region and depth Marshes Gulfs, lagoons, estuaries Shelf 0--50 m 50--200 m Continental slope 200--1000 m 1000--3000m Deep-water sediments Total for the ocean

Intensity of SOl- reduction (....gS kg-1 d-l)

(')34S (%0) -17.7 -16.5 -24.2

-33.2 -41.1

491.9

" Data insufficient to calculate average sulphate reduction intensity.


h

The figuresin parentheses give the areas of biogeochemically active sediments with bacterial

sulphate reduction.

and estuaries. It is, however, evident that the total quantity of sulphur buried in these sediments cannot exceed 10-15 Tg S ycl globally. Table 6.6 shows a lateral trend in sulphate reduction with the intensity decreasing from littoral towards pelagic sediments. 6.3.2 Rate of bacterial sulphate reduction and other factors influencing the distribution of sulphur compounds and their 8348 values in recent sediments Analysis of dissolved sulphate in silt and total reduced sulphur in columns of bottom sediments revealed two extreme patterns in sulphur isotope distribution (Figure 6.4). In sediments with intensive pore water sulphate reduction and poor filtering properties (Figure 6.2, st. 655), the isotopic compositions of sulphur of pore-water sulphate and total reduced sulphur approach equality with depth. In sediments with a constant high SOi- content (Figure 6.2, st. 663), reduced sulphur in lower horizons is appreciably depleted in the heavier 34S.

194

Stable Isotopes 8348 (%0)


0
-40 -20 0 +20

[SOlJ(ppm
+40 0
200 400 600

S)
800

SO 4 2- REDUCTION
.1

u '-'

100

Ifu 300 0 IZ W ~ 0 w
(/)

I200

'82-

..

( mg S k -1 -1 1 10 9 10yr10)

:"
1 11

SOi1 1

:! ,iii~~!i:~iw!
I

II

/ I~

Depth 43 m

~ /
J
!

/
{

~t. 665
Depth

3260

St.

663 1760 m

Depth

I \
\

St.

107 125 m

Depth

St. 59 Depth 4800 m

St. 668 Depth 140 m

Figure 6.4 Changes of the sulphate reduction intensity values (mg S kg-I d-I) of sulphate ion concentration in silt water (mg I-I) and a34s (0/00) of general reduced and sulphate sulphur value in contemporary ocean sediment columns

Both patterns of the distribution of 334Svalues point to the influence of sulphate reduction on fractionation of isotopes in reduced sediments. In the first case, the isotopic composition of sulphur is shifted by the loss of 3ZS in upper silt horizons. In the second case, production of HzS with a lighter isotopic composition is explained by slower sulphate reduction rates in silt, resulting in greater isotope fractionation as observed in laboratory

~ :::... ;j ~ ;:'" ~ Table 6.7 Average isotopic composition of reduced sulphur, sulphate, and total sulphur in world ocean sediments Sulphide Morphometric zones and depth Shelf, 0-200 m Continental slope and rise Ocean bed Total ocean Total Area *a
(106km2) Sulphate

Total sulphur
Tg S yr-Ib !)J4S (%0)

Tg S yr-1b 10.4 88.0

!)J4S (%0)

Tg S yr-1b 1.1' 5.3' 2.0" 19.4<1 27.8

!)J4S (%0)

27 76 257

-24.2 -33.2

+27.9 +22.9 +20.0 +20.0 +20.6

11.5 93.3 2.0 19.4 126.2

-19.2 -30.0 +20.0 +20.0 -20.5

360

98.4

-32.2

" Gershanovich el al. (1974). h According to Volkov and Rozanov (1983).


<' Sulphate
d

dissolved

in pore

water.

Solid phase sulphate.

>-' \D VI

196

Stable Isotopes

experiments (cf. Jones and Starkey, 1957; Harrison and Thode, 1958; Kaplan and Rittenberg, 1964; Chambers and Trudinger, 1979). The substantial increase in the fractionation of sulphur isotopes with decreasing intensity of sulphate reduction is expressed in the ranges of ;)34S data for various geomorphological zones of the world ocean system (Table 6.6; Figure 6.5). These data are of primary importance since they suggest that ;)34Svalues of reduced sulphur in rocks may enable an assessment of the intensity of sulphate reduction in the geological past. 6.3.3 Mass and isotopic balance of sulphur in recent sediments Table 6.7 contains data necessary for calculating the material and isotopic sulphur balance in modern oceanic sediments. The total sulphur buried annually (other than evaporite formation) is estimated to be 126 Tg, with an average ;)34Svalue = -20.5%0. The major portion of sulphur is buried in reduced forms (primarily pyrite) with an average ;)34Svalue of - 32.2%0. A comparison of data in Tables 6.6 and 6.7 shows that only 20% of newly formed hydrogen sulphide is buried as sulphide minerals in sediments, whereas the remainder is internally cycled. H2S migrates to the upper
834S (%0) 0

-40

-20 .

+20

+40

SULPHIDE
BAL TIC SEA CONTINENT AL SHELF 0-200 m CONTINENTAL SLOPE 200-1000 m 200-1000m 1000-3000 m

SULPHATE

CD

&"

I""'''

"'f
~!I!l1SI

ex5hg 00

o~fjId3!! 0~8T 00 0

&~

...
"1

t~ 00 0
"3 ot ..4

SLOPE SEAFLOOR TRANSITION ZONE > 3000 m OCEANIC TRENCH > 6000 m RED SEA RIFT

.~ 0
0
0

'ffi~BBo
0

0 00

02 .6

IS! 5

B 0
0

07 +8

e~
+

~ e .

Figure 6.5 Variations of 8]4Ssulphide and sulphate value in sea and ocean sediments. I-Baltic Sea, 2-South China Sea, 3-California Bay, 4-Tasman Sea, 5-Pacific Ocean Shelf near Mexico, 6--western part of the transocean geologic section, 7the area of the Peru upwelling activity, 8-abiogenic sulphides of the active ocean zones

Hydrosphere

197

horizons and is oxidized there to sulphate. This is confirmed by maxima in pore-water sulphate concentration just below the surface of the sediments. This sulphate is depleted in 34S (034S = -15.2 to -18.5%0) as compared with 21%0 for open-ocean SOi- (Lein et al., 1981). This demonstrates unambiguously the generation of excessive sulphate by the oxidation of biogenic hydrogen sulphide (Figure 6.2, s1. 668). In rare cases where one knows the rate of sedimentation as well as the total sulphur content and sulphate reduction intensity, it is possible to estimate the portion of reduced sulphur buried in the form of pyrite for individual sediment horizons (Table 6.8). Values of buried sulphur for two examples considered (15 and 16%, Table 6.8) are within the limits accepted for global calculations (about 20%, Tables 6.6 and 6.7). 6.3.4 Influence of the rate of sulphate reduction on sulphide mineral formation Sulphide minerals isolated from marine sediments with varying intensities of bacterial sulphate reduction are markedly different in composition and morphology. Dissolved sulphide and iron sulphide are always present in subcontinental sediments where there is intensive bacterial sulphate reduction. Sulphide ion is easily oxidized chemically and microbiologically (Jannasch et al., 1974; Gorlenko et al., 1977), yielding excess sulphur. Such conditions promote framboidal pyrite formation. In sediments with low rates of bacterial reduction, free hydrogen sulphide and sulphide ion are practically absent in pore water. These conditions

Table 6.8 Example of the calculation of sulphide sulphur buried in sediments with known rate of sedimentation Pacific ocean shelf near Mexico, depth 140 m" -17.5-25 750 0.56 37.5 6.0 16

Baltic Sea, st. 2656, depth 46 m" Horizon of silt (cm) Duration of the process (years) Intensity (mg kg-I d-I) Calculated content (g kg-I) True content (g kg-I) Buried sulphur (% of calculated value)
" Lein et al. (1982).
b Ivanov et al. (1976).

3.0-4.0 35 5.4 68.0 10.2 15

198

Stable Isotopes

promote formation of non-framboidal pyrite and pyrite aggregates with sulphide minerals of the mackinawite-greigite group. Therefore, the decrease in sulphate reduction rate evidently inhibits the transformation of unstable sulphide minerals to pyrite. Sulphides formed from abiogenic hydrogen sulphide in sediments of tectonically active zones of the ocean (Red Sea depressions, East Pacific upwelling, crater lakes, etc.) are represented by individual crystals of cubic, prismatic, and other habits or by drusy and clusterlike aggregates. Such sulphides are characterized by 534S values in the range -3.0 to +3.0%0, which is radically different from that of sulphides in other regions of the ocean (Lein et al., 1982; Lein, 1983).

6.4 LAKES* 6.4.1 Lakes: water column


Lakes have now been recognized as the critical receptors most susceptible to inputs of acid rain and associated long-range transported pollutants from the atmosphere (Almer et al., 1978; Wright and Snekvik, 1978). Potentially, measurements of the isotopic composition of both the sulphur and oxygen in the sulphate can be used to identify the sources and ascertain the behaviour and fate of pollutant sulphur in a given lake. The application of stable isotope techniques in acid rain studies has been only partially successful, however. The 534S0i- data for different types of lakes in many parts of the world are summarized in Table 6.9. The range is fairly wide although most of the data fall between + 5 and + 15%0.In addition to the source dependence, the isotopic composition of sulphur is also influenced strongly by the trophic conditions in the lake. 6.4. 1.1 Isotopic changes due to lake pollution The diagnostic and prognostic capabilities of 534S0i- measurements can be demonstrated using the data in Table 6.10. The measurements pertain to lakes in the Sudbury basin of Canada and show differing degrees of stress from the inputs of smelter-derived pollutants. On the basis of the observed pH, the lakes can be subdivided into (a) well-buffered lakes unresponsive to the input of acid precipitation and (b) poorly buffered lakes more easily affected by the smelter emissions. The SOl- concentration is correlated inversely with distance from the major smelter stack according to the relation [SOi-]
* J.O. Nriagu.

= 55 (distance) -0.38

Table 6.9 Isotopic composition of dissolved sulphate in selected lakes Lake Ontario Erie Western Basin Eastern Basin Huron Michigan Superior Turkey Lake watershed (four lakes) Lake WWI, Algoma, Ontario Lake WWII, Algoma, Ontario Fenton Lake, Algoma, Ontario Logger, Algoma, Ontario Ten lakes, Algoma, Ontario Linsley Pond, Connecticut Queechy, Connecticut Mt Tom Pond, Connecticut Green Lake, New York Fayetteville Green Lake, New York 0 m (surface), sulphate 50 m (bottom), sulphate 18 m, HzS 50 m, HzS Lake 'A', Ellesmere Island, Canada 3m 40 m [SO/-] (mg -')
---

Year sampled --1972 1972 1972 1972 1972 1985 1983-85 1982 1982 1982 1982 1983

0:14S (%0) ----+5.9 +5.4 +5.2 +6.3 +4.6 +4.4 +4.1 +5.3 +5.5 +4.5 +3.5 +4.5 +8.6 +5.8 +6.1 +7.5 +23 +28 -32 -27

Reference Nriagu (1973) Nriagu Nriagu Nriagu Nriagu Nriagu (1973) (1973) (1973) (1973) (unpublished results)

29 25 25 16 21 3.0 25 28 8.2 6.8 5.3 4.3 4.6 5.1

1983 1983

Nriagu (unpublished results) Thode and Dickman (1984) Thode and Dickman (1984) Thode and Dickman (1984) Thode and Dickman (1984) Thode and Dickman (1984) Nakai and Jensen (1967) Nakai and Jensen (1967) Nakai and Jensen (1967) Fry (1986) Fry (1986)

1982 30 1800 +11 +35

Jeffries et al. (1984)

68 m

1900

+37

.....

Lake 'B', Ellesmere Island, Canada 0 m (surface) 20 m 40 m Lake Tanganyika Lake Kinneret Surface, July Surface, April 20 m, July 20 m, April 42 m, July 42 m, April Solar Lake Sulphate H2S Lake Creteil 4 m (mid-column) 6 m (bottom) Sernoye (reservoir), USSR Baikal Ladoga Onega Onega Balkhash

1983 50 750 700 3.9 1976 67 63 66 68 90 68 +12 +16 +12 +17 +15 +13 +19 +39 +87 +11

Jeffries and Krouse (1984)

N 0 0

Monster (1973) Nissenbaum (1978)

Aizenshtat et at. (1981) +21 -16 590 620 Chesterikoff e/ at. (1981) +32 +27 + 12 to + 15 Pan kina and Mekhitiyeva (1969) +7.1 Rabinovich and Grinenko (1979) +5.9 Rabinovich and Grinenko (1979) +5.2 Rabinovich and Grinenko (1979) +8.8 Rabinovich and Grinenko (1979) +11 Rabinovich and Grinenko (1979)

1969 1958 1958 1972

5.4

Table 6.9 Continued [S042(mg -1)


834S

Lake Issyk Kul Sabundy-Kul Dzhasybay UI'kenonkol'l Pashennoye Shchuch'ye Chelbar (brackish) Itkol' Sakovo 2 m, sulphate 7 m, sulphate 7 m, H2S Yanda, Antarctica 4m 40 m 60 m 68 m

Year sampled --

(%0)
---------

Reference Rabinovich and (1979) Chukhrov et al. Chukhrov et al. Chukhrov et al. Chukhrov et al. Chukhrov et al. Chukhrov et al. Chukhrov et al. Matrosov et al. Grinenko (1975) (1975) (1975) (1975) (1975) (1975) (1975) (1975)

+15 +9.5 +15 +13 +4.4 +4.4 +10 +6.2 51 242 1973-3 8.7 24 237 611 + 15 +17 +22 +46 +13 +15 -11

Nakai et al. (1975)

N 0 ......

202 Table 6.10 Concentrations and sulphur isotopic compositions of sulphate in lakes around the copper-nickel smelter at Sudbury, Ontario (from Nriagu and Harvey, 1978) Lake Smelter distance (km) 3.2 3.9 4.5 5.2 5.2 5.2 6.5 7.7 8.4 9.7 10.3 11.0 11.6 12.9 13.5 14.8 20.0 20.0 32.3 32.9 37.4 39.4 41.9 45.2 45.2 47.1 47.7 49.0 50.3 52.3 57.4 58.1 58.1 61.3 62.6 67.0 73.6 75.5 76.7 83.9 85.8 90.3 pH Sulphate level (ppm) 30.5 53.0 35.4 36.1 23.5 39.5 28.5 34.7 25.7 27.4 22.3 20.9 24.4 19.3 21.8 87.8 15.8 13.7 19.9 7.8 12.9 12.9 11.0 12.8 12.3 14.6 11.9 13.6 20.3 9.5 8.9 8.9 10.8 17.6 8.1 9.7 12.1 12.1 11.2 13.2 8.1 834S (%0)

Robinson L. Hannah L. St Charles Nepohwin 'e' Middle Silver Ramsay Richard McFarlane Lohi Long Raft Clearwater Tilton Makada McCharies Wavy Little Panache Panache Broker Tyson Log Boom Johnnie Ruth-Roy Norway Perdix Carlyle Kakakise Lang Acid (Lum II) Lumsden I Lumsden III Apsey Frood Grab Evangeline Maple Cutler La Cloche Little La Cloche Owl

6.09 3.40 4.53 6.15 3.20 5.78 3.20 6.61 5.06 5.32 4.20 6.45 4.15 3.50 4.20 6.75 8.95 3.30 8.71 6.70 5.20 5.70 5.19 4.15 4.50 4.20 4.41 4.85 5.75 6.75 4.39 4.39 4.60 7.01 6.70 6.25 6.42 6.40 6.79 6.68 6.80 4.75

+8.46 +3.48 +7.32 +5.19 +2.66 +5.19 +7.34 +8.43 +6.76 +5.31 +2.84 +4.56 +4.50 +4.23 +3.57 +5.23 +6.08 +4.70 +4.75 +4.84 +4.80 +4.30 +4.05 +4.00 +4.99 +4.99 +4.67 +4.55 +4.48 +4.55 +5.71 +5.43 +4.82 +5.26 +5.36 +4.46 +4.22 +5.03 +4.34 +4.73

~
:::...

Table 6.11 Apparent isotopic fractionation factors for sulphate reduction in lakes Lake
-~--

~ :::'" ~

(S

[S042-] (mg {H)


--

[H2S] (mg -1)


.--

(ex - 1) 10-'
--~

Reference
--~

Water column Lake Yanda, Antarctica Fayetteville Green Lake, New York Lake Sakovo Chernyi Kichiyer Bol'shoy Kichiyer Black Sea Solar Lake Sediments (pore water) Linsley Pond, Connecticut Queechy Lake, Connecticut Mt Tom Pond, Connecticut Chernyi Kichiyer Bol'shoy Kichiyer Lake Kononyer Lake Kuznechikha Lake Sakovo

608 1340 780 86 43 2340

70 31 9.1 50 8.5 7.4

35 56 28 23 4.1 51 47 10 5.4 9.1

Nakai et al. (1975) Deevey et al. (1963); Fry (1986) Matrosov et al. (1975) Matrosov et al. (1975) Matrosov et al. (1975) Matrosov et al. (1975) Aizenshtat et al. (1981) Nakai and Jensen ( 1967) Nakai and Jensen (1967) Nakai and Jensen (1967) Matrosov et al. (1975) Matrosov et al. (1975) Matrosov et al. (1975) Matrosov et al. (1975) Matrosov et al. (1975) N 0
w

13 14 15 105 43 35 36 738 346 155 608 181 200

7.8 6.6 4.8 1.6 38

204

Stable Isotopes

with r = 0.81 (Nriagu and Harvey, 1978). There is also a significant correlation between the pH of poorly buffered lakes and the distances from the smelter stack. The 034S0/- for the well-buffered lakes show considerable scatter (see Fig. 3 in Nriagu and Harvey, 1978), suggesting (a) derivation of the sulphur from several sources and/or (b) complex interplay between the point source influence and sulphur transformations either in the watershed or within the lakes. For example, high acid loading may increase the amount of sulphur in these lakes by accelerating the weathering of the bedrock and the overburden. By contrast, the 034S042- data for the poorly buffered lakes are remarkably uniform, lying mostly between +4 and +5.5%0 (Table 6.10). This uniformity clearly suggests that the isotopic composition of these lakes is regulated by the influx of sulphur released from the smelters. In the Great Lakes system of North America, anthropogenic sources now account for 30% of the SO/- in Lakes Superior and Huron and over 70% in Lake Erie (Nriagu, 1984). There is some suggestion that the industrial inputs are affecting the isotopic composition of SO/- (Table 6.9), although the inter-lake differences do not seem to be consistent with the levels of pollution. Lake Superior, in particular, has a long water flushing time (about 180 years) and thus contains some sulphur of precolonial age. Like the other Great Lakes, Superior has a small watershed-to-lake area ratio, implying that direct atmospheric deposition accounts for a large fraction of the sulphur input. Its 034S0/- can thus be regarded as being representative of 'unpolluted' waters of the Great Lakes. As can be seen in Table 6.11, the 034S0/- data for the other Great Lakes with more pollutant sulphur deviate from this background value significantly. The stable isotope technique has been used with some success to assess the contamination of surface waters with sulphur emissions from sour gas processing plants at West Whitecourt, Alberta (Krouse et al., 1984). It was found that the dissolved sulphur became heavier with increasing organosulphur content. The 034S approached a limiting value of + 22%0 (Figure 6.6), which corresponded to the isotopic composition of the sulphur released from the sour gas plants. The results of the study suggest that the emission were affecting the biological uptake of sulphur in the boreal forest, which in turn resulted in a marked increase in the flux of isotopically labelled organosulphur compounds into the surface waters (Krouse et al., 1984). 6.4.1. 2 Oligotrophic lakes With the exception of Lake Erie, the Great Lakes of North America possess a characteristic feature of oligotrophic lakes, namely very homogeneous isotopic composition of sulphate. Differences in mean 034S0/~ values of the hypolimnion and epilimnion waters were generally less than 0.2%0

Hydrosphere

205

+20
-0
~

, OJ

0"
(/) -

I
I

1>
'" (.()

0 +1

00

10

20

DISSOLVED ORGANIC SULPHUR(JLg r') Figure 6.6 The 1)34S values for total dissolvedsulphur in water versusorganicsulphur concentration in selectedwater samplesfrom the West Whitecourtstudy area (Krouse et al., 1984) (Nriagu, 1973). Temperature (seasonal) changes apparently do not promote significant sulphur isotope fractionation; nor does assimilatory sulphate reduction by biota. For many hardwater and/or softwatwer oligotrophic lakes, the 034S0ivalues are very close to those of rainfall in their drainage basins, implying a common origin. Good examples are the Algoma Lakes (including those of the Turkey Lakes Watershed) of northern Ontario (Table 6.9). For many hardwater lakes of oligotrophic classification, the 034S0i- values reflect the combined signatures of sulphur from atmospheric fallout and from the weathering of the bedrocks. As noted already, most of the poorly buffered lakes in the Sudbury basin fall into this category. Recent studies of the isotopic composition of oxygen of the dissolved sulphate in oligotrophic lakes have yielded unexpected results (Caron, 1984; Fritz et al., 1986). The 0180 values of SO i- tend to be distinctly different from those of rainfall and often show complex patterns unrelated to the 034S data (Figure 6.7). The oxygen isotopic data thus point to extensive assimilatory reduction of sulphate, and suggest that the turnover of sulphur in oligotrophic lakes may be more rapid than has hitherto been realized. 6.4.1. 3 Eutrophic lakes The S042-, HzS, and ()34S0l- profiles in a typical eutrophic lake that develops an anoxic hypolimnion are shown in Figure 6.8. During the spring

206

Stable Isotopes

~+6 0
.,. Jt) (;()

~ U)+4

"g

+2 0 J J A SON 1983 D J F 1984

.'-----

+14
o.."

+12
+10 0 c- +8 0 ~ '0 +6 +4

/~" ,...' "," '",


", "
" -'

~
JJASONDJF 1983 1984

+2 0

Figure 6.7 Isotopic composition of sulphur and oxygen in sulphate in lake water and precipitation samples in six head waters of Quebec (Caron, 1984)

and autumn overturn, any HzS formed in the hypolimnion is oxidized and mixed with the sulphate in the epilimnion, resulting in a fairly homogeneous 034S value throughout the water column (Figure 6.8a). During the period of summer stratification, the SO/- is progressively converted to HzS in the anaerobic hypolimnetic waters. Ideally, the 034S0/- profile for lakes of this class should have three distinctive segments (see Figure 6.8b and c): (a) the epilimnion with a fairly constant 034S profile; (b) the chemocline where the partial reduction of sulphate results in o34S0/- values higher than that of the epilimnion-if the hypolimnion is completely anoxic, a maximum in the 034S04Z- profile should occur at this interface; and (c) the hypolimnion, which may display any profile depending on the oxygen concentration (compare Figure 6.8b

Hydrosphere

207

and c). If the system is completely anoxic, there is a severe shortage of sulphate and relatively little fractionation of the sulphur isotopes is expected. If the system is partly anoxic (due to leakage of oxygen across the chemocline), the 034S0i- should increase towards the sediment-water interface, in accordance with gradients of declining redox potential and increasing intensity of sulphate reduction. These two may be regarded as the ideal profiles. Many factors, however, can modify the 034S0i- profiles in the hypolimnion, such as (a) the thickness of the hypolimnion and the duration of stratification, (b) the sulphate concentration and internal biogeochemical processes in the water column, and (c) the exchange of sulphur isotopes across the sediment-water interface (Nriagu and Soon, 1985). If a large fraction of the product HzS is removed to the sediments by reaction with Fez+, the 034S0i- data in the anoxic hypolimnion may even show measurable seasonal changes. In view of the many compounding variables, few, if any, good models have been developed to describe the fractionation of sulphur isotopes in this lake type. It should be noted that few studies have included the measurement of the isotopic composition of the oxygen in the sulphate. 6.4.1.4 Meromictic lakes These differ from the eutrophic lakes in (a) being permanently stratified and (b) generally having a higher concentration of dissolved salts. Typical 034S0i- profiles in meromictic lakes can also be expected to show three distinct segments corresponding to the epilimnion, chemocline, and monimolimnion. Representative profiles are shown in Figure 6.9. In Fayetteville Green Lake, State of New York, the 034S0i- value is essentially constant near +23%0 in the epilimnion (1-15 m), but increases gradually with depth down the monimolimnion. As expected, the reduced sulphur species were depleted in 34Sand the isotopic difference between the SO i- and HS- was large and remarkably constant at about 56.6%0 (Deevey et al., 1963; Fry, 1986). By contrast, the profile of 034S0i- in Lake Yanda, Antarctica, shows a sharp increase at a depth of 60 m below the ice cover where the HzS becomes detectable. The maximum value of +49%0 was attained right at the sediment-water interface (Figure 6.9). The sharp increase in sulphate concentration occurs at about 10 metres above the point of rapid increase in o34S0i- values (Nakai et al., 1975).

Figure 6.8(a) 534S0/-, SO/-, and total dissolved sulphide profiles in Lake 223 on 15 June 1979, or following the spring overturn (Cook, 1981). (b) 5'4S0/-, SO/-, and total dissolved sulphate profiles in Lake 223 on 21 September 1979, or during stratification (Cook, 1981). (c) 534S042-, SO/-, and total dissolved sulphate profiles in Lake 227 on 1 September 1978, or during stratification (Cook, 1981)

208

Stable Isotopes

254 0 0 40 8 "8 (fLmoll 80

-1 ) 120

160 01

40 I

80

120

~ ..... Q. IV C 10

..
80:I +10 I f + 20

{
+30

-l
210
151
0
I

o:-l*

\
I

i
+20 +30

150

I~ Mn filtered t[;w w.."

+10

8345,50t(%0)

8345,5042- (%0) . -1 2:H25 (fLmol t )

2: H25 (fLmol

r1)

~
~

10f

"S.

'\'
15i!:. 0

::]0 100 200


(0 )
5042-

300

1::&<":
0 100
( b)

~O 200

300

(fLmoi r1)

00

8345,542- (%0) 0 +10 +20 Or-Q I I

2~

40)

f~*/

10

-1
2: H25 (fLmol ( )

20

40

i5~

~j {

'

Hydrosphere
-34
0

209
8348 (%0) -26 +22

-30

+26

+30

I Fayett~ville Gr~en -r.ak~. ' New York j

20 .s:: C. Q) c 40 60

~
0 0 +20 8348 (%0) -I- 40 + 60 +80

20

]
-5 a. 40
Q) C 60

180:-1
80

Figure 6.9 Profiles of sulphur isotopes in typical meromictic lakes (data for Fayetteville Green Lake from Fry, 1986, for Lake Yanda from Nakai et al., 1975, and for 'A' Lake from Jeffries et al., 1984)

The sulphate concentration in Lake 'A', Ellesmere Island in the Canadian Arctic, increases from about 30 mg -l just below the ice cover to well over 1900 mg -l in the bottom waters (Jeffries et ai., 1984). The 034S0i- and oU'O-S042- also increase with depth from the ice cover although samples closest to the sediments are slightly depleted in 34S. The isotopic profiles do not depict the freshwater-to-seawater transition zone located at depths of 15-25 metres. In the case of smaller Lake 'B', which is slightly further inland, Jeffries and Krouse (1984) found the 034S values of bottomwater S042- to be as high as +87%0 (Figure 6.7). Dissolved carbonate in the bottom waters had 034C values of - 21 and - 27%0for Lakes A and B respectively, which is in the range of organic matter. In summary, the sulphur, oxygen, and carbon isotope data collectively provide strong evidence of anaerobic bacterial S042- reduction with attending oxidation of organic nutrients. Because of

210

Stable Isotopes

preferential 3zS0i- reduction, the unreacted S04z- became highly enriched in 34S in the bottom waters. For example, the 034Svalue of +87%0 in Lake B corresponded to a much lower [S04Z-], 700 ppm, i.e. about one-third of that found in the bottom of Lake A. Insofar as a lake maintains a permanent meromixis, the hypolimnion can be regarded as a closed system. The fractionation of sulphur isotopes during sulphate reduction can hence be approximated by the Rayleigh distillation equation:

(
C
are the corresponding SO/-

Rt

= C
( Co )

(1 -

l/a)

Ro )

where Ro and Rt are the 3ZSp4S ratios at the start and after time t, Co and
concentrations, and (Xis the instantaneous

isotopic fractionation factor. Figure 6.10 shows typical Rayleigh plots for residual S04Z~ and the HzS produced, assuming (X= 1.005 and an initial

+30

0.8

0.6

0.4

0.2

+20
0 0 (f) <t r<)

ro

+10

Accumulated H2S 0
Figure 6.10 Rayleigh distillation plots for a closed system undergoing sulphate reduction assuming ex= 1.005, and initial 0 values of +10%0. F is the fraction of sulphate remaining in the system

Hydrosphere

211

834S0i- of + 10%0.The patterns of the calculated graphs generally match what have been observed in some meromictic lakes (compare Figures 6.9 and 6.10). Values of a for sulphate reduction in lake waters and sediments range from 1.004 to 1.056 (Table 6.11). As to be expected, the lowest a values are found in lakes with the lowest sulphate concentrations-the sulphate reducers become less discriminating as the available sulphate is depleted. It also means that in lakes with low SOi- content, the value a is not constant but will decrease as the available sulphate is biotransformed into reduced species. The available data suggest that when the sulphate concentration is less than 20 mg -1, the value of a is generally below 1.01. In contrast, a is much larger (1.03-1.07) in marine environments with abundant S042(Goldhaber and Kaplan, 1974). It is not surprising that some of the a values for lakes fall in the marine range, considering their sulphate concentrations (Table 6.11). 6.4.2 Lakes: sediments A detailed review of the distribution and diagenesis of sulphur in lacustrine sediments has already been given in a preceding SCOPE volume (Ivanov and Freney, 1983). The monograph includes a good summary on the use of sulphur isotopes in delineating the critical pathways of sulphur transformations in sediment ecosystems. To avoid duplication, the following section concentrates on the use of stable sulphur isotopes in understanding freshwater sediments as sinks for anthropogenic sulphur, especially in eastern North America. The reduction of anthropogenic sulphate in lake sediments is an important alkalinity-generating process. Measurements of the isotopic composition of sulphur in sediments can thus (a) provide some insight on the self-purification of a lake and its ability to recover from an acid stress and (b) be used for retrospective monitoring of past fluxes of pollutant sulphur into the lake basin. These are key questions relevant to acid rain. 6.4.2.1 Anthropogenic influence on the isotopic composition of sulphur in lake sediments of eastern North America The role of lake sediments as sinks for pollutant sulphur has received little attention. Several studies have now demonstrated that there has been a significant increase in the sulphur contents of the most recent sediments of lakes in many parts of the world (Table 6.12). For example, the enrichment factors (EF) for the Great Lakes vary from about 1.5 in the fairly pristine Lake Superior to over 3 in the most polluted, much shallower, Lake Erie. (EF is the ratio of the average S concentration in surficial sediments to that in the precolonial layers. The difference between the S contents of the two

212

Stable Isotopes
recent

Table 6.12 Average concentrations and enrichment factors for sulphur in


(surficial) lake sediments
Lake/location (Reference a)

Average [S] (%) 0.014 0.092 0.113 0.150 0.210 0.51 0.71 0.81 0.31 0.16 0.13 0.67 0.12 0.19 0.39

Enrichment

factor

Superior-Great Lakes (1) Huron-Great Lakes (1) Michigan-Great Lakes (1) Erie-Great Lakes (1) Ontario-Great Lakes (1) Algonquin Provincial Park, Ontario (2) Around Sudbury, Ontario (2) Wawa, downwind of smelter (3) Wawa, upwind of smelter (3) Experimental lakes area (4) Adirondack Mts, New York (5) Swiss lakes (6) English lakes (7) Unproductive Moderately productive Very productive
a

1.8 1.5 2.8 3.3 2.6 -2.8 4.5 > 3.0 1.7 -3.0 -2.5 4.2

(1) Nriagu (1984). (2) Nriagu and Coker (1983). (3) Derived from the data of Thode and Dickman (1984) using the average background concentration for Beaver Lake of 0.18 wt%. (4) Cook (1981). (5) Mitchell et at. (1983). (6) Deevey (1972). (7) Gorham et al. (1974).

sediment zones will be referred to as excess sulphur.) On the lake sediments near the smelters at Sudbury, which daily release about 2500 tonnes of S02 into the atmosphere, the EF values typically exceed 4 (see Table 6.12). The observed increase in the S contents of other recent sediments has also been attributed to increased sulphur inputs from anthropogenic sources (Mitchell et al., 1983; Nriagu and Coker, 1983). The suggestion that the input determines the sulphur concentration is inconsistent with the presumed general mobility of sulphur in sediments. As a matter of fact, the near-quantitative interconversion of one form of sulphur into the other at a given horizon (d. Goldhaber and Kaplan, 1974; Altschuler et al., 1983) suggests that sulphur may not be as mobile as is generally believed. Most lake sediments are deficient in S and any sulphate ions transferred into them are quickly reduced to sulphide below the oxidized microzone. Furthermore, typical lake sediments contain a large excess of reduced iron which quickly precipitates sulphide ions (Volkov, 1961; Berner, 1971; Nriagu, 1975). Further, recent studies (Smith and Klug, 1981; Mitchell et al., 1981; Nriagu and Coker, 1983) show that a large percentage of the S in many lake sediments is bound to organic material, and consequently also rendered immobile. The immobilization of sulphur can explain why the S contents of the sediments tend to chronicle the S flux into the lake.

Hydrosphere

213

The influence of inputs on sedimentary sulphur is further demonstrated by the strong correlation between the S contents of recent sediments and the sulphate concentrations in the overlying waters (Gorham et at., 1974; Nriagu, 1984). It should be emphasized that the source influence can best be seen in lakes with aerobic hypolimnion. Any development of bottom anoxia increases the transfer of S to the sediments, thereby confusing the relationship between sulphur input and accumulation in the sediments. In fact, the pollution component of S042- in the Great Lakes can be estimated using the shift in the S contents of the sediments. Such data for Lake Erie sediments suggest that anthropogenic sources account for about 70% (18 mg -1) of SOi- (Nriagu, 1984). The extrapolation of the historical graph of changes in SOi- concentrations (in the overlying water) to precolonial times shows a similar pollution input (Beeton, 1965; Pringle et at., 1981). A mass balance of the inputs from the weathering of the bedrocks and the glacial overburdens likewise suggests that 60-70% of the SOi- in Lake Erie comes from pollution sources (Nriagu, 1975). On the basis of the excess sulphur values, it has also been estimated that anthropogenic sources now account for about 30-40% of the SOi- in the waters of Lakes Huron and Superior, and about 60% in Lake Ontario and southern Lake Michigan. These percentages are in good agreement with the estimated background of precolonial SO/concentrations of 14 mg -1 in Lake Ontario, 5 mg -1 in Lake Michigan, 8 mg -1 in Lake Huron, and 2 mg -1 in Lake Superior (Beeton, 1965; Pringle et at., 1981). Potentially, the isotopic technique should provide a confirmation as to whether the sources of S in the recent lake sediments have indeed changed. In one of the first detailed studies, Nriagu and Coker (1976) showed that the surficial sediments of Lake Ontario are depleted in 34Scompared to the older sediments. They interpreted the 334S profiles in terms of sulphur diagenesis and preferential loss of 32Sto the overlying water; the interpretation was adapted from diagenetic models for S in marine sediments. A recent study of sediments as sinks for S in the Great Lakes (Nriagu, 1984) suggests, in fact, that the observed 334Sprofiles are more likely to be a reflection of increasing input of pollutant sulphur into Lake Ontario. Subsequently, the isotopic composition of S in the humic acids (HA) from Lake Ontario has been determined (Figure 6.11). What is remarkable is the depletion of 34S in the surficial samples and the close similarity with the 334S profiles of the other forms of sulphur in these sediments (see Nriagu and Coker, 1976). The 334S of the HA presumably includes the isotopic signature of the precursor organic matter, the isotopic imprints of the subsequent diagenetic changes, and the isotopic effects of any secondary enrichment reactions (Nissenbaum and Kaplan, 1972; Dinur et at., 1980). It is impossible to establish the relative influence of each of these three processes. In view of the little that is currently known about the fractionation

214
0345, D/..) +10

Stable Isotopes
034S,(%O) +10

+5 0

+15 0

+5

+15

E
~
Q) 0
.... Q)

.8. 10

10

:5
Qj

CENTRAL

BASIN

KINGSTON

~
20 20
Q)

.c 15.
0

(Eastern) BASIN

30

30

Figure 6.11 The 034Sprofiles for humic acid sulphur from Lake Ontario

of S isotopes in organic material, it would be ill advised to attribute the shift in isotopic composition solely to a change in the source of the sulphur in the humic acids. Nevertheless, the profile observed is very suggestive of a strong source influence. It may be noted that the HA is depleted in 34S in relation to the acid volatile sulphides and the total S in the same sediment horizons. A more recent study dealt with the concentration and isotopic composition of S in lake sediments located in two contrasting regions with respect to sulphur pollution. The first group of lakes is located near Sudbury, Ontario, and derives most of the sulphur from the big nickel-copper smelting complex nearby (Nriagu and Coker, 1978a; Nriagu and Harvey, 1978). This point source has a fairly constant isotopic signature and the sedimentary record of sulphur pollution in these lakes should be related to the local history of smelting activities. The strong influence of smelter emissions is manifested by the fact that many of the lakes in this area have become acidic (OME, 1981). The second group of lakes are located in the more remote Algonquin Provincial Park with no known major local sources of sulphur. The historical records and isotopic profiles in the sediments should

Hydrosphere

215

presumably reflect the flux of long-range transported sulphur into the lake basins. Representative profiles of total sulphur and 334Sin lake sediments in the two areas are shown in Figure 6.12. A prominent feature of the profiles is the spectacular enrichment of sulphur in the surficial sediments. For the lakes around Sudbury, the dramatic increase in the flux of excess sulphur began around 1890, coinciding with the initiation of smelting operations (Nriagu et at., 1982). It is interesting that the pronounced increase in the accumulation of S in Windy Lake sediments, located about 20 km from Sudbury, dates to roughly 45 years BP, and hence coincides with the installation of the taller 170 m stack at Copper Cliff in 1923. The accumulation of excess S in lake sediments in the Algonquin Park generally began around 1860-70, and may be related to the beginning of extensive industrialization of the Great Lakes basin. These data thus suggest that lake sediments are sensitive to the influx of pollutant sulphur from both local and distant sources. The changes in the 334Sprofiles closely parallel the input of excess sulphur into the lake sediments (Figure 6.12). Typically, the most recently deposited sulphur is isotopically lighter than the sulphur in the older sediments. More importantly, the shift in 334Scoincides with the onset of excess S in every case (Figure 6.12). The sulphur in the surficial sediments near Sudbury tends to be lighter (average 334S value, ~-8%0) than in Algonquin Park sediments of comparable age (mean 334S value ~-0.5%0). The apparent difference may be related to the disparity in the isotopic signature of sulphur from local versus regional sources (Nriagu and Coker, 1983). The highly contaminated sediments of Kelly Lake (a recipient of both domestic and industrial effluents) contain over 3% sulphur with very negative 334Svalues of -20 to -30%0 (Figure 6.12). Various diagenetic models, conceivably, can be invoked to account for the total S or 334S profiles. The time frame for the onset of excess sulphur accumulation in these lakes, however, lends support to their interpretation in terms of input control. The results do suggest that sedimentary sulphur can be used as a time tracer for acid rain deposition in these lakes. The quantification of the historical changes in the flux of acidic sulphur compounds into lakes would require a careful assessment of principal pathways of sulphur flux into the sediments. There are currently few reliable indices of past changes in the buffering capacity of lakes which can be attributed to acidic precipitation, and the sulphur isotopic technique does seem to provide the much-needed clues. A few other studies have addressed the 334S profiles in recent lake sediments. The total S content and 334Sprofiles which Cook (1981) reported in the Experimental Lakes Area of northern Ontario are very similar to those in Figure 6.12. The total S profiles in two lakes in the Adirondack

216

Stable Isotopes

TOTAL S (Dry wi %) 1 2 4

0 034S( -10 %.,) 0

0.2

TOTAL S (Dry wi '.IiI) 0.4 0.6 0.8 I I 0

E
.g w
()
L

0
5

-30

-25

-20

-15

E .g1Ow 0
L
ex:

ex:

~10 w

UJ

20 . .

j!:
::: 15

w30
j!:

9 w
In

40
Kelly Lake
..:.J

i!: a.
UJ Q

LohiLake

i!:50

a.

60 70

0
034S(%") -10 -5

TOTALS (Drywt ,,) 0.2 0.4 0.6 0.6

10,

w 0
L
ex:

20

J:: I-

!z ;;:;30

w In J:: SQ. Ia. w Q 60,

~40 J

Mcfarlane Lake

Figure 6.12 Profiles of total sulphur and o34Sin sediments of representative of Northern Ontario (Nriagu and Coker, 1983)

lakes

Hydrosphere

217

Mountains of New York likewise show an increasing flux of excess sulphur beginning around 1850 (Mitchell et at., 1983); the preliminary data furthermore show a depletion of 34S in the surficial sediments samples (M.J. Mitchell, personal communication, 1983). Thode and Dickman (1983) found the difference in 034S values between the recent and the underlying older sediments to be about 3%0and 9-11%0in lakes located downwind and upwind respectively of the smelters at Wawa, Ontario. In each instance, the 034Swas shifted to a lower value in the surficial sediments. These available data, all from lakes in eastern North America, show a very diagnostic pattern which needs to be confirmed in lakes in the other parts of the world. 6.5 ISOTOPIC COMPOSITION OF SULPHUR IN CONTINENTAL SEAS*

In contrast to the world's oceans, concentration and 034Svalues of sulphate in continental seas vary widely. Isotopic data for different sulphur forms in water and sediments of continental seas are rather sparse even for the largest, the Mediterranean Sea. The Black Sea has been studied comprehensively; less information has been obtained for the Azov, Caspian, Baltic, and Red Seas. From the environmental viewpoint, inland seas may serve as the recipients of anthropogenic sulphur. Alternatively, non-sulphur-containing pollutants might alter the sulphur cycle in these waters, perhaps affecting biota and emission of S compounds to the atmosphere. 6.5.1 The Black Sea The Black Sea is a unique water basin. Its surface layers (150-200 m) contain oxygen, whereas the bulk of its water body (about 87%) contains hydrogen sulphide with concentrations as high as 9.6 mg e-l (average 7.6 mg e-I). The water exchange between the Black Sea and the Mediterranean and Azov Seas via the Bosphorus and Kerch Straits was investigated by Scopintsev (1975). The contents of different sulphur forms in Black Sea sediments have been studied in detail (Ostroumov et at., 1961; Ostroumov and Volkov, 1964). The isotopic composition of sulphur in water and sediments of the Black Sea was first studied by Vinogradov et at. (1962) and later by Migdisov et at. (1974). All the reduced S forms in recent sediments, as well as hydrogen sulphide in the water column, are enriched in the lighter isotope with 034Svalues as low as -42%0. The isotopic composition of SO/- (+ 19.4%0)approximates the oceanic value and is rather homogeneous throughout the total depth.
* V.L. Mekhtiyeva.

218

Stable Isotopes

Some variations in the isotopic composition of SOl- (from + 16.7 to +21.2%0) are observed in the littoral zone (Mekhtiyeva and Pankina, 1968; Chukhrov et al., 1975). The isotopic composition of sulphate in the river system of the Black Sea Basin was investigated by Rabinovich and Grinenko (1979). From data on the contents of different sulphur forms in sediments and the annual sediment deposition rate on the shelf (145 Tg yr-l), the annual reduced sulphur flux to the Black Sea sediments has been estimated to be 2.4 Tg, of which 1.4 Tg reaches the shelf (Lein and Ivanov, 1983). Lein et al. (1983) have calculated the mass isotopic balance for sediments in the hydrogen sulphide zone of the sea. On the basis of the above data, a mass isotopic balance calculation for sulphur in the entire Black Sea has been attempted (Tables 6.13 and 6.14). As seen from Table 6.13, the isotopically light sulphur is buried in sediments. Despite the fact that this flux is relatively small, it establishes

Table 6.13 Flux of sulphur in the Black Sea sediments Average S content (%) HzS SO/S-ftux (l0f> tonnes yr-I) HzS SO/-

Zone Hydrogen sulphide Oxygen

Sediment deposition (106 tonnes yr-I)

/)J4S(%0) HzS SO/-

75 145

1.35 1.01

0.03 0.05

-23.9 20.0

25.6 23.0

1.01 1.46

0.02 0.07

Table 6.14 Mass isotopic 106 tonnes

balance
I

of sulphur

in the Black Sea

/)J4S

Inputs River runoff Flux from the Azov Sea Flux from Bosphorus Total

yr-I 3.8 16.5 158.7

('roo) +4.5 +17.0 +19.8

! Outputs
Efflux to Azov Sea Efflux to Bosporus Burial in sediments

106 tonnes yr-I 14.98 159.9 2.47

/)J4S ('roo) +19.0 + 19.4 -20.0

:
i i

179

+19.2

177

-18.8

Hydrosphere

219

that the isotopic composition for the total efflux from the Black Sea is somewhat lighter than that of the influx. Therefore, the calculated mass balance (Table 6.14) predicts that the 834S value of sulphate is gradually increasing in the Black Sea. 6.5.2 The Azov Sea The Azov Sea is a huge, shallow gulf of the Black Sea. Its salinity is as high as 13%0in the south and 10-12%0in the remainder, decreasing to 2-4%0 near river mouths. The sulphate flux to the Azov Sea includes river runoff, precipitation, fluxes via the Kerch Strait, and a small-scale water exchange with Sivash (Tsurikova and Shulgina, 1964). On the average, 27.5 Tg of sediments with a sulphur content of 1.1% reach the sea bottom annually (SCOPE 19, 1983). The isotopic compositions of water-insoluble sulphate and other sulphur forms in sediments of this water basin have been thoroughly studied (Rabinovich, 1971; Migdisov et al., 1974; Rabinovich and Veselovsky, 1974; Mekhtiyeva, 1975; Chukhrov et al., 1975). The 834S values for S042correlate with the salinity and directions of currents (Figure 6.13). The

360 I

SEA OF AZOV I .""I~P;ffj!j ""~0''.,,

i-

- - -460

+19.0Black Sea Figure 6.13 Map of the Azov Sea with currents (arrows), sampling sites, and 6o4S values of S042-

220

Stable Isotopes

influence of river runoff is clearly evident in zones adjacent to mouths. The admixture with Black Sea water is observed in the south and central parts of the basin. The 034S values of sulphate vary from + 12.8%0 in the middle of the Taganzanrog Gulf to + 18.5%0 in the open, southern part of the sea. The reduced S forms in sediments become gradually depleted in 34S in going from freshwater lagoon to normal sea sediments. These variations in the isotopic composition of sulphur are not related to the distribution of organic carbon in the sediments. On the contrary, they correspond to changes in salinity and sedimentation conditions. Data obtained by Migdisov et ai. (1974) show that sulphur with a 034S value of +9.2%0 is buried in sediments over most of the sea area. In the Kerch region, the 034Svalue is -22%0. If sulphur with isotopic composition similar to the Kerch region is assumed to constitute only 10% of the bulk of the sulphur, then sulphur buried in the Azov Sea sediments has a 034Svalue of about 6%0. The mass isotopic balance of sulphur from the Azov Sea shows that the expenditure of sulphur exceeds its income, with a net loss of 34S-depleted sulphur (Table 6.15). Therefore, SOi- in the Azov Sea is becoming progressively enriched in 34S. 6.5.3 The Caspian Sea The present isolation of the Caspian Sea from the world ocean system is a short period in its long history. This basin is one of the links in the
Table 6.15 Mass isotopic balance of sulphur in the Azov Sea 106 tonnes yr-I 1.15 0.07 0.16 14.98 1)34S (%0) +5.4 7.1 +19.0 +19.0 106 tonnes yr-I 16.50 0.42 0.30 1)34S (%0) +17.0 +16.0 +9.0

Inputs River runoff Precipitation Flux from Sivash Flux from the Black Sea Total

Outputs Efflux to the Black Sea Efflux to Sivash Burial in sediment

I I

16.36

+18.0

17.22

+16.8

Note: Flux data from Tsurikova and Shulgina (1964), Migdisov et al. (1974). Rabinovich and Veselovsky (1974), Scopintsev (1975); isotopic data from Mekhtiyeva (1975), Rabinovich and Grinenko (1979), Lein and Ivanov (1983).

Hydrosphere

221

Mediterranean Sea system, known in palaeogeography as the Thetis Ocean. Ocean waters last penetrated into the Caspian depression about 10 million years ago. Since that time, the Caspian Sea has undergone repeated salination and desalination. The present salinity ranges from 0.03%0in the northern part to 13.5%0in the south-east (Figure 6.14). As the salinity increases, the sulphur isotopic composition also changes. The northern part is influenced by river runoff (334S value of sulphate, +8.7%0). The 334S value increases to +10.5%0 in the middle and to + 13.3%0in the southern regions. The content and isotopic composition of sulphur forms in the Caspian sediments were studied by Mekhtiyeva (1974, 1980). Mekhtiyeva and Rabinovich (1975) calculated the isotopic balance of sulphur for 1973 (Table 6.16).

Kara
'-' Bogaz
,.'

Gol

CASPIAN SEA
Figure 6.14 Map of the Caspian Sea with currents (arrows), sampling sites, and 034S values of SOl-. For Kara-Bogaz-Gol, the average 034Sis based on 15 determinations

222

Stable Isotopes Table 6.16 Mass isotopic balance of sulphur in the Caspian Sea

Inputs River runoff

106 tonnes yr--l 6.682 0.364 0.833 7.879

834S (%0)

Outputs Efflux to
:

106 tonnes yr-l


10.0

8345 (%0)

~
+6.6 +6.2
+4.0

+ 13.0
-22.3

I
i
I

KaraBurial in
Bogaz-Gol 2.49 sediments

Precipitation Ground water Total

1 Wind___~_-=~~
j
I

+6.0

12.56

+6.0

The data show that the sulphur balance in the Caspian Sea is very dependent on the water supply to the Gulf of Kara-Bogaz-Gol, the largest gulf in this area. Kara-Bogaz-Gol is a huge lagoon to which water arrives over high rapids. Due to intensive evaporation, large amounts of salt are precipitated, making this gulf the largest salt deposit in the world. The annual water flux from the Caspian Sea to the gulf changed from 25 km3 in 1930 to 5.39 km3 in 1976. Calculations show that at a water flux to KaraBogaz-Gol of 10-11 km3 yr-1, a balance is established both in the concentration of sulphates in the Caspian Sea and the isotope composition of sulphur. At a water flux of 25 km3 yr-l, the sulphate content in the Caspian Sea would decrease in one year by 30 Tg or 13%, and its 1)34S value would increase by 0.13%0as the result of biogenic sulphide production. 6.5.4 The Baltic Sea The Baltic Sea is almost surrounded by land. The limited water exchange with the North Sea and significant runoff established its low salinity, which decreases drastically from west to east and from bottom layers to the surface. The Baltic Sea has a constant bilayer structure. Its southern and central parts are covered by a water layer with a salinity of 7-8%0. At the mouth of the Gulf of Finland, it decreases to 6.2%0;in Neva Bay, it drops to 0.5%0. The bottom water salinity ranges from 8 to 19%0, increasing towards the Danish Straits. Between the surface and bottom layers, there is an intermediate mixing zone (Chernovskaya et al., 1965; Mikulsky, 1974). The distribution of 1)34Svalues in silt sediment cores of Kiel Bay, over the depth interval from 0 to 40 cm, was studied by Hartman and Nielsen (1969). The sediments in the central, eastern and southeastern parts of the Baltic Sea were studied by Lein et at. (1982) and Migdisov et at. (1983).

Hydrosphere

223

The mass balance of sulphur calculated from these data (Lein and Ivanov, 1983) shows that 0.44 Tg of reduced sulphur with an average 334Svalue of -17.2%0 and -0.616 Tg of sulphate sulphur with a 334S value of +24.3%0 enter the reduced fine-grained sediments over the entire area of the sea annually. The total annual sulphur flux to the sediment of 0.46 Tg has an average 334S value of -15.7%0. A sulphur mass balance for the Baltic Sea is given in Table 6.17; since information is limited these calculations are tentative. Because of the lack of data on the isotopic composition of waters flowing through the Danish Straits, analogous information from the Azov and Black Seas has been used. The flux from the North Sea, varying in salinity from 10 to 12%0, consists of water from the Baltic Sea which mixes with ocean water in Kattegate. The salinity of the latter waters is lower than that of ocean water (15-25%0; Mikulsky, 1974). Preliminary calculations show that the average isotopic composition of sulphates in the Baltic Sea is strongly influenced by the ocean. Lesser effects are produced by continental runoff, sulphur burial in sediments, and fluxes from the atmosphere. 6.5.5 The Red Sea The Red Sea depression was formed as a result of complex rift processes beyond the limits of the continental shelf. The depth increases steeply down to 500--700 m. On the long axis, there is a crack with depths as much as 200--2600 m. It is a zone of young fractures, characterized by tectonic and volcanic activity. Strong thermal currents and the evolution of hot brines

Table 6.17 Approximate mass isotopic balance of sulphur in the Baltic Sea 106 tonnes yr-1 914 0.7 1.0 915.7 834S (%0) +19.5 +4.8
I

Input Flux via straits Precipitation and dry deposition River runoff Total

Outputs Efflux vIa straits Burial in sediments

106 tonnes yr-1 915 0.6

834S (%0) +18.0 -15.7

+6.7 +19.5 915.6 +18.0

" Precipitation anddry deposition, Jensen andNakai(1961)andChukhrovet al. (1977);river runoff, Rabinovich and Grinenko (1979);sediments, Lein and Ivanov (1983).

224

Stable Isotopes

have formed metal-bearing sediments. Sixteen underwater brine 'basins' have been discovered so far. The salinity is substantially higher than that of the ocean due to intense evaporation. It is 36.7%0in the southern and 40.5%0 in the northern parts of the sea. The salinity of thermal waters filling the depressions is much higher (up to 319%0in Atlantis II). This unusual hydrothermal regime of the Red Sea is reflected in the isotopic composition of sulphur in the sediments (Hartman and Nielsen, 1966; Kaplan et at., 1969). The isotopic composition of sulphate from pore waters in metal-bearing sediments in the Atlantis depression is the same as those of brine and seawater sulphates. On this basis and other evidence, it is concluded that bacterial sulphate reduction does not occur in these sediments (temperature about 60C and salinity about 300%0). The sediments in this depression differ from those normally found in the deep ocean. There are abnormally high concentrations of sulphides, mainly sphalerite and to a lesser extent pyrite, whereas in normal sediments, pyrite prevails. In Atlantis II, there is also gypsum, anhydrite, and elemental sulphur. On the basis of 034S values, sulphur of the Red Sea may be separated into four groups: sulphates of pore waters in normal sediments (+ 23%0); sulphates of brines, sulphate minerals, and some pore waters (+ 15 to + 23%0); sulphides, sulphate minerals, and elemental sulphur of metalbearing sediments with temperatures higher than 50C (+ 12 to + 15%0); and sulphides from normal sea sediments and depressions with brine temperatures of less than 50C (- 20 to - 37%0). The sulphur isotope data demonstrate the existence of two sulphur cycles in the Red Sea sediments. One is the normal type found in the water basins and includes both abiogenic and biogenic reactions. The second is only biogenic and typical of thermal waters and highly mineralized sediments. Both systems and their interactions have not been adequately studied and it is impossible to calculate a mass isotopic balance. Sulphate in the body of the Red Sea is isotopically homogeneous and similar to that of ocean water. 6.5.6 Conclusion The isotopic composition of sulphur has been summarized for five inland seas. With the exception of the Red Sea, they are distinguished by low salinities. Many factors determine the isotopic composition of SOi- in continental seas. The most important is water exchange with the World oceans, via straits. The Azov, Baltic, and Caspian Seas are strongly influenced by river runoff bearing SOi- that is isotopically lighter than that of the ocean.

Hydrosphere

225

Sulphate in the Caspian Sea is also affected isotopically by efflux to KaraBogaz-Gol. An important factor for all the basins is bacterial sulphate reduction in bottom silt, resulting in the enrichment of heavy isotopes in SOi-. The isotopic systematics of sulphur in sediments of continental seas is similar to those of the ocean. The distinction is that the initial concentration and isotopic composition of dissolved sulphate may be different.

6.6 RIVERS*
A comprehensive review of the isotopic composition of sulphur in rivers has already been published (Ivanov et al., 1983). This section therefore, will focus primarily on more recent reports. The ;)34S0i- in any river represents a mixture of the sulphur derived from the soil, bedrock, and the atmosphere. It can be modified by biotic processes in the river and its watershed. In principle, the ;)34Smeasurements can be used (a) to fingerprint the sources of sulphur and (b) as a natural tag on the key pathways of sulphur transformation in riverine ecosystems. Combined measurements of the isotopic compositions of sulphur and oxygen in sulphate further constrains the possible sources, key processes, and pathways. So far, however, relatively little work has been done on sulphur isotope systematics in the watersheds of the major rivers of the world. In an ambitious study, Longinelli and Edmond (1983) found fairly uniform 1)34S values in the Amazon basin (Figure 6.15). The sulphur isotope data suggest that a large fraction (over 60%) of the sulphate load of the river originates from the Permian evaporites (1)34S,+ 10%0)in the Andes source. region. About 35% of the dissolved sulphate was attributed to oxidative weathering of sulphides in bedrocks downstream in the drainage basin. The contribution of cyclic sulphate from rainfall was believed to be only 5%. The SOi- concentration in the Amazon River is only 3 mg -1, which is similar to the level one finds in many remote (unpolluted) headwater streams dominated by atmospheric inputs. The conclusion that little of the sulphate in the Amazon is derived from the atmosphere is thus surprising considering the high frequency of volcanic activities in South America. The ;)180 values of sulphate in the Amazon basin hover around +7.5%0 upstream of the Tefe confluence (Figure 6.15). Below this point, a divergence develops with downstream 1)34Svalues becoming less than +3%0 in 1976, but increasing to +90/00in 1977. These values presumably reflect yearly differences in rates of SOi- uptake (assimilatory sulphate reduction) by the local fauna and flora. The absence of discernable parallel effects in the

*J .R. Nriagu.

226
+17.70
.. +8,', "
0 0

Stable Isotopes

,/ 0/.

,"", 0

,/

0+ 1 1. 7

""0"
/ / 0/

/D

0- - -
~
+6
..........

"-8- ~

.e-

_e

"
0

" "01

"

co
.....

0
"0

I.

.\
./
TRIBUTARIES.
MAIN CHANNEL.

e, --e- - -- - --e \

1976
1976
... co 0 0 r/) ... a. 0 ... CJ) <l> z

:t
.--. 0
0 ..........

0 a

1977 1977

._-, \
e

.
~ 0> c: X a.

0 C. a1 Z

...,

a1

a1'> co ..., I

..., I-

... -<l>
I

co

<l>

a1 r/) ... co

u.Q C. co E co ::2' 0 I... I-:-

-0;;

0 co

r/)

... <l> co

co E

<l>

+9
0

(j) <t t<)

ro

a- - . '10-: +7-, -<:--cr

.
0

~
0

r-. ,
/ /',

0
0

.f'<. r--~'

''(J'''' " 'I

0/

-,-0.,

.";-0'

..<-::

0 --

-. ...,.
\
I

...

."''fl

.
and Edmond, 1983)

+5 Figure 6.15 8345 and 81HO values for dissolved sulphate as a function of location
along the main channel of the Amazon River (Longinelli

334Svalues supports this conclusion and argues against extensive dissimilatory sulphate reduction in the river basin. The histograph of 334S042- for the Amazon basin is compared against similar plots of the data for Mackenzie River, Canada, and the major rivers of the Soviet Union (Figure 6.16). The very tight clustering of data from the Amazon strongly suggests that most of the sulphur is derived from one dominant source. The Mackenzie data show two minor clusters in the range of -1 to +4%0 and +8 to 10%0.The faint bimodality and the wide spread in 334Svalues (which spans the isotopic range for the bedrocks) suggest that

Hydrosphere Amazon '77

227

i
Amazon '76

~
Mackenzie I:] ",I:]

0 f1f:Ir:Jr:J~r.r.r.1
...

'"

USSR

.,...,.,,,,, ...,.,.".,.
., """'...,.,.,.

-20

-10

+32

Figure 6.16 834S0/- histographs for the Amazon, Mackenzie, and the major rivers of the Soviet Union (data from Longinelli and Edmond, 1983; Hitchon and Krouse, 1972; Rabinovich and Grinenko, 1979)

the sulphur load in different segments of the river is controlled primarily by inputs from small streams draining geological terrains with isotopically distinct signatures (see Hitchon and Krouse, 1972). The data for the Soviet rivers show a slightly skewed normal distribution about +6%0, although the discharge weighted average was +8%0 (Rabinovich and Grinenko, 1979). The sulphur brought in by individual rivers, however, has distinctive isotopic imprints acquired from the bedrocks. The 034S0i- values of Soviet rivers often show pronounced seasonal changes which can result from (a) input of groundwater sulphate (the sole component during the base flow regime) and (b) the additional isotopically lighter sulphur from surface runoff. Figure 6.17 shows the distribution of sulphur concentrations in the major rivers of the world. The lowest mean SOi- concentration in a major river is about 0.6 mg -l in River Niger (Nriagu, 1986), whereas the arid Colorado

228

Stable Isotopes

.<9

119-24

.>24

Figure 6.15 Sulphate concentrations (mgl-1) for the major world river basins; random dots represent regions with no runoff (Husar and Husar, 1985)

River has the highest average concentration of over 100mg - I. Other


rivers with over 50 mg -l S042- include the Rhine, Vistula and Oder, Mississippi, Shatt al Arab, and Huang He (see Husar and Husar, 1985). The average SOi- concentration in African rivers is less than 3.0 mg -l (Nriagu, 1986). The sulphur load is derived mostly from the weathering of bedrocks dominated by Precambrian shield formations. The mean 034 SOifor African rivers is therefore believed to be +2%0. Sulphate concentrations in South Americn rivers also average under 3 mg -I. Since the Amazon is responsible for about 75% of the sulphur flux from the continent, its mean 034S042- of +7%0 can be regarded as the continental average.

Sulphate concentrationsin Asian riversrange from 4 mg -I (Brahmaputra


and Mekong) to over 70 mg -l (Huang He, Indus, and Shatt al Arab). Most of the sulphur burden comes from the weathering of carbonates and evaporites with only minor contributions from the degradation of aluminosilicates (Ming-Hui et al., 1982; Sarin and Krishnaswami, 1984). As expected, the isotopic composition of sulphate in Asian rivers should be highly variable. The 034S0i- value of + 10%0can be considered a realistic average for these rivers. On the basis of published isotopic data for surface waters of Japan and New Zealand, a 034S0i- average value of +5%0 is proposed for rivers and streams of the Pacific Islands. A large fraction of the dissolved sulphur in Australian rivers is derived from evaporites and from weathering of

Hydrosphere

229

sulphide minerals in bedrocks. On the basis of the few published data, the 034S0l- for Australian rivers can be assigned a provisional mean value of + 12%0. Sulphate concentrations in many rivers of Europe and central North

America now exceed 20 mg -I, the average values for the two continents
being 28 and 26 mg -1 respectively (Husar and Husar, 1985). Most of the sulphur burden is of anthropogenic origin, derived primarily from the burning of fossil fuels and smelting operations. On the basis of the available data in the literature, mean 034S0l- values for Europe and North America have been estimated to be +6 and +4%0 respectively. From available data, a global average for 034S0l- in rivers is estimated to be +7%0. This must be regarded as a very tentative value pending further studies on the isotopic composition of sulphur in the major rivers of the world. 6.7 6.7.1. Introduction The concentration and isotopic composition of aqueous sulphur compounds in ground waters is determined by a variety of inputs and reactions (Figure 6.18). In some areas, wet and/or dry fallout with subsequent evaporative enrichment at ground surface is the dominant source. In other areas, any or all of dissolution or precipitation of sulphate minerals, the oxidation of pyrite or organic matter, bacterial reductions, and sulphide mineralization may be important. Thus, the use of environmental isotopes for the recognition of anthropogenic sulphur in ground waters is a difficult task. Difficulties in interpreting such data is one reason why, until recently, only a relatively small number of papers on the subject have been published. However, recent investigations clearly show that the 34Sand 180 contents of inorganic, aqueous sulphur compounds (primarily SOl-, HzS, or HS-) can provide information on (a) the origin of sulphur; (b) biogeochemistry, evolution of ground waters; and (c) groundwater residence flow paths. The geochemistry of many young and economically important groundwater systems is controlled by the recharge environment, i.e. climatic conditions, precipitation and dry fallout, vegetation cover, soil and bedrock compositions, and depth to the water table. This environment determines how much and
'P. Fritz and R.J. Drimmie.

GROUND WATER*

230

Stable Isotopes

ATMOSPHERICFALLOUT
BIOLOGICAL RETURN FLUX

GYPSUM SUL PHA TES SULPHIDES -

REDOX CYCLING ORGANIC MATTER

Figure 6.18 The isotopic composition and abundance of sulphate in ground waters are determined by a variety of sources and processes, which usually includes: fallout, the dissolution or precipitation of sulphates or sulphides, the mineralization of organic matter, and bacterially mediated or inorganic redox reactions

what kind of sulphur species enter the groundwater regime and it is here that the magnitude of man's influence on the aqueous sulphur system is determined. Rock-water interactions within aquifers subsequently can add considerably more sulphate (or sulphide) than the amounts released in the unsaturated zone. Such 'aquifer' sulphur will reflect the isotopic composition of dissolving aquifer minerals or, where sulphide oxidation has occurred, the environment in which sulphate was formed. These additions may mask recharge input, but can yield information about flow paths and the geochemical evolution of a groundwater body. In the following discussion, recharge environment and ground waters sensu stricto are discussed separately in an attempt to describe the complexities of the sulphur system in the hydrosphere and to summarize the observations that can be derived from isotope analyses on aqueous sulphur compounds. 6.7.2 The recharge environment Where groundwater recharge takes place, man's influence on the sulphate/ sulphide concentrations and isotopic compositions occurs at several levels:

DORSET
T20r

WATERSHED, ONTARIO

PERCH LAKE BASIN, CHALK RIVER, ONTARIO -tIOr tf


t:J.
I

t:J.

TISf-

.
0

l:>
l:>

t:J.

.,.S

.
0

'8 TIO I 0

N",
0

(f)

ID

.0

0 0 t:J.. 0 0

oL
0

o.

.
.
--.

""00-t5

.
.

.
I

. .

-S

. .

oL

l:> PRECIPITATION 0 SOIL SULPHATE (SOLUBLE)

GROUND WATER,

RUNOFF

-10 -t5 8 34S (%0) TIO TIS 0

l:> PRECIPITATION 0 SOIL SULPHATE (SOWBLE) GROUND WATER

...S 834S (%0)

+10

+IS
and shallow ground

Figure

6.19 The oxygen and sulphur isotopic composition of aqueous sulphate in fallout, soil leachate waters in two small watersheds on the crystalline rocks of the Canadian Shield

N W ......

232

Stable Isotopes

(a) Fallout in precipitation or as particulate matter, from reservoirs that arc approximately equal in size (Kramer, 1975) (b) Surface 'contamination' from waste disposal or fertilizer (c) Geochemical or biological reactions initiated through land cultivation, irrigation, or drainage Fallout and its isotopic composition has been described in Chapter 5. The 0180 and 034Svalues of fallout are strongly dependent on the sulphur source and the meteorologic environment in which the oxidation to sulphate occurred. Thus, in coastal environments, marine sulphate tends to dominate (0180 = +9.5%0 SMOW and 034S = +20%0). At inland stations, industrial or bacterial sulphur released from soils and open waters determines isotopic compositions (Nriagu and Coker, 1978b). Terrestrial sources can be variable with 0180 and 034S values as high as +20%0 (Holt et al., 1978; Krouse, 1980). In general, anthropogenic sulphate has 034S values between 0 and +10%0 and 0180 values between +5 and +15%0. In the upper soil zone, fallout participates in a number of biochemical reactions. The nature of these is not yet fully understood, but the effects of such transformations are clearly seen at two small watersheds in northern Ontario, Canada. Fallout in Ontario is strongly influenced by industrial emissions and typical 0180 values vary between + 7 and + 15%0with extreme values as high as + 25%0.The 034Svalues vary between 0 and + 10%0(Novicki, 1976; Holt et al., 1978; Feenstra, 1980). Shallow ground waters, however, show much lower 0 values. This is shown in Figure 6.19 for the Perch Lake basin (Chalk River) and Plastic Lake watershed (Dorset). Both basins are heavily forested. In the Perch Lake basin, the unsaturated zone and the shallow aquifers overlying crystalline bedrock consist of up to 20 m of uniform sand with simple mineralogies. No measurable amounts of sulphate, sulphides, or carbonates are found in the unsaturated zone (2 m to water table). In the Plastic Lake basin, only a thin veneer of glacial till overlies the bedrock. The 034Svalues of fallout in both watersheds are very similar and typical for industrial pollution in southwestern Ontario. Unfortunately, in both studies, the sampling periods were too short to record seasonal variations similar to those noted by Nriagu and Coker (1978a, 1978b). The 0180 values for the fallout sulphate differ somewhat between the two basins, possibly because of differences in source and meteorological conditions. Remarkably similar 034S values were found for sulphate in fallout, soil zone (water soluble), and shallow ground waters (Figure 6.20). This contrasts sharply with the 0180 values in the different compartments, where significant depletions in 180 between fallout and ground water/runoff are recognized.

Hydrosphere +5

233

r.;-o

\ O' /
4

I"
o<t (06

-']

(0 0

. OJ . 0 \;

ro -10 \8\ 8 O/ -15-1

8 -;7 \ .... J!

.
-20 -15 -10

. Nonweathered
Weathered
till

till

0 Transition zone

-5

8345,5042- (%.)

Figure 6.20 The isotopic composition of water-soluble sulphate in glacial tills of south-west Alberta, Canada (after Hendry et af., 1986)

The limited number of samples from soils from the Plastic and Perch Lake basins strongly suggests that the 180 depletion occurs already in the uppermost soil horizons, most probably in the organic rich and biologically very active 'A' horizons. If, then, the 034S values are 'inherited' from the fallout, this would suggest that the sulphur is in steady state, i.e. no significant sources or permanent sinks exist in the soil zone as well as the shallow ground waters of those systems. However, for the sulphur budget of recharge areas, the release of biogenic, including bacteriogenic, sulphur may be important. Nriagu and Coker (1978b) provide data which suggest that 10-30% of fallout sulphur in the Great Lakes basin has such an origin. Reduced sulphur compounds emitted by vegetation are most probably depleted in 34S (Winner et ai., 1981) and it is believed, but not fully documented, that bacterial activities in soils, wetlands, and open-water bodies are mainly responsible for their production and release to the atmosphere. Thus, many detailed studies are required before the role of wetlands versus normal forest vegetation and soils in this cycle is fully described and understood. Minor enrichment in 34S in soils and, more importantly, in aqueous environments are most likely due to bacterial reduction which would cause

234

Stable Isotopes

a preferential enrichment in 34Sand 180 in the residual sulphate (Mizutani and Rafter, 1969). Uptake of sulphur by vegetation or bacterial activities and the mineralization of organic sulphur are expected to be important. This would involve reduction and oxidation reactions, which would cause a 'readjustment' of the 0180 value of sulphate. The origin of oxygen in 'secondary' sulphate depends on the environment in which oxidation takes place. Under aerobic conditions, about two-thirds of the oxygen will come from the water and one-third from atmospheric oxygen (Lloyd, 1968; see Chapter 2), but in oxygen-poor environments, the 0180 value of the sulphate can be very close to that of the water. In both cases, the observed depletion of 180 in soils and the unsaturated zone can be explained (Taylor et at., 1984; van Everdingen and Krouse, 1985). Land cultivation, irrigation, and possibly fertilizers (Chesterikoff et at., 1981) contribute to salinization problems in many parts of the world, and as sulphate is often a major dissolved constituent in the saline waters, isotopic studies might reveal important information about its source. For example, on the prairies of southern Alberta, Canada, saline soils and ground waters threaten large areas of agricultural land. Sulphate is a major constituent of these salts. Canadian prairie soils in Alberta are largely developed on tills which since deglaciation have been subject to weathering to a depth of several metres. Organic carbon in the form of coal chips (Cret?.ceous) is abundant in many areas. During weathering, this organic matter becomes oxidized and is believed to be the source of the sulphate. Isotope analyses on sulphate found in shallow ground waters in recharge environments support this interpretation (Wallick, 1981; Wallick and Huemmert, 1983; Hendry et at., 1986). However, locally and in deeper strata, reduction does take place. Since bacterial or inorganic reduction will cause enrichments in both 180 and 34S, a wide spectrum of isotopic compositions in the aqueous sulphate is found. The 034S values measured in this till aquifer vary between + 10 and -30 permil%o (Shakur, 1982), with the more negative values being attributed to the oxidation of reduced and isotopically depleted sulphur. On the other hand, where reduction is important, pyrite may precipitate a 034S value similar to this dissolved sulphide. Overall, this can result in a depth gradient with more positive 034S values in sulphur found at depth within partially reducing environments and lower 034Svalues near the top, where oxidation of reduced sulphur becomes important. Till deposits are typically very tight with very low hydraulic conductivities. Thus, it can take many weeks before piezometers yield enough water for chemical and isotopic analyses. Hendry et at. (1986) present, therefore, isotope data extracted directly from the tills of southern Alberta (Canada) by an aqueous extraction technique. Their findings (Figure 6.20) are strikingly similar to the data shown in Figure 6.19. In both cases, the 034S

Hydrosphere

235

values of sulphate in weathered and unweathered tills are very similar, but the secondary sulphates of the weathered till have considerably lower (')180 values. Contrary to the Chalk River-Dorset data, the weathered tills are characterized by substantially higher sulphate contents than seen in the unweathered material. Thus, although the physical settings and geochemical environments are different, the data indicate that secondary processes in soils and ground water can produce sulphate with negative 180 values approaching those of the local ground water. The oxidation of reduced sulphur can proceed via inorganic reactions or can be bacterially mediated. Of special interest is a denitrification reaction proposed by Kolle et at. (1983) on the basis of field data. The reduction (loss) of N03 - (often formed by oxidation of NH4+) can proceed with the help of Thiobacillus denitrificans by the reaction: 5 FeS2 + 14 N03 + 4 H+ --3> 7 N2 + 10 SO~- + 5 Fe2+ + 2 H2O The pyrite is thought to be microcrystalline and may have an organi<: matter precursor. The reaction suggests that all oxygen in the sulphate originates in the nitrate. However, a limited number of data from shallow ground water in a small watershed in northern Germany where this proceess appears to be operative indicates that the (')180 values are close to those expected by normal aerobic pyrite oxidation (Kolle et at., 1983; Dr O. Strebel, personal communication). Whether this is due to kinetic isotope effects or the participation of atmospheric oxygen is not yet decided. In conclusion, it is seen that recharge environments are characterized by very intense biological and geochemical processes which have a profound influence on the distribution of environmental isotopes. Thus, the (')180and (')34Svalues in aqueous sulphate and sulphide reflect the environmental conditions of the recharge environment and do not necessarily correspond to the fallout sulphate values. It is important to note that oxygen isotope analyses clearly point to the upper soil horizons as the loci where the transformation of fallout sulphur takes place. In deeper horizons of the unsaturated zone and shallow ground waters, addition of sulphate from pyrite and/or organic matter oxidation is potentially important. Such sulphate will reflect the isotopic composition of its precursor and the (')180 value will be determined by the environment in which the oxidation occurred. The highest (')180values are typically found in sulphate formed in systems open to atmospheric oxygen, whereas the lowest will form in oxygen-poor environments and approach the (')180 value of the water in which they are generated.

236

Stable Isotopes

6.7.3 Ground water


As discussed above, the geochemistry of many young and active groundwater systems is dominated by processes in the recharge environment. Evaporites are usually not important in recharge areas. Exceptions are evaporative environments such as those found in arid zones (Robinson and Ruwaih, 1985; see Section 8.6). In deeper aquifers, sedimentary strata are often important. The dissolution of gypsum and/or anhydrite proceeds with isotope selectivity and the sulphate in solution preserves the evaporite 0 values unless reducing conditions cause a partial loss and heavy isotope enrichments in the residual S042-. The ranges of 0180 and 034S values of marine evaporites as they were deposited through geologic time are shown in Figure 6.21 which also depicts the typical range of 'terrestrial' sulphates as they form through the oxidation of organic matter sulphur, biogenic/diagenetic pyrite, or fallout. The trend for isotope effects associated with bacterial reduction is indicated by the maximum and minimum relative enrichment trends. A summary of isotope analyses on sulphur compounds in groundwater systems has been presented by Pearson and Rightmire (1980).

+-25

+20

MARINE EVAPORATES
1000

:s: +15 0 ~
(J)

~
w f-

+10
+5

..

J: a. ..J ::J (/) 0 ~ co

-----

RANGE OF BACTERIAL REDUCTION TRENDS

-5

+ 15

+ 25

/) 34S SULP",ATE (%. CDT)

Figure 6.21 Range of isotopic compositions in terrestrial and marine sulphates. Trend lines expected for bacterial reductions and isotope exchange between sulphate and water are also indicated

Hydrosphere

237

A practical application was presented by Frank (1974) at the Asse salt dome in Germany where a nuclear waste disposal programme is in progress. Sulphur isotope analyses were used to identify flow paths of ground waters that had passed through different evaporitic sequences. Permian and Lower and Middle Triassic gypsum occur in various horizons. Since the 334Svalues of oceanic SOi- were different during each period, the origin of groundwater sulphate can be determined. Similar studies based on the observation that dissolved, marine sulphates behave conservatively in many aquifers and preserve the signatures of their evaporative precursors have been undertaken elsewhere in Germany (e.g. Nielsen and Rambow, 1969; Geyh and Michel, 1977; Michel and Nielsen, 1977; Nielsen, 1979). Correlations with evaporative sequences permitted determinations of flow paths and also recognized where marine sulphate was mixed with sulphate originating from the oxidation of pyrite. The dissolution of marine gypsum is often accompanied by a dramatic increase in [SOi-]. Figure 6.22 summarizes isotope data collected from a major aquifer in south-west Ontario where good-quality ground water is found in glacial deposits, whereas the underlying bedrock yields more saline water. Well installations have to be kept above these strata, a task that is rather difficult because of a very irregular bedrock topography. The isotope results from this Ontario aquifer show that the shallow ground waters are considerably depleted in heavy isotopes and have lower 3 values than precipitation fallout. This observation was discussed above and appears to be valid for many recharge areas with a biogeochemically active, unsaturated zone. The addition of Salina sulphate, i.e. gypsum from Silurian sequences, is clearly recognized and sulphur isotope data can be used to quantify bedrock water contributions to the water supply wells (Woeller, 1982). This could not be done with the techniques of physical hydrology or other geochemical techniques. In a somewhat different environment, Fontes and Zuppi (1976) studied sulphate sources in springs and ground waters of central Italy and recognized two sources for the dissolved sulphate, one being evaporites enriched in heavy isotopes and the other a 34S-depleted source of reduced sulphur associated with volcanic rocks. It is evident that in groundwater systems in which reduction becomes important such simple considerations are not possible. An interesting example for this is the Edwards aquifer described by Rightmire et al. (1974) and Rye et al. (1981). The latter report that:
The 1)34S values of dissolved sulfide and the sulfur isotope fractionation between dissolved sulfide and sulfate species in Floridan groundwater generally correlate with dissolved sulfate concentrations which are related to flow patterns and residencetime within the aquifer. The dissolved sulfide derives from the slow in situ biogenic reduction of sulphate dissolved from sedimentary gypsum in

238

Stable Isotopes

KITCHENER - WATERLOO AQUIFERS

illill
+30
DEEP
GROUNDWATER IN CONTACT WITH SALINA GYPSUM

. ..

+20
0

~ CL
'" 0'1:
!

.
10

If)

U)"t

ro

0'-

.
SHALLOW GROUND WATER

- .
. ..
0

. .. .

$
I I I I

CD

0 0

@)

-10

+ 10
18 0 3 SO 4 2-( YoO)

+20

Figure 6.22 The oxygen and sulphur isotopic composition of aqueous sulphate in the Kitchener-Waterloo aquifers. References for the ranges of isotope values in the fallout areas: I-Holt et al. (1978), 2-Feenstra (1980), 3-Nriagu and Coker (1978a), and 4-Novicki (1976). Salina gypsum has a 034S= +28%0 (Miles, 1982) the aquifer. In areas where the water is oldest, the dissolved sulfide has apparently attained isotopic equilibrium with the dissolved sulfate (034S = +65%0) at the temperature (28C) of the system. This approach to equilibrium reflects an extremely slow reduction rate of the dissolved sulfate by bacteria; this slow rate probably results from very low concentrations of organic matter in the aquifer.

Hydrosphere

239

In the reducing part of the Edwards aquifer, Texas, there is a general downgradient increase in both dissolved sulfide and sulfate concentrations, but neither the S34Svalues of sulfide nor the sulfide-sulfate isotope fractionation correlates with the groundwater flow pattern. The dissolved sulfide species appear to be derived primarily from biogenic reduction of sulfate ions whose source is gypsum dissolution although up-gradient diffusion of HzS gas from deeper oil field brines may be important in places. The sulfur isotope fractionation for sulfide-sulfate (~38%0) is similar to that observed for modern oceanic sediments and probably reflects moderate sulfate reduction in the reducing part of the aquifer owing to the higher temperature and significant amount of organic matter present; contributions of isotopically heavy HzS from oil field brines are also possible.

A Cambro-Ordivician system in the North Central USA was studied by Gilkeson et al. (1981). They present sulphur and oxygen isotope analyses on sulphate from an aquifer in which the recharge aqueous sulphate changes in 834S value from between 0 and + 7%0up to as high as + 30%0whereas 8180 values increase from between 0 and +5%0to about + 17%0.Interesting, however, is the fact that 'ahead' of the reduced zone is an oxidized zone in which the dissolution of marine carbonates is recognized. The authors explain this somewhat unusual phenomenon as being due to recharge during earlier, but still postglacial, times during which the reducing barrier had not yet been established. They also were able to document that excessively high barium concentrations in these waters could be directly linked to sulphur geochemistry. It has been proposed that oxygen atoms in aqueous sulphate can exchange with those of water (Figure 6.21). Lloyd (1968) presented experimental data which suggested half-times of between 104 and 105 years for this exchange. Thus, in very old groundwater sulphate, one might have expected a 8180 value in equilibrium with that of the water. Recently, however, on the basis of extrapolation from elevated temperature experiments, Chiba and Sakai (1985) argued that the exchange is much slower with a half-time of the order of 107 years. A comparison of published oxygen isotope data for groundwater sulphates and associated water strongly suggests that equilibration has not been significant. As indicated in Figure 6.21, it would cause an enrichment in 180 only and none of the aquifers analysed shows such phenomena. However, oxygen isotope exchange between sulphate and water apparently occurs during sulphate reduction. As indicated above (Figure 6.21), bacterial reduction causes both 180 and 34Senrichments. As the product, i.e. reduced sulphur, is removed from the system, often as Fe-S minerals, both isotopes should show Rayleigh behaviour. A 834Sversus 8180 plot should thus be a straight line, usually with a slope of 0.25-0.4. This, however, is not maintained in all systems and both laboratory (Mizutani and Rafter, 1973; Fritz et al., in preparation) as well as field data (Basharmal, 1983; Fontes

240

Stable Isotopes

et at., 1985) strongly suggest that the initial Rayleigh behaviour of the tHO enrichments does change and asymptotically approaches a constant value which is at least close to the expected value of the sulphate water equilibrium. This is shown in Figure 6.23. The reaction sequence which permits such 'exchange' is not clear but is possibly approximated by the following reaction paths:
H2O

H
S04Z-

+ enzyme ~ SO/-

enzyme complex ~ SO/-

-3>HzS

H
H2O Overall SO/~ S032- equilibration ex:: 1.049 S032- ~ H2O ex::?

This shows the enzyme complexing of sulphate and its reduction to sulphite and sulphide. Also indicated are possibilities for isotope exchange as proposed by Mizutani and Rafter (1973). This behaviour is shown in Figure 6.24 which also summarized unpublished data by Fritz et al. (in preparation). The laboratory data document that a 0180 difference between sulphates and water of about + 30%0is approached if this difference is plotted against the residual sulphate of a reducing system. This is close, if not identical, to the isotopic difference expected for isotopic equilibrium between water and sulphate. The field data (Figure 6.24) come from a study in central Sweden and the 'asymptotic' behaviour of OIHO,if compared to 034S, can be recognized. Again, the isotopic difference between sulphate and water is close to 30%0. The isotope geochemistry of sulphur compounds in ground waters depends strongly on existing redox regimes. When oxidizing conditions prevail, the oxygen and sulphur isotopic compositions of sulphate in ground waters are diagnostic of the origin of the sulphate. It is usually possible to distinguish between sulphates added through the oxidation of reduced sulphur and contributions from marine evaporites. The variable isotopic composition of evaporites deposited during different geologic periods often permits rather detailed analyses of flow paths. Under favourable circumstances, isotope data can be used in mass balance ca1culations that attempt to unravel mixing relationships of different ground waters. Bacterial reduction of sulphate causes both IHO and 34S enrichments. However, whereas the sulphur isotopic composition of the residual sulphate shows Rayleigh behaviour, the 0180 values approach those expected for

+30

VI BOREHOLE

l>. V2 BOREHOLE 0 M3,EI,NI BOREHOLES + SHALLOW GROUNDWATER

$ ~ C1 ~ ~ '" ~

.
"0

REGIONAL SURVEY

~ ~

~20

0 CD (.0

J: Q. :; en + 10

., 0

..

.
SULPHATE

.
~ OlIo~

.
0
l>.l>.l>.

0
0

../. .
+10

00
+20

STRIPA

GROUNDWATERS

--RQCK
+ 60

30

+ 40 (%0) 5UI Ph a t e

+ 50

.70

.80

8345

Figure 6.23 The isotopic composition of sulphate in ground waters in Central Sweden

N .j::. ......

242
+40

Stable Isotopes

--00---0-0
+30'
8 8

l
0

-0
+20
0

0 0

- - -~-o

:r:

'"

!.,I +10 0"


(f)

-------

0 -10 1 2 1 3
I 5

0-- -------

----cP----

-I 10 20

-..Q........0------

-50

-0-0

I 100

RESIDUAL FRACTION (%)

Figure 6.24 The isotopic difference in lXOcontents between water and sulphate in a reducing system is plotted against the residual fraction (% residual SO/-). Open squares are data from Mizutani and Rafter (1973), solid circles are unpublished data from Fritz et al. (in preparation). The figure shows convergence of the experimental data at ol~Owater - olXO,ulphatc = +30%0

sulphate-water equilibria. This bacterially mediated 'exchange' does occur in ground waters, but strict inorganic exchange as proposed by Lloyd (1968) does not playa significant role. Therefore 8180 values of sulphate cannot be used as an indirect dating tool to assess groundwater ages. 6.8 ISOTOPIC COMPOSITION OF SULPHATE IN GAS- AND OILFIELD FORMATION WATERS*

The isotopic composition of S042- in 'formation waters' of oil and gas deposits is appropriate to study for the following reasons: (a) During exploration and exploitation of oil and gas deposits, formation water comes to the earth's surface together with hydrocarbons. Sulphur isotopes can be used to detect its presence and trace its migration in shallow groundwater systems. (b) Formation waters may be contaminated by drilling fluids or water to enhance oil recovery.
*R.G. Pankina.

Hydrosphere

243

(c) Sulphate in formation waters can be biologically and thermochemically reduced to H2S. Oil and gas may be involved in the reduction process. Thode et al. (1958) were the first to report on the isotope composition of SOi- in formation waters of the Devonian Leduc field of Alberta, Canada (Table 6.18). The range in 834Svalues of 10%0can be explained by variations in the isotope composition of Devonian evaporites (see Section 4.3). Data obtained by the author on the sulphur isotopic composition of formation water SO/in gas- and oil-bearing strata of Devonian, Carboniferous, and Permian reservoir rocks in the Middle Volga region are shown in Figure 6.25. Waters in terrigenic deposits of the Middle and Upper Devonian are CaCh brines with very low [SO/-]. The gas does not contain H2S and the 834S values of SO/- vary from +9.7 to + 16.8%0 over the region. Since these values are lower than those of Devonian evaporites, there is a significant contribution of secondary sulphate. The mineralization in waters in Lower Carboniferous deposits varies from several milligrams to 280 g -1. However, the Na + ICl- coefficient is rather stable (0.72-0.85). The [SO/-] content usually varies from 100 to 2680 mg -1. Most samples are enriched in 34Swhereas some approach the Carboniferous evaporite value (+ 14.0%0). The enrichments in 34S result from intensive SO/- reduction, as shown by the presence of 34S-depleted H2S and the existence of sulphate reducing bacteria. Since the minimum 834Svalues were only slightly lower than those for evaporites, contributions from secondary sulphates leached from clays and sands prevailing in the region are minor. In some cases, the lower 834S values might reflect an incursion of Lower Permian waters. Waters from Middle Carboniferous deposits (mostly carbonate) are less mineralized with higher S042- contents (up to 1560 mg -l). However, there are regions where waters are more metamorphosed and contain less SO/-. H2S is found in all dissolved gases. The isotopic composition of
Table 6.18 Isotope composition of SO/- of formation waters in Leduc deposit (0-2) (after Thode et al., 1958) Well 314 527 308 271 153 Okalta Ireton 4 SOl-, ppm 834S, (%0) +19.5 +19.9 +13.2 +22.6 +18.8

743 541 550 676 625 572

+12.6

244

Stable Isotopes

O 5~CE:NE:
5

en 5

~ c..
:2

en LL 0
IX)

10

ffi

lL:
I

~
I

CRETACEOUS
JURASSIC

I I I I

PERMIAN

5 10
Z 0

LL:
:

CARBONIFEROUS

I I

5~VONIAN 0 0

+20

334S (%0)

Figure 6.25 Sulphur isotope composition of SO/- in formation waters of gas- and oilfields in the Middle Volga region. Vertical dashed lines are estimated mean 034S for contemporaneous oceanic S042-

SOi- in these waters varies from +2.6 to + 19.7%0.The significant depletions of 34S identify secondary sulphates as the dominant source. Even intensive SOi- can result in insignificant 34Senrichment if the SOi- concentrations are so high that the actual percentage of reduction is low. In general, Permian waters are highly mineralized with concentrations of SOi- up to 9400 mg -1 and HzS up to 601 mg -1. The isotopic composition of SOi- approximates that of Permian evaporites (+8.2 to + 10.0%0). Belyi and Vinogradov (1972) reported that SOi- in formation waters of Cretaceous and Jurassic deposits of the Amu-Daria syncline had 034Svalues of +10.4 to +15.0%0 and +10.1 to +21.4%0 respectively. Vredenburg and Cheney (1971) reported 034S values for SOiin formation waters in the Wind River basin. In reservoir rocks of different ages, they found +9.1 to + 26.7%0for Mississippian, +9.5 to + 36.6%0 for Pennsylvanian, + 14.5 to +23.4%0for Permian, + 12.4 to + 28.5%0for Triassic, +10.4 to +32.7%0 for Jurassic, and +6.4 to +14.3%0 for Cretaceous. In the Big Horn basin, Orr (1974) reported 034S values for SO/- from +8.3 to +25.3%0 and +10.7 to +27.6%0 in formation waters contacting Carboniferous (Tensleep) and Permian (Phosphoria) deposits respectively. The diversity of the isotopic composition of SO/- in formation waters is explained by the combination of water-rock interactions and biological

Hydrosphere

245

activity. There may be a SOi- remnant from the original water in sediments. Additional SOi- may arise from the dissolution of evaporites (enriched in 34S)or secondary sulphate minerals (often depleted in 34S). Surface or nearsurface S04Z- transported downwards by meteoric water intrusion tends to have lower 534Svalues than deeper evaporites. 6.9 GEOTHERMAL AREAS*

In volcanic and geothermal areas where fluids are derived from basic magmas at high temperatures, HzS is the dominant S-bearing gas and its isotopic composition reflects a magmatic source in the upper mantle. Typical 534S values are around 0%0:Kilauea +0.6 :t 0.2%0, Krafla (Iceland) +0.6 :t 0.3%0 (Robinson, 1986), and Yellowstone 0 :t 3%0(Truesdell et at., 1978). Acidic melts have higher SOzlHzS ratios, but as the temperature decreases HzS becomes the dominant S-bearing gas in low-temperature fumaroles 300C) and geothermal systems. Its isotopic composition often still reflects the magmatic source, but in this case there may have been an involvement of crustal rocks and crustal sulphur. The 534S values of net sulphur discharged from andesite volcanoes in Japan vary from +2.0 to + 12%0as a consequence of crustal contamination (Sakai and Matsubaya, 1977). Average 534S values for HzS in New Zealand geothermal systems range from +3.3 to +7.9%0, dominantly due to their setting in rhyolites/ andesites (Robinson, 1986). In the near-surface environment, the HzS oxidizes rapidly to [SOi-] in surface waters. Under the extreme conditions in volcanic crater lakes and fumaroles (S concentration 5-10 g kg-I; pH 1), such sulphate may have 534S values up to + 20%0which can be interpreted as approaching isotopic equilibrium with the HzS (Robinson, 1978, in New Zealand; Sakai and Matsubaya, 1977, in Japan). Normally, however, the sulphate represents complete or partial oxidation of HzS without any isotopic fractionation; e.g. Ketetahi hot springs in Mt Tongariro, New Zealand, have 534S values of +2.6 :t 0.7%0 for sulphate which represents complete oxidation of HzS. Sulphate is also present deep in geothermal systems but in much lower concentrations (30 mg kg-I at Wairakei, New Zealand). This primary sulphate has much higher 534S values (up to +24%0), representing an approach to or attainment of isotopic equilibrium with HzS at temperatures around 350-370C. Due to the slow rates of isotopic exchange, the hightemperature value is frozen in. If exchange rates are known, residence times in the geothermal system can be calculated (Sakai, 1983). In contrast, oxygen isotopes in the sulphate exchange with water in the geothermal reservoir and can be used as a geothermometer (McKenzie and Truesdeli, 1977). The
'B.W. Robinson.

246

Stable Isotopes

deep geothermal fluids are characterized by high chloride and low sulphate concentrations with high 034S values. However, near-surface sulphate produced by H2S oxidation characterizes waters that have low chloride and high sulphate concentration with low 034S values. Geothermal waters in many parts of the world (United States, New Zealand, and the Philippines) represent mixing between these two end numbers and data scatter about mixing lines as shown in Figure 6.24. In some coastal geothermal areas of Japan, sulphate 034Svalues of around +20%0 have been used as evidence for present-day sea water contaminating the geothermal fluids (Sakai and Matsubaya, 1977). Also in Iceland, 034S values up to +8%0 have been interpreted by Sakai et al. (1980) as due to some seawater interaction. Thermal waters of the Green Tuff type in Japan represent Miocene oceanic sulphate, which in some areas has undergone bacterial reduction, and sulphate 034S values range from + 20 to + 34%0. Similar high 034S values have been found in some New Zealand warm springs 50C) where bacterial reduction may have taken place.

+25

CJ Deep
+20

Sulphate

..

+15
i i
0 eft.

i 0-1
I

l
H2S

~~y

J:/~<v
",0

Mixing
deep sulphate

lines
with

of
sulphate

if) '<t M
c..O

~ i.

~t:.:J

N.Z.

5l~
H2S 0

" ~<v"

produced non-equilibrium

by near-surface
oxidation of H2S

l Y/ST ~ I
0
10 20 30 40 50 60

'

-5

ct/S04

(Wt)

Figure 6.26 Mixing of deep and shallow sulphate in geothermal systems. Data scatter about the mixing lines for Yellowstone (Truesdell et al., 1978) and New Zealand (Robinson and Sheppard, 1986)

Hydrosphere

247

Continental geothermal areas in India have sulphate with a wide range in 034S values: +6.5 to +37.5%0 (Giggenbach et al., 1983). The oxygen isotopic composition of this sulphate has equilibrated with the thermal water down to 100C. Little alteration in the near-surface environment appears to have taken place and the sulphate is thought to originate from sedimentary material at depth. 6.10 6.10.1 Introduction Sulphur isotope compositions have been measured for sulphur-containing components in a number of springs throughout the world. To a lesser extent, 0180 values for dissolved and related mineral sulphates have also been determined. The isotope data have proved useful in identifying strata with which these waters have interacted and elucidating associated physical, chemical, and biological processes. Processes occurring in springs are site specific. Two subsurface sulphur source extremes can be considered: sulphide minerals and sedimentary sulphate (Muller et al., 1966; Pankina et al., 1966). It is highly likely that the dissolved S-containing ions are subsequently redistributed in response to redox reactions and pH changes. Sulphur isotope variations in thermal and cold springs of the world have been reviewed by Krouse (1974). For this volume, the sulphur isotope geochemistry of a few springs where either sulphide or sulphate minerals were primarily dissolved will be examined. 6.10.2 Springs containing primarily dissolved sulphate In western Canada and probably world wide, the majority of dissolved sulphur in warm and cold springs has origimted from the solution of sulphate minerals. In many cases, this sulphate is in evaporites deposited in ancient ocean basins. Such springs often precipitate travertine. The isotope geochemistry of these springs is essentially that of ground water (Section 6.7) and its modification upon emergence. The concentration of the dissolved ions are altered by pressure reduction, more aerobic conditions, and escape of gases to the atmosphere. A dominant feature of these waters is that the dissolved sulphate often undergoes bacterial reduction and the sulphide product is markedly depleted in 34S. Data illustrating this isotopic selectivity for wells, boreholes, and springs south of Great Slave Lake, Canada, are shown in Figure 6.27. WARM AND COLD SPRINGS*

H.R. Krouse.

248

Stable Isotopes

HS-

8345

(0/00)

Figure 6.27 Distribution of /)34Svalues for dissolved sulphate and sulphide in wells, boreholes, and springs south of Great Slave Lake, Canada (after Weyer and Krouse, 1979)

Upon approaching the orifice, the sulphide tends to be reoxidized, although significant quantities of H2S are usually found escaping to the atmosphere. The extent of this reoxidation can seldom be determined from the sulphur isotope composition of SOi-. However, since oxygen is introduced from O2 and H2O during oxidation (Section 2.2.5), the &180 value of SO i- can be used effectively to assess the sulphide reoxidation. This technique is particularly effective at higher latitudes where meteoric water is more depleted in 180. For example, in spring waters near Paige Mountain, NWT, Canada, van Everdingen et at. (1982) demonstrated that up to 30% of the SOi- was reoxidized sulphide. Part of the oxidized sulphide may be precipitated as elemental sulphur and secondary sulphate, often in algal mats, near the spring orifice. These precipitates have &34S values near that of the dissolved sulphide, as shown for a few springs of western Canada in Figure 6.28. Some secondary sulphate may arise by oxidation of evolved H2S to S02 and H2SO4 which subsequently attacks carbonate minerals. This mechanism has been invoked to explain cave development in travertine deposits (van Everdingen et at., 1985). The secondary sulphates are depleted in both 34S and 180 as compared to the primary sulphate source. In some hydrological systems, secondary sulphate depleted in 34Sand 180 may be subsequently reduced again to render dissolved sulphide extremely depleted in 34S. This appears to be the case for some Flysch waters in Czechoslovakia (Smejkal et at., 1971).

Hydrosphere

249

LOCATION LA SALINE MISSION

oHS- .So- AALGAE oGYPSUM .S042I


I

. .
I

MIETTE
COLD (JASPER) UPPER l.L. KIDNEY

...

l.L.MIDDLE z CAVE <X:


aJ BASIN CANMORE

. .
...

VERMILION L.

HIGHWAY NO.1

.
o
I I I

LUSSIER R.
TURTLE MT.

-30

-20

-10
834

0
S (%0)

+10

+20 +30

Figure 6.28 Sulphur isotope variations in some selected springs from western Canada (after Krouse et al., 1970)

6.10.3 Springs containing primarily dissolved sulphide In volcanic and geothermal areas, HzS is usually the dominant S-bearing species in the fluids and often has 034S values near 0%0, reflecting its magmatic source (see Section 6.9). There are, however, rare occurrences of warm and cold springs where dissolution of mineral sulphides constitutes the main source of sulphur. Such springs often form colourful surface deposits of sulphate minerals such as jarosite. The ultimate origin of the sulphate from deeper mineral sulphide deposits is deduced from metal ion content, variable 034S values of the sulphate that are close to those of the deep sulphide deposit, and 0180 values of the sulphate that may be highly negative at high latitudes. A few such springs in western Canada were studied by Shakur (1982). Barite sinter deposits at Flybye Springs, NWT, Canada, were investigated by Cecile et ai. (1984). Although the 034Svalues ranged from + 15 to +23%0, they were not as positive as barites in vugs of nearby Early to Middle Devonian carbonate units (+29 to +32%0). Further, the 0180 values of the sinter barite were in the range -11 to -17%0, in contrast to the vug barite values of + 17 to + 26%0. The isotope data clearly show that the sinter deposits did not result from simple dissolution of barite, transport, and

250

Stable Isotopes

precipitation. Rather, sulphate reduction occurred afterdissolution, andthe sinter sulphate constitutes oxidized sulphide from that reduction.
6.10.4 Sulphur geochemistry of springs in relation to anthropogenic sulphur From the viewpoint of this volume, the sulphur isotope geochemistry of springs is relevant to many aspects of anthropogenic sulphur in the environment. Emissions of gaseous sulphur compounds from certain springs are substantial, even to the point that high HzS concentrations threaten life. Emission rates have seldom been measured for individual springs and the contributions to the atmospheric sulphur reservoir from springs world wide is not well known. There is the further complication that biogenic emissions from some springs may vary with the seasons. Parameters that may change seasonally include flow rate, temperature, pH, and EH. These in turn influence the concentrations of S-containing ions and microbiological conversions. Consequently, the emissions may change in rate and isotope composition. In local situations, seasonally dependent emissions from springs could conceivably complicate assessments of anthropogenic contributions of S compounds to the atmosphere. Conc~ntration measurements alone would prove inadequate. However, it is highly likely that the 334S values of the natural and industrial emissions will differ sufficiently so that, in combination with concentration measurements, sources could be duly apportioned. Another issue is the extent to which anthropogenic activities might alter the S content and/or microbiological conversions in springs. It is noted that anthropogenic perturbations need not be limited to the addition of sulphur compounds. Introduction of other chemicals, such as nitrogen compounds from fertilizer applications or organic nutrients, might upset physical chemical parameters and microbial balances. Such effects are potentially more likely in springs associated with shallow groundwater systems. 6.11 SUMMARY AND KEY AREAS FOR FURTHER RESEARCH*

The principal sulphur reservoirs in the ocean are shown in Figure 6.29. With the exception of the SO i- pool, reliable data on the sizes and isotopic compositions of these reservoirs are currently not available. The transfer of sulphur between the reservoirs involves physical, chemical, and biological processes, many of which are isotopically selective. Among the transformations, dissimilatory SOl- reduction has been recognized as being the most dramatic in terms of S isotope fractionation, and typically results in

J.O. Nriagu.

Hydrosphere Volatile S Compounds 0 to +5

251

Seasalt Spray +12 to +18

Figure 6.29 Isotopic composition of marine sulphur reservoirs and typical directions of isotopic shift along the transfer pathways

accumulation of the lighter 32S in the reduced sulphide reservoirs. Isotopic differentiation associated with SOi- assimilation is usually small under natural conditions. Little is known about the isotopic effects of the biologically mediated production of dissolved and volatile organic sulphur compounds. It is conceivable that the production of volatile sulphur compounds is an important pathway in the enrichment of ocean waters with 34S and the transfer of the lighter 32S to the continents. The isotopic composition of present-day sea water is uninspiring in that it deviates but little from a constant value of +21%0 (Section 6.2.1). The surprising feature, however, is that oceanic 034S0i- had pronounced secular trends (see Section 4.2). The available data in the literature suggest that the mean oceanic 034S value was similar to the present-day value in late Proterozoic, became heavier in Cambrian through Devonian, declined to a minimum in the Permian, rose again sharply during the Triassic, and has been increasing gradually since the mid-Cretaceous (Figure 6.30).

252
0
I .."1' I

Stable Isotopes

100

,
I I I I I I I

T
K J

~
~

200 300
400-1

R
p IP M D S

u
!) ~

I
~

/'
30-2

~;---~

500-1
600

!
I

, .~---s

n
I
I

~I P
2
(b)

+10

20
(0)

-0+

8345 sulpha,., (lc 00)

<>13 0 Ccarbana'. (0/00)

Figure 6.30 Isotopic records of marine sulphate and inorganic carbon. (a) The sulphur isotopic record of marine evaporites, in per mil relative to the Canyon Diablo Troilite standard. (b) The carbon isotopic record of marine carbonate, in per mil relative to the Peedee Belemnite standard. (Both curves taken from Klump and Garrels, 1986)

Furthermore, the mean of 034S-S0i- and o13Ccarbonate for the geological periods are negatively correlated (r = 0.839), with the linear regression, 034C = 3.074 - 0.131o34S (Veizer et al., 1980)

When examined in greater detail, however, cyclical patterns with clockwise and counterclockwise excursions can be discerned (Klump and Garrels, 1986). This relationship stems from the fact that bacterial SOi- reduction is a major pathway for organic matter decay in marine sediments. It implies that there is a close coupling of the sulphur and carbon cycles which in turn has controlled the levels of O2 in the atmosphere. The covariance of 034S and ol3C through time indicates that, during periods of organic carbon burial, oxygen derived from photosynthesis is partly consumed by the oxidation of pyrite to sulphate: carbon is translocated from the inorganic to the organic pool. On the other hand, during periods of intensive sulphate reduction, pyrite is generated and organic carbon is oxidized with the concurrent

Hydrosphere

253

regeneration of the inorganic carbon pool. The linkage of the marine sulphur-carbon-oxygen cycles can be expressed by the following reactions: 8 SO~- + 2 Fez03 + 15 CHzO + 16 H+ ~ 4 FeSz + 15 COz + 23 HzO or 2 Fez03 + SO~- + 16 COz + 8 HzO ~ 4 FeSz + 16 HC03 + 15 Oz The SO~- includes seawater sulphate or evaporitic gypsum and anhydrite, and CHzO represents organic carbon. The use of sulphur isotopes in deciphering the linkages between the sulphur cycle and the biogeochemical cycles of the other key elements (N, P, Si, Fe, 0, and C) remains a fruitful area for future research. In this respect, it should be noted that in a model in which the burial rate of organic carbon increased by about 30%, the response of the oceanic o\3C was nearly instantaneous whereas the 034S value of SOl- changed at a much slower rate, reflecting the long turnover time of the oceanic sulphate reservoir. Oxygen levels are predicted to respond fairly quickly in comparison to the transfer of material between the oxidized and reduced reservoirs of C and S. The latter approach their new steady-state masses and isotopic compositions over several millions of years (Klump and Garrels, 1986). The present-day biosphere shows a discern able dichotomy in the distribution of sulphur isotopes: 3ZStends to accumulate in terrestrial ecosystems whereas the marine environment is enriched with the heavier 34S. The available data are clearly inadequate to ascertain the implications of this arrangement on the 034S of the oceans (see Section 6.2.1). There is need for isotopic measurements on the sulphur transported by the major rivers of the world. Even less is known about the isotopic composition of the groundwater sulphur reservoir (Section 6.7). Without these data, it will be impossible to determine the overall isotopic balance which is necessary to evaluate published models of the global sulphur cycle. Lakes and rivers constitute a tiny sulphur pool in comparison to oceans. Small inputs of sulphur from anthropogenic sources can thus significantly change the concentrations and isotopic composition of the sulphur in many freshwater ecosystems. Indeed, lakes and rivers have become the critical target ecosystems most susceptible to stress from acid rain (Section 6.4.1). The reservoirs and principal transport pathways for sulphur in freshwater ecosystems are depicted in Figure 6.31. Since lacustrine environments are mostly sulphur deficient, the biota are less discriminating in their metabolism of sulphur isotopes and the fractionation by the various biogeochemical processes generally tends to be small (compare Figure 6.29 and 6.31). Sulphate reduction (equations above) has now been recognized as a major source of alkalinity in poorly buffered lakes (Kelly and Rudd, 1984; Carignan,

254
Rivers -5 to +15
Volatile S Compounds 0 to +5

Stable Isotopes
Atmospheric Precipitation 0 to +15
",

':::,:- 1
-~"""--,.-:.~ - " ~.'... Ground water' -5 to +5

Figure 6.31 Typical a34s values of sulphur reservoirs of the lacustrine environments

1985). The sulphur cycle therefore plays an indirect key role in determining the response of a susceptible lake to acid rain inputs and the rate of recovery of an acid-stressed lake. The reaction pathway for reduced sulphur in freshwater ecosystems differs from that of the marine cycle in one fundamental way: the HzS formed is tied up primarily by organic matter and, to a lesser extent, as 'FeS' (Nriagu and Soon, 1985; David and Mitchell, 1985). Formation of pyrite, typical of marine environments, is very much a minor process and presumably occurs by a radically different pathway involving the reaction of organic sulphur compounds with iron: S-(L) + So + Fez+ ~ FeSz + US-(L) + HS- + FeH ~ FeSz + HLwhere (L) is any organic ligand. Evidence for the above reaction comes from the isotopic compositions of pyrite in lake sediments which are usually very close to that of the organic sulphur fraction (Nriagu and Soon, 1985). Alternatively, organic Sand FeSz might arise from the same dissolved sulphide. The actual process of organosulphur formation and the role of these compounds in the lacustrine sulphur cycle remains an important area for research.

Hydrosphere

255 REFERENCES

Aizenshtat, Z., Stoler, A., Cohen, Y., and Nielsen, H. (1981). The geochemical sulfur enrichment of recent organic matter of polysulfides in the Solar-Lake. In: Bjoroy, M., et al. (Eds.) Advances in Organic Geochemistry, Wiley, New York, pp. 279-88. Almer, B., Dickson, W., Ekstrom, c., and Hornstrom, E. (1978). Sulfur pollution and the aquatic ecosystem. In: Nriagu, J.O. (Ed.) Sulfur in the Environment, Vol. II, Wiley, New York, pp. 271-311. Altschuler, Z.S., Schnepfe, M.M., Silber, c.c., and Simon, F.O. (1983). Sulfur diagenesis in the Everglades peat and the origin of pyrite in coal. Science, 221, 221-7. Atkinson, L.P., and Hall, J.R. (1976). Methane distribution and production in a Georgia salt-marsh. EST COAS M, 4 (6), 677-86. Atlas of the Oceans (in Russian) (1977). Vols I and II, Leningrad, MO 555R, VMF. Ault, W.V., and Kulp, J.L. (1959). Isotope geochemistry of sulphur. Geochim. Cosmochim. Acta, 16, 201-35. Banat, I.M., Lindstrom, E.B., NedwelL, D.B., and Balba, M.T. (1981). Evidence for coexistence of two distinct functional groups of sulfate-reducing bacteria in salt marsh sediment. Appl. Environ. Microbiol., 42, 985-92. Basharmal, M. (1983). The isotopic composition of sulphur compounds in landfills. M.Sc. thesis, University of Waterloo. Beeton, A.M. (1965). Eutrophication of the St. Lawrence Great Lakes. Limnol. Oceanogr., 10, 240-54. Belyi, V.M., and Vinogradov, V.I. (1972). Isotope composition of highly concentrated hydrogen sulfide gases in oil and gas bearing regions (in Russian). Geol. Nefti, Gaza, 7, 37-41. Berner, R.A. (1971). Principles of Chemical Sedimentology, McGraw-Hill, New York. Bouleque, J., Lord, c.J., and Church, T.M. (1982). Sulfur speciation and associated trace-metals (Fe-Cu) in the pore waters of Great Marsh, Delaware. Geochim. Cosmochim. Acta, 46 (3), 453-64. Brunold, c., and Erisman, K.H. (1974). HzS as sulfur source for Lemna-minor L influence on growth; sulfur content and sulfate uptake. Experientia, 30 (5), 465-7. Brunold, c., and Erisman, K.H. (1975). HzS as sulfate source in Lemna-minor L. 2. Direct incorporation into cysteine and inhibition of sulfate assimilation. Experientia, 31 (5), 508-10. Calvert, D.V., and Ford, H.W. (1973). Chemical properties of acid-sulfate soils recently reclaimed from Florida Marshland. Soil Sci. Soc., 37 (3), 367-71. Carignan, R. (1985). Quantitative importance of alkalinity flux from the sediments of acid lakes. Nature, 317, 158-60. Caron, F. (1984). Evaluation a l'aide des mesures des rapports des isotopes stables du soufre et de l'oxygine, de la proportion des sulfates mesures dans les lacs attributable aux depots atmospheriques. Thesis, Inst. National Recher. Scientif. Eau, University of Quebec, Quebec City. Cecile, M.P., Goodfellow, W.D., Jones, L.D., Krouse, H.R., and Shakur, M.A. (1984). Origin of radioactive barite sinter, Flybye springs, Northwest Territories, Canada. Can. J. Earth Sci., 21 (4), 383-95. Chambers, L.A. (1982). Sulfur isotope study of a modern intertidal environment and interpretation of ancient sulfides. Geochim. Cosmochim. Acta, 46 (5), 721-8. Chambers, L.A., and Trudinger, P.A. (1979). Microbiological fractionation of stable sulfur isotopes: a review and critique. Geomicrobiol. J., 1, 249-93.

256

Stable Isotopes

Chernovskaya,E,N" Pastukhova,N,M" Buinovich, A,G" ctal. (1965), Hydrochemical Regime of the Baltic Sea, Gidrometeoizdat, Leningrad, p. 168. Chesterikoff, A., Lecolle, P., Letolle, R., and Carbonnel, J.P. (1981). Sulfur and oxygen isotopes as tracers of the origin of sulfate in Lake Creteil (southeast of Paris, France). J. Hydrol., 54, 141-50. Chiba, H., and Sakai, H. (1985). Oxygen isotope exchange rate between dissolved sulfate and water at hydrothermal temperature. Geochim. Cosmochim. Acta, 49, 993-1000. Chukhrov, F.V., Churikov, V.S., Yermilova, L.P., and Nosik, L.P. (1975). Variations in the isotopic composition of sulphur in some natural water basins. Geokhimiya 3, pp. 343-56; Geochem. Int., 12, 20-3l. Chukhrov, F.V., Churikov, V.S., Yermilova, L.P., and Nosik, L.P. (1977). Isotopic composition of atmospheric sulphur and its possible evolution in the history of the Earth. Izvestiya Akad. Nauk SSSR, Ser. Geokh., 7, 4-13. Claypool, G., Holser, W.T., Kaplan, I.R., Sakai, H., and Zak, I. (1980). The age curves of sulfur and oxygen isotopes in marine sulfate and mutual interpretation. Chem. Geol., 28, 199-260. Cook, R.B. (1981). The biogeochemistry of sulfur in two small lakes. Ph.D. Thesis, Columbia University, New York City, New York. David, M.B., and Mitchell, M.J. (1985). Sulfur constituents and cycling in waters, seston, and sediments of an oligotrophic lake. Limnol. Oceanogr., 30, 1196-207. Deevey, E.S. (1972). Biogeochemistry of lakes: major substances. In: Likens, G.E. (Ed.) Nutrients and Eutrophication, Vol. I, American Society Limnol. Oceanogr., Special Symposia, pp. 14-19. Deevey, E.S., Nakai, N., and Stuiver, M. (1963). Fractionation of sulfur and carbon isotopes in a meromictic lake. Science, 139, 407-8. Dinur, D., Spiro, B. and Aizenshtat, Z. (1980). The distribution and isotopic composition of sulfur in organic-rich sedimentary rocks. Chem. Geol., 31, 37-5l. Ellis, R.J. (1969). Sulfate activation in higher plants. Planta, 88 (1), 34-42. Eriksson, E. (1960). The yearly circulation of chloride and sulfur in nature; meteorological, geochemical and pedological implications, part 2. Tellus, 12, 63-109. Faller, N. (1972). Sulfur dioxide, hydrogen sulfide, nitrous gases, and ammonia as sole sources of sulfur and nitrogen for higher plants (in German). Z. Pfianzenernaehr. Bodenk., 131 (2), 130-8. Feenstra, S. (1980). The isotopic evolution of sulphate in a shallow groundwater flow system on the Canadian Shield. M.Sc. thesis, University of Waterloo, manuscript. Fontes, J.C., and Zuppi, G.M. (1976). Isotopes and chemistry of sulphate bearing springs of central Italy. In: Interpretation of Environmental Isotopes and Geochemical Data in Ground Water Hydrology, IAEA, Vienna, pp. 143-58. Fontes, J.e., Michelot, J.L., and Fritz, P. (1985). Stable isotope geochemistry of sulphur compounds. In: Stripa Project, Hydrogeological and Hydrochemical Investigation in Boreholes, T.R., pp. 85-106, Ch. 7. Frank, H. (1974). Hydrogeologische und hydrogeochemische Untersuchungen an der Asse bei Wolfenbuettel. Gesellsch. fuer Strahlenforsch., Bericht R-87 , Miinchen, p.208. Friend, J.P. (1973). The global sulfur cycle. In: Rasool, S.I. (Ed.) Chemistry of the Lower Atmosphere, Plenum Press, New York, pp. 177-20l. Fritz, P., Fries, e.A., and Drimmie, R.J. (1986). Isotopic investigation of the sulfur cycle in selected Dorset watersheds. Final Report, Project No. 509-09. Department of Earth Sciences, University of Waterloo, Waterloo, Ontario, 32 pp.

Hydrosphere

257

Fritz, P., Drimmie, R.J., Ibsen, J., and Qureshi, R.M. (in preparation). Oxygen isotope equilibration during sulfate reduction. Fry, B. (1986). Sources of carbon and sulfur nutrition for consumers in three meromictic lakes of New York State. Limnol. Oceanogr., 31, 79-88. Fry, B., Scalan, R.S., Winters, J.K., and Perker, P.L. (1982). Sulfur uptake by salt grasses, mangroves, and seagrasses in anaerobic sediments. Geochim. Cosmochim. Acta, 46 (6), 1121-4. Fry, B., Gest, H., and Hayes, J.M. (1983). Sulphur isotopic compositions of deepsea hydrothermal vent animals. Nature, 306 (5), 51-2. Garrels, R.M., and Thompson, M.E. (1962). A chemical model for seawater at 25C and one atmosphere total pressure. Am. 1. Sci., 260, 57-66. Gershanovich, D.E., Gorshkova, T.I., and Konyukhov, A.I. (1974). Organic substance of modern suboceanic peripheral regions of continents. In: Organic Substance of Model and Fossil Sediments and Methods for Its Study (in Russian), Nauka, Moscow, pp. 63-80. Geyh, M.A., and Michel, G. (1977). Zur Isotopenchemie von Mineralwassern und sonstigen Grundwassern im Heilquellengebiet von Bad Oeynhausen und Bad Salzuflen. Fortschr. Geol. Rheinld. u. Westf., 26, 229-52. Giggenbach, W.F., Gonfiantini, R., Jangi, B.L., and Truesdell, A.H. (1983). Isotopic and chemical composition of Parbati Valley geothermal discharges, North-West Himalaya, India. Geothermics, 12, 199-222. Gilkeson, R.H., Perry, E.C., and Cartwright, K. (1981). Isotopic and geologic studies to identify the sources of sulfate in groundwater containing high barium concentrations. University of Illinois, Water Research Center Report 81-0165. Goldhaber, M.B., and Kaplan, I.R. (1974). The sulfur cycle. In: Goldberg, E.D. (Ed.) The Sea, Vol. 5, Wiley, New York, pp. 569-655. Gorham, E., Lund, J.W.G., Sanger, J.E., and Dean, W.E. (1974). Some relationships between algal standing crop, water chemistry and sediment chemistry in the English lakes. Limnol. Oceanogr., 19, 601-17. Gorlenko, V.M., Dubinina, G.A., and Kuznetsov, S.I. (1977). Ecology of aqueous microorganisms (in Russian), Nauka, Moscow, 288 pp. Granat, L., Rodke, H., and Hallberg, R.D. (1976). The global sulphur cycle. In: Soderlund, R. (Ed.) Nitrogen, Phosphorous and Sulphur-Global Cycle, SCOPE Report 7, Ecol. Bull. (Stockholm), 22, 89-134. Grinenko, V.A., and Grinenko, L.N. (1974). Geochemistry of Sulphur Isotopes, Nauka, Moscow, 272 pp. Harrison, A.G., and Thode, H.G. (1958). Mechanism of bacterial reduction of sulphate from isotope fractionation studies. Trans. Faraday Soc., 54, 84-92. Hartman, M., and Nielsen, H. (1966). Sulfur isotopes in the hot brine and sediment of Atlantis II deep (Red Sea). Marine Geol, 4, 4. Hartman, M., and Nielsen, H. (1969). 34SWerte in rezenten Meeresedimenten und ihre Deutung am Beispiel einiger Sedimentprofile aus der westlichen Ostsee. Geol. Rundschau, Bd., 58 (H3), 621-55. Hendry, J., Cherry, J.A., and Wallick, E.I. (1986). Origin and distribution of sulfate in a fractured till in Southern Alberta, Canada. Water Resource Res., 22, 45-61. Hitchon, B., and Krouse, H.R. (1972). Hydrogeochemistry of the surface waters of the Mackenzie River drainage basin, Canada: stable isotopes of oxygen, carbon, and sulfur. Geochim. Cosmochim. Acta, 36, 1337-57. Holland, H.D. (1973). Systematics of the isotopic composition of sulfur in the oceans during the Phanerozoic and its implications for atmospheric oxygen. Geochim. Cosmochim. Acta, 37, 2605-16.

258

Stable Isotopes

Holser, W.T., and Kaplan, I.R. (1966). Isotope geochemistry of sedimentary sulfates. Chem. Geol., 1, 93-135. Holt, B.D., Cunningham, P.T., and Kumar, R. (1978). Seasonal variations of oxygen-18 in atmospheric sulphates. Int. J. Environmental Analyt. Chem. (manuscript). Horn, R.A. (1969). Marine Chemistry. The Structure of Water and Chemistry of Hydrosphere, Wiley-Interscience, New York, 568 pp. Howarth, R.W., and Giblin, A. (1983). Sulfate reduction in the salt marshes at Sapelo Island, Georgia. Limnol. Oceanogr., 28 (1), 70-82. Howarth, R.W., and Teal, J.M. (1979). Sulfate reduction on a New England salt marsh. Limnol. Oceanogr., 24 (6),999-1013. Howarth, R.W., and Teal, J.M. (1980). Energy flows in a salt-marsh ecosystem: the role of reduced inorganic sulfur compounds. Am. Nat., 116 (6),862-72. Howarth, R.W., Giblin, A., Gale, J., Peterson, B.J., and Luther, G.W. III (1983). Reduced sulfur compounds in the porewaters of a New England salt marsh. Environ. Biogeochem. Ecol. Bull. (Stockholm), 35 135-52. Husar, R.B., and Husar, J.D. (1985). Regional river sulfur runoff. J. Geophys. Res., 90, 1115-25. Ivanov, M.V. (1983). Major fluxes of the global biogeochemical cycle of sulphur. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE Report 19, Wiley, New York, pp. 449-63. Ivanov, M.V., and Freney, J.R. (Eds.) (1983). The Global Biogeochemical Sulphur Cycle, SCOPE Report 19, Wiley, New York, 470 pp. Ivanov, M.V., Lein, A. Yu., and Kashparova, E.V. (1976). Intensity of production and diagenetic transformation of reduced sulphur compounds in the Pacific Ocean sediments. In: Biogeochemistry of Diagenesis of Ocean Sediments (in Russian), Nauka, Moscow, pp. 171-8. Ivanov, M.V., Grinenko, V.A., and Rabinovich, A.L. (1983). Sulphur flux from continents to the ocean. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE Report 19, Wiley, New York, pp. 331-56. Jannasch, H.W., Triiper, H.G., and Tuttle, J.H. (1974). Microbial sulfur cycle in the Black Sea. In: Degens, E.T., and Ross, D. (Eds.) The Black Sea: its geology, chemistry and biology, American Association of Petroleum Geologists, No. 20. Jeffries, M.a., and Krouse, H.R. (1984). Isotope Geochemistry of Stratified Water Bodies on Northern Ellesmere Island, Canadian Arctic, ZFI-Mitteilungen, pp. 159-69. Jeffries, M.a., Krouse, H.R., Shakur, M.A., and Harris, S.A. (1984). Isotopic geochemistry of stratified Lake 'A', Ellesmere Island, N.W.T., Canada. Can. 1. Earth Sci., 21, 1008-17. Jensen, M.L., and Nakai, N. (1961). Sources and isotopic composition of atmospheric sulphur. Science, 134, 2102--4. Jones, G.E., and Starkey, R.L. (1957). Some necessary conditions for fractionation of stable isotopes of sulfur by Desulfobibrio desulfuricans. Appl. Microbiol., 5, 111-18. Kaplan, LR., and Rittenberg, S.c. (1964). Microbiological fractionation of sulphur isotopes. J. Gen. Microbiol., 34, 195-210. Kaplan, LR., Emery, K.O., and Rittenberg, S.c. (1963). The distribution and isotopic abundance of sulfur in recent marine sediments off Southern California. Geochim. Cosmochim. Acta, 27, 297-331. Kaplan, LR., Sweeney, R., and Nissenbaum, A. (1969). Sulphur isotope studies on the Red Sea geothermal brines and sediments. In: Degens, E.T. and Ross, D.A.

Hydrosphere

259

(Eds.) Hot Brines and Recent Heavy Metal Deposits in the Red Sea, pp. 263-71, Springer, New York. Kelly, e.A., and Rudd, J.W.M. (1984). Epilimnetic sulfate reduction and its relationship to lake acidification. Biogeochem., 1, 63-77. Kester, D.R., and Pytkowicz, R.M. (1969). Sodium, magnesium and calcium sulfate ion-pair in seawater at 25e. Limno!. Oceanogr., 14, 686-92. King, G.M., and Wiebe, W.J. (1980). Regulation of sulfate concentrations and methanogenesis in salt marsh soils. Estuarine Coastal Mar. Sci., 10 (2), 215-23. Klump, L.R., and Garrels, R.M. (1986). Modeling atmospheric O2 in the global sedimentary redox cycle. Am. J. Sci., 286, 337-60. Kolle, W., Werner, P., Strebel, 0., and Bottcher, J. (1983). Denitrification by Pyrite in a reducing aquifer. Vom Wasser, 61, 125-47. Kramer, J.R. (1975). Atmospheric loading of the Upper Great Lakes. Acers Report Vol. 2, Canada Centre for Inland Waters, Burlington, Ontario. Krouse, H.R. (1974). Sulphur isotope variations in thermal and mineral waters. In: Proceedings of International Symposium on Water-Rock Interaction (Int. Assoc. Geochem. Cosmochem.), Prague, Czechoslovakia, 9-19 September 1974 (Prague: Geological Survey 1976), pp. 340-7. Krouse, H.R. (1980). Sulphur isotopes in our environment. In: Fritz, P., and Fontes, J.e. (Eds.) Handbook of Environmental Isotope Geochemistry, Elsevier, Amsterdam, Ch. 11, pp. 435-71. Krouse, H.R., and Ueda, A. (1987). Sulphur isotope analysis of trace sulphide and sulphate in various materials using Kiba reagents. In: Studies of Sulphur Isotope Variations in Nature, International Atomic Energy Agency, Vienna, 1987. STII PUBI747, ISBN 92--0-141087-5, pp. 113-21. Krouse, H.R., Cook, ED., Sasaki, A., and Smejkal, V. (1970). Microbiological isotope fractionation in springs of Western Canada. In: Ogata, K., and Hayakawa, T. (Eds.) Recent Development in Mass Spectroscopy, Proceedings of International Conference on Mass Spectrometry, Kyoto, Japan, September 1969, University Park Press, Chicago, pp. 629-39. Krouse, H.R., Legge, A.H., and Brown, H.M. (1984). Sulfur gas emissions in the boreal forest: the West Whitecourt case study, V. Stable sulfur isotopes. Water, Air, Soil Pollut., 22, 321-47. Lein, A. Yu. (1983). Biogeochemistry of anaerobic diagenesis of recent Baltic Sea sediments. Eco!' Bull., 35, 441-61. Lein, A. Yu., and Ivanov, M.V. (1983). Reduced sulphur accumulation sediments of marine basins with high rates of sulphate reduction. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, Wiley, Toronto, pp. 413-20. Lein, A. Yu., Namsaraev, RB., Trotsyuk, V. Ya., and Ivanov, M.V. (1981). Bacterial methanogenesis in Holocene sediments of the Baltic Sea. Geomicrobiol. J., 2 (4), 299-315. Lein, A. Yu., Vainstein, M.B., Namsaraev, B.B., Kashparova, E.V., Matrasov, A.G., Bondar, V.A., and Ivanov, M.V. (1982). Biochemistry of anaerobic diagenesis in recent sediments of the Baltic Sea. Geokhimiya, 3, 428-40. Lein, A. Yu., Grinenko, V.A., and Migdisov, A.A. (1983). Isotope sulphur balance in oceanic sediments. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, Wiley, New York, pp. 423-48. Lew, M. (1981). The distribution of some major and trace elements in sediments of the Atlantic Ocean (DSDP samples). I. The distribution of sulfur, sulfur

isotopes,and Mn, Fe, Zn, and Cu. Chern.Geol., 33(3-4), 205-24.

260

Stable Isotopes

Lloyd, R.M. (1968). Oxygen isotope behaviour in the sulfate-water system. J. Geophys. Res., 73, 6099-110. Longinelli, A., and Edmond, J.M. (1983). Isotope geochemistry of the Amazon basin: a reconnaissance. J. Geophys. Res., 88, 3703-17. Luther, G.W. III, Giblin, A., Howarth, R.W., and Ryans, R.A. (1982). Pyrite and oxidation in salt marsh and estuarine sediments. Geochim. Cosmochim. Acta, 46, 2665-9. McKenzie, W.F., and Truesdell, A.H. (1977). Geothermal reservoir temperatures estimated from the oxygen isotope compositions of dissolved sulfate and water from hot springs and shallow drill holes. Geothermics, 5, 51-61. MacLeod, D.A. (1973). Iron, manganese, and sulphur forms in salt marsh soils of the Wash, E. England and changes resulting from reclamation. In: Pseudogley and Gley, Transactions of Commissions 5 and 6 of the International Society of Soil Science, pp. 647-56. Matrosov, A.G., Chebotarev, N. Ye, Kudryavtseva, A.J., Zyakun, A.M., and Ivanov, M.V. (1975). Sulfur isotopic composition in freshwater lakes containing HzS. Geokhimiya 6, pp. 943-7; Geochem. Int., 12, 217-21. Mekhtiyeva, V.L. (1971). Isotope composition of sulfur of plants and animals from reservoirs of different salinity. Geokhimiya, 6, 725-30. Mekhtiyeva, V.L. (1974). Application of sulphur isotope composition of fossil shells to determine paleohydrochemical conditions in ancient water basins. Geokhimiya, 11, 1682-7. Mekhtiyeva, V.L. (1975). Isotope composition of sulphur of water soluble sulphate and living organisms from the Azov Sea. In: Chemico-oceanographic Studies of Seas and Oceans, Nauka, Moscow, pp. 130-7. Mekhtiyeva, V.L. (1980). Isotope composition of sulphur of sediments and brines in ancient and recent Kara-Bogaz-Gol. Geokhimiya, 5, 745-53. Mekhtiyeva, V.L., and Pankina, R.G. (1968). Isotope composition of sulphur of water plants and sulphates of water basins. Geokhimiya, 6, 739-42. Mekhtiyeva, V.L., and Rabinovich, A.L. (1975). Isotope sulphur balance in the Caspian Sea. Okeanologiya, XV (1), 66-73. Mekhtiyeva, V.L., Pankina, R.G., and Gavrilov, E. Ya. (1976). Distribution and isotope composition of sulfur in animal and plant tissues-hydrobionths. Geokhimiya, 9, 1419-26. Michel, G., and Nielsen, H. (1977). Schwefel-Isotopenuntersuchungen an Sulphaten ostwestfaelischer Mineralwasser. Fortschr. Geol. Rheinld. u. Westf., 26, 185-227. Migdisov, A.A., Cherkosvsky, S.L., and Grinenko, V.A. (1974). Relation of the isotope composition of humid sediments and conditions of their formation. Geokhimiya, 10, 1482-502. Migdisov, A.A., Ronov, A.B., and Grinenko, V.A. (1983). The sulphur cycle in the lithosphere. Part I. Reservoirs in the global biogeochemical sulphur cycle. In: Ivanov, M.V. and Freney, J.F. (Eds.) The Global Biogeochemical Sulphur Cycle, SCOPE Report 19, Wiley, New York, pp. 25-94. Mikulsky, Z. (1974). Water balance of the Baltic Sea. Vodnye Resursy, 5, 3-14. Miles, M. (1982). A study of the isotopic composition of the Salina Group gypsum deposits at Caledonia (Ontario). B.Sc. thesis, University of Waterloo, 1982. Ming-hui, H., Stallard, R.F., and Edmond, J.M. (1982). Major ion chemistry of some large Chinese rivers. Nature, 298, 550-3. Mitchell, M.J., Landers, D.H., and Brodowski, D.F. (1981). Sulfur constituents of sediments and their relationship to lake acidification. Water, Air, Soil Pollut., 16, 351-9.

Hydrosphere

261

Mitchell, M.J., David, M.B., and Uutala, A.J. (1984). Sulfur distribution in lake sediment profiles as an index of historical depositional patterns. Hydrobiology, 121, No.2, 121-27. Mizutani, Y., and Rafter, T.A. (1969). Oxygen isotopic composition of sulphatespart 4. N .2, J. Sci., 12, 60--8. Mizutani, Y., and Rafter, T.A. (1973). Isotopic behaviour of sulphate oxygen isotopes in the bacterial reduction of sulphate. Geochem. J., 6, 183-91. Monster, J. (1973). Lake Tanganyika: sulfur isotopic composition of a sedimentary core and dissolved sulfate of lake water. Report 91, Department of Chemistry, McMaster University, Hamilton, Ontario. Muller, G., Nielsen, H., and Ricke, W. (1966). Schwefel-Isotopen-Verhaltnisse in Formations-Wassern und Evaporiten Nord- und Siiddeutschlands (in German). Chern. Geol., 1, 211-220. Nakai, N., and Jensen, M.L. (1967). Sources of atmospheric sulfur compounds. Geochern. J.., 1, 199-210. Nakai, N., Wada, H., Kiyosu, Y., and Takimoto, M. (1975). Stable isotope studies on the origin and geological history of water and salts in the Lake Vanda area, Antarctica. Geochem. J., 9, 7-24. Nedwell, D.E., and Abram, I.W. (1978). Bacterial reduction in relation and sulfur geochemistry in two contrasting areas of saltmarsh sediments. Estuarine Coastal Mar. Sci., 6, 341-51. Nedwell, D.B., and Abram, I.W. (1979). Relative influence of temperature and electron donor and electron acceptor concentration on bacterial sulfate reduction in salt marsh sediment. Microbiol. Ecol., 5 (1), 67-72. Nedwell, D.B., and Banat, I.M. (1981). Hydrogen as an electron donor for sulfatereducing bacteria in Hurries of salt marsh sediment. Microbio/. Eco/., 7, 305-13. Nguyen, B.C., Gaudry, A., Bonsany, B., and Lambert, G. (1978). Reevaluation of the role of dimethyl sulfide in the sulfur budget. Nature, 275, 637-9. Nielsen, H. (1965). Schwefelisotope im marinen Kreislauf und das 834Sder friiheren Meere. Geo/. Rundsch., 55, 160--72. Nielsen, H. (1979). Sulfur isotopes. In: Jager, E., and Hunziker, J.C.S. (Eds.) Lectures in Isotope Geology (course), Springer-Verlag, Berlin, Heidelberg, New York, pp. 283-312. Nielsen, H., and Rambow, D. (1969). S-Isotopenuntersuchungen an Sulphaten hessischer Mineralwasser. Notizbl. hess. L. Amt Bodenforsch., 97, 352-66. Nissenbaum, A. (1978). Sulfur isotope distribution in sulfates from surface waters from the northern Jordan Valley, Israel. Environ. Sci. Tech., 12,962-5. Nissenbaum, A., and Kaplan, I.R. (1972). Chemical and isotopic evidence for the in situ origin of marine humic substances. Limno/. Oceanogr., 17, 570--81. Novicki, V. (1976). An investigation of the Kitchener aquifer system using the stable isotopes 180 and 34S. M.Sc. thesis, University of Waterloo. Nriagu, J.O. (1973). Sulfur isotope abundance in the Great Lakes waters: a preliminary report. In: Proceedings of 16th Conference on Great Lakes Research, International Association of Great Lakes Research, Ann Arbor, Michigan, pp. 1038-43. Nriagu, J.O. (1975). Sulfur isotopic variations in relation to sulfur pollution of Lake Erie. In: Isotope Ratios as Pol/utant Source and Behavior Indicators, IAEA, Vienna, pp. 77-93. Nriagu, J.O. (1984). Role of inland water sediments as sinks for anthropogenic sulfur. Sci. Total Environ., 38, 7-13. Nriagu, lO. (1986). Chemistry of River Niger. I: Major ions. Sci. Total Environ., 58, 81-8. II: Trace metals. Sci. Total Environ., 58, 89-92.

262

Stable Isotopes

Nriagu, 1.0., and Coker, R.D. (1976). Emission of sulfur from Lake Ontario sediments. Limnol. Oceanogr., 21, 485-9. Nriagu, 1.0., and Coker, R.D. (1978a). Isotopic composition of sulfur in atmospheric precipitation around Sudbury, Ontario. Nature, 274, 883-5. Nriagu, J. 0., and Coker, R.D. (1978b). Isotopic composition of sulfur in precipitations within the Great Lakes Basin. Tellus, 30, 365-75. Nriagu, J.O., and Coker, R.D. (1983). Sulfur in sediments chronicles past changes in lake acidification. Nature, 303, 692--4. Nriagu, J.O., and Harvey, H.H. (1978). Isotopic variation as an index of sulfur pollution in lakes around Sudbury, Ontario, Nature, 273, 223--4. Nriagu, J.O., and Soon, Y.K. (1985). Distribution and isotopic composition of sulfur in lake sediments of northern Ontario. Geochim. Cosmochim. Acta, 49, 823-34. Nriagu, J.O., Wong, H.K.T., and Coker, R.D. (1982). Deposition and chemistry of pollutant metals in lakes around the smelters at Sudbury, Ontario. Environ. Sci. Techno/., 16 (9), 551-60. OME (1981). Studies of Lake and Watersheds near Sudbury, Ontario, Ontario Ministry of the Environment, Toronto, Ontario. Off, W.L. (1974). Changes in sulfur content and isotopic ratios of sulfur during petroleum maturation-study of Big Horn basin Paleozoic oils. Bull. Am. Assoc. Pet. Geol., 50, 2295-318. Ostroumov, E.A., and Volkov, 1.1. (1964). Sulfates in the bottom deposits of the Black Sea. In: Proceedings of the Institute of Oceanology, USSR Academy. Ostroumov, E.A., Volkov, 1.1., and Fomina, L.S. (1961). Distribution of sulphur compounds in bottom sediments of the Black Sea. Proceedings of the Institute of Oceanology, USSR Acad. Sci., 50, 93-129. Pankina, R.G., and Mekhtiyeva, V.L. (1969). Isotopic composition of different forms of sulfur in water and sediments of Sernoye Lake (Sergiyevsk mineral waters). Geokhimiya 5, pp. 589-94; Geochem. Int., 6, 511-16. Pankina, R.G., Mekhtiyeva, V.L. Grinenko, V.A., and Churmenteyeva, M.N. (1966). Sulphur isotope composition of sulphates and sulphides in sulphidic waters of various regions-caucausis in relation to their genesis (in Russian). Geokhimiya 9, pp. 1087-94. Pearson, J.F., and Rightmire, CT. (1980). Sulphur and oxygen isotopes in aqueous sulphur compounds. In: Fritz, P., and Fontes, J.C (Eds.) Handbook on Environmental Isotope Geochemistry, Elsevier, Amsterdam, Ch. 6, pp. 227-58. Peterson, B.I., Steudler, P.A., Howart, R.W., Friedlander, A.I., Juers, D., and Bowlers, F.P. (1983). Tidal export of reduced sulfur from salt marsh ecosystem. In: Hallberg, R. (Ed.) Environmental Biogeochemistry-Proceedings of 5th International Symposium (ISEB), 1981, Ecology Bulletin (Stockholm) 35, pp. 153-65. Pringle, CM., White, D.S., Rice, CP., and Tuchman, M.L. (1981). Biological effects of chloride and sulfate with special emphasis on the Laurentian Great Lakes. Publication 20, Great Lakes Research Division, University of Michigan, Ann Arbor, Michigan. Rabinovich, A.L. (1971). Isotopic composition of sulphate ions of some ground waters and factors of its formation. Candidate thesis, Novocherkassk. Rabinovich, A.L., and Grinenko, V.A. (1979). Sulfate sulfur isotope ratios for USSR river water. Geokhimiya 3, pp. 441-54; Geochem. Int., 16, 68-79. Rabinovich, A.L., and Veselovsky, N.W. (1974). Isotopic composition of sulphate ions of some surface ground waters and factors of its formation. Candidate thesis, Novocherkassk. Rees, CE. (1970). The sulphur isotope balance of the ocean: an improved model.

Hydrosphere

263

Earth Planet. Sci. Lett., 7, 366-70. Rees, CE. (1978). Sulphur isotope measurements using S02 and SFo. Geochim. Cosmochim. Acta, 42, 383-9. Rees, CE., Jenkins, W.J., and Monster, J. (1978). The sulphur isotopic composition of ocean water sulphate. Geochim. Cosmochim. Acta, 42, 377-81. Rightmire, CT., Pearson, J.F., Back, W., Rye, R.O., and Hanshaw, B.B. (1974). Distribution of sulphur isotopes of sulfates in groundwaters from the principal artesian aquifer of Florida and the Edwards aquifer of Texas, USA. In: Isotope Techniques in Groundwater Hydrology, Vol. 2, IAEA, Vienna, pp. 191-207. Robinson, B.W. (1978). Sulphate water and H2S isotopic thermometry in the New Zealand geothermal systems. U.S. Geological Survey Open File Report 78-701, pp. 354-6. Robinson, B.W. (1987). Studies on Sulphur Isotope Variations in Nature, IAEA., Vienna, pp. 31-48. Robinson, E., and Robbins, R.C (1970). Gaseous sulfur pollutants from urban and natural sources. J. Air Pollut. Control. Assoc., 20, 233-5. Robinson, B.W., and Ruwaih, 1. (1985). The stable isotopic composition of water and sulfate from the Raudhatain and Umm Al Aish freshwater fields, Kuwait. Chemical Geology-Isotope Geoscience, 58, 129-36. Robinson, B.W., and Sheppard, D.S. (1986). A chemical and isotopic study of the Tokaanu-Waihi area, New Zealand. J. of Volcanology and Geothermal Res., 27, 135-51. Romankevich, E.A. (1977). Geochemistry of Organic Matter in the Ocean (in Russian), Nauka, Moscow, 256 pp. Rye, R.O., Back, W., Hanshaw, B.B., Rightmire, CT., and Pearson, J.F. (1981). The origin and isotopic composition of dissolved sulfide in groundwater from carbonate aquifers in Florida and Texas. Geochim. Cosmochim. Acta, 45, 1941-50. Sakai, H. (1983). Sulfur isotope exchange rate between sulfate and sulfide and its applications. Geothermics, 12, 111-17. Sakai, H., and Matsubaya, O. (1977). Stable isotope studies of Japanese geothermal systems. Geothermics, 5, 97-124. Sakai, H., Gunnlaugsson, E., Tomasson, J., and Rouse, J.E. (1980). Sulfur isotope systematics in Icelandic geothermal systems and influence of seawater circulation at Reykjanes. Geochim. Cosmochim. Acta, 44, 1223-31. Sarin, M.M., and Krishnaswami, S. (1984). Major ion chemistry of the Ganga-Brahmaputra river system, India. Nature, 312, 538-41. Sasaki, A. (1972). Variations in sulphur isotopic composition of oceanic sulphate. 24th IGC, 1972, Sec. 10, pp. 342-5. Schidlowski, M., Junge, CE., and Pietrek, H. (1977). Sulfur isotope variations in marine sulfate evaporites and the Phanerozoic oxygen budget. J. Geophys. Res., 82, 2556--{j5. SCOPE 19 (1983). Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, Wiley, Chichester, 470 pp. Scopintsev, B.A. (1975). Formation of the Recent Chemical Composition of the Black Sea Waters. Gidrometeoizdat, Leningrad, 336 pp. Shakur, M.A. (1982). 334S and 3180 variations in terrestrial sulfates. Ph.D. thesis, University of Calgary. Shink, D.R. (1979). Review of marine geochemistry. In: U.S. National Committee Report: 17th General Assembly, IUGG Review of Geophysical Space Physics, Vol. 17, No.7, pp. 1447-73. Skyring, G.W., Oshrain, R.L., and Wiebe, W.J. (1979). Sulfate reduction rates in Georgia marshland soils. Geomicrobiol. J., 1 (4), 389-400.

264

Stable Isotopes

Skyring, G.W., Chambers, L.A., and Bauld, J. (1983). Sulfate reduction in sediments colonized by cyano-bacteria, Spencer Gulf, South Australia. Australian J. of Marine and Freshwater Res., 34, 359-74. Smejkal, V., Michalicek, M., and Krouse, H.R. (1971). Sulphur isotope fractionation in some springs of the Carpathian Mountain System in Czechoslovakia. Cas. Mineral Geol., 16, 275-83. Smith, R.L., and Klug, M.J. (1981). Reduction of sulfur compounds in the sediments of a eutrophic lake basin. Appl. Environ. Microbiol., 41, 1230--7. Stround, D.W., and Paynter, M.J.B. (1980). Enumeration of methanogenic and sulfate-reducing populations in a salt-marsh. In: Abstract of Annual Meeting of American Society of Microbiology, Washington, D.C., p. 203. Taylor, B.C., Wheeler, M.C., and Nordstrom, D.K. (1984). Isotope composition of sulphate in acid mine drainage as measure of bacterial oxidation. Nature, 308, 538-41. Thode, H.G., and Dickman, M. (1983).> Sediment sulfur distribution and its relationship to lake acidification in four lakes north of Lake Superior. Unpublished Contractual Report, Canada Center for Inland Waters, Burlington, Ontario. Thode, H.G., and Dickman, M. (1984). Sulfur isotope distribution patterns and their relationship to acid rain in selected lakes north of Lake Superior. Research Report, Department of Chemistry, McMaster University, Hamilton, Ontario. Thode, H.G., and Monster, J. (1964). Distribution of sulfur isotopes in evaporites and ancient oceans. In: Chemistry of the Earth's Crust (in Russian), Vol. 2, Nauka, Moscow, pp. 589-600. Thode, H.G., and Monster, J. (1965). Sulfur isotope geochemistry of petroleum, evaporites and ancient seas. In: Fluids in Subsurface Environments-A Symposium, American Association of Petroleum Geology Memo, 4, pp. 367-78. Thode, H.G., Monster, J., and Dunford, H.B. (1958). Sulphur isotope abundances in petroleum and associated materials. Bull. Am. Assoc. Petrol. Geol., 42, 2619--41. Thode, H.G., Monster, J., and Dunford, H.B. (1961). Sulphur isotope geochemistry. Geochim. Cosmochim. Acta, 25, 159-74. Thomas, M.D., Hendricks, R.H., Collier, T.R., and Hill, G.R. (1943). The utilization of sulfate and S02 for the S nutrition of alfalfa. Plant Physiol., 18, 345-71. Truesdell, A.H., Rye, R.O., Whelan, J.F., and Thompson, J.M. (1978). Sulfate chemical and isotopic patterns in thermal waters of Yellowstone Park, Wyoming. U.S. Geological Survey Open File Report 78-701, pp. 435-6. Tsurikova, A.G., and Shulgina, E.F. (1964). Hydrochemistry of the Azov Sea, Gidrometeoizdat, Leningrad, 258 pp. Ueda, A., and Sakai, H. (1983). Simultaneous determinations of the concentration and isotope ratio of sulfate- and sulfide-sulfur and carbonate-carbon in geological samples. Geochem. J., 17, 185-296. van Everdingen, R.O., and Krouse, H.R. (1985). Isotope composition of sulphates generated by bacterial and abiological oxidation. Nature, 315, 395-6. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1982). Isotope geochemistry of dissolved, precipitated airborne and fallout sulfur species associated with springs near Paige Mountain, Norman Range, NWT. Can. J. Earth Sci., 19, 1395--407. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1985). Role of corrosion by H2SO4 fallout in cave development in a travertine deposit-evidence from sulfur and oxygen isotopes. Chem. Geol., 49, 205-11. Veiser, J., Holser, W.T., and Wilgus, C.K. (1980). Correlation of 13CI2Cand 34S/

Hydrosphere

265

32Ssecular variations. Geochim. Cosmochim. Acta, 44 579-87. Vinogradov, A.P. (1953). The Elemental Chemical Composition of Marine Organisms, New Haven Sears Foundation for Marine Research, New York, 687 pp. Vinogradov, A.P., Grinenko, V.A., and Ustinov, V.I. (1962). Isotopic composition of sulphur compounds in the Black Sea. Geokhimiya, 10, 851-73. Volkov, 1.1. (1961). Iron sulfides, their interdependence and transformations in the Black Sea bottom sediments. Tr. Inst. Okeanol., Akad. Nauk SSSR, 50, 68-92. Volkov, 1.1., and Rozanov, A.G. (1983). The sulfur cycle in Oceans-part I. Reservoirs and fluxes. In: Ivanov, M.V., and Freney, J.R. (Eds.) The Global Biogeochemical Sulphur Cycle, The Scope Report 19, Wiley, Chichester, pp. 357-423, 440-8. Vredenburg, L.D., and Cheney, E.S. (1971). Sulfur and carbon isotopic investigation of petroleum, Wind River Basin, Wyoming. Bull. Am. Assoc. Pet. Geol., 55, 1954-75. Wallick, E.I. (1981). Chemical evolution of groundwater in a drainage basin of Holocene age, East Central Alberta, Canada. J. Hydrol., 54, 245-83. Wallick, E.I., and Huemmert, M.RR. (1983). Overburden weathering studyBattle River Site. First year status summary. In: Assessment of Reclamation Potential and Hydrologic Impact of Large-scale Surface Coal Mining in Plains Areas of Alberta, ARC, Edmonton, March 1983. Weyer, K.U., and Krouse, RR. (1979). Investigations of regional geohydrology of south of Great Slave Lake, NWT, Canada, utilizing natural sulphur and hydrogen isotope variations. In: Isotope Hydrology, Vol. 1, Proceedings of International Symposium on Isotope Hydrology, Neuherberg, West Germany, 19-23 June, 1978, IAEA, Vienna, pp. 251-64. Winner, W.E., Smith, c.L., Koch, G.W., Mooney, H.A., Bewley, J.D., and Krouse, RR. (1981). Rates of emission of H2S from plants and patterns of stable sulphur isotope fractionation. Nature, 289, 672-3. Woeller, R.M. (1982). Greenbrook Well Field Management Study 1981-1982. M.Sc. thesis, University of Waterloo. Wright, R.F., and Snekvik, E. (1978). Acid precipitation: chemistry and fish populations in 700 lakes in southernmost Norway. Verh. Int. Verein. Limnol., 20, 705-75.

CHAPTER 7

Pedosphere and Biosphere


H.R. KROUSE, l.W.B. STEWART AND V.A GRINENKO

7.1

PEDOSPHERE

7.1.1 Forms of sulphur in soil In soils, most of the S occurs as organic forms. However, there are a few special exceptions to this statement. Inorganic sulphate can dominate in saline and acid sulphate soils. Generally the inorganic pool of soluble and adsorbed sulphates (in well-drained soils) and sulphides (in poorly drained anaerobic soils) account for less than 10% of the total sulphur. The sulphur cycle in soil has been extensively reviewed by Freney and Williams (1983) and will only be described briefly to provide a framework for discussion of the isotope data. 7.1.1.1 Inorganic sulphur In most well-drained soils, the dominant form of inorganic sulphur is sulphate with a negligible percentage of compounds of lower oxidation state present (d. reviews by Biederbeck, 1978; Bettany and Stewart, 1983; Freney and Williams, 1983). Sulphate occurs as water-soluble salts (mainly calcium, magnesium, and sodium sulphates) and as 'adsorbed sulphate' where the sulphate ion is bonded to the soil inorganic components. Adsorbed sulphate is an important fraction in leached tropical soils containing hydrous oxides of aluminium and iron as it prevents the rapid loss of sulphate out of the plant rooting zone. Under anaerobic conditions such as occurs in flooded rice soils, sulphides are the principal stable form of inorganic sulphur. It is generally believed that the water-soluble plus the majority of adsorbed sulphate is the 'available' fraction for plant uptake. The amounts of water-soluble sulphate vary widely, both within the soil profile itself and among soils developed from different parent materials and in different climatic regions. However, within a given area, it is the factors that relate to man's activities (e.g. crop management, irrigation, fertilization, and industrial processing) that appear to have the greatest effects on sulphate levels.
Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko (Q 1991 SCOPE, Published by John Wiley & Sons Lid

268 7.1.1.2 Organic sutphur

Stable Isotopes

Current analytical techniques permit only a broad fractionation of organic sulphur compounds in soils. Two major groups have been identified: one in which sulphur is directly bonded to carbon (C-bonded S) and another in which sulphur is linked to carbon through an oxygen or nitrogen atom (organic sulphates). The former group would include free or combined sulphur-containing amino acids and sulphonates, and the latter the sulphate esters, thioglucosides, and sulphamates. Problems with the analytical determination of these fractions have been discussed elsewhere (Freney et at., 1967, 1972). Most workers determine organic sulphate by reduction with hydriodic acid (HI-reducible S) and calculate C-bonded S as the difference between total S and HI-reducible S. Alternately, a substantial proportion of C-bonded S can be determined directly by reduction to HzS by Raney-nickel-HCI (Lowe, 1965; Freney et at., 1975). It is obvious that these broad fractionation techniques are subject to error and lack 'biological significance'. However, they take on meaning when considered in terms of the soil-forming factors and available sulphur (Bettany and Stewart, 1983). Generally, the HI-reducible S fraction is considered to be more biologically active or 'labile' (Fitzgerald, 1976), and accounts for 27-78% of the total sulphur in surface soils, with mean values of approximately 50%. The percent HI-reducible S shows considerable variation across environmental gradients and cultivation chronosequences, generally increases with depth in the soil profile, and is concentrated in the clay and fulvic acid fractions of soils (Bettany and Stewart, 1983). The C-bonded S fraction contains both the labile sulphur amino acids, protein sulphur and the more resistant sulphones, sulphonic acids, and heterocyclic compounds. It is likely that a portion of this fraction is highly labile, since tracer experiments show that recently immobilized sulphate is remineralized from both HI-reducible and C-bonded fractions (Freney et at., 1971; Saggar et at., 1981b; Maynard et at., 1984). The soil organic fraction comprises a complex heterogeneous mixture of soil organisms and residues of partially decomposed plant, animal, and microbial cells. This is further complicated by the intimate association of the organic fraction with soil minerals. It is therefore not surprising that determination of the two sulphur fractions (C-bonded and organic sulphates) on whole soils gives little information on the dynamics of sulphur turnover in organic matter. Attempts have been made to subdivide soil organic matter into biologically meaningful fractions prior to sulphur analyses. The various approaches that have been taken include: (a) Isolation, identification, and determination belonging to definite chemical groups of individual compounds

Pedosphere and Biosphere

269

(b) Chemical extraction followed by physical chemical separations of various fractions, such as humic and fulvic acids and humin (c) Physical separation of soils into various organomineral sized fractions (d) Biological techniques aided with radio- and stable isotopes, with the object of isolating and identifying fractions that are significant in terms of nutrient cycling Artefact formation during chemical extraction and purification of organic matter fractions is inevitable but can be minimized. For example, the physical separation of whole soils into organomineral complexes of various particle sizes by densimetric separation and/or centrifugation yields fractions that have minimal chemical alteration. This enables the study of the role of inorganic constituents in the stabilization and transformation of humus (Turchenek and Oades, 1979; Anderson et at., 1981). Distinct differences in the forms and distribution of organic nitrogen and sulphur among particle size fractions have been observed, suggesting that sulphur follows different pathways in humus formation and transformation than carbon and nitrogen. Consideration of the biologically significant fractions of soil organic matter inevitably involves the role of the soil biomass. Complete understanding of the sulphur cycle requires an understanding of the forms of sulphur in soil microorganisms and the mechanisms involved in microbial transformations. Methodology for the determination of biomass sulphur in soils has just become available and current assays are directed towards the forms of sulphur in microbes and their role in sulphur cycling. Although only 2-3% of the total sulphur in soils is contained in the microbial biomass, it assumes greater significance because of its rapid turnover (mainly bacteria and fungi) (Saggar et at., 1981a). In bacteria, the major portion of sulphur appears to be in the form of amino acids, with about 10% as HI-reducible S. The sparse literature suggests that sulphur compounds other than amino acids and related compounds are rare, although small amounts of organic sulphates (choline sulphate) have been found in bacterial cell walls and a bacterial sulpholipid has also been identified (Fitzgerald and Lushinski, 1977). It has also been shown that some species of chemoautotrophic and photosynthetic purple sulphur bacteria are capable of accumulating elemental sulphur. Like bacteria, the major portion of sulphur in fungi is in the amino acid form. However, fungi have a larger portion of organic sulphates, such as choline sulphate. It appears that fungi are able to store sulphur as choline sulphate when sulphur is in adequate supply. Recent experiments show that the relative percentage of HI-reducible S increases significantly with increase in the concentration of sulphate in the growing medium. There is also evidence that significant amounts of various aryl sulphates may be present in fungi, and elemental sulphur is contained in some species (cf. the review by Bettany and Stewart, 1983).

ATMOSPHERIC H2S

tV --.] 0

STABLE I N 0 R G ~ I C S
ADSORBED + OCCLUDED S04 PRECIPITATED S2-

STABLE
/ / \ \ \ / / / / /

LABILE
INORGANIC S RESINEXTRACT

LEACHING TO GROUND WATER

LABILE ORGANIC S
(C-Q-S) (C-S) -10yr

CLAY PROTECTED ORGANIC S (C-Q-S) (C- S) -50 yr

0 R G A N I C S
(C-Q-S) (C-Q)
VJ

-1500yr

Figure 7.1 Conceptual flow diagram of the main forms and transformations of sulphur in the soil-plant system

.g '" '"

E; Q~ t;;< a

Pedosphere and Biosphere

271

The origin of the large organic sulphate pool in soils is still open to question. It has been suggested that in addition to fungal synthesis, a considerable portion of ester sulphate is introduced into soils through animal excretions, decaying animal matter, extracellular synthesis by certain bacteria, and synthesis and subsequent decay of plant tissue, algae, and lichens (Fitzgerald, 1976, 1978). 7.1.1.3 Transformations of sulphur in soils Soil organic sulphur (both C-O-S and C-S) can be subdivided into three compartments representing labile organic S consisting of compounds that are easily mineralized and have a short half-life (~1O years), similar compounds with a longer lifetime (-50 years) due to chemical and physical protection, and relatively stable compounds in the soil matrix with a halflife of hundreds of years (Figure 7.1). The transformation of sulphur in soils is considered to be mainly microbial (Freney and Williams, 1983), although'some strictly chemical processes (e.g. the oxidation of iron sulphide) are also possible (Bloomfield and Coulter, 1973). The microbial processes may be grouped into four distinct processes: (a) Mineralization: a process in which large organic molecules containing sulphur are broken down to smaller units and eventually to inorganic sulphate. (b) Immobilization: simple inorganic S molecules (mainly sulphate) are converted to organic compounds. (c) Oxidation: inorganic sulphur of lower oxidation states (elemental sulphur, thiosulphates, polythionates, sulphides, etc.) are generally converted to sulphate as a stable end-product. (d) Reduction: sulphate and other partially reduced sulphur-containing anions are converted to sulphide. Although there are organisms capable of performing all these processes in all soils, mineralization and immobilization are the dominant transformations that occur in well-aerated arable soils of the world. 7.1.2 Sources of sulphur compounds in soil The pedosphere is the site of many complex interactions between the lithosphere, biosphere, atmosphere, and hydrosphere. In terms of the sulphur cycle, the major inputs are dead organic matter, dry and wet deposition of atmospheric sulphur compounds, groundwater leaching of subsurface minerals, and, in cultivated regions, fertilizer applications (Freney and Williams, 1983; Bettany and Stewart, 1983). The main outputs are river runoff and biogenic emissions of gaseous sulphur compounds.

272

Stable Isotopes

The sources of sulphur compounds in the soil embrace not only inputs from the atmosphere, biosphere, and lithosphere via the hydrosphere, but also transformations within the soil. Solution SOi- constitutes the central pool in Figure 7.1. It can 'flow' to the microbial S pool, plants, and various inorganic forms. Most foliar sulphur returns to the soil either directly (e.g. falling leaves) or indirectly (e.g. death of insects, excrement). Volatile sulphide compounds are lost to the atmosphere but they can in turn be oxidized to S02' The S02 can return to the soil intact (dry fall), be oxidized to SOi- which falls as particulates or precipitation, or be incorporated into plants through the stomata and returned to the soil upon plant decay. Anthropogenic sulphur can enter soil by many of the pathways shown in Figure 7.1. When a given flux is augmented by the pollutant S, other fluxes attempt to adjust accordingly. For example, SOi- added to soil is rapidly converted to organic sulphur. Consequently, concentration determinations of soil SOi- cannot provide an accurate assessment of uptake of pollutant sulphate from the atmosphere. However, if the isotopic composition of the indigenous sulphate differs from that of the added pollutant sulphate, the 334S value of the mixture is a direct measurement of the relative amounts of each component, even if the SOi- pool is depleted by conversion to organic S. The accuracy of the isotopic approach depends upon whether there is isotopic fractionation during the SOiconversions. During assimilation, the isotopic selectivity is usually small (Sections 7.1.4 and 7.2.3), whereas large kinetic isotope effects accompany dissimilatory (respiratory) sulphate reduction (Section 1.3.1.2). Fortunately most soil situations of interest are aerobic and isotope fractionation during sulphate reduction should not present problems. 7.1.3 Sulphur and oxygen isotope variations in soil Soils range widely in their 334S values because of isotopic variations in parent materials and sulphur gains from, and losses to, the atmosphere, hydrosphere, and biosphere. A world survey of soil 334Svalues has not been conducted. In most cases, data were obtained in areas where concerned environmental agencies wished to evaluate the effects of existing industrial pollution. However, there are sufficient data (Figure 7.2) to draw some conclusions regarding sources of sulphur compounds and their transformations. The large range of 334S values in Figure 7.2 (from below -30 to above + 30%0)can be related to a variety of natural and anthropogenic sources of sulphur. For instance, soils from various locations in New Zealand have 334S values approaching those of seawater S042- (Kusakabe et ai., 1976). Thus the authors concluded that sea spray as the principal source of soil sulphur even in inland basins with little rainfall. Whereas New Zealand soils

Pedosphere and Biosphere

273

CALIFORNIA

ITALY

USSR

10 SAMPLES

T 1

~
-10 0 +10 +20 +30 +40 834S (%0)

-30

-20

Figure 7.2 Natural sulphur isotope abundance variations in soils (see text for sources of data)

appear to derive marine sulphate from modern sea spray, other soils such as the few studied in Tunisia by Kusakabe et ai. (1976) derive sulphate from gypsum occurrences related to the ancient ocean. The data for the Soviet Union in Figure 7.2 are those casually mentioned in the thorough study of plants in various regions by Chukhrov et ai. (1978). There is a considerable range in sulphur isotope composition with many data near + 3%0. These authors attributed the wide range in 334Svalues to essentially three processes which varied with locality. These are (a) local weathering of rocks containing evaporites and sulphide ores, representing extreme situations, (b) inputs of sulphate from the atmosphere, and (c) bacterial SOl- reduction

274

Stable Isotopes

in less aerobic environments leading to differentiation of 0345 values with soil depth. As this study was primarily concerned with 334S values in vegetation, specific data will be discussed in Section 7.2.2. As in the Soviet Union, data from many other locations in the world fall near 0%0, which is considered to approximate the terrestrial mean (Figure 7.2). In some cases, these signify natural sources but in other situations (e.g. Wawa, Ontario, and Thompson, Manitoba) smelting of sulphide ores of deep crustal origin is involved. Soils very depleted in 34S (- 30%0)are often encountered (d. California, USA, and Alberta, Canada; Figure 7.2). Surface or shallow occurrences of Cretaceous shales, found in Alberta, other parts of western Canada, and Montana, USA, have pyrite which is very depleted in 34S. Therefore the soil S is believed to be derived from weathering of sulphide minerals and perhaps organic sulphur in the shales. This process may have been ongoing since the last continental glacial retreat some 10 000 years ago. Organic soils developed under anaerobic and therefore reducing conditions may exhibit depletions in 34S. An example is a peat deposit in Northern Alberta (Krouse, 1980). At 1.2 m depth, pyrite was the dominant sulphur species with a 334S value near -24%0. Upon approaching the surface

and more aerobic conditions, elemental sulphur (334S= - 28 to - 30%0) dominated. Sulphate salts on the surface had a 834Svalue near - 33%0
whereas coexistent So had a 334S value near -30%0. This sulphur isotope pattern is diagnostic of sequential kinetic isotope effects whereby 32Scontaining sulphide and intermediates were oxidized preferentially. Quite positive 834S values are found in soils near travertine (CaC03, minor CaS04 . 2H2O) depositing springs reflecting evaporite strata which were dissolved at depth by these waters. In many locations in Alberta, such positive 334S values in soil may also be the consequence of emissions from sour gas processing (see below and Section 8.2). Thus the wide variations in 834S values in soil embrace both natural and anthropogenic sources. Sulphur isotope data in combination with concentration measurements can be used effectively to delineate those sources. In the Teepee Creek area of Alberta, more negative 834S values of soil specimens were found for higher S contents (Krouse and Case, 1981). The most negative values approached those found for shallow sulphate mineral deposits, believed to have been derived from weathering of Cretaceous shales. Therefore elevated sulphur levels in the soil were attributed to a subsurface source rather than nearby minor industrial emiSSiOns. A subsurface sulphate mineral source is also indicated in Figure 7.3, where the 334Svalue has been plotted against the inverse of concentration. Again, high total S concentrations are identified with quite negative 834S values even though the two locations, Teepee Creek and Twin Butte, are over

Pedosphere and Biosphere

275

+5
0

~
(f)

SOIL B HORIZON

<!

-5

~
u) -10
<::t r<1 00

-15

-200

0005 0.010 0.015 CONCENTRATION-1 (ppm-1)

0.020

Figure 7.3 Plot of 034S values versus the inverse of total sulphur concentration in the B horizon of soils near Twin Butte, Alberta. A dominant sulphur source with 034S near - 17%0 is indicated along with two minor sources (from Krouse and Tabatabai, 1986)

800 km apart. The two trends in Figure 7.3 suggest that three sources are present with the dominant one corresponding to the point of intersection of the trends and a 1)34S value of -17%0, again indicative of derivation from Cretaceous shales. The presence of pollutant sulphur in soil enriched in 34S has been frequently reported in the vicinity of sour gas processing plants in Alberta (Section 8.2). Data from one of the latter locations are summarized in Figures 7.4 and 7.5. In Figure 7.4, the 1)34S values for soluble S in individual soil horizons are plotted as a function of distance from a gas flare stack which had been operative for 28 years. Within 9 km of the flare, there is evidence that sulphur of industrial origin has penetrated to the C horizon. Beyond 9 km, the 1)34S values of the lower horizons are reasonably constant, suggesting that the preindustrial 1)34S value for soil was about -12%0. The trend of decreasing 1)34S values with distance is maintained beyond 17 km in the upper horizons, attesting to industrial fallout beyond that distance. Using Figure 7.4, statements can be made regarding the relative contributions of industrial and natural sulphur to the SOi- pool. For example, in the Ap horizon at 17.5 km, 25% of industrial sulphur with a 1)34Svalue of + 20%0combined with 75% of natural sulphur with a 1)34S value of -12%0 yields the observed 1)34S value of -4%0. In Figure 7.5, data for total sulphur in only one soil horizon are plotted for the Valleyview study. The 834Svalues of total S display a similar distance

276

Stable Isotopes

+20

.
.
.

SOIL
SOLUBLE S

C/) v r<>
/JO

cf2. I.. ..
0

<3 +10f-

.
0

.
..
. Q

HORIZONS Q Ap LH Ae 0 Bt A BC D. Ck

-10
0

-20 0

5 10 15 DISTANCE TO STACK (km)

20

Figure 7.4 Presence of sulphur of industrial origin in soil, documented by sulphur isotope data for soluble sulphur, Valleyview, Alberta (from Krouse and Case, 1982)

+201-

TOT A L S LFH 0 0 0

o:t r<)

C/)
/JO +t

0 0 CO 0

-100

5 10 DISTANCE TO STACK

15 (km)

Figure 7.5 Presence of sulphur of industrial origin documented by sulphur isotope data for total sulphur in upper LH horizon, Valleyview, Alberta (from Krouse and Case, 1982)

Pedosphere and Biosphere

277

dependence, although they are 2 to 9%0depleted in 34S compared to the soluble S (see Section 7.1.4). Although HzS emission from soil and its subsequent oxidation have not been studied isotopically, sulphur and oxygen isotope data have been obtained for analogous biogenic HzS emission from thermal springs (van Everdingen et al., 1982; Krouse and van Everdingen, 1984). These emissions after oxidation contributed secondary sulphate to the soil which was depleted in 34S and 180 by as much as 30 and 20%0respectively, in comparison to primary SO~- in the spring waters. The depletion in 34S in the secondary sulphate minerals relates to the kinetic isotope effects during bacterial HzS generation (Section 1.3.1.2), whereas the 180 depletion arises from incorporation of oxygen atoms from the water molecule during oxidation (Section 2.2.1). Similar isotopic data were reported for SOi- in groundwater (Shakur, 1982; Hendry et al., 1989) and soil (Chae and Krouse, 1987) at depths near the water table in southern Alberta. Positive 334S and 3180 values were found in the deepest samples, whereas shallower samples were characterized by 334S and 3180 values as low as -12 and -18%0 respectively. This distribution can arise through changing levels of the water table coupled with SO i- reduction and mineralization of the product sulphide in the anaerobic zone. If the water level dropped, unreacted SOi- enriched in 34S would remain in solution. The mineralized sulphide depleted in 34S would become exposed to a more aerobic environment above the water table. It would tend to be reoxidized, incorporating some oxygen from the l80-depleted water of meteoric origin (either vapour or penetrating rainfall). Meanwhile, SOi- below the water table becomes more enriched in 34Sand 180 as reduction proceeds. Therefore the sulphate tends to be depleted in heavy isotopes in shallower environments with progressive enrichment with depth. This depth trend will be moderated as the soil acquires SOi- from precipitation and/or dissolution of primary sulphate minerals. The combination of sulphur and oxygen isotope abundances constitutes a unique tool for understanding soil SO i- transformations and for differentiating primary and secondary sulphate. A precautionary note is appropriate. Dissolved sulphate in standing water in wells or boreholes may not be representative of that in soil. For example, microbial contamination during drilling might alter concentrations and isotope compositions of sulphur compounds over time. If possible, standing water should be pumped out and water entering from the soil horizon sampled immediately. 7.1.4 Relationships among 0348 values of different soil extracts The 834$ variations among different S compounds, compound classes, or extracts can be attributed to isotope selectivity in biochemical transformations

278

Stable Isotopes

or different sulphur sources. The extracts usually obtained in soil analyses are mixtures of compounds that cannot be readily related to specific biological processes (Section 7.1.1.2). Although the soluble fraction is mainly SOi~, some soluble organic S compounds may be included. Since the insoluble S fraction is the remnant of extraction by water or salt solutions, it will contain varying amounts of sulphate inversely related to the soluble S recovery efficiency. For many locations in Alberta, Canada, the 034S values for total soil S are systematically lower than those for soluble S (Figure 7.6). It is unlikely that this resulted from kinetic isotope effects during the conversion of SOi- to water-insoluble organic S. An alternate explanation is based on mixing phenomena. At these locations, most of the soluble S, particularly in the upper LFH soil horizon (litter), came from the atmosphere which contained 34S-enriched sulphur compounds of industrial origin. Deciduous leaves and conifer needles contributed significantly to the litter. Some of this insoluble

SUB
SURFACE SURFACE
...

SOIL

.
+20 0
en

t::,.

NEWZEALAND
ALBERTA Valleyview Other locations

+10r

-'

0 en en <t !<)
(,()

CD :) -'

-10

-20

-30 -30

-20

-10

+10

+20

834S, TOTAL S(%o)


Figure 7.6 The 834Svalues for soluble S versus the same for total S in soil samples of Alberta and New Zealand (New Zealand data from Kusakabe et al., 1976)

Pedosphere and Biosphere

279

organic S was derived from atmospheric S02, but the bulk was acquired from soil water SO/- more depleted in 34S(see Section 7.2.5). Hence, the isotopic difference between the two soil components is readily explainable by two sources coupled with the lack of isotopic homogenization. This in turn implies that the different sulphur forms had not interconverted extensively since entering the soil. In some deeper soil horizons in Alberta the isotopic distribution between soluble and insoluble S is opposite to that described above; i.e. the data fall below the line of unity slope in Figure 7.6. This can be explained by an argument similar to the above except that the dominant source of soluble S is a subsurface mineral depleted in 34S (Section 7.1.3). At depth, organic S is much less likely to be oxidized to sulphate under the anaerobic conditions. In the New Zealand soils studied by Kusakabe et at. (1976), the data fall close to the line of slope unity, passing through the origin in Figure 7.6. These authors conclude that there is negligible isotopic selectivity during SOi- to organic S transformations and that the soil components which they studied were homogeneous in S isotope composition. The homogenization need not take place in the soil per se since sulphate assimilation by plants is accompanied by minimal isotopic selectivity, i.e. the organic S in litter has essentially the same isotopic composition as SOi- available to the plant roots if the influence of atmospheric sulphur is minimal. In summary, if all soil sulphur is derived over a long time from one isotopically uniform source, all S fractions acquire similar isotopic compositions; i.e. kinetic isotope effects involved in the transformations are small. A possible exception is dissimilatory SOi- reduction in water-logged soils. If there are a number of isotopically different sources or, alternately, temporal variations in the isotopic composition of one source, there may not be sufficient time for significant interconversions either within the soil or through soil-plant cycling. Then, isotopic variatons will persist between extracts or spatially with the same extract. Whereas the microbial or botanical conversion of SO i- to organic S can be rapid, oxidative turnover of the much larger accumulated organic S pool in soil requires more time. Therefore, the latter process would appear to be rate controlling in establishing isotopic uniformity among sulphur pools. Anthropogenic So may enter the soil as applied fertilizer or dustfall from loading operations. The sulphur isotope selectivity was found to be very small during chemical and bacterial oxidation of Soin potted soil (McCready and Krouse, 1982). This implies that the kinetic isotope effect in the ratecontrolling step is close to unity. This is consistent with the overall fractionation being determined by reaction on the sulphur surface. Dependent upon the concentration of background SO i- and the difference in the 334S value of the So and background S042-, S isotopes can be used to determine

280

Stable Isotopes

the oxidation rate. Further, during oxidation, oxygen atoms are incorporated from both HzO and atmospheric Oz. Therefore, both the sulphur and oxygen isotope compositions can be used to delineate natural sulphate from that generated by the oxidation of applied So. 7.1.5 Transfer of sulphur into the soil as revealed by sulphur isotope data There are a number of pathways by which sulphur may enter the soil (Figure 7.1). Of these, the dominant mechanism is leaching of SOi-. In acid soils, downward movement will be impeded by sorption of SOi- by Al and Fe compounds. Other factors of importance include movement of organic compounds and biological transformations. In forest ecosystems, flow of sulphur compounds from the atmosphere to the subsoil may be blocked by moss and litter (see Figure 7.13). The trapping mechanism might be physical (low permeability) or S04Zassimilation might occur. Groundwater behaviour is relevant. Downward transport of sulphur compounds should be enhanced in recharge areas and suppressed in discharge areas. Penetration of atmospheric pollutants into the soil does not depend simply upon accumulated fallout. The rates and duration of the emissions are very relevant. Episodes at high SOz concentrations for short periods can chemically destroy the integrity of the surface and radically upset bacterial balances. In contrast, the same amount of SOz deposited at a constant rate might be retained in litter and/or botanical cover. A controversial concept is the possible migration of sulphur compounds to the B or even C soil horizons by downflow within plants to the roots. The evidence, however, suggests that this flux is not substantial (Section 7.2.5). The downward movement of sulphate in soil is influenced by texture. It has been found that sand has a much higher filtration rate than glaciolacustrine clay (Figure 7.7). At Valleyview, Alberta, industrial flaring over many years generated SOz with 634S values around +20%0. Figure 7.7a shows the concentrations of soluble sulphur (mainly S04Z-) in the Ae horizon as a function of distance from the stack. Site 1 is a thin sand veneer overtill, whereas site 15 is a thin sand veneer over a glaciolacustrine clay veneer overtill. Clay cover exists at all other sites. The soluble sulphur contents at sites 1 and 15 are lower than the trend since the sandy veneer permitted penetration to lower horizons. This interpretation is further supported by the sulphur isotope data (Figure 7.7b and c). In all three plots, movement of points 1 and 15 to higher concentrations align them with the trends found for the clay veneer soils. Figures 7.4 and 7.7b and c collectively demonstrate how 634S data can very effectively assist in interpreting the movement of sulphur in the soil

Pedosphere and Biosphere E 30 Q. Q. (/) 20 Ae HORIZON

281

~
0
.(/)

10
/

/"

::::>

...J

00
/'

/ 1 SAND {15 /"

t~" -i1,

CLAY
0

0 0

5 ,

10

15

(a)
(/) +10

DISTANCE FROM STACK (km) Ae HORIZON


\

...J

,_S~~~~/ m+ 5 0 CLAY

(15.

3 0
(f)

DO

en 0 f-!=!- 0 10 20 v r<> SOLUBLE S (ppm) ro (b) LFH HORIZON " (f) +25 / \ .. /15. w / ...J+20 0 m t I SAND 1 3+15 .. 1. I I 0 0 \ ./ (/)+10

--

,_

en ~ + 5 ro
0 0
(c)

0 0 CLAY I 30

DO I 10

I 20

I 40

SOLUBLE

S (ppm)

Figure 7.7 Concentration and sulphur isotope data relevant to retention of soluble S in sand and clay, Valleyview, Alberta (from Krouse and Case, 1982). Arrows indicate where data would plot if soluble S had not readily moved downward through upper sandy horizons. (a) Concentration of soluble S in the Ae horizon versus distance from the flare stack. (b) The 334Svalue versus concentration of soluble S in Ae horizon. (c) The 334Svalue versus concentration of soluble S in LFH horizon. The higher 334Svalues in the uppermost LFH horizon attests to a larger proportion of SOl- of industrial origin than in the underlying Ae horizon

profile. It must also be stressed that concentrations determined by expedient sampling of soil to a preselected depth, as advocated by some environmental agencies (typically 5-10 em), may not be reliable indicators of the pollutant S content. Sulphur in each soil horizon in a longer core should be examined and the depth profile fully characterized.

282

Stable Isotopes

Whereas total soil S is most readily analysed, isotope data from specific compounds or compound classes may prove more informative. For SOz emissions and So dustfall, sulphate is the logical choice for two reasons. The added sulphur is converted to SO/- and the natural soil sulphate content is usually low (except where gypsum is found in the parent material) compared to that of organic S. In contrast, Chae and Krouse (1987) found that the presence of sulphur in lower soil horizons from manure application could not be discerned isotopically with sulphate because of the high natural background. It was demonstrated that analysis of C-bonded S was more appropriate.

7.2 VEGETATION
7.2.1 Forms of sulphur in vegetation Plants contain on average about 0.25% S (by dry weight). A wide variety of S-containing compounds have been found in plants, but only a few are known to be required for normal cell function. The essential compounds include the S amino acids, glutathione, thiamine, vitamin B, biotin, ferredoxin, lipoic acid, coenzyme A, and the sulpholipid of the chloroplasts. Compounds of non-essential or unknown functions in plants include glucosinolates, which are found in the cruciferae, choline sulphate, and penicillin. A variable portion of the S in plants is in the form of inorganic SO/-. Generally, the S amino acids contain approximately 90% of the S found in plants (Blair, 1979). However, in some species, a considerable portion of the total S may be contained in secondary plant products (Anderson, 1975). The total CIS ratios of plant material ranges from 150:1 in rape seed grain to 450:1 in straw. The largest portion of S in plants is in protein either as cysteine, cystine, or methionine. Plant proteins generally contain 1 % Sand 17% N (Dijkshoorn et al., 1960; Pumphrey and Moore, 1965). For any plant species, the composition of a given protein, which is controlled by genetics, is constant. Therefore, environmental factors such as the Nand S supply, age of plant, etc., should have no influence on the N/S ratio in plant proteins. However, the total N/total S ratio can be greatly affected by environmental factors. When S is adequate, non-protein S (mainly SO/-) will accumulate in the plant and the total N/total S will be less than the N/S ratio in protein. When S is deficient, protein formation is suppressed and non-protein N accumulates. The resulting bulk plant N/S ratio is greater than the N/S ratio of the proteins. In most plants, inorganic SO/- is the main component of the HI-reducible S fraction. Plants absorb S as SO/-, which is reduced and incorporated

Pedosphere and Biosphere

283

into amino acids as a precursor for protein synthesis. Any excess S04Zthat is not metabolized accumulates unchanged in the tissue. When S is in short supply, most of this SOi- is incorporated into protein and little remains free in the tissue (Janzen and Bettany, 1984). Studies on the decomposition of plant material indicate that net mineralization of S is mainly related to the content of S in the material added to the soil (Freney, 1967). No net mineralization of S occurred from plant material if it contained less than 0.13% S. 7.2.2 Sulphur isotope distributiou iu vegetation Variations in 834Svalues for vegetation samples from different parts of the world are summarized in Figure 7.8. The bulk of the data, which include many species and parts of individual plants, displays a range in 834Svalues from -20 to + 30%0, comparable to that found for sedimentary rocks (Chapter 4). There are isolated extremes outside this range (Krouse and Tabatabai, 1986). For example, plants growing near a biogenic HzS emitter had 834S values as low as -33%0 (Krouse and van Everdingen, 1984). Whereas parts of some trees were found to vary by less than 1%0,in other cases variations of over 15%0were encountered (Section 7.2.5). In some studies, vegetation had uniform 834S values over a large area whereas in the vicinity of natural or industrial emitters variations could be significant over distances of a few metres. The average and most frequently encountered 834S value for vegetation over extensive areas of the Soviet Union was found to be near + 3%0 (Chukhrov et ai., 1975, 1978; Figure 7.8). These authors discussed a number of factors that influence the sulphur isotope composition of vegetation. These include lithospheric sources, groundwater leaching and transport, bacterial redox reactions in soil solutions, and the relative contribution of SOi- in precipitation. In the Lena River in the Republic of Central Yakutia, SOi- with 834S values as high as +26%0 was identified with evaporite dissolution. The 834S values for plants in the vicinity were as high as + 17%0,even at considerable distances from the river. Plants in the flood plain of the Lena River and nearby marshy lowlands possessed higher 834Svalues. Whereas ground water played a role in transporting 34S-enriched sulphur to the vegetation, the authors considered local precipitation to be a major factor. They believed that the 834S values ranging from + 5 to + 22%0in rain and snow reflected atmospheric uptake from this vast water body. At some locations, metal sulphide deposits influenced the isotopic composition of plants whereas in other cases there were no correlations. For example, at Zhosaly Sopka, pyrite deposits apparently contributed negative 034S values to vegetation whereas, at Alaygyur, there was no

284

NEWZEALAND TUNISIA ITALY

CJ [s c:J

.-'l{1n,

10 SAMPLES .

USSR

100

(f) W -.J 0..

(f)
LL

0 a::: w 50 aJ 2 ::J Z

0 -20

-10

0 8345

+10 (%0)

+20

+30

Figure 7.8 The ()34Svalues for vegetation from various locations in the world. (see text for sources of data)

Pedosphere and Biosphere

285

evidence of such a contribution. This inconsistency was attributed to groundwater circulation and the effects of elevation. Sites below ore deposits could be influenced by leachate carried by ground water whereas sites above the ore occurrences would not be affected. Negative 1)34S values in vegetation near copper sulphide occurrences at Kounrad peaks were in part attributed to ore dust which entered the atmosphere during pit-blasting operations. In some locations, trees growing in igneous rock fissures had foliar 1)34S values close to the atmospheric average and the seasonal variations were quite small (+1 to +4%0) as compared to the average value of +2.1%0 found in snow. Precipitation SO~- should be rapidly lost through crevices where extensive humic matter had not built up to provide a soil sulphur reservoir. In contrast, large temporal 1)34Svariations were found in locations where plants derived most of their sulphur from soil solutions, particularly under anaerobic conditions. As the consequence of bacterial SOi- reduction, the 1)34Svalues of residual SO/- in stagnant water in a marsh location was found to increase from 0 to 5%0in the spring to as high as +29.6%0 in late summer. Birch and willow trees in the same area some distance from the marsh had 1)34Svalues near 0%0whereas grass growing near the stagnant water acquired 1)34S values as high as + 15.4%0. The data from Alberta contrast with those of the Soviet Union in that the former display two peaks. The highly negative values are found near Cretaceous shale exposures, particularly in the Peace River area (Krouse and Case, 1981). It is believed that these plants assimilated 34S-depleted SOigenerated by weathering of reduced S species in the shales. Occasionally, a highly positive 1)34S value relates to groundwater dissolution of evaporites. However, in most cases, the vegetation is enriched in 34S because foliar S has been derived from atmospheric SOz emitted during sour gas processing (Section 8.2). In contrast to the above, limited data from other locations emphasize single sources of soil sulphur (Figure 7.9). Plants in New Zealand are enriched in 34S in coastal areas because of marine S04Z- in precipitation (Kusakabe et al., 1976). The positive 1)34S values for the few specimens from Tunisia relate to gypsum occurrences in the soils (Kusakabe et al., 1976).

7.2.3 Assimilation of sulphate by plants and algae It is important to know the isotopic fractionation during SOi- assimilation to assess uptake of pollutant sulphur by plants. This can be best measured in the laboratory to avoid problems with multiple sources of sulphur in the field. In the case of algae, the isotopic composition of total sulphur as well as insoluble organic sulphur is typically depleted in 34Sby 1-2%0with respect to SOi- in the medium (Section 6.2.2; Section 1.3.1.1).

286

Stable Isotopes

+25

~ 0

0:: :J :I: a.. -I +20 :J


C/)

+
NW.T. CANADA

IZ c::{ -I a.. +15


C/) v

.
NEW ZEALAND

+10 +10

+15

+20

+25

8345, SOIL 542- (%0)


Figure 7.9 The 834S values for plant sulphur versus those for soil sulphate in situations where uptake from the atmosphere is minor compared to derivation of foliar S from the soil (data sources: New Zealand and Tunisia, Kusakabe et al., 1976; Northwest Territories, Canada, Krouse et al., 1984)

With higher plants, either negligible fractionation or an average depletion of 34Sby 1 to 2%0in the organic sulphur has been realized (Section 1.3.1.1; Section 6.2.2). In studies with very high SO/or HS03 - concentrations, the foliar S may become enriched in 34S due to release of reduced S compounds which are depleted in 34S (Section 7.2.7). For land plants, the isotopic selectivity during assimilation of S042- in soil water is difficult to evaluate since the foliage may also acquire gaseous atmospheric S (Section 7.2.5). Therefore, one should examine data from areas where the ambient sulphur oxide concentrations are low compared to the soil SOl- content. Further, data for soil water sulphate should be obtained in the vicinity of the roots. Isotopic similarity of total or organic sulphur in plants and S042- in the soil has been frequently reported. For example, Mekhtiyeva et al. (1976) found the foliar total sulphur to be slightly depleted in 34Swith respect to soil sulphate for six perennial plants growing near Moscow. There were temporal trends with the conifers becoming more depleted and the deciduous trees (linden, poplar, birch, and elm-leaved spirea) becoming more enriched in 34S during the period May-September. This isotopic similarity is usually

Pedosphere and Biosphere

287

found where soils are high in S content and ambient SOz is low. Examples are locations shown in Figure 7.9 where soil sulphur is derived from dissolution of evaporites or sea spray fallout. An alternative scenario is where the sulphur available to the plants is derived almost entirely from precipitation. This was found to be the case for plants growing in granitic rock fissures (Chukhrov et al., 1975). In highly polluted areas, the ()34S value of atmospheric sulphur may be the same as the soil and in turn the foliar sulphur in the absence of sulphur stress (Section 7.2.7). An alternate approach for determining the isotopic selectivity during S04Z- assimilation is to compare the ()34Svalue of foliar SOl- to the total or insoluble S. Data from one such study are shown in Figure 7.10. The bulk of the samples had sulphate levels (extracted by sonication in 0.1 N LiCI) lower than 0.02% whereas the total sulphur content by oxidation with a Parr combustion bomb was as high as 0.2%. With a few exceptions, the sulphur isotope composition of the sulphate was similar to that of total sulphur and averaged -7%0. The soil had similar ()34S values. Since atmospheric sulphur compounds had an average 334S value of + 10%0, the higher 334Svalues for some sulphate extracts in Figure 7.10 probably relate to uptake from the atmosphere. This illustrates how sulphur from different sources make it difficult to determine kinetic isotope effects in natural settings. In Figure 7.11, data for water-soluble S in plants are plotted against total plant S for specimens from various parts of Japan. It is noted that many

+16 0 +81-0

0 Li C Extractable Sulphur 0 Total

0 0
0

0 0 0 0 00 Cb 0

00 00 0 0 I 0.20 I 0.25 Q

00

0.05

0.10

I 0.15

S (% by wI)

Figure 7.10 The 534Svalues versus S content for LiCI extractable and total S in Populus tremuloides leaves, Teepee Creek, Alberta (from Krouse and Case, 1981)

288
+10
0 ~ ~
(f)

Stable Isotopes

w +5
..J OJ :::> --1 0 (f)

(f) <t 1'0

01-0

eo

LAND PLANTS JAPAN

-5 -5

0 8348 I

+5

+10

TOTAL 8 (%0)

Figure 7.11 The 334S values for water-soluble S versus those for total sulphur in plants from Japan (data provided by N. Nakai)

points fall close to the line of unity slope going through the origin. There is a slight preference for points to plot above the line, suggesting that, on the average, foliar organic sulphur is 1-2%0 lighter than the soluble S. Thus, in situations where ambient SOz is low and uptake of soil sulphate dominates, the soluble and insoluble sulphur in plants have similar sulphur isotope compositions. If the uptake of atmospheric sulphur compounds is extensive, then different extracts, and indeed different locations in a plant (Section 7.2.5), may differ markedly in isotopic composition. It is interesting to note that the soluble S component is more mobile and on average should have been acquired more recently than the organic S. Hence, deviations from the line of unity slope in Figure 7.11 could reflect temporal changes in the isotopic composition of the source SOl-. Therefore, comparisons of foliar SOl- and insoluble S over time can potentially serve to monitor changes in pollutant S levels and/or sources in the soil or atmosphere. In summary, as with aquatic plants (Section 6.2.2) and bacteria (Section 1.3.1.1), sulphur isotope selectivity during SOl- assimilation by land plants is small with a tendency for the organic S to be relatively depleted in 34S by 1-2%0. There is also evidence that when atmospheric SOz is incorporated by lichens, the transformation to organic sulphur involves very little isotopic selectivity (Krouse, 1977).

Pedosphere and Biosphere

289

7.2.4. Uptake of sulphide by plants and algae Fry et at. (1982) found submerged and emerged rooted plants in anoxic sediments to have 334S values near that of the sulphide species (-24%0) rather than the SOl- in the interstitial pore water (+15 to +17%0). They could not ascertain whether sulphide was incorporated directly or whether it was oxidized at the root-sediment interface. Carlson and Forrest (1982) found the salt marsh cordgrass, Spartina atterniflora, to be isotopically similar to sulphide in the marsh pore water. They concluded that sulphide had been oxidized within the plant. Sulphur in marine algae is generally slightly depleted in 34S with respect to seawater sulphate (Section 6.2.2). Lacking roots, algae do not have access to sulphide in the sediments. These observations contrast to studies with algae in thermal springs (Krouse et at., 1970; H.R. Krouse and A. Sasaki, unpublished data). Rather, algae tend to acquire the 334Svalue of dissolved sulphide even though the concentration of SOl- may be two or three orders of magnitude higher (Figure 7.12). Sulphate crusts were found on some specimens and their 334Svalues were similar to the dissolved sulphide species. The sulphur-oxidizing bacteria Beggiatoa sp. were also found associated with some algae (F.D. Cook, personal communication). The above data are consistent with the ability of cyanobacteria to carry out anoxygenic photosynthesis using sulphide (Castenholz, 1973; Cohen et at., 1975). This topic will be thoroughly reviewed in another SCOPE publication.

SPRING LOCATION FAIRMOUNT

.6HS 0 SorJ
I

WATER ... O Sor rganic ALGAE

(HS-Free)
LUSSIERR.
ALBERTA
BANFF ALBERTA

.0
I I I I

.
["pp" CAVE
MIDDLE
I

t:. ... .
I t:.
I I

0
I I

0 0
I I 0 I 0 I

....... IILXl

... ...

I LABORATORY.........
CULTURE
I

.
I I I

-20

-10

+10

+20

+30

834S (%0)
Figure 7.12 The ;334S values of algal growths, HS-, and SOl- in thermal springs of western Canada. The laboratory culture comprises algae which developed in a clear stoppered glass bottle of spring water during storage on an open shelf for several months (data provided by A. Sasaki and H.R. Krouse)

290

Stable Isotopes

For the purposes of the current volume, it is significant that this process represents a mechanism by which the biosphere may respond to pollutant S in the form of sulphide. 7.2.5 Relative incorporation of soil and atmospheric sulphur by vegetation Vegetation incorporates sulphur rather directly from the atmosphere, e.g. gases entering stomata, or by assimilating SOi- and other S-bearing ions in soil or water bodies. Relative uptakes of atmosphere S gases and soil S042- can be readily ascertained if the sulphur isotopic compositions of these two sources differ significantly (Krouse, 1977). In some situations, one uptake mechanism dominates. Epiphytic lichens have 534S values close to that of ambient S02 (Figure 7.13) unless under severe stress (see Section 7.2.7; Figure 7.20). Further, there may be topographical effects. Lichens on the leeward side of hills were found to contain less S and lower 534Svalues than those facing a source of industrial emissions enriched in 34S (Case and Krouse, 1980). In the former case, the [S02] was lower and a greater proportion of it came from natural sources. Another example of one dominant uptake mechanism is shown in Figure 7.14. In this case, bark acquired sulphur dust introduced to the atmosphere during loading operations. Some of the dust was oxidized, and the sulphate in turn converted to insoluble organic S. However, large quantities of elemental S were also entrapped. The decrease in 534S values with lower concentrations of So and derived compounds was coincident with greater distances from the source of So dust. (The rationale for plotting 534Sversus

concentration

-1

is discussed in Section 5.5.1.)

Conifer needles and deciduous leaves have 534Svalues between those of soil SOi- and atmospheric S gases (Figure 7.13). The proportion of sulphur derived from each source depends upon many factors, including availability of suitable sulphur forms and plant response. Although high concentrations of S02 may damage higher plants, there is also evidence that atmospheric sulphur compounds can provide nutritional needs (Faller et af., 1970; Cowling et af., 1973). The concentration at which S02 exposure is injurious is not well defined. It depends upon factors such as duration and intensity of exposure, susceptibility of the species, stage of growth, humidity, soil moisture, and light intensity. On the basis of experiments of Zahn (1961), 0.15 ppm (420 f-Lg S02 m-3) is gaining acceptance as a threshold below which prolonged exposure will not be detrimental. The two S uptake mechanisms are also expressed in the isotopic variability among different parts of an individual plant. In the studies of Chukhrov et af. (1975, 1978), plant parts were found to be uniform in isotopic composition. In other studies, considerable variability has been found (Figure 7.15). The roots and tops for bull rushes were compared in a number of shallow-water

Pedosphere and Biosphere

291

AIR S02

i
(f)

-l a..
0 (f)

LICHENS

I.l...

0:: W
(I)

::) Z

PINE NEEDLES

CONTROL

SITE

SOIL 0
LOOSE LITTER GREEN MOSS DEAD MOSS ORGANIC MINERAL

I.......
W~

Ia.. E u
5

I I +10

+20 834 S(%o)

+30

Figure 7.13 The (334S values for atmospheric SOz, vegetation, and soil, Ram River Area, Alberta, 1971-2. Note how vegetation reflects S uptake from the atmosphere and soil. Note also that moss cover impeded the transport of atmospheric S to the soil subsurface

292 +22L 52

Stable Isotopes

j
+201,

AI All

TAm

0 ::Q. 0

+18

r
+16

11
1 [TOTAL

1"

]513

+14
0

fY
2 S]-1X10-a 3

Figure 7.14 The 534Svalues versus the inverse of concentration for sulphur in bark specimens, West Whitecourt, Alberta, The bars signify the ranges in concentration and 534Svalues for several specimens at a given site. The 534Svalues decreased with increasing distance from the industrial operation (data provided by A. Legge and H. R. Krouse)

bodies around a sour gas processing plant near Twin Butte, Alberta. The isotopic composition of SO/- in the water and associated flooded soil (or sediment) varied, with higher 834Svalues corresponding to greater S inputs from the industry. Roots tended to have the same isotopic composition as surrounding soil. For some specimens, the tops of the plant had nearly the same 334S value as the roots. These corresponded to situations which had been influenced very little by the industrial emissions or those containing high [SOi-] with a considerable anthropogenic component. The ambient [SOz] was small in comparison so the upper parts of the plant derived sulphur dominantly from SOi-. With ponds containing low [SO/-] the plant tops acquired considerable atmospheric S. These locations correspond to points significantly above the line, 334Stops = 334Sroots' The concept of uptake of both soil and atmospheric S was pursued with potted plant experiments (Winner et ai., 1978). Soil and plants grown therein

Pedosphere and Biosphere BULLRUSH +20

293

+10

"IuJ ;1;

-~ 0
'"

-10

.
-10
834 5,.", (%.)

-20 -20

+10

Figure 7.15 Comparison of B34Svalues in the tops and roots of bullrushes (Typha latifolia) near a sour gas processing plant (from Krouse and Tabatabai, 1986). Points above the line of slope unity correspond to proportionally greater uptake of atmospheric S compounds by the exposed upper parts of the plant

had initial 534S values near -8%0. In early June, they were exposed to ambient S02 with 534Svalues near +20%0, near an industrial operation. The 534S values of both the plants and soil increased uniformly throughout the growing season if the soil surface was exposed. If charcoal, peat, or moss were placed on the surface, the soil tended to retain the 534S value established in the laboratory. The foliage of plants grown in covered soils did not increase in 534S values as quickly as those in uncovered soils. In the bare soil surface case, pollutant S was acquired ultimately from the atmosphere since atmospheric S entered the soil. In the covered soil case, sulphur of industrial origin was acquired by the plants from atmospheric gaseous S but little atmospheric S had entered the soil. Moss cover has also been shown to be effective in blocking the movement of sulphur compounds from the atmosphere to the subsoil in natural situations (Figure 7.13). The potted plant experiments imply that in the environments examined by Chukhrov et at. (1975, 1978), either the uptake of atmospheric S compounds by foliage was minimal or the soil and air had similar isotopic compositions. The relative rates of uptake of sulphur from the soil and atmosphere may result in a height dependence for the isotopic composition of foliage in a

294

Stable Isotopes

single plant. Whereas upper foliage in a stand of trees tends to interact with sulphurous compounds in the atmosphere, lower foliage may take up its sulphur almost entirely from the soil. This is a consequence of the upper foliage exerting a canopy action on the lower branches. This has been noted on many occasions in forest ecosystems exposed to emissions from the sour gas industry in Alberta (Figure 7.16 and Figure 7.21 in Section 7.2.7). The atmospheric sulphur had 1)34S values near + 20%0whereas that of soil sulphur was closer to 0%0. It should be noted that uptake of SOz by the upper growth may not be solely due to conductance through stomata. Some SOz may be oxidized and washed down as SOi-. Other species such as fungi may utilize atmospheric S. The 1)34S height gradient becomes more pronounced if the vegetation emits HzS under high sulphur stress (see Section 7.2.7; Figure 7.21). The relative incorporation rates of the atmospheric and soil sulphur is also found to vary with different plant species at a given location. In the data of Figure 7.17, where industrial emissions are minimal, the 1)34S values

16
MATURE OVERSTORY
PINE TREE

E f::r: S2 w ::r: 8

-J
0 10

SHORT UNDERSTORY
PINE TREE

I
30 40

20 834S (%0)

Figure 7.16 Variation of 034Svalues in conifer needles with height at one location in a forest stand near a sour gas plant operation, West Whitecourt, Alberta (from Krouse et al., 1984)

Pedosphere and Biosphere

295

+20

.
00

USNEA PICEA

.
0

.;
(f) <T 0<')

+10

.
..*~0
0

00

-10

1000

2000

3000

TOTAL S (ppm)
Figure 7.17 Comparison of 1)34S values for total S in lichens and spruce as a function of concentration in an area of Alberta where industrial emissions are minimal. Note that the mean values are similar (in contrast to Figure 7.13)

for both lichens and pine needles tend towards a similar 834Svalue at high concentrations. This is not the case in Figure 7.13 where the difference of 20%0in the 834S values of the air and soil is reflected in the different 834S distributions for lichens and needles. There are topographical factors that indirectly influence the relative amounts of sulphur acquired by the plants from soil and air. This concept is illustrated by 834Sdata for needles from conifers growing in rolling terrain around a sour gas plant operation near Zama, Alberta, Canada (Figure 7.18). High foliar sulphur concentrations are not consistent with the positive 834S values of the industrial emissions. Further, conifer needles at higher elevations contain less sulphur but have 834S values closer to that of the emissions. This apparent inconsistency can be resolved by considering the soil water behaviour. Subsurface SOi- (negative 834Svalues) leached from knolls has migrated and accumulated in depressions. Trees growing in the depressions with high [SOi-] near the roots have high foliar S contents. Some of these trees are under sulphur stress and vulnerable to small additions of industrial S. However, their needles with a high S content would retain a 834S value close to that of the soil sulphate and not reflect the minor industrial addition. In contrast, trees growing on exposed knolls with sulphurdeficient soil may be effectively fertilized by pollutant S02' Their foliage would acquire (j34S values approaching near to those of the industrial emiSSIOns.

296

Stable Isotopes

+4
0
0

9
0 1200 0 0

~ 0

0
@// %

./

(/) -4 <t r<> ro

E 0~

-8

I 350

)60 I I

800 ..:; en
400 380 (m)

-12 L-L.Z:.
400 450 500 ELEVATION (m) 550 360 370 ELEVATION

Figure 7.18 Variations in 034Svalues and total S concentrations in conifer needles with elevation in rolling terrain near a sour gas plant operation in Alberta

7.2.6 The influence of biological parameters on the sulphur isotope composition of plants Sulphur isotope abundances can be used very effectively for determining sources of sulphur in the environment. Assessing the environmental consequences of pollutant sulphur is not straightforward. One approach is to search for correlations between biological parameters and the isotope data. One biological indicator of possible environmental damage is areal extent (coverage) of species. Data from mosses growing in the Fox Creek area of Alberta are shown in Figure 7.19. Traverse A was downwind from the industrial source which emits SOz enriched in 34S. The second, angular traverse B is designated as the direction in degrees from downwind. The coverage increases with distance from the stack and greater angle from the dominantly downwind direction. For both traverses, the isotope data attest that reduction in coverage of mosses corresponds to an increase in the proportion of sulphur of industrial origin in the vegetation. Biological damage indicators can often be related to sulphur isotope compositions and, in turn, excessive pollutant sulphur. However, this coincidence is not proof that any damage was due to the anthropogenic sulphur. Other pollutants may be the cause, or their presence promoted damage by the sulphur stress. There may also be secondary effects of the sulphur increment, such as lowering of the Se/S ratio or undesirable chemical reactions, e.g. dissolution of metals.

Pedosphere and Biosphere

297 B
100

+261-A

~ 0

t
+22

.
W W

-+r<>
t()

<.9 <! 0:::

.
3 6 9 0 60

t tr
120
0

w U 0::: W CL

DISTANCE (km) (o)

DEGREES UPWIND (b)

Figure 7.19 Variation in 034Svalues of total S in mosses with coverage: (a) traverse with distance downwind, (b) traverse with direction in degrees from downwind of a sour gas refining operation (data from Winner et al., 1978)

7.2.7 Emission of reduced sulphur compounds by vegetation Plants have been shown to emit HzS (DeCormis, 1968; Spaleny, 1977; Wilson et ai., 1978). Sulphur stress, light, and other factors such as root damage have been shown to enhance these emissions. The process was found to be isotopically selective in laboratory experiments with cucumber plants, Cucumis sativus (Winner et ai., 1981). Whereas the S34Svalues of the leaves were similar to that of SOi- and HS03 - available to the roots, the emitted HzS was comparatively depleted in 34S by about 15%0. Indirect evidence that such gaseous emissions occur in nature is provided by stable sulphur isotope data. In areas of high S stress, foliar S has S34S values that are considerably higher than any available S sources. For example, data for the epiphytic lichen, Usnea scabrata, sampled downwind from a sour gas processing operation near Fox Creek, Alberta, fit a plot of S34Sversus concentration-1 (see Section 5.5.1) with the exception of two specimens designated by asterisks in Figure 7.20. These two samples with S34Svalues of +31.9 and +27.2%0contain 2394 and 1430 ppm S respectively, and were collected within 2.2 km of the emission stack of a sour gas processing plant. Thus the high 34Senrichments, which equal or exceed the maximum for ambient atmospheric S compounds, are identified with elevated sulphur contents and high exposure to SOz. There is a 'reservoir effect'

298
Usnea scabrata spp. nylanderiana

Stable Isotopes

a a (f) r<J

+30

.
r2 = 0.97

+20

+10 0 10

20

[Sr\104
Figure 7.20 The 8345 values versus the inverse of total 5 concentration for Usnea scabrata sampled near Fox Creek, Alberta, Canada (data from Case and Krouse, 1980)

(Section 1.2.3) in the foliage whereby the loss of 34S-depleted (or 32S-enriched) sulphur as emissions progressively increases the relative concentration of foliar 34S. The 834Svalues for total S in conifer needles may increase with height in the vicinity of sour gas processing operations (Section 7.2.4). This effect is usually less pronounced than that found for a Lodgepole X jack pine tree in the West Whitcourt region (Figure 7.21). For this specimen, the 834S value for the uppermost crown exceeded the maximum 834S value of atmospheric S by 10%0,consistent with the reservoir effect described above. Further evidence is suggested by the isotopic composition of salt crystals which occur occasionally on plant leaves. Two situations were examined by Shakur (1982). Evaporite crystals on vegetation near Kelly Lake Springs (6524' N, 12606' W) were found to have 8180 and 834Svalues very similar to those for SOi- in the spring waters. In these springs, the [S042-] ranged from 250 to 800 ppm. This contrasts markedly with an unusual natrojarosite, NaFe3(OHMS04)z, depositing spring (6512' N, 12438' W) near Norman Wells, NWT, Canada. The [SO/-] is much higher, 5300 ppm, in these waters. The 834S values of the mineral deposits are very close to that of S042- in the spring. Crystals of various sulphates occur on vegetation around this spring. In contrast to the brilliant yellowish-orange colour of the jarosite deposit, the crystals on plants are white and do not contain iron. They are also enriched in 34Sby over 3%0as compared to the spring SO~- and mineral deposit (Figure 7.22).

Pedosphere and Biosphere 25


1975 1976 20 E I15
/

299

jJ1975 --\ \ \

-----

--0 1976

~
w I

10
5 0 +20

KROUSE eta/. (1983) +22 +24 +26 +28


(%0)

+30

+32

+34

834S

Figure 7.21 The 834S variations in needles of two Lodgepole x Jack Pine trees (indicated by closed and open symbols) as a function of height, West Whitecourt area, Alberta (from Krouse et al., 1984)

WHITE

BLOEDITE a<1

8345

=-19.1 ~
=- 22.3

8345

NaFe3(OH)6 (S04)2 NATRO-JAROSITE(orange) 2- -504 5300 ppm --8345=-22.8 -=-~-

~&
CRYSTALS \

MgS04'Na2S04'4H20
HEXAHYDRITE MgS04' 6H20 EPSOMITE

~-

MgS04' 7H20 KAINITE MgS04. KC'. 3 H2O

Figure 7.22 The 834S values for crystals on vegetation and minerals at a natrojarosite depositing spring, Norman Wells, NWT, Canada (data from Shakur, 1982)

Thus, needles, sources process

in areas of high 5 stress, 0345 values have been found for lichens, and salt crystals on vegetation that are higher than those of sulphur in the vicinity. This phenomenon is consistent with a reductive in which the 325 species react at a faster rate. 7.3 FOOD WEBS AND HIGHER ANIMALS

Higher animals cannot assimilate 5042-, i.e. to convert it to amino acids such as cysteine and methionine. They must ingest these essential organosul-

300

Stable Isotopes

phur compounds and interconvert them as required. Isotopic selectivity during assimilation by aqueous plants and animals tends to be small (Section 6.2.2). Similarly, assimilation of sulphate by terrestrial plants appears to be accompanied by minimal isotopic fractionation (Section 7.2.3). Sulphur isotope fractionation or 'shifts' in a food web must be evaluated where one source of sulphur dominates. In marine food webs, S042- with a 034Svalue near +21%0 is the primary S source. Shifts to lower 034Svalues by as much as 3%0are found for marine algae and plants. Salmon off the coast of western Canada have 034Svalues of + 17 to + 18%0 for inner flesh and + 19%0 for the skin (Krouse and Herbert, 1988). These same authors reported that the fur of polar bears in the Canadian Arctic had 034Svalues near + 17%0as compared to the range of + 16 to + 18%0for the fur of seals, their main food source. Thus sulphur of marine origin has progressively moved up this food chain. With higher animals, it is difficult to differentiate small kinetic isotope effects during metabolism from preferential metabolism of different food constituents (carbohydrates, fats, proteins, etc.) which had slightly differently 034Svalues. Heron Island on the Great Barrier Reef of Australia has a large bird population because of the absence of predators. The combination of sea spray sulphate and fertilization by birds that are primarily fish feeders has resulted in the soil, vegetation, and feathers of birds having a remarkably uniform 034S value near + 19%0(Krouse and Herbert, 1988). The fur of koala bears at the Lone Pine Sanctuary, Queensland, Australia, was found to be about 2%0 depleted in 34S as compared to their diet of eucalyptus leaves. Hair clipped from kangaroos at the same location had 034S values close to the mean of their known food sources (Krouse, 1989). Sulphur isotopic compositions have been determined for hair, nails, blood, urine, and cystine kidney stones in humans. The overall variation in 34S of these components in a given human is about 2%0and appears to approximate the mean sulphur isotope composition of the diet. Data for residents at Calgary, Alberta, Canada, were reported by Krouse et at. (1987). The 034S values for available foods range from -7 to +20%0. Locally produced meat, dairy products, and grain constitute a large part of the diet and have 034S values close to 0%0.It was found that hair, nails, blood, urine, and cystine kidney stones in Calgary residents also have 034S values close to 0%0. Incidentally, there was no discernable difference in the sulphur isotope composition of kidney stone formers and non-formers. In contrast to the narrow range of isotopic composition within one individual, large variations have been found in the isotopic composition of hair, nails, and kidney stones with geographical location (Krouse and Levinson, 1984; Krouse and Herbert, 1988; A. Sasaki, unpublished data). For example, the 034S values for nails of attendees at the 1983 SCOPE/ UNEP Sulphur Workshop in Pushchino ranged from +3 to +14%0.

Pedosphere and Biosphere

301

From the above, it follows that if a human increases the marine food content of his diet, the &34S value of his tissues and fluids should increase. Whiskers of a Calgarian with &34S values ranging between + 1.5 and + 2.0%0 increased to + 3.1%0after he resided for one month in Japan. Upon returning to Calgary, they returned to the pre-visit values after about six weeks (Table 7.1). The oxygen isotope data for SOi- in urine from the above study are also interesting (Table 7.1). They reflect the greater 180 depletion in drinking water of Calgarians in comparison to Japanese and imply that some oxygen from body water molecules is incorporated into SOi- during biochemical oxidation. Whereas the above study relates to changes in diet with location, the food sources at a given location may be seasonally dependent and might be reflected in the sulphur isotope composition of animal tissues. This may be the case for the data from consecutive fingernail clippings shown in Figure 7.23. The ability to incorporate reduced inorganic sulphur compounds appears not to be limited to certain plants and algae (Section 7.2.4). Dense animal assemblages near deep sea hydrothermal vents, including vestimentiferan worms, brachyuran crabs, and giant clams, were found to have &34Svalues near 0%0,close to the range for nearby sulphide minerals (Fry et at., 1983). It seems unlikely that the vent fauna utilized sulphide directly. Primary productivity was probably due to chemoautotrophic bacteria utilizing sulphide as an energy source. In some instances, sulphur oxidizing bacteria may have functioned symbiotically within tissues of host animals. Preliminary isotopic data attest to sulphur originating from industrial processing, moving up to higher members of food webs. The feathers of
Table 7.1 Changes in isotope composition of a Calgarian who visited Japan for one month Sample time Whiskers Pre-visit End of visit Back in Calgary 3 days 7 days 14 days 28 days
034S (%0)

olga (%0) Urine SO}Urine S042-

+1.5 to +2.0 +3.1


+2.4 +2.0 +2.0

-0.1 +2.1
+1.5 +0.6 +0.4 +0.2

-5.8 -2.3 -3.0 -3.7 -4.2 -5.8

35days

302
+5.0
0 0

Stable Isotopes

.8, I 1'0" 0I -"""' 8, """'.Q

.-

0 8

FINGER NAIL TOENAIL

~+4.5
(f) <t r<> GO

II

---'8

+4.0
I

-------0 8 -L

0 ,,/
"'"

8 ,.0 " ,,..,.

JUL. 1982

OCT.

JAN. 1983

APR.

JUL.

Figure 7.23 Variation in sulphur isotope composition of fingernail clippings over a one-year period (unpublished data of A. Sasaki, 1983)

grouse living downwind of a sour gas plant (834S value of emissions near + 22%0) in Alberta, Canada, were found to have 834S values near + 10%0 whereas specimens not exposed to the emissions had 834S values near 0%0 (W. Winner and H.R. Krouse, unpublished data). This suggests that about one-half of the sulphur in the former can be attributed to industrial sulphur emisisons which were directly or indirectly incorporated into vegetation of the bird's diet. Environmental studies of higher members of the food chain have logistic advantages in the sense of integration and therefore the possibility of fewer measurements. For example, the isotopic composition of one bee should represent the average for thousands of plants in an area. Ideally, the animal should not roam over a large region. It must be emphasized that such data are only relevant to long-term phenomena. Although a short-term exposure to low levels of HzS may be lethal, it would not be measurably reflected as a change in isotopic composition because of the large mass of biological sulphur in the animal. Further, unlike plants and soil, one should not expect to observe radical variations in sulphur concentration. For example, the sulphur content of hair is about 4% as dictated by its molecular composition. Sulphur isotope data may prove useful for studying related environmental problems. The extent to which the Se/S ratio is depressed in the environment by industrial sulphur additions should be discernable by plotting this ratio against the 834S value. Since Se deficiency incurs such problems as white muscle disease in sheep, this plot should be constructed with data from soils, grass, and animals. The isotopic composition of metals, with the exception of lead, are difficult to determine. However, if excess metals arise in the environment from a process where sulphur emissions are involved, e.g. ore roasting, then plots of metal concentrations versus 834Svalues should prove informative. The use of stable isotopes of several elements can prove very useful for complex food web analyses. A combination of stable isotopes of sulphur,

Pedosphere and Biosphere

303

carbon, and nitrogen were found to eliminate many ambiguities in tracing organic matter flow in salt marshes and estuaries (Peterson et at., 1985). Although birds on Heron Island, Great Barrier Reef, had a narrow range of 034Svalues (+ 17.9 ::!:: 0.4%0),there was a wide range in the carbon isotope composition of their feathers (Krouse and Herbert, 1988). More restricted ranges in carbon isotope composition accrued when the data were assigned to the categories: tree top dwellers, ground nesters, fish eaters, and waders. In the case of human hair, populations can be differentiated in a plot of 034S values versus carbon isotope composition more effectively than with data from either element alone. Residents of Canberra and Brisbane, Australia, have similar 034S values (+ 12 to + 13%0) which are markedly different from those of residents of Calgary, Canada (near 0%0). However, the residents of Canberra can be distinguished from those of Brisbane because diets of the latter are more enriched in the heavier carbon isotope Be. On the other hand, the mean carbon isotope composition of Calgarians is not too significantly different from that of residents of Canberra (Krouse and Herbert, 1988). Clearly, extension to include stable isotope data for other elements, such as N, H, and 0, would further remove ambiguities. REFERENCES
Anderson, J.W. (1975). The function of sulphur in plant growth and metabolism. In: McLachlon, K.D. (Ed.) Sulphur in Australian Agriculture, University Press, Sydney, 87-97. Anderson, D.W., Saggar, S.K., Bettany, J.R., and Stewart, J.W.B. (1981). Particle size fractions and their use in studies of soil organic matter. 1. The nature and distribution of carbon, nitrogen and sulphur, Soil Sci. Society of America Jour. 45, 767-72. Bettany, J.R., and Stewart, J.W.B. (1983). Sulphur cycling in soils. Sulphur 82 Proc. Int. Conf-, II, 767-85. Biederbeck, V.O. (1978). Soil organic sulphur and fertility. In: Schnitzer, M. and Khan, S.U. (Eds) Soil organic matter, Elsevier, Amsterdam, 273-310. Blair, G.J. (1979). Sulphur in the Tropics. International Fertilizer Development Center, Muscle shoals, Alabama. Bloomfield, c., and Coulter, J.K. (1973). Genesis and management of acid sulphate soils. Advances in Agronomy 25, 265-326. Carlson, P.R. (Jr.), and Forrest, J. (1982). Uptake of dissolved sulphide by Spartina alterniflora. Evidence from natural sulphur isotope abundance ratios, Science 216, 633-5. Case, J.W., and Krouse, H.R. (1980). Variations in sulfur content and stable isotope composition of vegetation near a SOz source at Fox Creek, Alberta, Canada. Oecologia 44, 248-57. Castenholz, R.W. (1973). The possible photosynthetic use of sulphide by the filamentous phototrophic bacteria of hot springs. Limnol. Oceanogr. 18, 863-76. Chae, Y.M., and Krouse, H.R. (1987). Alteration of sulfur-34 natural abundance in soil by application of feedlot manure. Soil Sci. Am. 1. 50, 1425-30. Chukhrov, F.V., Yermilova, L.P., Churikov, V.S., and Nosik, L.P. (1975). K

304

Stable Isotopes

biogeokhimii izotopov sery (sulphur-isotope biogeochemistry). lzv. An SSR, sa. Geol. 8 (in Russian). Chukhrov, F.V., Yermilova, L.P., Churikov, V.S., and Nosik, L.P. (1978). Sulphurisotope phytogeochemistry. Geokhimiya 7, 1015-31 (in Russian). Geochem. Int. 1978, 25-40. Cohen, Y., Jorgensen, B.B., Padin, E., and Shilo, M. (1975). Sulphide-dependent anoxygenic photosynthesis in the cyanobacterium Oscillatoria limnetica. Nature 257, 489-91. Cowling, D.W., Jones, L.H.P., and Lockyer, D.R. (1973). Increased yield through correction of sulphur deficiency in ryegrass exposed to sulphur dioxide. Nature 243, 479-80. DeCormis, L. (1968). Hydrogen sulfide production by plants placed in an atmosphere containing sulfur dioxide. C.R. Acad. Sci. Ser. D. 266, 7, 683-5. Dijkshoorn, W., Lampe, J.E.M., and van Burg, P.F.J. (1960). A method of diagnosing the sulphur status of herbage. Plant Soil 13, 227-41. Faller, N., Herwig, K., and Kuhn, H. (1970). Die Aufnahme von (S350Z) Schwefeldioxyd aus der Luft. Plant and Soil 33, 177-91. Fitzgerald, J.W. (1976). Sulphate ester formation and hydrolysis: A potentially important yet often ignored aspect of the sulphur cycle in aerobic soils. Bacteriological Reviews 40, 698-721. Fitzgerald, J.W. (1978). Naturally occurring organosulphur compounds in soil. Sulphur in the Environment, Part II, Ecological impacts. In: Nriagu, J.O. (Ed.)

Wiley - Interscience, New York, 391-443.

Fitzgerald, S.W. and Lushinski, P.O. (1977). Further studies on the formation of choline sulfate by bacteria. Can. 1. Microbiol., 23, 5, 483-90. Freney, J.R. (1967). Sulfur-containing organics. In: McLaren, A.D., and Peterson, G.H. (Eds) Soil Biochemistry, Marcel Dekker, New York, 229-59. Freney, J.R., Melville, G.E., and Williams, C.H. (1967). The determination of carbon-bonded sulphur in soil. Soil Science 109, 310-18. Freney, J.R., Melville, G.E., and Williams, C.H. (1971). Organic sulphur fractions labelled by addition of 35S-sulphate to soil. Soil Biology and Biochemistry, 3, 133-41. Freney, J.R., Stevenson, F.J., and Beavers, A.R (1972). Sulphur-containing amino acids in soil hydrolysates. Soil Science 14, 468-76. Freney, J .R., Melville, G.E., and Williams, C.H. (1975). Soil organic matter fractions as sources of plant-available sulphur. Soil. Bioi. Biochem., 7, 217-21. Freney, J.R., and Williams, C.R (1983). The sulphur cycle in soil. In: Ivanov, M.V. and Freney, J.R. (Eds) The Global Biogeochemical Sulphur Cycle, Scope Report No. 19, 129-202. Fry, B., Scalan, R.S., Winters, J.K., and Parker, P.L. (1982). Sulphur uptake by salt grasses, mangroves and seagrasses in anaerobic sediments. Geochem. Cosmochim., 46, 1121-4. Fry, B., Gest, H., and Hayes, J.M. (1983). Sulphur isotopic compositions of deepsea hydrothermal vent animals. Nature 306, 51-2. Hendry, M.J., Krouse, H.R., and Shakur, M.A. (1989). Interpretation of oxygen and sulfur isotopes from dissolved sulfates in Tills of Southern Alberta, Canada. Water Resources Res. 25, 567-72. Janzen, RH., and Bettany, J.R. (1984). Sulphur nutrition of rapeseed 1. Influence of fertilizer nitrogen and sulphur ratios. Soil Sci. Soc. America Jour. 48, 100-7. Krouse, H.R. (1977). Sulphur isotope abundances elucidate uptake of atmospheric sulphur emissions by vegetation. Nature 265, 45-6.

Pedosphere and Biosphere

305

Krouse, H.R. (1980). Sulphur isotopes in our environment. In: Fritz, P. and Fontes, J.e. (Eds) Handbook of Environmental Isotope Geochemistry, Vol. I. The terrestrial Environment, Elsevier, New York, 435-72. Krouse, H.R. (1989). Sulfur isotope studies of the pedosphere and biosphere. In: Rundel, P.W., Ehleringer, J.R., and Nagy, K.A. (Eds). Stable Isotopes in Ecological Research, Springer-Verlag, New York, Inc. Ecological Studies, 68, 424--4. Krouse, H.R., Cook, F.D., Sasaki, A., and Smejkal, V. (1970). Microbiological isotope fractionation in springs of Western Canada. In: K. Agata and T. Hayakawa (Eds) Recent Developments in Mass Spectroscopy. Proc. Int. Cont. on Mass Spectroscopy. Kyoto, Japan, Sept. 1969. Univ. Tokyo press, 629-39. Krouse, H.R., and Case, J.W. (1981). Sulphur isotope ratios in water, air, soil, and vegetation near Teepee Creek Gas Plant, Alberta. Water, Air and Soil Pollution 15, 11-28. Krouse, H.R., and Case, J.W. (1982). Sulphur isotope tracing of acid emissions into the environment in Alberta. Proc. Symp. Acid Forming Emissions in Alberta and Their Ecological Effects. Edmonton 1982, Alberta Dept. of Environment, Canadian Petroleum, Association, Oil Sands Environmental Study Group. Krouse, H.R., and Herbert, H.K. (1988). Sulphur and carbon isotope studies of food webs. In: Kennedy, B.V., and LeMoine, G.M. (Eds). Diet and Subsistence: Current Archaeological Perspectives CHACMOOL, The University of Calgary, 315-22. Krouse, H.R. and Levinson, A.A. (1984). Geographical trends of carbon and sulfur isotope abundances in human kidney stones. Geochim. Cosmochim. Acta 48, 187-91. Krouse, H.R. and Tabatabai, M.A. (1986). Stable sulfur isotopes. In: Tabatabai, M.A. (Ed.) Sulfur in Agriculture, Amer. Soc. Agron., Crop. Sci. Soc. Amer., Soil Sci. Soc. Amer. Madison, Wisconsin, USA, 160-205. Krouse, H.R., and van Everdingen, R.O. (1984). 34Svariations in vegetation and soil exposed to intense biogenic sulphide emissions near Paige Mountain, N.W.T., Canada. Water, Air, and Soil Pollution 23, 61-7. Krouse, H.R., Legge, A.H., and Brown, H.M. (1984). Sulphur gas emissions in the Boreal Forest: The West Whitecourt Case Study V. Stable Sulphur Isotopes. Water, Air, and Soil Pollution 22, 321-47. Krouse, H.R., Levinson, A.A., Piggott, D., and Ueda, A. (1987). Further stable isotope investigations of human kidney urinary stones: comparison with other body components. Appl. Geochem. 2, 205-11. Kusakabe, M.T.A., Rafter, J.D., Stout, and Collie, T.W. (1976). Isotopic ratios of sulphur extracted from some plants, soils and related materials. N.Z.J. Sci. 19, 433-40. Lowe, L.E. (1965). Sulphur fractions of selected Alberta profiles of the chernozenic and podzolic orders. Can. 1. Soil Sci. 45, 297-303. Maynard, D.G., Stewart, J.W.B., and Bettany, J.R. (1984). The effect of plants on soil sulphur transformations. Soil Biology and Biochemistry 16 (in press). McCready, R.G.L., and Krouse, H.R. (1982). Sulphur isotope fractionation during the oxidation of elemental sulphur by Thiobacilli in a solonetzic Soil. Can. J. Soil Sci. 62, 105-10. Mekhtiyeva, V.L., Pankina, R.S., and Gavrilov, Ye. Ya. (1976). Distributions and isotopic compositions of forms of sulphur in water animals and plants. Geochem. Int., 82-7. Peterson, B.C., Howarth, R.W., Garritt, R.H. (1985). Multiple stable isotopes used

306

Stable Isotopes

to trace the flow of organic matter in estuarine food webs. Science227, 1361-3. Pumphrey, F.V., and Moore, D.P. (1965). Diagnosing sulphur deficiency of alfalfa plant analysis. Agron. J. 57, 364-6. Saggar, S., Bettany, J.R., and Stewart, J.W.B. (1981a). Measurement of microbial sulphur in soil. Soil Biology and Biochemistry 13, 493-8. Saggar, S., Bettany, J.R., and Stewart,. J.W.B. (1981b). Sulphur transformations in relation to carbon and nitrogen in incubated soils. Soil Biology and Biochemistry 13, 499-511. Shakur, A. (1982). 34Sand 180 variations in terrestrial sulphates. Ph.D. thesis, Univ. Calgary. Calgary, Alberta. Spaleny, J. (1977). Sulphate transformations to hydrogen sulphide in spruce seedlings. Plant and Soil 48, 557-63. Turchenek, L.W., and Oades, J.M. (1979). Fractionation of organomineral complexes by sedimentation and density techniques. Geoderma 21, 311-43. van Everdingen, R.O., Shakur, M.A., and Krouse, H.R. (1982). Isotope geochemistry of dissolved, precipitated, airborne, and fallout sulphur species associated with springs near Paige Mountain, Norman Range, N.W.T. Can. J. Earth Sci. 19, 1395-407. Wilson, L.G., Bressan, R.A., and Filner, P. (1978). Light-dependent emission of hydrogen sulphide from plants. Plant Physiol. 61, 184-9. Winner, W.E., Bewley, J.D., Krouse, H.R., and Brown, H.M. (1978). Stable sulphur isotope analysis of S02 pollution impact on vegetation. Oecologia 36, 351-61. Winner, W.E., Smith, c.L., Koch, G.W., Mooney, H.A., Bewley, J.E., and Krouse, H.R. (1981). Rates of emission of H2S from plants and patterns of stable sulphur isotope fractionation. Nature 289, 672-3. Zahn, R.K. (1961). The action of S02 on plants. Staub 21, 56-60.

CHAPTER 8

Case Studies and Potential Applications


H.R. KROUSE, L.N. GRINENKO, V.A. GRINENKO, L. NEWMAN, J. FORREST, N. NAKAI, Y. TSUJI, T. YATSUMIMI,U. TAKEUCHI,B.W. ROBINSON, M.K. STEWART,A. GUNATILAKA,L.A. CHAMBERS,J.W. SMITH, L.A. PLUMB, F. BUZEK, J. CERNY, J. SRAMEK,A.G. MENON, G.V.A. hER, V.S. VENKATASUBRAMANIAN, B.E.C. EGBOKA, M.M. IROGBENACHI, AND C.A. EUGWE

8.1

INTRODUCTION*

In previous chapters, sulphur and oxygen isotope systematics were summarized for the lithosphere, hydrosphere, atmosphere, and biosphere. Processes of isotope fractionation have been examined theoretically, in laboratory experiments, and in situ. Attempts have been made to calculate isotopic balances regionally and globally. In some chapters, the identification and quantification of anthropogenic sulphur by isotope techniques have been discussed. However, pollutant sulphur from a given source impacts on a variety of environmental receptors. Therefore, in a given study, a number of components in the ecosystem need to be examined. One approach for this book would have been an integration of the concepts from previous chapters from which prescriptive methodologies may have been recommended for various ecosystems. Although this is addressed briefly in Chapter 9, it was deemed better to summarize a number of actual case studies throughout the world. In this way, the reader should benefit from the first-hand experience of other colleagues trying to use sulphur isotopes to characterize anthropogenic sulphur in the environment. The case studies differ greatly in terms of number of environmental receptors, size of area (local or regional), number of sources of anthropogenic sulphur, uniformity of 834Svalues among sources and unpolluted receptors, and the 'isotopic leverage', i.e. the difference in 834S values of pollutant emissions and receptors. Examples are also given where only the 834Svalues of potential sources have been measured (Section 8.10) and where the amount and mean of 834Svalues of known emissions have been calculated
* V.A. Grinenko, H.R. Krouse.

Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko @ 1991 SCOPE, Published by John Wiley & Sons LId

0 00

VJ ~

0Figure 8.1 Sour gas processing plant, West Whitecourt, Alberta. Note elemental sulphur stockpile in the background (photo courtesy A. Legge)

r;;t;:;o

~ '" '"

Case Studies and Potential Applications

309

(Section 8.8). There is even a case where no isotope data are given but potential applications of sulphur isotope techniques are identified (Section 8.11). 8.2 SULPHUR ISOTOPE TRACING OF THE FATE OF EMISSIONS FROM SOUR GAS PROCESSING IN ALBERT A, CANADA*

8.2.1 Overview Sulphurous emissions, mainly SOz, from sour gas (high HzS content) processing in the Province of Alberta, Canada, are significantly enriched in 34Scompared to environmental receptors. This 'isotopic leverage' has proved to be an effective technique for tracing the fate of these emissions in surrounding ecosystems. These studies, which have been ongoing since 1970, have enhanced our basic understanding of the response of various environmental receptors to sulphur pollutants. A number of examples have been cited in Chapters 5 to 7. Data from the atmosphere, hydrosphere, pedosphere, and biosphere are available for locations within the Province which have been subjected to varying amounts of these emissions (Figure 8.12). Unfortunately, environmental sulphur isotope and often concentration data were not obtained prior to the onset of the emissions. Therefore, it is difficult to accurately evaluate the proportion of industrial sulphur in the environment. An effective approach has been a composite diagram in which 834S values of environmental receptors are plotted in order of decreasing elevation, i.e. atmosphere, tree lichens, ground lichens and mosses, leaves and needles, litter, and increasing soil depth. This order corresponds to that of decreasing 834S since the emissions have high 834S values. However, this trend may not hold in rolling terrain, where sulphur may be redistributed and the subsoil modified through leaching of sulphur-containing salts from knolls and redeposition in depressions (Section 7.2.5, Figure 7.18). Elemental sulphur is the product of sour gas processing. Although it is remarkably stable stored as large blocks (Figure 8.1), there is some input to nearby soil through bacterial oxidation to SO/-, and sulphur dust can be released to the environment during loading operations. 8.2.2 The sour gas industry of Alberta The sulphur recovery industry in western Canada, based on the processing of sour natural gas, is currently the largest in the world. For example, in
. H.R. Krouse

310

Stable Isotopes

1976,Canadaproduced 45% of the world's recoveredsulphur (Hyne, 1981). Production and reserve statistics as of 1982 are given in Table 8.1. Gas streams with as high as 53% HzS are currently processed and the technology to use streams containing up to 90% HzS is being developed. Wells with such high HzS concentrations can be found at Panther River, Hunter Valley, and Bearberry, Alberta. Initially, clean gas was the primary objective and sulphur was a recovered by-product. Historically, there have been times when the commercial value of sulphur could not justify its production and the methane contents of the wells with excessive HzS were too low for economic viability. More recently, higher financial returns from the sale of sulphur have ushered in an era of reservoir development primarily directed towards sulphur recovery. Sour gas is found mainly in Devonian carbonate reservoirs in western Canada. The percentage of HzS may increase with depth or temperature (Figure 8.2). The extremely high HzS concentrations are not found in lowTable 8.1 Western Canada HzS statistics, 1982 (Hyne, 1983) Processed daily Plant capacity Number of plants Reserves
10.5 x IOn m3 d-I 7.7 X 10~ m' yr-I 50; 75% of HzS is processed by 9 plants, one with a maximum capacity of 3 x IOn m3 d-I 9.5 X 1010 m3 equivalent to 1.35 x 1O~tonnes S

1001- DEVONIAN 397 WELLS


(f) (\J

:r:

I-

w
n..

Z W U a::

50

'

',' , ,"

, .
,

50

100 (OC)

TEMPERATURE

Figure 8.2 Percent HzS in natural gas versus reservoir temperature, Alberta, Canada (data from Energy Resources Conservation Board, Alberta)

Case Studies and Potential Applications

311

temperature shallow reservoirs. The isotopic composition of the HzS shows a marked transition at reservoir temperatures of about 80 DC(Section 4.5.4). Lower 834S values below this transition are interpreted as reflecting biological processes with large associated kinetic isotope effects, in shallower environments. Above 80 DC, thermochemical reduction of evaporites are believed to occur with negligible isotopic selectivity. The lack of isotopic discrimination might reflect complete sulphate reduction at localized sites or a step such as anhydrite dissolution assuming rate control. On the basis of carbon isotope data, Krouse et at. (1989) concluded that, in some environments, light hydrocarbon gases served as reducing agents. The net result is that HzS in gas-prone reservoirs of western Canada have 834Svalues in the range + 15 to + 30%0. The HzS is scrubbed from the gas by a variety of reagents with separation efficiencies of over 99%. Modified Claus process plants are used to convert the separated HzS into elemental sulphur (Figure 8.1). In this reaction, some HzS is converted to SOz which in turn is reacted with more HzS to produce the So end-product. This process emits SOz to the atmosphere. Flaring has also been a source of SOz. Despite increased production, the emissions have decreased as the result of more efficient processing (Table 8.2). The maximum emission rate, 5.4 x 105ton per annum, or nearly 1500 tonnes d-I, occurred in 1972. This flux is comparatively low, considering the size of the region and the fact that a number of individual smelters and power stations around the world have exceeded these rates. Nevertheless, with the increased emphasis on environmental pollution, it was important to determine the fate of the SOz emissions from the sour gas industry of Alberta. Traditional approaches used sulphur concentration measurements in the atmosphere, hydrosphere, pedosphere, biosphere, meteorological data, and models of plume behaviour. Since the emissions are very enriched in 34S compared to the preindustry receptors, stable sulphur isotope

Table 8.2 Improvement Date

in overall sulphur plant recovery efficiency in Alberta, Canada (Hyne, 1983) Annual production (106 tonnes) 1.59 4.31 6.72 7.57 5.93 (5.50) To atmosphere (103 tonnes yr-l) 190 430 540 330 237 (165)

Recovery (%) 88 90 92 95 96 (97)

Pre 1965 1970 1972 1975 1980 1983 (est.)

312

Stable Isotopes

abundances have proved to be very effective in supporting many environmental investigations. 8.2.3 Sulphur isotope composition of the atmosphere in Alberta 8.2.3.1 Regional pattern Since the atmosphere is the primary conveyor of S emissions and So dust from these industrial operations, a regional survey of the sulphur isotopic composition of S02 was an appropriate starting point. Initial measurements were carried out with lead peroxide treated cylinders which had been exposed for one-month intervals. The ;)34Svalues were found to range from +7 to +28%0 (Figure 8.3). The two major cities, Edmonton and Calgary,

ALBERTA

GRANDE PRAIRIE
(+ 18.6) .-( + 18.6)

.-( +27.1 ) .-( +15.5)

. EDMONTON
(+ 10.6)

(+ 129)-.

...(+120)

. RED DEER (+20.6)


.-( + 14.0) (+19.4}e .-(+141) (+16.7>-::{+172) (+16.6)-.CALGARY
.-(+18.2)

. (+10.9)

CREEK

Figure 8.3 The 834Svalues for atmospheric S02 trapped by lead peroxide exposure cylinders at various locations in Alberta, 1971

Case Studies and Potential Applications

313

had 834S values near + 11%0,whereas sites near sour gas plants had much higher values. Some specific gas plant locations relevant to the data of Figure 8.3 and later discussion are given in Figure 8.13. Data in Table 8.3 display a number of features. Often, but not always, the isotopic composition remained the same for consecutive samplings at the same site (Red Deer 4, Windfall 5, Okotoks 2, Coleman 8). For a given month, different sites at the same location may have similar 334Svalues (Red Deer 4 and 5, Okotoks 2 and 4). These consistencies suggest that lead peroxide exposure cylinders can provide reliable isotopic data. There were also inconsistencies. Consecutive samplings sometimes yielded significant isotopic shifts (Carstairs 4). A site at a given location could be isotopically different from other sites (Red Deer 6 and Okotoks 3). The overall isotopic behaviour was consistent with localized emissions which had not mixed effectively with the broader meteorological systems. The 334S value for an individual sample depended upon the distribution of emitters and wind behaviour. Since the lead peroxide cylinders provided integrated data for one-month periods, data for shorter terms were required

Table 8.3 Isotopic composition of atmospheric S02 near ground level in Alberta, 1971 Location Site Date S34S (%0) +20.9 +20.3 +20.7 +14.0 +27.1 +28.1 +15.6 + 15.4 +21.3 +21.5 +21.6 +18.2 +18.3 + 18.4 +12.7 +12.8 +16.7 +19.0

Red Deer

4 4 5 6 2 2 5 5 2 2 4 3 1 2 8 8 4 4

13 Aug. 15 Sept. 15 Sept. 15 Sept. 12 Aug. 16 Sept. 6 Aug. 9 Sept. 4 Aug. 1 Sept. 1 Sept. 1 Sept. 8 July 8 July 27 July 2 Sept. 11 Aug. 16 Sept.

Lone Pine Creek Windfall Okotoks

Bigstone Coleman Carstairs

314

Stable Isotopes

to relate 034Svalues to concentration and wind direction. Therefore, a highvolume sampling programme was developed using techniques described by Forrest and Newman (1973). 8.2.3.2 Relationship of 034S to concentrations of atmospheric S compounds Near a given sour gas plant operation, more positive 034S values usually correspond to higher concentrations. This was demonstrated convincingly during the monitoring of the startup of the Crossfield plant (see Section 5.5.4). It was also emphasized in a study near a gas refinery at Valleyview, Alberta, where flaring ceased after 25 years of operation (Krouse and Case, 1983). The a34s values for both atmospheric SOz and particulates decreased markedly when flaring stopped (Figure 8.4). It was shown in Section 5.5.1 that when a given site is sampled randomly under different meteorological conditions, the a34s values for high concentrations of ambient SOz approach that of the dominant industrial emitter in the area. Minor sources with various a34s values are evident at low concentrations. For some locations, the minor sources could be distant gas plant operations and possess even higher a34s values than the dominant emitter. This appears to be the case for Balzac where the a34s values for emissions from the nearest gas plant are about + 16%0but a34s values as high as +22%0 are found in the vicinity (Figure 8.5). Nearby, at Crossfield

Cf)

PRE - SHUTDOWN 0 POST-SHUTDOWN

C:{ -.J+ 20 ::) <.) Ia:::

--

- -.... 0

- 0

- 0

ct
<j'

Cf)

+10

r<>

ro

W8
0 0

+10

+20

+30

8345, 52(%0)
Figure 8.4 The 034Svalues for particulates versus the same for S02 at Valleyview, Alberta, 1982. Emissions from flaring ceased after 28 years of operation

Case Studies and Potential Applications

315

+24
........

+20
0

cf!- +16 (f) ~ ,., GQ +12

. . . . . . ..
.
.

"-

. '..............
-......-

. .
./ / /./ ./
/'

.--

-. ..... -/

,./

BALZAC 1972

+8

.
+4
0

/
/

0.2

0.4

0.6

53 (equiv. mg d-'100cm-2) Figure 8.5 The 8J4S values for SOz versus concentrations of the same at Balzac, Alberta. Data from lead peroxide exposure cylinders

(See Figure 8.12), the 034S value for the emissions is about + 29%0 (Section 5.5.4). Rarely have 034S values for ambient SOz in Alberta exceeded this value. Consequently, the 034S values for SOz collected near Crossfield increase to +29%0when the industrial contribution dominates. In contrast, at Balzac, increasing industrial S corresponds to the 034S values converging towards + 16%0 from more positive values (corresponding to contributions from Crossfield) and more negative values ('natural' sources). 8.2.3.3 Dependence of 034S values of atmospheric S on wind direction Since atmospheric gases from various sources reaching an observation point depend upon the wind pattern, sulphur isotope data acquired as a function of wind direction can serve to apportion sources. Arrays of four and eight high-volume samplers, wherein each sampler is triggered by a preselected wind direction, have been successfully tested near sour gas processing plants in Alberta. Winds coming from the direction of the industry tend to be the most enriched in 34S. This is illustrated in Figure 8.6 where winds from the north to north-east and south-west to north-west sectors contain SOz from two nearby industrial emitters. Other examples of data from high-volume sampling arrays are given in Section 5.3.2. A criticism of the ground-based directional array sampling is that winds near the ground are unlikely to be consistent with those aloft which may

316

Stable Isotopes

S
Figure 8.6 The 534Svalues of S02 obtained at site influenced by three gas processing plants using a wind directionally triggered high-volume sampling array. A ninth directionally independent sampler functioning at low wind speeds collected S02 with a 534Svalue of +9.30/00

be primarily responsible for the S02 transport. However, the presence of environmental receptors, including man, near the ground warrant such measurements. It is also clear that traditional meteorological data are required to establish atmospheric S02 flow patterns. 8.2.3.4 Isotopic composition of atmospheric particulate matter Occasionally, particulates in the atmosphere of Alberta have been found to have 034S values close to those of associated S02 (Section 5.3.2). The similarity is expected for elemental sulphur dust from storage blocks or loading operations at sour gas processing plants. Agreement is also expected when sulphate aerosols have formed from S02 oxidation (Section 5.2). In many instances, however, the particulates are isotopically quite different from the S02' The reason lies primarily in the nature of the particulates and more numerous sources, e.g. road dust. The diverse nature of atmospheric particulates can be shown by scanning electron microscopy (SEM). An SEM photograph depicting abiological dust particles as well as biological matter consisting of broken plant debris and pollen grains can be found in Section 5.3.3. The isotopic composition of the latter reflects a mixture of sulphur derived by the living plant from the atmosphere and soil (Section 7.2.5). Isotopic analyses of different aerodynamically sized fractions, coupled with SEM examination of particulate matter, can provide effective source

Case Studies and Potential Applications

317

identification. In studies near sour gas plant operations, particles of less than 1 J.Lmin size are typically depleted in 34S (by as much as 15%0) in comparison to larger sized fractions ranging up to greater than 7 J.Lm (Section 5.3.2.2). This difference possibly arises because the smaller particles can potentially arrive from a more distant source. 8.2.4 Sulphur isotope composition of surface waters Surface water SOi- in a given stream entering the Mackenzie River system of north-west Canada was found to vary widely in its 34Scontent dependent upon the surrounding geology (Hitchon and Krouse, 1972). Waters in the Peace River area were found to be very depleted in 34Sconsistent with the isotopic composition of native soils in the region (Lowe et at., 1971; Krouse and Case, 1981). Deep formation waters in Alberta tend to be enriched in 34Sbecause of association with marine evaporites. Occasionally S042- from this deep source appears at the surface in spring waters (Krouse, 1976). However, in most locations remote from industrial activity, the 034S values of surface water SOi- tend to be near 0 or negative. Therefore, the extent of sulphur of industrial origin in the hydrosphere can usually be seen by plotting 034S values versus [SOi-], as shown for Valleyview, Alberta, in Figure 8.7. For [SOi-] less than 10 ppm, the isotopic composition is similar to that of soil horizons beneath the litter layer. At 40 ppm S042-, the 034S value is close to that of nearby industrial emissions. In environmental studies, the relating of cause to effect is often difficult. Sulphur isotopes are very effective for calculating the proportion of industrial +20 0 0 ; +10 ::R 0
~

<T r<)

en

ro

B 0

- 10I-CtD
-20 0

20

40

DISSOLVED SULPHUR (ppm) Figure 8.7 The 534Svalues versus concentration of dissolved sulphur species in surface waters, Valleyview,Alberta, 1982

318

Stable Isotopes

S in an environmental receptor. However, measurements of other parameters are required to determine the effect. Even if a correlation is found between 1)34S and a damage parameter, it does not necessarily follow that pollutant S was the cause. However, at West Whitecourt, Alberta, the 1)34S value for SOi- was found to increase with the concentration of dissolved organic sulphur in surface waters surrounding a sour gas plant operation (Krouse et ai., 1984; see Figure 6.4). Industrial emissions introduced 34S-enriched sulphur into the environment, partly as SOi- in the hydrosphere. The industrial S also generated greater quantities of organic sulphur compounds in the environment, probably through biological assimilation processes, although chemical reactions should not be discounted. 8.2.5 Sulphur isotope abundances in soils and vegetation of Alberta Most native soils in Alberta range in 1)34S values from - 30 to + 5%0(Lowe et ai., 1971; Krouse and Case, 1981). Values near 0%0are found in southeastern Alberta, whereas the negative values are encountered in the Peace River area and the south-western corner. In the foothills, values near -10%0 are frequently found. The very negative values are associated with subsurface salt occurrences which in turn were probably derived from the weathering of reduced S compounds in Cretaceous shales. Since the emissions from sour gas processing have an average 1)34S value of +20%0, the presence of industrial sulphur in soil and vegetation is readily detected by sulphur isotope analysis. In Chapter 7, many examples from studies in Alberta were used to illustrate a number of fundamental concepts concerning sulphur in soil and vegetation. For this reason, only a few data from one study will be used to illustrate how measurements of the isotopic composition of many receptors at a given site enhance the interpretation (Figure 8.8). Decreases in 1)34S with distance from the stack in the West Whitecourt area are clearly shown for pine needles, litter, and different soil depths. The sites were chosen to be as ecologically equivalent as possible. This meant that they were not in a straight line with respect to the dominant NNW wind direction. This, coupled with topographical effects, can explain departures from smooth trends in Figure 8.8. As with many industrial operations, insufficient measurements were made on the ecosystem prior to the onset of emissions. The question then arises as to the extent the trends in isotope composition reflect varying industrial inputs and natural variations. Certain features of Figure 8.8 imply that the variation of 1)34S values is almost entirely due to diminishing amounts of industrial sulphur with distance from the stack. The mean 1)34S value of needles 1 km from the stack is the same as that of the emissions. The litter is very similar isotopically to the live needles.

Case Studies and Potential Applications


+25

319

+23
+ 21

-.. +191-

5cm ~ +17f-15cml\
30cmJl,1 \ 'v \1 \ +151II \ //

~,

"g "'-- +13 en ... r<> "" + 11 +9 +7

"

J \V"'~--.< 40Cf1J-."',~\ I~ .-\" ~-""'--"


I

;/

~I

/
/ /; 1 //

/.

/"'

/\ II 1

S~I

/"\

\+'

,.a--'\tI"/

~\(1...

1
S6

11 \1 II
II

+5'+3 +1

- 10

. . , . . , .
2 4 6

\\

S13
44

10

12

14 42

DISTANCE

FROM SULPHUR EMISSION SOURCES


(km)

Figure 8.8 A comparison of the mean 834Svalues of litter, Lodgepole x Jack Pine foliage, bark, and selected soil depths as a function of distance from S emission sources in the West Whitecourt study area (data provided by A. Legge and H. R. Krouse)

The differences in 834Svalues (834S) between two components at a given site and the variations in 834S for the same component among sites are particularly diagnostic. For needles at the nearest and furthest sites, the range is about 10%0,whereas it is nearly twice this value for either the 5 or 15 cm soil samples. This is expected since atmospheric sulphur compounds can interact quite readily with the needles, but their penetration into the soil is a delayed process. This is further verified by the 834Svalues for either needles or litter and surface soil at the nearest and further sites (1 and 12%0 respectively) . It is also noted that the soil samples at 40 cm depth have less isotopic variation with distance than more shallow samples since the former have been less contaminated with industrial sulphur. It is difficult to ascertain the background 834Svalue for the soil since it is not approaching an asymptotic limit with distance. There may in fact be two values, one near 0 and the

320

Stable Isotopes

other near +7%0.This is in markedcontrastto the data for Valleyview, Alberta, where a limit near -12%0 was found within 10 km of the flaring stack (Figure 7.5).
8.2.6 Isotopic modelling of uptake of industrial sulphur by the environment near sour gas processing plants in Alberta Sulphur isotope data for soil and foliage from many studies in Alberta can be combined into various composite diagrams. In areas that have not been influenced by industrial sulphur emissions (natural range), the (')34Svalues are generally near 0%0or quite negative. The (')34S values of the soil in the vicinity of the roots versus those of the corresponding foliage should fall on the line of unity slope if ambient SOz levels are low. This reflects leaves deriving the bulk of their sulphur from the soil during growth and returning sulphur to the soil upon decay. In an intermediate situation where some industrial sulphur enriched in 34S has been introduced, the soil tends to be isotopically lighter than vegetation for two reasons: (a) The leaves contain isotopic mixtures of sulphur derived from the soil and 34S-enriched air. (b) Litter and botanical cover may limit the penetration of atmospheric compounds into mineral soil horizons (Section 7.2.5). Consequently, when the (')34S results for vegetation are plotted against those for soil, data plot above the line of unity slope (Figure 8.9). This trend was also found for analogous plots with the tops and roots of bullrushes (see Section 7.2.5). In industrial areas where sulphur compounds have penetrated the soil, data tend to plot on a line of slope unity. In this case, soil, atmosphere, and foliage have similar but higher (')34S values because of the prevalence of sulphur of industrial origin. In preliminary studies in the Ram River area, it was concluded that in contrast to needles and leaves, epiphytic lichens acquire a sulphur isotopic composition similar to atmospheric S (Krouse, 1977). This relationship does not hold true for all type of lichens and mosses. In particular, species dwelling near the ground appear to acquire some sulphur from the litter layer and may have (')34Svalues intermediate to the epiphytic lichens and leaves. These observations suggest another synthesis of isotopic data in which vegetation is further subclassified in the conceptual model of Figure 8.10. Since studies to date suggest that preindustrial (')34Svalues of the environment were zero or lower, it is postulated that (')34Svalues will

Case Studies and Potential Applications


+40

321

+20
(; lJ.J (D <1

0 l.1.. <t '" C()

::J en

-20

-40 -40

-20

0
8345, SOIL (%0)

+20

+40

Figure 8.9 The 834S values of foliage versus the same for soil as a function of exposure to sour gas processing emissions and soil cover

progressively increase from the deep soil horizons upwards to epiphytic lichens in response to industrial sulphur deposition. The 'intermediate' situation corresponds to the levels of sulphur of industrial origin being comparable to those of the natural environment. This situation should not be considered in terms of absolute sulphur content. For example, if the natural sulphur content were low, the environment could remain depleted in sulphur even with industrial contributions. In this case, sulphur isotope data could display the intermediate pattern even though an environmental response to the additional sulphur may not be obvious. In the intermediate situation, it is expected that some of the 34Senriched industrial sulphur compounds will have penetrated the soil. The concentrations of these compounds and their &34Svalues tend to decrease with depth. The &34S value for atmospheric compounds will be slightly less

W N N
INDUSTRIAL INPUTTO ENVIRONMENT RELATIVETO NATURAL LEVELS 0 NONE () INTERMEDIATE

. HIGH

AIR
t:I: !2 w :I:
TREELICHENS

---

0
0 0 0 0

. - ---- - - - - - - - - -()()()()<L - 1-------------()() ()() ()()(j()()

------

.........

NEEDLES
GROUNDLICHENS

LEAVES

-- 0
0

I- -----------()()()()

()

()()

.. .. ... ------...
. .. .
. .
!

n
't
:I: t....

ANDMOSSES

0
SOIL 0 0

0 0
() ()

()

()

()

O()

PREINDUSTRIAL

EMISSIONS

Figure 8.10 Trends in the sulphur isotope compositions of foliage and soil near sour gas processing plants in Alberta, Canada

VJ is' <:J" ~ ~ ~ ~ ~

Case Studies and Potential Applications

323

than that of the industrial emitter because of dispersion and mixing with background sulphur sources. Epiphytic lichens tend to isotopically resemble the atmosphere. Trees tend to be isotopically intermediate to the soil and the atmosphere and usually display a height trend with the uppermost needles/leaves possessing a higher component of atmospheric sulphur than the lower ones (Krouse et ai., 1984). This is attributed to a canopy action exerted by the upper foliage. The ground lichens and mosses tend to display j)34Svalues between those of trees and epiphytic lichens. In the high industrial impact situation (Figure 8.11), the isotopic composition for all components of the environment is essentially that of the emissions. This does not necessarily imply that the sulphur content of the environment is excessive but rather the majority of sulphur is of industrial origin. If excessive environmental sulphur is present, j)34Svalues for epiphytic lichens and pine needles may be higher than that of ambient SOz. This is attributed to the isotopically selective emission of reduced sulphur compounds by the vegetation under sulphur stress (Section 7.2.7). In Figure 8.11, data from West Pembina and Brazeau are tested in the conceptual model of Figure 8.10. In both investigations, the data fit the intermediate situation of Figure 8.10. Lichen possessed the highest j)34S values near +20%0, consistent with derivation of S from the atmosphere. There are no excessively positive j)34Svalues for either lichens or needles which would indicate S stress. Except for the moss data, the j)34Svalues for vegetation and soil tend to be slightly more positive at West Pembina. In the West Pembina area, there was possibly more effective upper story foliage cover so that the moss received less exposure to atmospheric S compounds. Penetration of sulphur of industrial origin into the upper A horizon is evident at all sites since the j)34Svalues are more positive than for underlying layers. The Band C horizons show similar ranges in both study areas but different mean values. This suggests that penetration below the A horizon is minimal and the higher j)34Svalues generally present in the West Pembina area reflect a more 34S-enriched background. In both studies, it must be emphasized that the S contents are low even with industrial S contributions, i.e. the data fit the intermediate case of Figure 8.11 because relatively small amounts of industrial S have been added to low S backgrounds. It is important to note that data for soil horizons were used for Figure 8.11. Data for selected depth intervals were also obtained but the horizon data were judged to be better in these and other studies. Prescribed depth intervals containing varying contributions from different horizons tend to obscure the isotopic variations. 8.2.7 Discussion and recommendations The use of sulphur isotope abundances in environmental research in Alberta has been ongoing for two decades. However, there are many unexplored

-5

+5

+10

+15

+ 20

+ 25

+30

W N ....

LICHENS

MOSS

VEGETATION

NEEDLES

LITTER

SOilS

BC IC

-5

0
DOME WEST

+5
PEMBINA BRAZEAU

+ 10

+ 15

+20

+25

+30

8348 (%0)
c,.,

I (x)

I PETRO-CANADA

NUMBER OF ANALYSES

Figure 8.11 Composite diagram comparing the sulphur isotope composition of environmental components as a function of elevation at West Pembina and Brazeau, Alberta

~ (;;

is' ~ '" ~ C

Case Studies and Potential Applications

325

research avenues. Isotopic compositions have successfully documented the presence of sulphur of industrial origin in the environment without the ambiguities that have arisen with other techniques. The link between cause and effect has only been established in a few cases (e.g. Figures 6.4 and 7.21). There is every indication that if isotopic investigations are carried out with individual sulphur compound groups, transformations of sulphur compounds in the environment will be better understood and the relation of 'cause' to 'effect' better established. For years, elemental sulphur was stockpiled. With recent demand, these piles have dwindled, but in the course of loading, sulphur dust in the environment in some areas is perceived as more problematic than gaseous emissions. There is a challenge to distinguish between these two sources of pollutant S which have similar 034S values at a given industrial site. A comparison of 034Sdata for bark and foliage shows some promise. Elemental S can be entrapped in bark whereas foliage acquires sulphur either as SOi- from the roots or by permeation of gaseous S compounds through stomata. It also appears that emission of reduced S compounds under S stress does not occur with bark. This can explain why the bark is not as enriched in 34S as needles in Figure 8.8. Oxygen isotope analyses of SOimight be another approach to distinguishing that derived from S02 or So. Over the years, the sour gas industry has become more sympathetic to sulphur isotope studies. Whereas background data are preferred, a survey at any point in the history of the industry operation is useful. For example, the data in Figure 8.11 provide a reference to which future data may be compared. Since the S content of the environment was low, 034S values should be quite sensitive to future industrial S inputs. Colleagues in industry frequently ask: 'What is the minimum number of analyses that can provide satisfactory interpretation of industrial S uptake by the environment?' There is the related question as to how many specimens can be examined on a limited budget. The data of this section and Chapter 7 show that answers to these questions are not straightforward and depend upon many environmental factors. A few general comments can be made. Epiphytic lichens provide a cumulative sulphur isotopic record of atmospheric S. Depth profiles of 034Svalues can provide information on the penetration of industrial S into soil. If detailed profiles are not possible, then sampling by horizon is preferred over sampling one or two prescribed depths. Site selection is important for realizing effective data and should be based on ecological criteria as well as predicted regional sulphur deposition patterns. A perceived problem in agriculture in Alberta is depression of the Se/S ratio by the industrial S additions to the environment. As a precaution, cows are injected with selenate solution prior to calving. Plotting of this ratio against 1)34Svalues for foliage and tissue samples should serve to ascertain whether sour gas emissions noticeably influence the selenium cycle.

326

Stable Isotopes

Environmental studies of higher members of the food chain have the logistic advantage of reducing the number of required samples (Section 7.3). However, there are relatively few sulphur isotope data for higher animals in Alberta. A current isotopic study of bees generally reveals the areal extent of pollutant S. However, the data are not ideal, presumably because bees may selectively feed on agricultural crops that reflect cultivation and fertilizer practices. Data from wild grazing animals who refrain from eating agricultural crops may prove more effective in regionally mapping the influence of industrial S throughout Alberta.

*ell

~
,,-,.'1:111::1818111:1.

EDMONTON

<8>

_:,.'~.:n'J::I:.
.l8l:18","1:\::II::I.'J

'_:,.,rl.'..
<8>

CALGARY

100 km """'II~.:\lIIiI::ll

Figure 8.12 Location of study sites near sour gas plants in Alberta

Case Studies and Potential Applications

327

8.3

THE ISOTOPIC COMPOSITION AND CONTENT OF SULPHUR IN SOILS OF KANSK-ACHINSK FUEL POWER GENERATION COMPLEX*

8.3.1 Introduction In the 1970s the first phase of a large coal-burning power complex was brought into operation in the Kansk-Achinsk region of the Soviet Union. The target consumption for this complex is 50 x 106 tonnes yr-1 and most of this is supplied from local open-p;t mines. The emission of S02 from this plant caused concern as to its impact on the local environment and prompted a study of the sulphurous depositions. Monitoring the areal extent and distribution of the deposited S and its fate in the environment can be a mammoth task involving continual measurement of many biological and physical factors including atmospheric concentrations, wind speed, emisison rates, and forms of S in soil, plants, and animals. A possible alternative for evaluating the amount, distribution, and fate of the emitted S in the ecosystem is stable sulphur isotope analyses since different coals and S02 from their combustion often have 034S values differing from those of soils and plants in the region (Grinenko and Grinenko, 1974). The study reported here was initiated to see if the 034Svalues for emissions differed significantly from the natural environment. Assuming that the difference was sufficient, a second objective was to document changes that have occurred in the sulphur isotope compositions of the environmental receptors over time. Soil samples from 23 sites were collected and analysed for S content and 034S values. Coal, cinder, and ash collected from furnaces and S02 from the combustion products were similarly analysed. 8.3.2 Results Concentration data for various sulphur forms and isotopic compositions in coal samples are given in Table 8.4. Sample 1 contained only sulphide sulphur, whereas sample 2 contained both sulphide and organic sulphur. In sample 2, the total sulphur content was found to be twice that of sample 1. The isotopic compositions of sulphur in the samples differ. In sample 2, it is just slightly enriched in 34S compared to the meteorite standard. Sulphide sulphur in sample 1 is markedly enriched in 34S (+8.6 to +9.4%0). Thus the sulphur in these coals has different origins. The sulphur contents in three samples of cinders remaining in furnaces after combustion were similar (about 0.4%), but 3 to 5 times higher than the raw coal. Their sulphur isotopic compositions were also similar, approximating that of coal sample 2 but differing significantly from coal sample 1. This is attributed to the sampling of ash after combustion of coal . with sample 2 characteristics.

L.N. Grinenko and V.A. Grinenko.

328

Stable Isotopes

Samples of the fly ash taken from different parts of a connecting pipe were also analysed (Table 8.5). As seen from the table, fly ash accumulates large amounts of sulphur (up to a few percent), i.e. 4-9 times higher than in cinders and 25-30 times higher than in coal. The sulphur isotope composition of fly ash was intermediate between coal samples 1 and 2. The locations where soils were sampled around the factory are shown in Figure 8.13. At a number of sites, samples were taken in depth increments representing several horizons. The results of analyses are presented in Table 8.6 and Figure 8.14. The total sulphur content in the soil samples varied from 0.06 to 0.31 % which is typical of most soils (Builov and Builova, 1976). The inorganic sulphur content for the majority of samples was greater than the amount of organic S present (this is not usually the case; see Section 7.1.1). There are no clear trends in the content of either sulphur forms with depth of sampling or location of sampling sites along the Nazarovsky transect (Figure 8.13). The isotopic compositions of sulphate and organic S in most samples fall within a narrow range (+0.7 to + 3. 7%0, except for two, +4.8 and 5.8%0, both of which occurred in surface horizons) (Figure 8.14). The isotopic compositions of organic sulphur and sulphate are rather similar
Table 8.4 Content and isotopic composition of different sulphur forms in coal and ash samples Sample Sulphide S (%) Coal 2 Coal 1 Ash 1 Ash 2 Ash 3 0.08 0.06 0.34 0.35 0.32 8'4S (%0) +0.7 +9.4 +2.2 +1.3 +3.6 Sulphate
---

Organic S (%) 0.123 Not detected 8'4S (%0) +2.4

S (%) Not detected Not detected 0.34 0.39 0.38

834S(%0) +1.8 +0.9 +3.3

Table 8.5 Content and isotopic composition of total sulphur in 'fly ash' samples Sample Section of connecting pipe (cm)
~---

S (%)

8'4S (%0)

-----

1 2 3 4 5

0- 30 30- 60 60- 90 90-120 120-150 Average

2.03 1.48 1.75 2.38 3.55 2.24 i: 0.8

+5.0 +3.4 +4.4 +4.1 +3.8 +4.1 i: 0.6

Case Studies and Potential Applications

329

82

88

87

SAMPLING SITE 8 SETTLEMENTS Ii(iI

Figure 8.13 Map of the Kansk-Achinsk study area showing sampling locations

.
0 +1 +2 +3 +4

80:8

DORGANIC

+5

+6

8348 (0/00)
Figure 8.14 Distribution of 0345 values of sulphate and organic sulphur in soils

Table 8.6 Content

and isotopic

composition content

of sulphur (%)

in soils 814S (%0) -Sulphate


-------

tH tH

0 Sulphur Sample Site Type of soil Sample depth (cm) Sulphate Organic

Organic

1 2 3 4 5 6 7 8 9 10 11 12 14 15 16 17 18 19 20 21 22 23

2a 2b 3a 4a 4b 5a 5b 6a 6b 1 1a 1b 12a 12b 12c 11a lIb lOa lOb lOc 9a 9b 8a 8b 7a 7b 7c

Grey forest soils Grey forest soils Grey forest soils Grey forest soils Grey forest soils Alluvial grassland Meadow soils Meadow soils Meadow soils Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem Chernozem

0-3 7-13 0-3 0-3 7-13 0-3 7-13 0-3 7-13 0-3 7-13 0-3 7-13 20-23 0-3 7-13 0-3 7-13 20-23 0-3 7-13 0-3 7-13 0-3 7-13 20-23

Trace Trace 0.23 0.12 0.22 0.12 0.23 0.19 0.17 Trace 0.1 Trace 0.07 Trace Trace 0.14 0.2 Trace Trace 0.13 0.14 0.22 0.25 Trace 0.06 0.10

0.06 0.08 0.06 0.08 0.08 0.08 0.06 0.06 0.07 0.08 0.06 0.11 0.11 0.12 0.18 0.08 0.11 0.08 0.06 0.01 0.08 0.06 0.06 0.10 0.06 0.09

Not detected Not detected +2.5 +5.8 +2.2 +1.7 +1.6 +4.8 +1.4 Not detected +2.5 Not detected Not detected Not detected +2.3 +1.9 Not detected Not detected +2.6 Not detected +2.4 +1.9 +1.6 +1.9 +1.4

+1.4 +3.7 +2.5 +1.4 +2.5 +3.1 +3.5 +1.1 +1.6 +2.6 +1.3 +1.1 +1.0 +1.7 +1.4 +1.2 +1.2 +0.7 +1.1 detected detected detected detected +2.0

Not Not Not Not

v, is' 0-

.g '" '"

::;;:::;()

Case Studies and Potential Applications

331

but appear to differ more as f>34S increases. Kusakabe et at. (1976) did not find a significant difference in the isotopic composition of such sulphur forms in soils of some regions of Tunis and New Zealand, in spite of the fact that f>34S values varied there from +12.6 to +20.4%0 (Section 7.1.4). It would seem that sulphur in soil samples of the current study originated from decayed plant residues which accumulated for many years prior to the power plant operation. There is little difference in the isotopic composition of sulphur from different soils. With the exception of surface SOi- samples at sites 4 and 6, f>34S values for forest, alluvial grassland, and chernozem soil sulphate are: +2.2 to +2.5, +1.4 to +1.7, and +1.4 to +2.6%0 respectively; and for organic sulphur: + 1.4 to + 3.7, + 1.1 to + 3.5, and +0.7 to 2.6%0respectively. There is no trend in f>34S values along the vertical soil profiles that would suggest bacterial SOi- reduction. This is consistent with the relatively good permeability of the soils and hence aerobic conditions. 8.3.3 Conclusion The patterns of sulphur isotopic ratios and S contents in the studied soils in most cases are characteristic of natural occurrences, and as yet have not been noticeably affected by the power plant S emissions. Some industrial S may be present in the two surface soil samples taken from sites 4 and 6 near the power station in the downwind direction. Sulphate in these samples is enriched in 34Scompared to other locations and is about 4%0heavier than the coexistent organic sulphur. This suggests that the sulphate in the surface layer at sites 4 and 6 was affected considerably by 34S-enriched rainwater sulphate and S02 dryfall arising from coal combustion. It is noteworthy that the concentrations of sulphates enriched in 34S are not statistically higher than at other locations. Further, isotopically heavy sulphate does not occur in deeper soil horizons. Therefore, anthropogenic S contamination is very localized in this area and insignificant to date. 8.4 SULPHUR ISOTOPE MEASUREMENTS RELEVANT TO POWER PLANT EMISSIONS IN THE NORTHEASTERN UNITED STATES* 8.4.1 Introduction A body of sulphur isotope data has been collected for fuels, power plant emissions, and atmospheric sulphur in the northeastern part of the United States. This information is now presented in part to document its existence but more importantly to stimulate the reader to consider how such

L. Newman and J. Forrest.

332

Stable Isotopes

information might be used to gain a better understanding of the fate of sulphur in the environment. The specific sampling techniques and related considerations are described elsewhere (Forrest and Newman, 1973, 1977; Forrest et at., 1973; Newman et at., 1971, 1975a, 1975b; Tucker 1969). 8.4.2 Fuels in use A program was undertaken to obtain information on the isotopic composition of sulphur in fuel oils supplied to users (mostly power plants) in the states of New York, Connecticut, and Florida. The results are presented in Table 8.7. These data are insufficient to compute exact regional averages but serve more to show representative values at these locations. The average (334S value was found to be approximately + 5%0 with a range of about -1 to +9%0. Fuel oils from anyone supplier or used at anyone power plant can show a variation in (334S of more than 3%0. This prompted a study of the variation with time of the isotopic composition of fuel oil used at two representative power plants over a two-year period. The results are shown in Table 8.8. Both these power plants obtained their fuel from the same source during the study. The average (334Svalue, found at the Northport plant, was +4.3%0 with a range from -0.4 to +5.4%0. The results at English Station were +5.1%0with a range from +3.5 to +6.2%0. However, it is noted that the (334S values could have daily variations of greater than 2%0. The isotopic composition of sulphur in coals used at two representative power plants, one in the state of Missouri and the other in the state of Pennsylvania, was also examined. A representative (334Svalue for coal is difficult to determine since coal by its very nature is not uniform and the large stockpiles at a given plant can be a mixture from a variety of suppliers. Obtaining a sample of a few grams that is representative of many tons of a highly variable material is a formidable task. Nevertheless, sample feedstock to burners were sampled at various times and the results are presented in Table 8.9. It is not surprising that within-day changes were almost 2%0 and day-to-day variations were as much as 6%0. The resulting variability in the isotopic composition of the sulphur emitted into the atmosphere was documented by sampling flue gases at two oilburning and two coal-burning plants over periods of 1-3 years. The results are presented in Table 8.10. Although the range in (334S values for the oilfired plants was less than 3%0 over a two-year period, in the case of coal it was found to be almost 9%0 over one year at Labadie. During combustion, generally less than 1% of the sulphur is converted to sulphur trioxide. The isotopic ratio of the emitted S03 was measured and the results are also presented in Table 8.10. The (334S values were found to be up to 3%0 more positive than that of coexisting S02' This fractionation has been discussed (Forrest et at., 1973) and is a direct measure of the

Case Studies and Potential Applications

333

Table 8.7 Sulphur isotope composition of fuel oils sampled at various sites in the eastern USA Date New York 10/02/68 10/17/68 10/02/68 10/17/68 10/02/68 01/09/70 01/29/70 03/09/70 12/08/73 11/19/74 09/11/75 Connecticut 10/17/68 05/14/70 10/17/68 05/14/70 10/17/68 05/14/70 10/17/68 05/14/70 05/14/70 05/14/70 05/14/70 05/14/70 05/14/70 Florida 08/18/76 08/18/76 08/19/76 Name
834S (%0)

Arthur Kill Power Plant Arthur Kill Power Plant Astoria Power Plant Astoria Power Plant Ravenswood Power Plant BNL Power Plant BNL Power Plant BNL Power Plant Albany Power Plant Port Jefferson Power Plant Roseton Power Plant Wyatt Oil Supplier Wyatt Oil Suppier Tad Jones Oil Supplier Tad Jones Oil Supplier Metro Oil Supplier Metro Oil Supplier Connecticut Refining Oil Supplier Connecticut Refining Oil Supplier Seamless Rubber Industrial User Federal Paper Industrial User Sinking Industry Industrial User Olin Matheson Industrial User Yale Central Power Plant Anciote Power Plant Anciote Power Plant Anciote Power Plant

+0.2 +0.9 +1.8 +3.5 +0.5 +6.5 +6.5 +8.7 +5.6 +1.5 -1.1 +6.5 +8.5 +7.4 +9.2 +5.8 +9.1 +5.1 +5.9 +8.9 +7.5 +7.0 +9.1 +9.1 +4.7 +4.8 +4.5

temperature at which the oxidation occurs during the combustion cycle. The emitted S03 is immediately hydrated in the atmosphere and serves as a primary source of aerosol sulphate. 8.4.3 Sulphur isotope compositions of ambient ground-level S02 and sulphate Almost 400 samples of sulphur dioxide and sulphate at a number of locations in the northeastern United States have been analysed over a period of 8 years. Samples were taken using high-volume samples with the filter pack

334

Stable Isotopes

Table 8.8 Variations in sulphur isotopiccompositionof oils sampled at powerplants in northeastern USA
Date 834S(%0) Long Island Lighting Co., Northport, N.Y. +0.1 -0.4 +5.0 +4.1 +5.2 +5.2 +5.0 +5.0 +5.4 +4.7 +4.2 +5.0 +4.9 +5.1 +5.0 +4.3 +4.9 English Station, New Haven, Connecticut +3.9 +6.2 +5.4 +3.5 +5.4 +5.3 +6.1

09/17/68 09/18/68 OS/22/69 06/05/69 04/09/70 04/14/70 05/14/70 06/10/70 07/07/70 06/23/70 12/08/70 11/10/70 04/21/71 05/12/70 11/10/70 12/08/70 04/21/71 02/26/68 03/07/68 04/18/68 10/17/68 07/31/69 05/12/70 05/13/70

described by Forrest and Newman (1973). The sampling times ranged from 1 to 24 hours. Averages and ranges for S02 at each location are given in Table 8.11. The overall average 834S value for S02 is +2.2%0 with a range from -2.1 to +8.7%0. Results for sulphate samples, obtained simultaneously with S02, are given in Table 8.12. The overall average for the sulphate 834S value is +5.1%0 with a range from +1.1 to +13.7%0. Therefore, the spreads in 834S values for S02 and sulphate were found to be similar, 11 to 12%0. All the data for which there are measurements of both sulphate and sulphur dioxide are plotted in Figure 8.15. One notes that most of the data lie above the y = x or 'one-to-one' line, meaning that the 834Svalue of an individual sulphate sample is almost invariably higher than that of cosampled S02' The average difference is 3%0,which is slightly higher than the average increase of approximately 1.5%0 (see Table 8.10) for the sulphate emitted

Case Studies and Potential Applications

335

Table 8.9 Variations in the sulphur isotopic composition of coals sampled at power plants Date Keystone, Pa. 05/07/68 05/09/68 05/01/69 09/02/70 OS/24/71 OS/25/71 Labadie, Mo. 06/12/74 06/13/74 06/14/74 06/16/74 06/17/74 06/18/74 07/30/74 8'4S (%0)

+1.9, +1.3, +0.1 + 1.3, + 1.3, + 1.3 + 3.4, +4.2 +2.9 +2.4 +3.2 -1.3 -4.3 +2.3 -0.7 +0.8 -1.9 -2.3

Table 8.10 Sulphur isotope composition of power plant flue gas samples Power plant Fuel type Periods 8'4S, S02 (%0) Mean BNL Northport Keystone Labadie Oil Oil Coal Coal 1969-70 1969-71 1970-71 1975 Range 8'4S, SO, (%0) Mean Range Number of samples

+7.4 +6.2 to +8.4 +4.8 +4.1 to +5.4 +2.9 +1.3 to +3.7 -2.7 -5.0 to +3.6

+8.9 +8.2 to +9.4 +6.6 +5.4 to +8.4 +4.3 +3.0 to +4.9

7 15 27 8

Table 8.11 Sulphur isotope composition of ground-level S02 samples in northeastern USA Location Period Mean
----

8,4S (%0) Range -2.1 to +8.7 -2.0 to +4.7 -1.9 to +3.9

Number of samples

New Haven, Conn. Long Island, N.Y. New York City Whiteface Mountain, N.Y.

1968-70 1970-75 1968-72 1972

+2.6 +2.1 +1.5

150 177 50

+ 1.8

-1.6 to +2.0

336

Stable Isotopes

Table 8.12 Sulphur isotopiccomposition of ground-level sulphatesamplesin northeastern USA


Location Period Mean Atlantic Coast, N.Y. Whiteface Mountain, N.Y. New York City, N.Y. Long Island, N.Y. New Haven, Conn. . Next highest value +6.9. 1974 1972 1971-72 1970--75 1969-70 +4.2 +5.9 +8.5 +5.3 +4.2 (')34S(%0) Range + 3.3 +4.1 +3.2 + 1.9 +1.1 to +7.3 to +7.9 to +13.7 to +9.8 to +13.3* 6 3 11 175 78 Number of samples

+14

.
+12
+10
0
"0

Ground level samples 0 Aircraft samples

'.

"g 0
<t

.
+8

if) ro

'"

.'

. .. '

-" ".. .' , . ..

.."

UJ 1-+6

J:
-I

Do ::> (/)+4

'.'

. ,,:', :. , " '.. ',' ., """'1,:... . .. . "':..'1:0', , '0 . ":' . : '-:;', :':,:",=': .: . '::' /.' '.. '..
0

. .

0
0

, Q' ":~',: "I"'":'::,.( ": ~:.::'0 '; ,,::::~' '. , . ",. ,


Q,

+ 2

-3

-2

-I

+1

+2

+3

+4

+5

SULPHUR

DIOXIDE

834S

(%0)

Figure 8.15 Plot of the (')34S of sulphate versus those of simultaneously collected S02

Case Studies and Potential Applications

337

as a result of combustion. One should refrain from trying to ascribe any simplistic interpretation to these observed differences in the isotopic ratios of sulphur dioxide and sulphate. Many factors contribute to these differences, including different sources, fractionation during combustion, a variety of oxidation reactions in going from S02 to sulphate, and fractionation arising during removal processes. Samples were also taken within the mixed layer of the atmosphere, with a high-volume sampler mounted on board an airplane (Romano et al., 1975). In the vicinity of power plants, care was taken to be upwind so as to avoid sampling the high concentrations from local emissions. Since there is more time for atmospheric mixing and less likelihood of local contamination these samples can be considered as more representative of a given area than the ground-level samples. Sampling times ranged from 0.5 to 1 hour. The results for sulphur dioxide are presented in Table 8.13. The overall average 034S for more than 100 samples is +2.9%0 with a range from -5.1 to +12.5%0. Noteworthy is that this airborne average is not significantly different from the value of +2.2%0 found at ground level. Since atmospheric sulphate concentrations are relatively low compared to S02, only a few suitable sulphate samples were obtained with available aircraft time (Table 8.14). Although the overall 034S value of +3.8%0 certainly cannot be considered representative of the whole north-east, it is not very different from the value of + 5.1%0obtained with the larger number of ground-level samples. When the 034S values for sulphate are plotted against those for simultaneously collected S02, the aircraft data are not significantly different from the ground-level data (Figure 8.15).

Table 8.13 Sulphur isotope composition of S02 sampled by aircraft Location Period Mean Long Island, N.Y. Roseton, N.Y. Albany, N.Y. New Haven, Conn. Northern, N.J. Keystone, Pa. St Louis, Mo. Sudbury, Canada Charleston, W.Va. 1968-74 1975 1973 1969-74 1970 1968-71 1974 1974 1972 +2.1 +1.6 +4.5 +3.8 +1.9 +2.4 +3.2 +2.6 834S (%0) Range -2.0 to +10.9 -5.1 +0.4 -0.6 +2.7 to +12.5 to +2.6 to +7.0 to +4.1 25 1 1 44 4 21 11 1 Number of samples

+0.9

- 2.5 to +1.6

338

Stable Isotopes Table 8.14 Sulphur isotope composition of sulphate sampled by aircraft Location Period
Mean 0'45 (%0) Range -0.2 to +9.6

Number of samples

Keystone, Pa.

1971

+3.8

13

8.4.4 Available fuels and other sources The isotopic compositions of sulphur in fuels that are potentially available for use in power plants were also determined. One motive was to search for fuels that could be used as unique isotopic tracers. There are four ways of utilizing sulphur isotopes tracers in power plant studies: (a) The sulphur in the power plant stack gas may have a (')'4S value sufficiently different from the background to permit tracer studies with current plant operations. (b) Power plants could be operated solely with special fuel from natural sources, selected on the basis of a sufficient difference in (')34S from the surrounding background. (c) Sulphur from natural sources with a (')34Svalue differing significantly from background can be added to the fuel or to the stack gas before emission. For example, either elemental sulphur or sulphur derived from seawater sulphate with a (')34S near +21%0could be used as an injectable tracer. Alternatively, sulphur with negative (')34S values might be derived from biogenic pyrites. (d) Isotopically enriched sulphur from commercial separation processes could be added to the stack gas as a tracer. From method (a) to method (d) the cost becomes progressively greater. Method (a) requires no change in the plant operation and no additional cost beyond that of sampling and analysis, which is common to any method. For method (b), locating a source of fuel with an appropriate (')34S would add little to the cost but would involve a number of special problems, described below. Method (c) would be more costly because of the processing required to get the sulphur into a form suitable for injection. For method (d), at the time of writing, the cost of isotopic enrichment is very high and supplies are limited. The criteria for a fossil fuel suitable for use as a tracer are as follows: (a) The difference between the (')34Svalue in the fuel and that in the background must be sufficient to overcome variations in the background.

Case Studies and Potential Applications

339

Since in the areas under investigation the background (')34Svaries between 0 and +5%0, a substance with a (')34S value of ::':: 10%0would be sufficient to trace a stack plume out to a distance of about ~20 km. (b) The sulphur content of the fuel should not be so low that dilution by background sources becomes a problem, nor so high that the fuel cannot be used in areas where sulphur emission is restricted. A sulphur content of 1-2% should be suitable for the metropolitan areas. (c) Enough fuel stock should be available for a power plant run of about one day's duration, including storage, startup, switchover, and shutdown. For a 300- MW (e) plant, this could amount to about a million gallons of oil. (d) The quality of the fuel should conform to the specifications used by the power plant. These include, for oil, the viscosity and flash point, and for coal, the grade or rank. A search was initiated for sources of fuel meeting these criteria. Because of the multitude of sources, the search was begun by analysing fuel from oil- and coalfields. The crude fuel could then be traced through the refineries to a supplier for the power plant. Since oil is more likely to be homogeneous in a given deposit than coal, effort was concentrated on finding a suitable fuel oil. Furthermore most coal-burning plants are designed for simple conversion to oil, a suitable fuel oil tracer could be used in both types of plants. With the aid of the American Petroleum Institute, the US Bureau of Mines Petroleum Research Center in BartIesville, Oklahoma, and the Petroleum Research Center in Laramie, Wyoming, a number of crude oil samples from various oilfields around the world were obtained for analysis. Subsequently, several oil, coal, and power companies provided samples of fuels in use. The results of the analyses for (')34S values and sulphur content for crude oil samples are listed in Table 8.15, for fuel oil in Table 8.16, and for coal and coke in Table 8.17. In view of the selection criteria, the following crude oils may be suitable tracer fuels for power plants in the northeastern USA: River Bend, Wyoming (sample PC-66-39), Reno, Wyoming (sample PC-66-48), Woodrow, Montana (sample PC-58-399), and Sho-Vel-Tum, Oklahoma (sample 59171). The Red Wash oils from Utah have interestingly high (')34S values (+ 12.8 and + 16.0), but their low sulphur contents preclude them from further consideration. Two foreign crude oils warranting further investigation are Wafra, Kuwait (sample 67065), and Safoniya, Saudi Arabia (sample 67066). Humble's Esso Refinery oils derived from Venezuelan crude oils have about the same 034S values (+5.0 to +5.8%0) as the New York City background value (+5.0%0) observed in 1968 when these fuels were known to have been used there.

340

Stable Isotopes Table 8.15 The ;)34Svalues in a selection of crude oils Sample Geographical location and source Sulphur content (wt %)
-~--------

;)34S (%0)

Bartlesville Petroleum Research Center California 66096 Wilmington 64075 Summerland Oklahoma 59171 Sho-Vel-Tum 67074 Konawa-Dora Utah 67131 Red Wash Laramie Petroleum Research Center Wyoming PC-66-39 River Bend PC-66-48 Reno PC-67-75 Cowley PC-67-76 Middle Dome PC-67-73 Beaver Creek PC-67-75 Cowley North Dakota PC-58-138a Newburg PC-59-187a Rival Montana PC-58-397a Pennel PC-58-399a Woodrow Nebraska PC-59- 368a Ackman Chevron Oil Company Red Wash crude Number 6 fuel oil (Salt Lake City Ref.) Rangely crude American Oil Company (AMOCO) YSN 6796A Middle East YSN 6976B Middle East 67065 Wafra, Kuwait 67066 Safoniya, Saudi Arabia Reno crude Beaver crude
a

1.51 0.23 1.44 0.44 0.31

+6.8 -7.9 +14.5 +2.5 +12.8

2.52 0.77 3.25 2.58 0.54 3.25 0.77 0.44 0.70 0.90 0.86 (0.2% )

-6.6 -4.5 +6.9 +5.1 -2.5 +2.0

+3.7 +2.6 -4.8 +4.3 +16.0 +1.4 -0.6

2.5 1.4 3.91 2.97

-7.1 +3.2 -9.8 -8.5 -4.8 -3.3

Oils represented BaS04 samples from Parr bomb sulphur determination.

Case Studies and Potential Applications Table 8.16 The &34S values for sources of fuel oils Sample Geographical location and source Sulphur content (wt %)

341

&34S (%0)

Power plants 59-84 RTF-28

Cod Ed (Hess Bunker C) Con Ed (Nepco Bunker C) LILCO (Venezuela residual) English Station N.H. Conn. (Number 6, 4/19) English Station N.H. Conn. (2/21) English Station N.H. Conn. (3/8) Arthur Kill Station S.I. Number 2 United Illuminating English Station (Conn. (10/17/68) English Station (Conn.) (4/19/68) Todd Jones (Conn.) (10/17/68) Wyatt Oil (Conn. (10/17/68) Connecticut Refinery (10/17/68) Metro Oil (Conn.) (10/17/68) 0.22 1.10 0.90 0.80 1.40 2.20 1.05 1.26

+4.9 +2.7 +3.3 +5.4 +3.0 +5.3 +1.4 +5.5 +5.4 +7.4 +6.5 +5.1 +5.8 +11.8 + 12.4 -4.7 -1.7 -2.2 -2.3 +13.7 +12.1

Kerr McGee Oil Company 242 Number 2 diesel oil 242 Number 6 fuel oil Shell Oil Company OSS 48182 Reno Unit, Wyoming (1/29/69) OSS 48183 Cracking crude OSS 48184 Topped crude OSS 48185 Vacuum flashed crude Sun Oil Company, DX Div., Duncan, Oklahoma 37-S-69 Vacuum residual fuel (Sho-Vel37-S-69 Turn) Sho-Vel-Tum grade Humble Oil Refinery Company ESSO F -901 Esso fuels sold in New York area ESSO F-950 ESSO F-954 ESSO F-962 ESSO F-750 ESSO F-958 Delivered to BNL power plant 1969

1.39 0.99 2.26 1.00 1.99

+5.8 +5.3 +5.2 +5.0 -1.4 +7.9

342 Table 8.17 The 034Svalues for coal and coke Sample
Coals AG-297 ER-167 Geographical location and source

Stable Isotopes

034S (%0)

Con Ed, Gauley Eagle Con Ed, Lancashire 24 Cod Ed, Moss 2 Con Ed, Moss 3 Con Ed, Kentland Arthur Kill coal 10/2/68 Arthur Kill coal Ravenswood coal 10/2/68 Astoria coal 10/2/68 V.I. coal 10/17/68 New Haven Power Plant

+2.4 +4.5 +4.1 +4.1 +2.4 +0.2 + 1.8 +0.5 +1.8 +4.7 +1.7

Coke

None of the coal 034Svalues is sufficiently different from the New York and New Haven area background values to be suitable as a tracer fuel. However, the change in oxidation state of sulphur through a coal-fired plant can still be followed by measurement of the 034Svalues. Fuel oil for power plants is usually heavy residual oil (No.6 or Bunker C) remaining after distillation of lighter oils from the crudes. Two problems arise in preparing a tracer: (a) refineries usually process a blend of crude oils to obtain a maximum light-fraction yield and (b) the residual oil might differ in 034S value substantially from that of the crude oil. A run in a refinery with a specified crude would be expensive, though not necessarily out of the question. Alternatively, residual oil can be obtained from refineries running a high proportion of desirable crudes. This was done for the Reno (and Shell Oil Company Unit) and Sho-Vel-Tum (Sun Oil Company, Duncan, Oklahoma) crudes. The Reno crude, with a 034Svalue of -4.7%0, gave a residual oil (designated as cracking, topped, and vacuum flashed crude) with values from -1.7 to -2%0, which indicated that the Reno crude was possibly mixed with other crudes having higher 034S values. The Shell Oil Company stated that on a monthly average the Reno crude comprises only about 2% of the total cracking stock. Based on a steady background 034Svalue of about +5%0 and a measurement reproducibility of 0.2%0, fuel with a 034Svalue of -2%0 would allow tracing of one part of sulphur in the plume with 35 parts of background sulphur. This is sufficient, normally, to

trace plumes out to

10 km.

The residual fuel oil from the Sun Oil Company refinery at Duncan, Oklahoma, has a high positive 034S value, + 12.9%0. This would be very

Case Studies and Potential Applications

343

suitable for areas where background 034Svalues range between 0 and +5%0. Table 8.18 lists fuel oils that could possibly be used for tracer studies in the northeastern USA. 8.4.5 Conclusions A body of data has been obtained for the sulphur isotope composition of fuels that are in use in the northeastern part of the United States. The variation in 034S values with time for emissions from typical oil- and coalfired power plants have been examined. The isotopic composition of S02 was measured for an extensive set of samples collected both at ground level and at elevations within the mixed layer, over a period of eight years at a number of sites in the north-east. There is no significant difference between ground-level and airborne data and the overall average of all results yields a 034S value of +2.4%0 with more than 90% of the values within the rang~.of 0 to +4%0.The isotope ratio of sulphur in sulphate in the atmosphere was found to be significantly more positive than SOz, with an average 034S value of +5.0%0 and a range containing 90% of the values lying between +3 and +8%0. Sources of fuels suitable for use as short-range sulphur isotope tracers for power plant plumes have been identified.
Table 8.18 Possible tracer fuel oils Sulphur content (wt %) Sun Oil Co., DX Division, Duncan, Oklahoma Shell Oil Co., Reno Unit Refinery, Wyoming Kerr McGee Oil Co., Wynnfield Refinery, Oklahoma 1.2 2.2 1.1
1)34S(%0)

+12.9 -2.1 +12.1

8.5

ENVIRONMENTAL SULPHUR ISOTOPE STUDIES IN JAPAN

8.5.1 Introduction*
Japan, consisting of four main and several hundred small islands, is geologically quite young. Its features include mountains, rivers, and heavy vegetation. High population densities exert enormous pressures on the land, and the surrounding sea is a major source of sustenance. Therefore, maintaining the integrity of the environment is very important.
* H.R. Krouse and V.A. Grinenko.

344

Stable Isotopes

From the viewpoint of the sulphur cycle, volcanism and biogenic sulphide constitute significant natural emissions to the atmosphere. Rapid industrial and economic growth since the end of the Second World War have increased anthropogenic sulphur inputs to the environment. Sulphur isotope studies are actively pursued in Japan. Initial facilities were established in the early 1960s by H. Sakai. Therefore, it is not surprising that the potential of sulphur isotopes in environmental investigations was recognized relatively early, as demonstrated by N. Nakai and his colleagues in the following sections. 8.5.2 The quantity and sulphur isotopic composition of hydrogen sulphide released from tidal flats* 8.5.2.1 Sampling Samples were collected at various seashore sites at Mikawa Bay, Japan, from November 1979 to May 1981 in order to examine the quantity and isotopic composition of hydrogen sulphide released from coastal belts containing high anthropogenic organic materials. The HzS was sampled below and above the water line as shown in Figure 8.16. The air, including the released HzS, was circulated for 10 to 60 min at the rate of 0.3-0.5 min-I by means of a 'Mini Pump', trapping HzS into the aqueous solution containing 1 M NaOH and 0.01 M ascorbic acid. The sulphide concentration of the solution was determined in situ by the sulphide ion electrode method, with a detection limit of 0.1 mol -l (corresponding to a release rate of 0.1 mol HzS m-Zh-I). For isotopic analysis, HzS was collected as CdS, then converted to AgzS, and finally to SOz by heating mixtures of CuO and AgzS in a vacuum system. 8.5.2.2 Observations and discussion The observed HzS emission rates observed are presented in Table 8.19. The rates were about 1-2 J.L mol HzS m-zh-I from organic-rich, black, and muddy tidal flats (organic carbon contents of sediments, 0.2 to 2.2%), and less than 0.1 J.Lmol HzS m -zh -I from organic-poor, white, and sandy flats. On one tidal flat particularly enriched in organic matter (12.4% in sediments), the emission rate was observed to be quite high, 23 mmol HzS m-zh-l. In Figure 8.17, the average diurnal variation in the emission of HzS is shown for the Gamagori tidal flat. It is obvious that HzS emission increased up to 1 mmol m-zh-I during the ebb tide. The seasonal variation of 834S values for both HzS released to air and the SOi- remaining in pore water of the sediment are presented in Table
* N. Nakai, Y Tsuji, and T. Yatsumimi.

Case Studies and Potential Applications MiniPump

345

Sea Water Bottom


(0)

Sediment

Mini Pump

~.~I~

(b)

Figure 8.16 Schematic diagram of the apparatus for capture of H2S released from tidal flats

Table 8.19 Average and maximum emission rates of hydrogen sulphide Locality Average H2S emiSSIOn (fLmol m-2 h-I) 1.4 1.6 76 <0.1" 0.9 <0.1" Maximum H2S emission (fLmol m-2 h-I) Number of samples

Nishiura Takeshima Gamagohri Ohtsuka Umeyabu Shiraya

3.8 2.1 23 000 5.4 6.6 <0.1"

4 2 61 4 4 3

a b.d., below detection limit < 0.1 J.l.molm-2 h-'.

8.20 and Figure 8.18. The 834Svalues of H2S vary widely (-21.9 to + 12.0%0), being most negative in winter and most positive in summer to autumn. This variation of 834Svalues with season may be caused by temperature-dependent variable SOl- reduction rates. There is also a difference in the 834Svalues

346
10.2

Stable Isotopes

~:.c
N

10-3

, E 100' U1 N J:
10-5

"0 1006

E
10-7
I I

06

I- I10 12

06

lit

16

I- I- I18 20

22

'-

I04

24

02

06

h
Figure 8.17 Average diurnal variation in the emission rate of hydrogen sulphide at Gamagohri tidal flat in winter, 1979-80. The points represent statistical analyses of many observations over longer time spans

Table 8.20 The 834Sof H2S released from a tidal flat and of SOlof sediment (Gamagori tidal flat) Sampling dateil Temperature (C) 834S (%0) H2S released 18 Nov. 1979 (N) 22 Nov. 1979 (N) 23 Nov. 1979 (D) 14 Dec. 1979 (D) 11 Dec. 1979 (N) 26 July 1980 (N) 27 July 1980 (N) 8 Sept. 1980 (D) 8 Sept. 1980 (N) 23 Oct. 1980 (N) 24 Oct. 1980 (D) 18 Mar. 1981 (D) 4 May 1980 (N) 17 14 15.5 14 11 22.6 27 25.2 24.5 15.5 18 8 12.6 -6.3 -10.1 -9.3 -17.9, -19.5, -1.8 +5.8 +11.2, +8.3, -1.0 +9.5 -9.3 +8.5,

in pore water

SOl+23.6 +23.3 +23.9, +23.8 +24.0 +23.7 +25.8 +37.2 +31.5 +24.0 +24.1 +26.1, +26.6

-16.4 -21.9 +11.5 +7.7

+ 12.0

D, sampled at day time; N, sampled at night.

Case Studies and Potential Applications

347

+30 +20

--0
cf!. + 10 ---(j) 'it !()

H2S released to atmosphere ,8- '.

0 -10 -20

. ,~ /'

'

---~

'.
\ \

,\

ro

of HzS released in the day and night (Figure 8.18). The mean annual 034S value of released HzS is probably close to 0%0because the emission rates are higher in summer to autumn than they are in winter. The 034S values of SO/- in pore water of the sediment also indicate a seasonal dependence which can be attributed to SO/- reduction. 8.5.3 Secondary production in sulphuric acid from volcanic ash in surface water after a volcanic eruption* 8.5.3.1 Introduction Mt Ontake, which is situated about 80 km north of Nagoya in Central Japan, erupted for the first time in human history on 28 October 1979. Volcanic ash fell steadily for several days around Mt Ontake, accumulating not only on land but at the bottoms of rivers and lakes. After the eruption ended, the ash remaining at the bottom of rivers and lakes produced sulphuric acid due to bacterial oxidation of indigeneous sulphide minerals. This can be called the 'secondary disaster' of the volcanic eruption.
* N. Nakai.

348 8.5.3.2 Results and discussion

Stable Isotopes

As shown in Table 8.21, the water-soluble sulphate and pyrite (FeSz) in fresh ash collected from the volcano area have &34Svalues of + 14.5 to + 14.7%0and -6.0 to -5.6%0, respectively. This isotopic difference between SOi- and pyrite can be interpreted in terms of isotopic exchange equilibrium under hydrothermal conditions. Because it is water soluble, sulphate could not be detected in ashes suspended in river water or piled on the bottom sediments. The volcanic ash was found to contain a high concentration of sulphide minerals (Table 8.21), and because of this it was expected that sulphuric acid would be produced, resulting in the acidification of river and lake water. Laboratory experiments on the bacterial oxidation of sulphide to sulphate were performed using the volcanic ash collected from the bottom of a river and stream water containing a negligible amount of SOi-. Fifteen grams of wet volcanic ash and 50 ml of stream water were placed in several flasks and kept at 25 e. Sulphate contents and &34Svalues, together with pH, were measured at appropriate intervals. The results, indicated in Figure 8.19, show that SOi- (sulphuric acid) was produced rapidly, with an almost linear relation between SOi- concentration and incubation time. The S04Z- concentration reached 3.0 g -1 (0.03 M) after 27 days and pH decreased to 2.5. The isotopic composition of SOi-, however, shows constant values of about -5.0%0, approximately equal to that of pyrite initially contained in the ash. These experiments show that if SOi- is produced from the oxidation of ash pyrite, its &34Svalue should be much less than that for sulphate originally contained in the volcanic ash. Figure 8.20 shows sampling locations of river and lake water. Figure 8.21 shows pH value variations before and after the volcanic eruption for inflow
Table 8.21 Sulphur compounds and (')34S values for the volcanic ash erupted from Mt Ontake on 28 October 1979 Sampling location Soluble S042Content (%) Kaidamura village, site 1 Kaidamura village, site 2 The mid-slope of Mt Ontake Mitake (Nakanoyu) Suspended ashes in Nigorigawa River 1.28 1.41 1.57 1.94 0.00 (')34S (%0) +14.6 +14.5 +14.7 +14.7 Pyrite Content (%) 5.03 4.99 5.40 6.19 6.06
(')34S

(%0) -5.6 -5.9 -6.0 -6.0 -5.9

Native S content (%) 0.75

0.81

Case Studies and Potential Applications

349

160 5042120
I

E 0 10 0> 80 E
I NO
(/)

8 345,(%0)

40
8345

-3

0---5

--0

-6
10 15 20 25 30 Incubation Time (days) Figure 8.19 H2SO4-production experimentsusingvolcanicash from Mt Ontake and river water
N

\ \ 00

\_~--

P
__I

/ /

/
Makio Dam/I'

"*

river water (Station 5) to the dammed lake and the lake bottom water (60 m in depth) (Station 9). After the eruption on 28 October 1979 pH values decreased from 6.5 to 4.5 within two months. Subsequently the bottom water of the lake became progressively acidified. The pH value reached 3.4 in September 1980 and all the living organisms died.

350
Eruption of 7

Stable Isotopes

Mt Ontake

~,

~,

- . \
""x

X I ' "-

/
\

~5
4
3 2

-'. x

"

'"

x ...

,x""'"

""

x- St.5 --x

\ x/

Oct. Nov. Dec.' Jan. Feb. Mar. Apr. May Jun. Jut. Aug. Sept. Oct. -1979 -11980 Figure 8.21 Decrease in pH at stations 5 and 9 (see Figure 8.20) after the eruption of Mt Ontake

In Figure 8.22a and b, SOi- contents and the 834Svalues for stations 5 to 9 are plotted from December 1979 to August 1980. For each station, SOi- contents of water increased gradually and reached 90 mg -1, which is unusually high for fresh water. On the other hand, the 834S values decreased unidirectionally during this time. In August 1980, SOi- at Station 8 had 834Svalues of -5.0 and -4.8%0, which are close to that of the original pyrite in the volcanic ash and also to that produced by SOi- in the laboratory oxidation experiments. Thus it is clear that the SOi- which caused the acidification of water and the extinction of living organisms originated from the bacterial oxidaton of pyrite contained primarily in volcanic ashes-the 'secondary disaster'. As described above, SOi- having a constant 834Svalue of about -5.0%0 is supplied to river and lake waters. Further, the 834S values of river and lake waters are known for the different stations at various times after the volcanic eruption. Therefore the contribution of SOi- added to the natural SOi- in the waters can be estimated by using measured 834S values. This contribution was calculated as follows:
A 834S2 + C 834Sp = 834Sh A + C = 1.0

Case Studies and Potential Applications

351

100

13December1979
80

24March1980

11June 1980

10 July 1980

7 August 1980
I I I I

......

'0 en 40

'a,

60
I I I I I I I

.
A
I I I

,I

. :
I I

A~
20 0 (a) +10

.Y

.~

-.. "..

"" A-"'/ '

~
I A'

#<>0

+5

,'

'

'

'

.
'

Jr-.

L
-

~
I I

A_

-5 (b)

~
_",

",

,,"

-0 ~

100 80
60

r: 0
en i::' as "tI

en
() Q)

40
20

~
L...& L...JJ
56789
Water flow (c)

A--

. - ,, ~

A_-fI
\ \ \ \

~
/
I I

' "

'.
. I .
56789

.
} Lake

0 Surface water 0 Bottom water L..& & J. ... River water I I I

01 I I I I I Station 5 6 7 8 9

..

56789

56789

Figure 8.22 Temporal variations of solution parameters at stations 5 to 9 from 13 December 1979 to 7 August 1980: (a) SO/- content, (b) 834S of SO/-, (c) contribution of secondary SO/- (%).

352 where
;)34Sa

Stable Isotopes

;)34Svalue of naturally occurring S04Z-. This is equal to + 13.3%0, as observed in the upper reaches of the river ;)34S value of SOiat the station where secondary S04Z- contribution is to be estimated ;)34Svalue of secondary SOi- produced by bacterial oxidation of sulphide. This is equal to -5.0%0 fractions of natural and secondary SOirespectively the

;)34Sb

;)34Sp

A and C

The results obtained from the above calculation are shown in Figure 8.22c. The contribution of secondary SO i- increases from the upper reaches to lower reaches of the river including the dammed lake. At each station, the contribution increased continuously, reaching 100% at Station 9 (bottom water) in August of 1980. It is apparent that a large amount of sulphuric acid was produced in the water system. Even in 1983, the river and lake were still acidic with pH values lower than 5.0, especially in bottom lake waters. 8.5.4 Sources of atmospheric sulphur compounds based on the sulphur isotopic composition of SOi- in precipitation in Japan, 1960-79* 8.5.4.1 Introduction Following the Second World War, Japan has gradually been industrialized and by 1965 air pollution by sulphur compounds had become a major environmental problem. Tanaka (1978) reported the concentrations of SOz and HzS (Table 8.22) and pointed out marked differences in the concentrations of SOz for urban and rural areas. Japan should represent a relatively simple case study since it is surrounded by a marine environment. It should be possible to use ;)34S values to determine the relative contributions of industrial and oceanic SOi-. Since 1960, samples of S04Z- in rain and snow have been collected in urban and rural areas of Japan and ;)34Svalues have been determined on samples collected between 1960 and 1979.

* N. Nakai, Y. Tsuji, and U. Takeuchi.

Case Studies and Potential Applications

353

Table 8.22 SOz and H2S concentration in the atmosphere of Japan (1975, 1978) (after Tanaka, 1978) Location Urban Yokohama (summer, 1975) Yokohama (winter, 1975) Omuta (summer, 1977) Nagoya (summer, 1978) Nagoya (winter, 1978) Rural Niitsu (autumn, 1976) Arao (summer, 1977) HzS (ppb) S02 (ppb) S02/H2S

2.7 0.87 2.5 1.1 2.5 2.3

11.7 43.6 36.0 17.7 49.3 4.9 5.2

4.3 50.1 7.1 44.8 2.0 2.3

8.5.4.2 Atmospheric 50z and Hz5 and pH of precipitation in Japan In order to estimate the extent of rain and snow S pollution in Japan, especially in industrial and urban areas, pH values and SO/- contents were determined for rain water collected on 4 and 5 July 1974 in Kanto plains around Tokyo and Tokyo-Yokohama Industrial sites (Table 8.23). On these dates, acid rains over the area had the lowest pH values recorded, ranging from 3.0 to 3.9 for all stations.
Table 8.23 pH values and S042- contents of rain waters taken on 4 and 5 July 1974 in the Kanto plains around Tokyo Sampling location Tokyo Ota-ku Bunkyo-ku Chofu City Musashino City Mitaka City Kanagawa Pref. Yokohama City Kawasaki City Gunma Pref. Tatebayashi City Ibaragi Pref. Furukawa City pH SO/(ppm)

3.9 3.5 3.0 3.3 3.8 3.0-4.4 3.7 3.6 3.9

16.7 22.6 22.6 14.9 16.7

27.0

354

Stable Isotopes

In Figure 8.23, a frequency distribution of pH for the wholeyear is shown


by a histogram (%) for the industrial and urban cities, Tokyo (1973) and Nagoya (1976), and the rural city, Sendai (1973). The pH values for Tokyo and Nagoya were lower than those for Sendai. Precipitation had pH values in the range 4.0-4.9 for 50-75% of the time. Surprisingly, values in the range 3.0-3.9 could be found with a probability of 10-18% in Tokyo and Nagoya. In Table 8.24, SOi-/Na+ and SOi-/Cl- ratios for rain and snow samples from Tokyo and Nagoya are compared to those of sea water, which is an

80 70 60
Urban Rural

~ D

NAGOYA. 1976
TOKYO, 1973 SENDAI, 1973

*
()
::::I

, 50 >-

~ 40
cO)

It 30 20 10 0
3.0-3.9 4.0-4.9 5.0~5.9 6.0-6.9 7.0-7.9 pH Figure 8.23 Frequency distribution of pH values for rain and snow in Japan Table 8.24 Annual mean SO/- /Na+ and SO/-/Cl- weight ratios of rain and snow samples taken in urban areas of Japan Sampling location Nagoya (1946, 1947) Nagoya (1973) Tokyo (1971) Tokyo (1973) Tokyo (1978) Sea water SOl-/Na+ 0.82 10.7 5.2 12.3 3.8 0.25 S042- /Cl0.47 3.4 2.7 5.4 1.6 0.14

Case Studies and Potential Applications

355

important source of atmospheric compounds. The data indicate that, since 1971, these ratios in urban areas are about one to two orders of magnitude higher than in sea water. Furthermore, for the period 1971-78, they are much higher than in 1946-7 when Japanese industrial activity was at its lowest just after the Second World War. This suggests that after the War, anthropogenic S04Z- in rain waters increased with Japanese industrial growth. 8.5.4.3 Sulphur isotopic composition of SO/in rain and snow in Japan

Figure 8.24 shows sampling locations for isotopic studies of rain and snow taken from 1960 to 1979. The isotopic data for S04Z- in precipitation are compared to seawater SO/-, atmospheric SOz, and aerosol SO/- in Figure 8.25. In Table 8.25, the 534S values of the SO/- are presented as yearly ranges and averages from 1960 to 1979. As seen in Figure 8.25, even in 1960, two sources of rain SO/- could be ascertained by sulphur isotope data. One had 534S values ranging from +3.2 to +7.3%0 whereas the other varied from +12.3 to +15.6%0. The former belongs to samples from industrial and urban sites and the latter to

~~~ OX'::,'?'
'?><v~

Tokyo
~

- Kelhin Industrial Site

Kurume Co

Osaka

V
Nagoya
Nara ChukyoIndustrial Site OU

~
~~

Co

g J...

$"

<J

~
.

fP

Wakayama

Hanshin Industrial . Site

Kitakyushu Industrial Site

<l..~

f>

Figure 8.24 Rain and snow sampling locations in Japan, 1960-79

356

Stable Isotopes

-10
1960
1971-72
1973

r---I

0
I

TOKYO NAGOYA
OSAKA-KOBE

TOKYO
NAGOYA TOKYO
NAGOYA

1974-75 1978
1979

NAGOYA TOKYO
NAGOYA TOKYO NAGOYA

m a:
::>

t--+--I
~ ........... ~

I. ~

..
I.

-.. . ...
I
I

+10

+20

I I
S042-

1960

TOTTORI
MATSUMOTO

KURUME 1971-72 TOTTORI KURUME


AKITA

-oJ

1973

KURUME
AKITA SUWA
ONTAKE

a:
::>

a:
~

..
r~ I

.
f

.. .. .

. . .

..

1975-76

ACHI

1975-76

NAGOYA

Aerosol
S02

1971-77

OSAKA KASUGAI NAGOYA


WAKAYAMA

1+1 1-8-1 1-8-4

Sea Water. .
I

-10

+10
834S (0/00)

+20

Figure 8.25 Isotope composition of atmospheric sulphur in urban and rural locations in Japan

samples from non-industrial and rural sites. Since 1960, the 0345 values decreased in industrial and urban sites and showed a minima of -1.3 to +4.2%0 in 1971 (Tokyo) and -1.0 to +2.7%0 in 1973 (Nagoya). The yearly lowering of 0345 values may indicate an increase in the use of fossil fuels for industrial activity and automobiles. The authors also sampled rain and snow in the United States in 1960. These samples also had two ranges of isotopic composition: +5.9 to +6.2%0 for industrial sites in New Haven, Connecticut, and + 18.9 and + 19.0%0for rural sites in Pittsfield, Massachusetts.

Case Studies and Potential Applications

357

Table 8.25 The 034S values of rain and snow S042- taken from 1960 to 1979 in Japan Sampling location Urban 1960 Tokyo Nagoya Osaka-Kobe 1971-2 Tokyo Nagoya 1973 Tokyo Nagoya 1978 Tokyo Nagoya 1979 Tokyo Nagoya Rural 1960 Tottori Matsumoto Kurume 1971-2 Tottori Kurume Akita 1973 Kurume Akita Suwa Ontake 1975-6 Achi 034S (%0) Annual mean 034S (%0)

+5.0 to +7.3 +3.2 to +5.9 +6.3 to +7.3 -1.3 to +4.2 -1.1 to +4.0 -1.0 to +4.9 -1.0 to +2.7 +0.8 to +5.5 +0.5 to +5.0 + 1.3 to +4.8 +2.0 to +5.1

+6.5 +4.5 +6.8 + 1.0 (1971), + 1.3 (1972) +2.1 (1971), + 1.7 (1972) +1.2 +0.5 +2.5 +2.3 +2.6 +2.8

+13.2 to +15.1 +12.3 + 12.8 to + 15.6 +12.0 +11.7 to +12.3 +11.1 to +12.1 +11.8 to +12.0 +5.5 +0.1 +0.4 -0.8 to +3.6

+14.2 +12.3 +14.7 +12.0 +12.0 +11.9 +11.9 +5.5 +0.1 +0.4 +2.1

All samples from both industrial and rural sites indicate that rain and snow SOi- are depleted in 34S compared with the 834S values of the seawater S042-. Even if some of the rainwater S042- is marine in origin, additional sources of S042- depleted in 34Sare required. Further, industrial and urban sites require additional sources of SOi- more depleted in 34S than at rural sites. It is clear that this additional 34S-depleted source is anthropogenic sulphur emissions in the form of S02 which is oxidized to SOi-.

358

Stable Isotopes

In order to determine the 534Sof anthropogenic SOl in the atmosphere, the Pb02 collection method was used at industrial sites from 1971 to 1977. Samples from Osaka, Kasugai, Nagoya, and Wakayama were found to have an average (')34S value of -4.0%0 (Figure 8.25). On the other hand, estimates of the anthropogenic (')34S value based on analyses of coal and oil samples have a relatively large spread. During 1971-2, the Japanese government regulated SOx and NOx emissions. After 1971 and 1973, (')34S values increased in Tokyo (+ 1.3 to +4.8%0 by 1979) and Nagoya (+2.0 to +5.1%0 by 1979). This is consistent with decreasing relative contributions of anthropogenic sulphur to the atmosphere. In contrast, with decreasing anthropogenic pollution in industrial and urban sites, the extension of pollution to rural and non-industrial sites can be clearly seen from the (')34S values after 1973, which reached +0.1 to +0.4%0 for Akita, Suwa, and Ontake in 1973 and an average of +2.1%0 for Achi during 1975. 8.5.5 Estimation of the contribution of anthropogenic sulphur to the atmosphere in Japan* Various sources for SOiin the atmosphere over Japan have been suggested: (a) sea spray S042-, (b) biogenic sulphur, i.e. S042- resulting from the atmospheric oxidation of H2S primarily produced by anaerobic bacteria in coastal belts, marshes, soil, and lakes, (c) anthropogenic sulphur, i.e. SOi- resulting from atmospheric oxdation of anthropogenic S02, and (d) volcanic sulphur, consisting of S02, H2S, etc. Of the above-mentioned sources, volcanic sulphur is comparatively insignificant. If the (')34S values for each individual source are known, it is possible to calculate the contribution of each to the atmosphere. Estimated (')34S values for three sources are as follows: (a) Sea spray SOi-, (')34Ss = +20.3%0 (b) Biogenic H2S (divided into two sources, coastal H2S and inland H2S): (i) Coastal H2S, (')34Sc= -2.6%0. This value was estimated on the assumption that H2S is produced at coastal flats by bacterial reduction (kinetic effect k32/k34 = 1.023for the followingreactions of seawater S042- with a constant (')34Sof +20.3%0 in an open system)
k32 32S042k34 7 H232S

34S0/* N. Nakai, Y. Tsuji, and V. Takeuchi.

3> Hl4S

Case Studies and Potential Applications

359

followed by the oxidation of HzS to sulphate in the atmosphere without isotopic fractionation. (ii) Land HzS, j)34Se = +6.0%0. The land HzS is released to the atmosphere through the decomposition of sulphur-containing organic materials. By assuming that no isotopic fractionation occurs in the decomposition process, the j)34Svalues can be estimated from those of organic sulphur. Combining (i) and (ii) with the emission ratio of coastal/land HzS of 3:7 (Robinson and Robbins, 1970; Eriksson, 1963), the j)34S value of total natural HzS origins is estimated to be +3.4%0. (c) Anthropogenic SOz, j)34Sa = -4.0%0. As described in the previous section, the value can be estimated from observed data. It is assumed that no isotopic fractionation occurs in the oxidation of SOz to SOl-. The above estimates are summarized as follows: (a) (b) (i) (b) (ii) (c) Land HzS Coastal HzS Sea spray Anthropogenic SOlj)34Sc = -2.6%0 j)34Sf = +6.0%0 SOz j)34Ss = +20.3%0 j)34Sa = -4.0%0 Coastal HzS/land HzS = 3:7 (b) Natural HzS
j)34Sn =

+3.4%0

The contributions of the individual sources can be calculated for 'polluted sites' by the following equations. For polluted rain and snow SOl-:
j)34Sp

= a

j)34Ss

+ b j)34Sn + C j)34Sa

(1) (2)

a + b + c = 1.0

where a b c

= fraction of sea spray SOl= fraction of SOl- from oxidation of natural HzS = fraction of anthropogenic SOl-

sites' .

In order to estimate the value of c, it is necessary to know the value of bfa, the ratio of natural HzS to sea spray SOl-. This is estimated to be 0.58 on the basis of the j)34Svalues of rain and snow SOl- at 'unpolluted

360

Stable Isotopes

Table 8.26 Estimates of the contributions from various sources to rain and snow SO/Contribution (%) Sampling location Annual mean OJ4S(0/00) Coastal and land HzS
b

Sea u spray

Anthropogenic
c

1960 Tokyo Nagoya Osaka-Kobe 1971 Tokyo Nagoya 1972 Tokyo Nagoya 1973 Tokyo Nagoya 1978 Tokyo Nagoya 1979 Tokyo Nagoya

+6.5 +4.5 +6.8 +1.0 +2.1 +1.3 +1.7 +1.2 +0.5 +2.5 +2.3 +2.6 +2.8

34 27 35 16 20 17 19 17 15 21 20 21 22
I

24 20 25 12 14 12 13 12 10 15 15 15 16

42 53 40 72 66 71 68 71 75 64 65 64 62
I

0 8

Nagoya Tokyo

"
2 ~ ...... ~
,/ ,/

--'",,,

"

saka-Kobe

if
0 1960
1970 1971 19721973 1 1978 1979

Figure 8.26 The anthropogenic to natural sulphur ratio, 1960--79, in industrial and urban sites of Japan

Case Studies and Potential Applications

361

For unpolluted rain and snow SO i-,


i)34Sup = ao i)34Ss + bo i)34Sn ao + bo = 1. 0

(3) (4)

where ao = fraction derived from sea spray bo = fraction derived from natural H2S If it is assumed that SOi- samples collected in non-industrial and rural sites in 1960 were not affected by anthropogenic SOb the average i)34Svalue of + 14.0%0substituted into i)34Ssand i)34Sngives a calculated b(/ao value of 0.58. It is reasonable to assume that b/a = b(/ao = 0.58 (5)

Finally, by combining equations (1), (2), and (5) and substituting the estimated i)34S,i)34Sn,and i)34Savalues and measured i)34SSOi- (p) values into the equations, the percentage contribution of each of the three sources a, b, and c can be calculated as presented in Table 8.26. Furthermore, 'anthropogenic/natural sulphur' ratios (c/(a + b can be calculated from Table 8.26 and plotted for industrial and urban sites (Figure 8.26). As seen in Table 8.26, the contribution of anthropogenic S02 was 40-50% of the total atmospheric sulphur at industrial and urban sites in Japan in 1960. The values then increased up to 75% in 1973, corresponding to the maximum air pollution. After 1973, the pollution has been reduced to between 60 and 65%. As described above, an attempt has been made to calculate the relative contributions of anthropogenic S02, sea spray SO i-, and natural H2S to the atmosphere by using i)34Svalues of rain and snow SOi-. Although there is still some uncertainty in the i)34Svalue of natural H2S, the estimated contribution of anthropogenic S02 (c value) is probably a valid indicator of annual variations. 8.6 THE USE OF SULPHUR AND OTHER STABLE ISOTOPES IN ENVIRONMENTAL STUDIES OF REGIONAL GROUNDWATER FLOW AND SULPHATE MINERAL FORMATION IN KUWAIT* 8.6.1 Introduction The State of Kuwait occupies part of the sandy and gravelly desert plain of eastern Arabia (Figure 8.27). This plain slopes towards the north-east and

B.W. Robinson, M.K. Stewart, and A. Gunatilaka.

362

Stable Isotopes

\0

20

30 Km. /

Figure 8.27 Kuwait topography and drainage together with the locations of the aquifer and sabkha water samples. For sample legend, see Figure 8.30

is broken only by low hills, depressions, escarpments, and shallow wadis. Elevations range up from sea level to almost 300 m in the south-west. Wadi Al Batin is a broad, shallow depression (6 km x 50 m) marking the western boundary of the state. There are no perennial rivers or streams in Kuwait and the wadis and channels carry only occasional runoff. The drainage is mainly internal: sometimes into large depressions such as Raudhatain and Umm Al Aish, where perched lenses of fresh groundwater occur, and sometimes into playas. Rainfall in Kuwait is characterized by small amounts (average 100 mm y-l) and much variability (40-240 mm) since most of it falls during intermittent thunderstorms from November to April. In this desert climate the average winter lows are 9 C and summer highs are 45C. High evaporation and low rainfall formed a continental getch layer or 'duricrust' of gypsum and calcite within recent eolian and alluvial sand and gravel of the desert. Also, in the coastal areas of Sabiyah and Al Khiran

Case Studies and Potential Applications

363

supratidal salt flats occur on carbonate and clastic muds respectively. These sabkha areas feature saline ground waters (up to 300%0TDS), algal mats, and the formation of evaporite minerals such as gypsum and halite. Water wells in other parts of Kuwait contain brackish ground water with 2-10%0 TDS. Before the effects of anthropogenic pollution can be assessed in this area, a framework of background data on the waters and their dissolved constituents must be established. Oxygen and deuterium isotopic compositions of water and oxygen, and sulphur isotopic compositions of sulphate were measured to determine the origin of the water and sulphate and develop a regional model for the formation of the sabkhas. 8.6.2 Geological setting Wadi channel and dune sands occurring in the area are Recent to Pleistocene in age. Finer grained sediments were deposited during the last marine transgression (maximum sea level 4000 years ago). These carbonate muds and mixed carbonate-quartz sands host the sabkhas, whereas the desert sands host the duricrusts of gypsum and calcite (getch layer and caliche). Fossil getch layers and caliche are found in the underlying Kuwait group sediments. The stratigraphy of this group and the underlying Hasa group is summarized in Table 8.27. These rocks dip gently to the north-east (Figure 8.28) and form part of the Arabian Gulf synclinorium, a vast thickness of sediments deposited on a foreland shelf of the Arabian shield to the south and west. The hydrology of Kuwait is shown in Fig. 8.28. A regional groundwater flow from west to east is observed together with an increase in salinity to the east. Supporting perched freshwater lenses at Raudhatain and Umm Al

SAUDI

ARABIA

KUW AIT

IRAN

7tf 1

11
"":""Kuwait Group ~ ""::" Dammam

~ Precipitation
-Groundwater flow

j Evaporation

":""" Formation - 7 Rus -Formation Umm er Radhuma Formation

Stagnant saline groundwater

Figure 8.28 Schematic geological cross-section of Kuwait showing groundwater flow

W 0\ .j::>.

Table 8.27 Kuwait stratigraphy

Age

Group

Formation

Dominant lithology

Range of thickness

>. .... '" I': ....


<])

Recent

Marine, eolian, and fluvial deposits

Sands, silts, clays, and gravels Up to 110 m

;:

Pleistocene

0 ....
00

0;:

Pliocene

......

'c;j ~ ;: ~

Miocene

>. ....
.g ....
E-<
<])

Oligocene Eocene

;: 0 .... 00

0'" <J>

Palaeocene

'" ::t

Gravels and sands, mainly conglomerate sandstone, siltstone, shale Lower Fars Calcareous sandstone, fossiliferous limestone, gypsiferous Ghar Quartzose sandstone, some shale in lower part Erosional unconformity Dammam Discontinuous chert cap, chalky and siliceous limestone, dolomite Rus Anhydrite, limestone, marl Umm er Radhuma Marly limestone, dolomite, anhydrite

Dibdibba

110m

1-280 m

185-215 m

75-125 m 185-430 m

v,

<:; CJ~ <:;c

~ '" "'

Case Studies and Potential Applications

365

Aish, the Kuwait group of porous sands, silts, and gravels can be an excellent aquifer. Also the Dammam group is a good aquifer due to jointing, faulting, and karstification. Considerable leakage may take place between the two aquifers. The Dammam dome outcrops in Saudi Arabia to the west of Kuwait. It has a present-day areal outcrop of about 100 km2 and dips to the north-east to reach a depth of 300 m below sea level in northern Kuwait. This aquifer underlies the whole of Kuwait and is underlain by the Rus formation which is an effective aquiclude. No leakage takes place from the lower Umm er Radhuma formation into the Dammam formation. The Dammam aquifer is the main water source for the brackish water supplies in Kuwait. 8.6.3 Sampling and analytical techniques Water and mineral samples were collected as summarized in Table 8.28. All water samples were analysed for their oxygen and deuterium isotopic compositions by standard techniques (Hulston et al., 1981). Separate I-litre water samples were collected and sterilized for all of the above except rain water, and the sulphate was precipitated in BaS04. The gypsum and anhydrite samples were also converted to BaS04' These BaS04 samples were analysed for their oxygen and sulphur isotopic compositions using the techniques given by Robinson (1976). The isotopic compositions of all the samples are shown in Table 8.28 and the sample locations are given in Figures 8.27 and 8.29. 8.6.4 Isotopic composition of the waters The isotopic compositions of the waters are shown in Figure 8.30. Included on this plot are data from Kuwait group and Dammam formation waters from western Kuwait (Abusada, 1981) and Dammam and Neogene aquifer waters in Saudi Arabia (Shampine et al., 1978). The rainfall analyses plot close to a meteoric water line (oD = 80180 + 15). Fresh waters at Raudhatain and Umm Al Aish also lie along this line, indicating that they are essentially rain water. There appears to have been little or no mixing with Dammam waters and little or no evaporation. The latter fact appears very surprising in the desert environment until it is considered that the rainfall is very heavy for short periods of time. Hence, water that recharges the aquifer does so quickly in depressions and wadi bottoms and will be little changed by evaporation compared to stagnant water and water that only penetrates to shallow levels. The Raudhatain and Umm Al Aish waters are also close in isotopic composition to waters sampled near the outcrop of their aquifers in Saudi Arabia (Shampine et at., 1978). Tritium analyses of a few Neogene aquifer waters in Kuwait

366

Stable Isotopes

Table 8.28 Oxygenand deuteriumisotopic compositions of Kuwait waters and


sulphur and oxygen isotopic compositions of dissolved sulphate and sulphate minerals Date Sample Water Sulphate lKO (%0) --

IHO (%0) D (%0) 34S (%0) Rain water Winter 1982/3 Winter 1982/3 Winter 1982/3 October 1982

Wadi Al Batin Khaldiya Al Khiran Southern Al Khiran

-0.5 -2.9 -2.2 -0.9

+10.3 -6.0 -7.9 +2.9

Freshwater wells (0.4-2%0 TDS; SOi-, Cl-, HC03) Raudhatain November 1981 Well R1 November 1981 R3 November 1981 R5 November 1981 R6 November 1981 R9 November 1981 R16 November 1981 R20 November 1981 R27 November 1981 R58 November 1981 R61 November 1981 R64 Umm AI Aish November 1981 Well U20 November 1981 U22 November 1981 U54 Gulf sea water (45%0 TDS) December 1981 Pits in Sabkha Muds (0.5-2 m; 20-300 TDS) Sabiyah S1 May 1982 S2 May 1982 S3 May 1982 Al Khiran December 1981 BIO December 1981 B9

-3.5 -3.2 -3.6 -3.2 -3.4 -3.3 -4.1 -3.7 -3.6 -3.2 -3.3 -3.3 -3.0 -2.5

-9.5 -8.1 -14.4 -9.3 -10.5 -11.5 -11.1 -12.4 -11.5 -11.4 -10.6 -13.0 -10.5 -6.1

+16.0 +15.7 +15.3 +15.6 +15.8 +15.7 +15.5 +15.5 +15.8 + 15.4 + 15.4 + 15.3 +15.5

+14.6 + 15.1 + 14.4 + 14.3 +14.9 + 15.0 +14.5 +15.8 + 14.0 +15.2 +14.0 +13.6 +13.7 +13.8

+2.4

+14.7

+20.9

+10.4

+0.8 +6.5 +6.2 +3.1 +6.1

-11.1 +21.0 +21.0 +16.1 +23.1

+14.9 + 19.4 + 19.7 +21.3 +20.2

+13.2 +9.8 +8.5 +11.0 + 10.1

Case Studies and Potential Applications

367

Table 8.28 Continued Date Sample Water


180 (%0)

Sulphate 180 (%0) +10.8 +14.3 +13.8 +13.6 +13.6 +10.5 +12.8 +13.1 +13.4 +14.6 + 13.3 +13.7 +13.9 +12.9 +13.4

D (%0) 34S(%0) +21.2 -8.8 -9.1 -15.0 -12.4 +4.6 -7.5 -9.1 -7.6 -16.7 -15.2 -16.2 -16.1 -14.8 -3.3 -8.2 +18.9 +17.6 +17.2 +15.8 +14.6 +17.0 +13.8 +14.0 +13.4 +13.7 +13.8 +13.5 +13.8 +12.9

October 1982 December 1981 October 1982 December 1981 October 1982 December 1982 December 1982 December 1982 December 1982 April 1982 October 1982 April 1982 April 1982 October 1982 April 1982 April 1982

B9 B-1 B-1 B-8 B-8 HZ K2 Kl K-l Cl Cl C2 C3 C3 C4 C5

+6.4 +1.2 +0.6 +0.2 +0.7 +4.2 +0.8 +0.8 +0.3 -0.5 -0.8 -0.9 -1.0 -0.8 +0.9 -0.3

Wafra Brackish Water wells (4-90/00 TDS; CI-, SOi-) C6 April 1982 C7 April 1982 Sulphate minerals December 1982 Rus anhydrite Wadi AI Batin getch

-1.9 -1.9 -

-13.9 -14.1 -

+14.9 +14.6 +18.6 +15.6

+14.5 +14.3 +14.2 +15.6

Sulphate minerals from Al Khiran sabkha 1980 A, primary discoid gypsum 1980 B, anhydrite (cottage cheese form) 1980 C, bassanite 1980 D, alabastine gypsum 1980 E, anhydrite 1980 F, anhydrite 1980 G, cemented gypsum crust

+15.8 +17.4 +12.1 +15.3 +15.8 +14.0 +16.2

+14.9 +13.9 +14.6 +14.5 +14.2 +13.8 +15.6

368

Stable Isotopes

B9, B10 Channel

Oolitic

Limestone Holocene Pleistocene

~ III
~

Carbonate Sediments (mud) Mixed Carbonate - Quartz Sand and Dune Sands -

Lillill Wadi Channel


Tertiary

:Jm o~

0 "5001000

Bedrock

Figure 8.29 An east-west cross-section of the Al Khiran sabkha. The area from the east of C1 to the coast is underlain by marine sediments; that to the west of C1 is underlain by desert sands

UMM AL AISH ~ +201RAUDHATAIN KUWAIT and


+101SAUDI ARABIA

Neogene 0 Damman.

01-1
0 ~ ~

-20

~ ,~ OX ' ~

f/I
0""-0

. ~ . ~ ~
S M/

./"./"

B9

B9'

+./"AB10 <'.-I'3-9: Gulf v./"


A./"
~~''''''''/

Coastal

./"./"

./"

.,..,..""'"

S3 S2

~.

C4 KI'

C5

t.-I~./"

HZ

f
-30

'b
~'b

-40 1- 'b O ,."../" ,.".

--0

.0 -5

C7 C3 <'.-1'0-9''''''''C2 C1 ~ v.,..,.../"

656 C ,04

cl

3 Cl' B1'

A B8,S1

AB!

.~-::::-~ABKHA SAMPLES' KUWAIT RAINWATER. n carbonate/

(autumn)

au!umn)

SEA WATER

In desert san~~88t1C muds A(sprlng) A( (spring) 0 +1 +2 +3 +4 +5 +6

-7

-6

-4

-3

-2
&!80

-1
(%0)

Figure 8.30 A plot of the water isotope

data for the sabkha

and ground

waters

indicate that variable amounts of modern recharge do take place (Abusada, 1981). Kuwait group and Dammam formation waters from western Kuwait (Abusada, 1981), together with Dammam and Neogene aquifer waters in Saudi Arabia (Shampine et at., 1978), are isotopically much lighter than the fresh waters. Published 14C analyses of these waters in Saudi Arabia place most of them between 8000 and 32 000 years old. Similarly, 14Cand tritium analyses of Kuwait waters suggest some waters may be as old as 40 000 years (Abusada, 1981). The lower isotopic values of these waters (plotting in the bottom left corner of Figure 8.30) are consistent with a cooler and

. - d-

Case Studies and Potential Applications

369

wetter climate than today. Chapman (1971) proposed a cool moist pluvial period between 11 000 and 60000 years B.P. Presumably these waters fell in Saudi Arabia onto the Dammam aquifer and have been slowly penetrating and moving towards the Gulf where they can now be tapped by drilling. Their isotopic composition was probably around oD-40%0, 0180, -6%0. These older formation waters show a marked isotopic shift from the present-day meteoric water line and may be closer to a line oD = 80180 + 8. Shifts from this line as noted in other areas, e.g. Fleischer et al. (1977) and Gat and Issar (1974), may arise from membrane filtration as the water passes through clay layers in the aquifers, slight oxygen isotope exchange of the waters with carbonate minerals in the rocks, or a combination of these effects. Waters sampled from the pits dug in the sabkhas and desert areas are seen as a board linear group in Figure 8.30. A line through this group and the old formation waters discussed above has a slope of 5, as theoretically predicted for water evaporating. Variable evaporation of the old formation waters along a line of slope 5 in Figure 8.30 could produce the observed trend. Furthermore, the samples taken from the sabkha areas underlain by marine muds show the most evaporation. This can be explained by the muds producing a capillary pumping effect on the waters (McKenzie et al., 1980) and causing more evaporation and enrichment in 180 and D. In addition, the coastal Gulf sea water exhibits a similar evaporation trend from open ocean water (SMOW). Sample BlO was taken from a tidal channel in the sabkha and appears to be derived from sea water. Other samples taken close to the areas of occasional sea water flooding of the sabkhas (B9, S3, and S2) have isotopic compositions that suggest that they are a mixture of highly evaporated formation water and highly evaporated sea water. The transverse taken for samples Cl to C6/7 at Wafra (see Figure 8.27) shows some mixing of Cl to C3 types of water with recent recharge water. Wafra well waters (C6 and C7) may represent aquifer water plus recharge from irrigation and rainfall. Samples C4 and C5 show more evaporation, which can be explained by the fact that the water table is much closer (0.5 m) to the surface in these desert pans. 8.6.5 Isotopic composition of the sulphate The sample of Gulf sea water and the sabkha samples collected in tidal channels or areas of occasional seawater flooding group around normal isotopic values for seawater sulphate in the bottom right corner of Figure 8.31. In samples such as B9, S2, and S3, the main water source is evaporated formation water but SOi- comes mainly from sea water brought into the area during flooding. Sample HZ represents sea-/continental-sulphate mixing.

370

Stable Isotopes

Raudhatain
+16 Umm AI Aish

'V

w"i8G +14~ Continental C30" B1


. Sulphate

8B

. c~6 C3' a.
C1'

G~G

--

C10 0 B8 C40 C2AK1 S1


K2

B1

A-- 6. 6. ~A
B1'

--

8R

'0 +12!---1 cfi Neogene


I

(Shampine 8 R W
A

et al., 1978) AHZ

I Aquifer

0 co

~ B9
S26.

'

B10

B9
6.

CO

...

+10

Rus Anhydrite Wadi AI Batin gypsum getch Anhydrite


Bassanlte Gypsum AI Khlran

GUlf

6.

Sea Water

S3

+8

B G

6.
Sulphate

}
+14 8348 +16 (%0)

Marine

+12

Figure 8.31 A plot of the sulphate isotope data for dissolved sulphate in the sabkha and ground waters and sulphate mineral samples. The sample legend is the same as in Figure 8.29

All the other samples have more positive 0180 values and lower 034S values, consistent with a continental origin. It was initially thought that the Rus anhydrite aquiclude may contribute sulphate to the formation waters. However, its sulphur isotopic composition is more positive than the SOi-. Also, drilling tests indicate that waters from deeper aquifers do not cross the Rus formation and sulphate contribution from any deeper source, as suggested by Rouse and Sherif (1980) for the sabkhas of the western Gulf of Sirte, is not applicable to this area. Rather, the sulphate must be from a relatively shallow source. Sulphate in the freshwater samples and the Wadi Al Batin getch gypsum were probably enriched in the heavier isotopes during bacterial reduction in the algal mats beneath the sabkhas. It should be noted that the source of sulphate for the sabkha minerals probably had isotope values similar to samples Sl and CI because a fractionation of up to a few per mil accompanies mineral precipitation (Butler et at., 1973). Sulphate samples from waters of Neogene aquifers in Saudi Arabia have 034Svalues around 12 :t 1%0and Dammam aquifers 12 :t 4%0(Shampine et

- __d

- - -

- .- .

Case Studies and Potential Applications

371

ai., 1978). These averages are a little lower than that for the Kuwait samples. The origin of this continental sulphate is rock-soil sulphate plus any rainwater contribution. Rainwater sulphate may have become concentrated in the getch layers in the upper part of the Kuwait group. Its sulphur isotope composition would represent a mixture of airborne seawater sulphate plus continental oxidized HzS on top of which is imposed a heavy isotope enrichment due to bacterial reduction. 8.6.6 Conclusions The Kuwait ground waters are mainly old formation waters with low ~D and ~18a values. Some areas, e.g. Umm Al Aish, have more positive isotope values because of local groundwater recharge. The sabkha waters are derived from formation waters by evaporation and isotopic enrichment. Some well waters (Wafra) show mixing of evaporated formation water and recent recharge. Similarly, the sabkha samples influenced by seawater incursion probably represent a mixture of highly evaporated formation water and highly evaporated sea water. Sulphate in the other waters and minerals is continental in origin. It may be derived in part from leaching of getch layers where concentration of rainwater sulphate has taken place for a long period of time. 8.7 NATURAL SULPHUR ISOTOPE DISTRIBUTIONS FOR SULPHATE AND SULPHIDES IN SEMI-ARID, MARGINAL MARINE ENVIRONMENTS IN AUSTRALIA*

8.7.1. Introduction
The peritidalt zone occupies a special position in biogeochemical sulphur cycling because it is an interface between the major geospheres and their sulphur reservoirs. Rivers and ground waters carry sulphate from precipitation and weathering processes to this zone, and marine incursions can leave high sulphate concentrations to form evaporitic deposits. Intertidal sediments are recognized as one environment in which bacterial sulphate reduction occurs, often in close proximity to cyanobacterial mats (Jorgensen and Cohen, 1977; Nedwell and Abram, 1978; Bauld et ai., 1979; Skyring et ai., 1983; Lyons et ai., 1984; Howarth and Marino, 1984). Such an environment has the potential not only to produce sulphide-rich sediments but also to contribute to the atmospheric sulphur budget by emissions of volatile sulphur compounds including HzS, CSz, cas, CH3S, OMS, and DMOS (Zehnder and Zinder,

. L.A.
;

Plumb (Chambers) Peritidal refers to sedimentological environments including supratidal, intertidal, and shallow
1973).

subtidal (Folk,

372

Stable Isotopes

1980). A further indication of the importance for sulphur cycling of this cyanobacterial, tidal-flat type of environment has come from the geological record. Features of these modern sediments have been observed in sediments of Proterozoic stratiform copper deposits (Smith, 1976)-economically important continental reservoirs in the global sulphur budget. These evaporite flats form along the landward margins of regressive seas and are extensively colonized by cyanobacterial mats. Classic examples are found on the Trucial Coast, Persian Gulf, along the shores of Spencer Gulf, South Australia, and within Shark Bay , Western Australia. This paper will present and compare sulphur isotope data from the Australian sites with relevant information from the other locations in an attempt to establish a framework for this type of environment against which anthropogenic introductions could be assessed. 8.7.2 Study sites 8.7.2.1 Spencer Gulf, South Australia Spencer Gulf is a funnel-shaped extension of the Southern Ocean into the semi-arid margin of the Australian continent (Figure 8.32B,C). It is an area of temperate carbonate sedimentation with relatively high salinities in the most northern section. A powerful tidal regime and wind-generated wave action have contributed to progradation of the shoreline and the formation of regressive sequences. Cyanobacterial mats colonize parts of the high intertidal and supratidal zones. The evolution of groundwaters and their interaction with the sediments has been studied at four areas along the north-eastern shore (Figure 8.23E). In the north, at Red Cliff and Mambray Creek, groundwater is localized in channels beneath alluvial fans flanking the relatively close Flinders Ranges. At Wood Point and Fisherman Bay the coastal plain is broader and the groundwater input to the peritidal zone is more diffuse, resulting in the formation of ephemeral lakes in addition to localized springs. Abundant gypsum is found associated with all the emergent waters, and biogenic sulphides form both in the lakes and in intertidal zones colonized by cyanobacterial mats. 8.7.2.2 Shark Bay, Western Australia Shark Bay is a large marine embayment on the west coast of Australia (Figure 8.32A, B). It is an area of active carbonate sedimentation in a semiarid environment where rainfall is of the order of one-tenth the evaporation rate. Within Hamelin Pool and Lharidon Bight (Figure 8.32D) hypersalinity develops as a result of restricted influx from the Indian Ocean because of sea-grass sills at their northern limits. Cyclonic summer depressions and

_u

H-

Case Studies and Potential Applications

373

severe winter storms produce extreme northerly winds which have produced prograding beach ridges on a regressive shoreline. This activity within the barred Hamelin Basin also produces an abnormal high-tide range and extended tidal flats, most notably in the southern region at Nilemah Embayment. The area is particularly famous for its columnar stromatolites which are analogous to the fossil structures found in many Proterozoic sedimentary sequences. Cyanobacterial mats also colonize large expanses of the peritidal sediments, and where these are dissected by tidal channels the fill sediments display an algal-laminated structure. Meteoric groundwaters emerge from Pleistocene dune formations and mix with brines of marine origin across as much as 2 km of tidal flat. Gypsum is the principal evaporite and is formed in the high intertidal zone. 8.7.3 Isotope data for water and sulphates Groundwater sulphate 034S values of peritidal areas studied range from +13.5 to +20.7%0 (Tables 8.29 and 8.30). Although there is some continental sulphate, marine sulphate is the dominant influence. The presence at Port Pirie of a smelter, which has treated Pb-Zn ores (034S = 0%0)from Broken Hill for 94 years, raises the possibility of anthropogenic sulphur in the environment. A study of heavy metal contamination of soils in the area (Cartwright et at., 1976) suggests that the effect is minimal. For the metals, the greatest influence was in the immediate vicinity of the smelter and on the western slopes of the Flinders Ranges in an area NNW of the smelter. This is consistent with wind patterns for the area which are predominantly from the SSE. If there had been considerable aerial transport of effluent it would most likely have been to the north, during the rainy season, and since sulphate in this region is more 34S enriched than in the south, contamination from a source with 034S = 0%0seems unlikely. In the Spencer Gulf areas, most groundwater evolves during passage through either a water table Pleistocene aquifer or a confined Tertiary carbonate aquifer (Ferguson et at., 1982). The latter commonly contains pyrite and a smell of HzS has been noted in some emerging waters (Clarke, 1975). A small Pleistocene gravel pressure aquifer, which may be influenced by a local reservoir, has been noted in the Baroota area (Clarke, 1975). Figure 8.32E shows the extent of the tertiary pressure aquifer within the Pirie-Torrens basin and denotes boreholes for which 034Svalues of dissolved sulphate are available. In the Red Cliff Point area, a drill hole and resurgences have also been marked because their waters are closely related, both spatially and chemically. Data for these are included in Table 8.29 for waters with> 25 meq .(;-1SO i- . The isotopic composition of hydrogen and oxygen in water varies, largely as a result of fractionation caused by evaporation and condensation.

V,) I ~

A
1130

c
136 0
I 330
-

50 km
J

26

.
TASMANIA

~,

21

V
c"" is' (J~ t::;' ~

.g ~

Q '"
~' ;:, :::
:::...

'" v, ~ :::...

~
Te'''''v sand
p.essu.. .auile,
:J:.

~ ~ .:

~ ;:;.
~ 0' ::: '"

I \
101m J ,

\ \ \ \ \

1(':\:',] Sublmoral pla//orm

\ 0 80'" hol.
"",

Inf.",d.I,

Supr.lid.1

I -,__I
0

Siudy

"'n

101m

Figure 8.32 Location of study areas: A-Shark Bay, Western Australia in B; CSpencer Gulf, South Australia in B; D-individual sites in Shark Bay region; Eindividual sites in Spencer Gulf region

VJ -.J V1

(,;.) ) 0\

Table 8.29 Isotopic and chemical data for groundwater sulphate and gypsum from study sites at Spencer Gulf, South Australia Location ground waters Reddiff Point (5) +18.6to +20.7 -4.4 to -2.9 -37 to -18 0.93 to 1.03 6.7 to 8.5 35 to 718 +19.9 to +21.1 Mambray (8) +16.4to +20.6 +4.7 to +5.5 +16 0.75 to 0.86 9.7 to 23.5 79 to 319 +19.2 to +21.4 Wood Point (11) + 13.5 to +20.6 +3.1 +2.0 0.75 to 0.84 8.9 to 14.8 204 to 377 +18.0 to +22.8 Fisherman Bay (8) +15.0to+19.6 -2.5 to +3.9 -20 to +19 0.79 to 1.04 7.6 to 15 63 to 386 +17.8

a34sS0~a1HoH20

(%0) (%0)

aD (%0) Na+/CI- (meq) CI-/SO/- (meq) SO/- (meq -1) Gypsum a34s (%0)
( ) Number of sites. Detailed data in Appendix B.

v, i:; s:: '"


<::;C)

.g '" ""

Case Studies and Potential Applications

377

Hydrogen isotopes are fractionated in proportion to the oxygen isotopes and there is a worldwide correlation between D and 180 contents of precipitation (* 5D = 85180 + 10) where both delta values are relative to the standard mean ocean water (SMOW), taken as zero. The extent of fractionation is related to latitude and is principally controlled by temperature. Rainfall near the equator is likely to show the least amount of fractionation with increasingly light values towards the poles. In the Spencer Gulf region, average values were 5D, -26.0%0 and 5180, -4.7%0 for rainfall from 1973 to 1975. The isotopic composition of waters subject to evaporation become enriched in deuterium and 180 such that highly saline brines of marine origin would be expected to have positive 5D and 511'0 values. 8.7.3.1 Continental waters, Spencer Gulf Most of the bore waters have relatively low sulphate content (Table 8.31) and, by comparison with regional precipitation values, oxygen and deuterium isotope distributions suggest a meteoric origin for the waters themselves (Ferguson et al., 1982). The range of 534Svalues for sulphate then suggests a mixture of sources being dissolved by the ground water. Only one water (north-east of Port Pirie) showed meq ion ratios (Cl-/SO/-, 9.6; Na+/Cl-, 0.82) resembling sea water (CI-/SO/-, 9.7; Na+/Cl-, 0.85), and had a 534S value of +20.8%0 which could be interpreted as a sea spray influence. The absence of consistent relationships between sulphate content, pertinent ion ratios, and the 534S values suggests that these dilute solutions are formed by interaction of the meteoric waters with aquifer sediments. Furthermore, there appears to be no clear distinction between ground waters in the Tertiary carbonates and those of the Pleistocene sediments (Ferguson et al., 1982), except that in the area east of Wood Point the sulphate is at the low end of the 534S range. 8.7.3.2 Peritidal waters, Spencer Gulf Although the waters at the landward margin of the peritidal zone are more saline than the continental bore waters, there are similarities between the 534S values of their dissolved sulphates (~ + 15 to + 18%0). The dissolved sulphate at Red Cliff Point is an exception. There, although the oxygen and deuterium distributions are clearly characteristic of continental waters, the sulphate in solution is predominantly marine in origin. This is interpreted as dissolution of marine evaporites which is consistent with Na+/Cl- meq ion ratios approaching unity and lower Cl-/SO/ratios. By contrast, obviously marine 534S sulphate values at the other Spencer Gulf sites are

oD

= lO3(D/H,,,",p,)D/HsMow

1).

1)IRO

= 103(1RO/\('O,,,mp,j'RO/160SMOW

-1).

00

--:J

Table 8.30 Isotopic and chemical data for groundwater sulphate and gypsum from study sites at Shark Bay, Western Australia Location ground waters Solar Ponds (6) +20.8to +22.0 +0.7 to +8.1 -17 to +27 0.87 to 0.98 11.2 to 8.8 101 to 252 +22.4 Nilemah (10) + 18.5 to +20.9 -1.1 to +4.0 - 24 to 0 0.85 to 0.97 13.9 to 9.8 46 to 272 +22.3 (intertidal) Playford (1) +22.0to +27.7" -1 +3.2 0.95 9.3 98 Flagpole (2) + 13.6" to 17.9 -44 to -23 -4.8 to -1.5 1.15 to 0.98 5.6 to 10.0 14 to 67

334SS0~318OH20

(%0) (%0)

3D (%0) Na+/Cl- (meq) Cl-/S042- (meq) SO/- (meq -1) Gypsum 334S (%0)

( ) number of sites. "Includes values from stations 2 and 3 (Bauld el al., 1979). h Bore water. Detailed data in Appendix C. '" ~ C) .g '" '"
~ ;::; :?::

Case Studies and Potential Applications

379

associated with strongly evaporated sea waters (8180, +3.1 to +5.5%0; 8D, +2 to +19%0). Figure 8.33 shows the location of drill holes and transects for these sites (Mambray Creek, Wood Point, and Fisherman Bay). At Mambray Creek dissolved sulphate along the southern transect is

under marine influence (834S +20%0)while the drill holes and northern
~

transect show 834S values < + 18.3%0. These waters have salinities greater than sea water and springs in the area form gypsum fans. At Wood Point a series of ephemeral lakes have formed as a consequence of groundwater discharge. Salinities and water table levels are variable and gypsum and biogenic sulphides form in the sediments. Sulphate (834S of + 15 to + 16%0)in waters landward of the lake is reminiscent of bore waters to the east, while in the intertidal drill hole sulphate shows an increase in 834S to +19.1%0; this trend is continued along the transect, presumably as a result of increased proportions of sea water. Within the lake sediments reflectinga groundwater sulphate is enriched in 32S (834S = +13.5%0), contribution from oxidation of biogenic sulphides during water table rises. Sulphides with 834Svalues of -18 to - 35%0have been found in the vadose zone. A strong continental influence is also evident in both transects at Fisherman Bay where sulphates from the drill holes and salt lake have 834S of + 15 to + 16%0. Even close to the sea on the southern transect a 834S value of + 17 to + 18%0is found. Gypsum 834Svalues are typically 1-2%0higher than groundwater sulphate in the immediate vicinity, except for one sample on the seaward side of the ephemeral lake. Gypsum with 834S = +23%0was found in proximity to groundwater sulphate with a 834S value of + 16.8%0, indicating an earlier domination by marine brines. 8.7.3.3 Shark Bay In the Shark Bay area only limited data are available from the Solar Ponds and Nilemah Embayment. The former show no evidence of continental sulphate (Table 8.30) despite an indication of continental waters from deuterium and oxygen isotopes. The gypsum lakes 1-2 km inland from the present peritidal zone have a marine source, but no hinterland bores or drill holes have been investigated. Similarly, at Nilemah Embayment, studies have concentrated on the peritidal region with lighter sulphate being found in continental inflow. Samples from two other sites reflect the very active sulphate reduction in sediments at Playford Transect (Bauld et al., 1979) and the possible influence of bore water mixing with sea water at Flagpole Landing. Table 8.32 compares the Australian data with those for the Persian Gulf. The Australian groundwater sulphate tends towards lighter values as a result of localized contribution from oxidation of biogenic sulphides. The high

w 00 0

Table 8.31 Distributians far sulphate in bare waters and Broughtan River, Spencer Gulf Mambray Creek (5) 334S (%0) 3180 (%0) 3D (%0) Na+/Ci- (meq) Ci-/SOl- (meq) SOl- (meq -1) +16.4 ta +20.1 -4.4 ta -5.9 -34 ta -39 0.83 ta 2.9 2.6 ta 6.8 1 ta 6 Baraata (4) +17.8 ta +19.7 0.61 ta 1.97 4.3 ta 12.0 1 ta 3 Part Pirie (3) +15.5 ta +20.8 -4.3 ta -5.2 -29 ta -32 0.75 ta 0.82 7.1 ta 10.1 4 ta 10 Waad Paint (4) +9.0 ta +14.0 -3.6 -26 0.71 ta 0.93 4.4 ta 13.0 1 ta 25 Braughtan River

+15.6 0.89 4.2 5

( ) Number of sites. Detailed data in Appendix A.

VJ

i:;
S2::

'" t:;o

~ '" '"

Case Studies and Potential Applications

381

11.

Wood

Point

Buch

ndges plam Holocene

~ D
~

Sup,""dal In"'Malllat Ple"tocene

EJ
~

Appw, seaw.,d m.,gm 01 cyanobacle"al mat colo",s"'on Flu...1 plam

Sup,.tidalll.,
aeohan dunes

tZZJAncient

.
D

Saltbush II.,
Buch Gypsum la"

0
~
0

[2ill Mohandun..
Sup,.t,dal manh
Onll holas

T,.nsact

Figure 8.33 Environmental

settings and transect and borehole locations at three Spencer Gulf sites

382

Stable Isotopes

Table 8.32 Comparison of sulphur isotope distributions in Australian locations with Persian Gulf
0345 (%0) Groundwater Australian Persian Gulfb sulphate" Gypsum

+13.5 to +27.7 +17.6 to +23.9

+17.8 to +22.4 +17.9 to +26.4

" Sulphate concentration exceeds 25 meq t-I. h Butler et al. (1973) (see also Section 8.6).

values of the Persian Gulf gypsum have also been attributed to bacterial sulphate reduction, depleting the brine sulphate in 32S (Butler et al., 1973). 8.7.4 Isotope distribution in peritidal sulphides The principal source of sulphide in this environment is from bacterial reduction of seawater sulphate and, because light sulphur is used preferentially in this process, the sulphides are 32S enriched. The extent of this discrimination is considered to depend principally on the organism's specific reduction rate, providing that there is an unlimited sulphate supply (Chambers and Trudinger, 1979). Depletion of the sulphate reservoir results in 34S enrichment and smaller apparent fractionation. In addition, if the sulphide is not precipitated immediately, a degree of mixing or loss can modify the range of values. Thus, small natural variations in environmental conditions could produce a range of sulphur isotope distributions. Within the peritidal zones of Spencer Gulf and Shark Bay there are a number of slightly different environments in which bacterial sulphate reduction occurs in association with cyanobacterial mats. Chambers (1982) has compared data from the high intertidal zone north of Mambray Creek with analyses of sulphides from beneath a comparable smooth mat and a compacted, gelatinous sediment at Playford Transect, Shark Bay (Bauld et al., 1979). The limited data from Shark Bay fell within the range of values found in Spencer Gulf. The more moderate values for sulphide from the compacted sediment at Playford Transect were, however, recognized as the product of an unusual environment with a restricted sulphate reservoir (Bauld et al., 1979). Tidal channels represent another variant of the intertidal environment. Cyanobacterial-layered sediments develop as the mat covering the floor of the channel is buried by successive deposits. Two such environments have been studied in channels incising tidal flats marginal to Lharidon Bight and Gladstone Embayment (Figure 8.32). The Lharidon channel is subject to

Case Studies and Potential Applications

383

intermittent desiccation, and when sampled the sediments were capped by halite crystals. It is unlikely that this would occur at Gladstone Embayment because the channel is directly connected with the embayment. Sulphides from these sediments showed 834S values of -17 to - 32%0to a depth of 50 cm at Lharidon Bight and from - 20 to - 27%0to a depth of 30 cm in the Gladstone channel. The distribution for these sites has been compared with that of the intertidal sediments of Spencer Gulf and Shark Bay in Figure 8.34. The tidal channel data extend the range of 834Svalues expected

TIDAL CHANNElS

[ill

It::::::::::::::':] lliJ

INTERTIDAL

2 ~ ~ 1 [ f/) 0

[] EJ
-20 --L -10 I

I
6345(%0)

-30 I

Figure 8.34 Sulphur isotope distributions for sulphides from channel-fillsediments compared with those of intertidal sediments

Table 8.33 Sources,

tonnages,

and isotopic

composition

of sulphur contributed Australia"

to the atmosphere

from utilization

of fossil fuels,

w 00

...".

New South Wales Black coal S (tonnes x 103) 334S (%0) Brown coal S (tonnes x 103) 334S (%0) Lighter refinery products (LRP) M spirit S(tonnes x 103) ATF S(tonnes x 103) ADO S(tonnes x 103) IDO S(tonnes x 103) Total (LRP) S(tonnes x 103) 334S(%0calc.) Fuel oil S(tonnes x 103) 334S(%0calc.) Total oil products S(tonnes x 103) 334S(%0calc.) Total coal + oil S(tonnes x 103)
334S (%0 calc.)

Victoria

Queensland

South Australia

Western Australia

Tasmania

Northern Territory

101.7 +4 0.6 0.3 4.6 1.0 6.5 +6 13.8 0 20.3 + 1.9 122.0 +3.6

28.4 +20 0.5 0.2 2.7 0.5 3.9 +6 11.4 0 15.3 +1.5 43.7 +13.6

31.5 +4

10.3 +11

14.4 +4

0.5 +4

0.3 0.1 3.8 0.2 4.4 +6 18.8 0 23.2 +1.1 54.7 +2.8

0.2 <0.1 1.3 0.3 1.8 +6 5.2 0 7.0 +1.5 17.3 +7.8

0.2 0.1 3.0 0.3 3.6 +6 24.9 0 28.5 +0.8 42.9 +1.8

<0.1 0.1 0.4 0.2 0.8 +6 6 0 7.6 +0.6 8.1 +0.8

0.1 0.2 0.5 0.1 0.9 +6 10.2 0 11.1 +0.6 11.1 +0.5

c"

i::; C!"'

t::;' c

"From 'Black coal in Australia 1981-82', 'Australian Energy Statistics 1982', and data supplied by the Australian Institute of Petroleum Ltd and the State Electricity Commissionof Victoria in August 1983.

Case Studies and Potential Applications

385

in the peritidal zone to more negative values (- 9 to - 32%0). The mean of all values is -19%0. 8.7.5 Summary Natural sulphur isotope distributions in arid or semi-arid peritidal environments are strongly influenced by discrimination during bacterial sulphate reduction. Examples of 32S-enriched groundwater sulphate in sediments where there are biogenic sulphides suggests a source for the light sulphate emerging in continental waters in the supratidal. This determines the lower range of 834S values for gypsum. Although no gypsum as 34S-enriched as Persian Gulf samples was observed in Australia, 34S-enriched groundwater sulphate was noted. Sulphides exhibit considerable variability but all are apparently fractionated by > 30%0(834S, seawater sulphate = +21%0). In regarding these data as natural distributions the contribution from any anthropogenic source must be assessed. As has already been mentioned, contamination from industrial gases in the Port Pirie area seems unlikely, particularly in view of the isotopic similarity of sulphates in other regions remote from any such activity. The same argument applies to agricultural practices which could cause perturbations in the natural patterns. 8.8 CONTRIBUTIONS OF ANTHROPOGENIC SULPHUR TO THE ATMOSPHERE IN AUSTRALIA

8.8.1 Coals* Contributions of sulphur to the atmosphere from coal consumption in major industries, e.g. electricity generation and the metallurgical and cement industries, are shown in Table 8.33. In the States of Queensland, New South Wales, South Australia, and Western Australia black coal is the main or a major energy source. From data presented elsewhere in this volume (Section 4.5.2), an average 834S value of +4 ::t 3%0 is generally assigned to the sulphur in this coal. However, the few available measurements suggest a 834S value of + 15%0for Leigh Creek coal (Smith and Batts, 1974), which is mainly used for electricity generation in South Australia. In calculating the tonnages of sulphur released annually in the individual States, the sulphur content of the coal utilized is assumed to be 0.5% of which 90% is voided to the atmosphere as gases. In Victoria, the situation is different as the main energy source for electricity generation is brown coal. Not only has this a characteristic 834S value of +20 ::t 3%0(Section 4.5.2) but inorganic constituents ensure a high
* l.W. Smith.

V-) 00 0\

100 x 103

tonnes -1 S yr
~f;e;~~'

'sOxlO3 WESTERN AUSTRALIA tonnes -1 S yr

@ @/

// , /,'/'

10x 103 tonnes


S yr -1

+4
//.

034S values

a "

"

500 km
V)
;:; <:J-

AUSTRALIA
Figure

+4
and isotopic composition are shown for each State

"
a
o:, '"

;;:;-

.g-

8.35

Contribution

of sulphur to the atmospherefrom combustion of coal in Australia, Annual sulphur contribution

Case Studies and Potential Applications

387 coal has a in ash on dry basis) coal data in Figure

degree of retention of sulphur in ash (O'Heare, 1980). Yallourn sulphur content of 0.28% (dry basis) of which 24% is retained combustion. Morwell coal has a higher sulphur content (0.45% but 50% of this is retained in ash. These factors were taken into account when compiling the presented in Table 8.33, which are also shown diagrammatically 8.35. 8.8.2 Crudes and condensates

In 1981-2 a total of 17800 x 103 tonnes of crude oil containing 15500 tonnes of sulphur with an estimated 334Svalue of + 7 ::!:2%0was produced locally. Corresponding values for imported crude were 10 900 x 103 tonnes of crude and 142000 tonnes of sulphur. Middle East crudes comprised 74% of these total imports with 51% coming from Saudi Arabia alone. Measurements on Arabian mixed crudes from a Sydney refinery gave a 334S value of -6.4%0. For purposes of calculation, in the absence of other information, Australian and all imported crudes, together with their corresponding refinery products, are credited in this report with 334Svalues of + 7%0and -6%0 respectively. In terms of both tonnages consumed and sulphur contents, the five most significant refinery products in 1981-2 were motor spirit (M/Spirit), aviation turbo fuel (ATF), automotive diesel oil (ADO), industrial diesel oil (IDa), and fuel oil (Flail). The tonnages, sulphur contributions, and calculated 334S values for these products are shown in Table 8.34. Clearly only the last of these products reflects the much higher 34S content and 334S value of the imported crude. All other products appear not to differ significantly in isotopic composition from local crudes which are dominant in the refining. The consumption of these refinery products in individual States and the calculated 334Svalues for sulphur released on combustion are shown in Table 8.33. These data for total oil refinery products are shown diagrammatically in Figure 8.36. Also included in Table 8.33 are the total tonnages and calculated 334S values of sulphur released to the atmosphere from fossil fuel (coal + oil) consumption. These data are presented diagrammatically in Figure 8.37. A comparison of Figures 8.35, 8.36, and 8.37 allows the variability in sulphur contributions to the atmosphere over the Australian content to be visualized. Corresponding 334S values for this sulphur should be regarded with reserve in view of the extremely limited isotopic information on crude oil products.

00 00

Table 8.34 Tonnages, sulphur contents, and isotopic compositions of major refinery products, 1981-2" Product Imports Quantity (tonnes x 103) M/Spirit ATF ADO IDa Flail 295 83 439 9 1452 Sulphur (tonnes) 88 84 1 759 76 43 574 Local Quantity (tonnes x 103) 10 959 1 674 6139 524 2817 Sulphur (tonnes)
-----

Total
--------

034S (%0calc.) Sulphur (tonnes) 1 782 739 16569 2467 91 146 +6 +5 +6 +7 0

Consumption (tonnes x 103)


-------

1694 655 14 810 2391 47 572

11 254 1 758 6579 534 4270

" From data supplied by the Australian Institute of Petroleum Ltd, August 1983.

" ~ a .g

V) ;::; (:J-

L:;

Q '" '" v, ~ ~ ~. I:> ;: ~ :;? ~ ~ ;:s. ~ ~ ::;. ~ c' ;: '"

a ,

500 , km

AUSTRALIA
Figure 8.36 Contribution of sulphur to the atmosphere from combustion of oil in Australia. Annual sulphur contribution and isotopic composition are shown for each State w 00 \0

VJ 'D 0

NORTHERN TERRITORY

I r I I
i

Mt Isa

I QUEENSLAND

AUSTRALIA

'-..?'~'

,,~, . 'f;(

0
L_,_, -"J

500 km

v, ;::;

AUSTRALIA
Figure 8.37 Contribution of sulphur to the atmosphere from combustion of coal and oil in Australia. Annual sulphur contribution and isotopic composition are shown for each State

"
""

.g ~

~ a

Case Studies and Potential Applications

391

8.8.3 Natural gases* No information on the isotopic composition of sulphur in natural gases exists; however, previous estimates of sulphur entering the environment from this source (800 tonnes) suggest its effect to be insignificant (Manuell, 1980). 8.8.4 Oil shales* The analyses of single, possibly unrepresentative samples of a variety of oil shales are shown in Table 8.35, together with estimates of the reserves of such shale, their total sulphur contents, and 534S values. These shales are not currently utilized; however, the size and isotopic composition of these shale reserves suggest that, if they were used for liquid hydrocarbon production in the future, combustion of the products might produce major changes to the existing scene. 8.8.5 Isotopic composition and fate of sulphur mined in Australia with copper, lead, zinc, and nickel orest Australia's mineral wealth lies principally in massive sulphide deposits whose mining and subsequent treatment will alter the S content of the environment. As a result of extensive research on many deposits, a considerable sulphur isotope data base exists for the economic minerals. This information, coupled with relevant production and export data, permits an estimate of the

Table 8.35 Quantity" and isotopic composition of sulphur associated with oil shale Source Estimated reserves (tonnes x 109) 18 3800 ? 4.7 0.03 Total sulphur Sulphur (%) (tonnes x 106)
834S (%0)

Condor Julia Creek (Toolebuc Formation) Rundle Tasmanite Nagoorin


a

0.61 2.11 0.90 3.35 1.09

110 80 000 42 1 -

+8.0 -7.9 +21.7 +12.7 +12.1

From Gibson (1981).

* l.W. Smith. :\: L.A. Plumb (Chambers).

VJ

Table 8.36 The O:J4S (%0) values for sulphides of copper, lead, and zinc ores mined in Australia Mine Pyrite, pyrrhotite -------Mt Isa Mt Lyell Tennant Creek Mt Gunson Cobar Woodlawn Dianne Mt Morgan Mt Chalmers +9.4 +8.1 +2.3 +9.2 +6.7 +7.7 +4.4 +9.5 +6.9 +2.0 +15.6 +7.2 +13.3 +6.4 +12.1 Chalcocite, chalcopyrite, bornite Galena Sphalerite -+16.7 +14.5 +6.9 +7 to +10 -9.0 +9.1 +6.8 +7 to +10 +2.4 Solomon (1965) Solomon ef al. (1969); Walshe and Solomon (1981) An expected value Knutson ef al. (1983) Sun (1983) Ayers ef al. (1979) An expected O:J4S value Eadington ef al. (1974)(average includes some pyrrhotite values) Stanton and Rafter (1966); Solomon ef al. (1969); Green ef al. (1981) Scott ef al. (1985) Collins (1981) Scott ef al. (1985) Both and Smith (1975); Stanton and Rafter (1966) Average References and comments

\0 N

Rosebery

+12.0

+11.1

+11.2

+11.2

Mammoth Cleveland Gunpowder Broken Hill

-12.5 +1.9 +11.9

-7.5 +2.1 +0.5

+ 1.9 +1.4

-10.1 +2.0 +11.9 +1.0

V)

;;:;

0.g

t:;c

Table 8.37 Stable isotope content of sulphur associated with Australian copper ores and concentrates" smelted and exported
during 1980 (Perkin, 1980) Mine 034S(A)
(%0) Mt Isa +14.5

D '"
'"

v,

::: To smelter (B)


(k tonnes) 145.1

To export (C)
(k tonnes) 13.3

(A) x (B)
(%0 x k tonnes) 2104

(A) x (C)
(%0 x k tonnes) 193

I:>..

I:>..
'"1::1 Q

;::, ;::s

Mt Lyell Tennant Creek"

+6.9 +7

1.0 10.9

18.9 -

Mt Gunson Cobar Woodlawn

Dianne"

-9.0 +9.2 +6.8


+7 +2.4 +11.2 -10.1 +2.0 +11.9

7 76

130

6.5 3.0
3.3

8.7 2.9
4.2

20
8

-78 +60 20
29 28 -10 1

:
::t.

:g --.

Mt Morgan Mt Chalmers Rosebery Mammoth Cleveland Gunpowder" Totals

'" ;::, ... c' ;::s '"

2.5 1.0 0.4 1.5 171.3 223.2 51.9 2233 2546 18 ----

313

Average 034S (%0)

+11.4 + 13.0 +6.0

Smelted Exported

" Assumes all copper is chalcopyrite. /> Anticipated &,"5 value.


(' Bacterial leaching.

W \0 W

W 1.0 .j:>.

Table 8.38 Stable isotope content of sulphur associated with Australian lead ores and concentrates" smelted and exported during 1980 (Mock and Roarty, 1980a) Mine 834S(A) (%0) +14.5 +0.5 +7.7 +6.8 +11.2 ----Totals 52.9 57 Average 834S (%0)'" Assumes all lead is galena. +6.3 +6.2 +6.9 4.1 +329.6 357.8 +28.2 To smelter (B) (k tonnes) 21.0 30.5 0.3 1.1 To export (C) (k tonnes) 0.8 0.1 2.0 1.2 (A) x (B) (%0x k tonnes) + 304.5 + 15.3 +2.3 +7.5 (A) x (C) (%0 x k tonnes)
---

Mt Isa Broken Hill Cobar Woodlawn Rosebery

+0.4 +0.8 + 13.6 +13.4

Smelted Exported
V)

E; t:J-

t:;o .g ~

"

Q '"
'" v, i2 Table 8.39 Stable isotope content of sulphur associated with Australian zinc ores and concentrates" refined, smelted, and exported during 1980 (Mock and Roarty, 1980b) Mine &34S (A) (%0) To refinery (B) (k tonnes) To smelter (C) (k tonnes) To export (D) (k tonnes) (A) x (B) (%0 x k tonnes) 248 74 273 23.2 216 98.7 595 1332 153 584 (A) x (C) (%0 x k tonnes) 3.4 118 32 (A) x (D) (%0 x k tonnes) 428 82 64 10
:::>... (;;. '" ." ;:: :::>...

~ ~ :? ~

Mt Isa Broken Hill Woodlawn Cobar Rosebery Totals

14.5 +1.4 +6.8 +9.5 +11.2

17.1 52.6 24.4 94.1

2.4 17.4 3.4

29.5 58.8 9.4 1.0

:J:. ~ c:;' E? 5' ;:: '"

Average &34S (%0)

+6.2 +6.3 +6.6 +5.9

Refined Smelted Exported

" Assumes all zinc is sphalerite.

W \0 VI

396

Stable Isotopes

proportion and isotopic composition of anthropogenic sulphurcompounds in the environment. Tables 8.36 and 8.37 represent a compilation of 334S values for sulphur associated with the ores mined and treated in Australia during 1980. Data for the sulphides of Cu, Pb, and Zn are listed in Table 8.36 with the average comprising all values reported in the quoted references. In Tables 8.37, 8.38 and 8.39, this average is used to calculate bulk 334Sof Cu, Pb, and Zn ores and concentrates mined, smelted, and exported from all mines except Broken Hill and Cobar. For these, the relevant values for sphalerite and galena were considered more appropriate. Agreement between the calculated sulphur of Pb and Zn concentrates treated at plants practising recovery (Table 8.40) and production values of H2SO4 quoted by Driessen (1980) suggests 100% recovery efficiency, which is unlikely. Thus the estimates of losses for these treatments must be considered as minimal. No recoveries as H2SO4 are reported in refining and smelting of Cu ores. The isotopic distributions of the nickel ores are given in Table 8.41 and these data are combined with production, refining, and smelting figures in Table 8.42. A formula of Fe2NiS3 has been adopted as representative of the nickel ore. This presents a more conservative sulphur budget than assuming a 32% sulphur content of nickel concentrates (Driessen, 1980) which is equivalent to a 4: 1 ratio for S/Ni and the nickel-poor end of the naturally occurring pentlandite solid solution (Harris and Nickel, 1972). Material not accounted for as matte, Ni-Co, or Cu sulphides and (NH4)zS04 is considered to have been lost to the atmosphere. Finally, a compilation of losses is presented in Table 8.43. The absence of tabulated losses during treatment of zinc ores reflects the assumption of complete recovery as H2SO4' In summary, in 1980, 52 k tonnes of sulphur with an average 334S = +6.0%0 were exported with Cu ores (Table 8.37), 4 k tonnes of sulphur with an average 334S = +6.9%0were exported with Pb ores (Table 8.38), and 99 k tonnes of sulphur with an average 334S = +5.9%0 were exported with Zn ores (Table 8.39). Sulphur recovered as H2SO4 (Table 8.40), assuming 100% efficiency, would have been 32 k tonnes with an average 334S = +0.8%0 from Pb ores and 117 k tonnes with an average 334S = +6.4%0 from zinc ores. For ores smelted and refined in Australia, at least 229 k tonnes of sulphur with an average 334S= + 11.3%0were lost to the environment (Table 8.43).

Case Studies and Potential Applications

397

Table 8.40 Sulphur (k tonnes) in lead and zinc ores and concentrates treated at refineries and smelters which manufacture H2SO4 from effluent (from Tables 8.36 and 8.37) Sulphur in lead ores Broken Hill Cobar Woodlawn Totals Average Sulphur in zinc ores Mt Isa Broken Hill Woodlawn Cobar Rosebery Totals Average &34S(A) %0 +0.5 +7.7 +6.8 +0.8 Quantity (B) k tonnes 30.5 0.3 1.1 31.9* (A) * (B) +15.3 + 2.3 + 7.5 +25.1

* Compare to 36 k tonnes S recovery as H2SO4 (Driessen, 1980) +14.5 +1.4 +6.8 +9.5 +11.2 +6.4 17.1 55.0 17.4 3.4 24.4 117.3* +248 + 77 +118 + 32 +273 +748

* Compare to 122 k tonnes S recovery as H2SO4 (Driessen, 1980)


u Driessen, 1980.

Table 8.41 The &34S values of sulphides associated with nickel ores mined in Western Australia Location
&34S (%0)

Reference

Kambalda Perseverance Windarra Various

+2.0 +2.4 -2.2 -0.9 +1.7 Average +0.6

Donnelly et al. (1978) Seccombe et al. (1981a) Donnelly et al. (1978) Donnelly et al. (1978) Seccombe et al. (1978) Donnelly et al. (1978)

398

Stable Isotopes

Table 8.42 Stable isotope content of sulphur associated with Australian nickel ores and concentrates" refined and smelted during 1980 (Pratt, 1980)
Mine
834S (%0)

Production (k tonnes) 57.0 15.1

834S produced (%0 x k tonnes) + 125.4 -33.2

Kambalda Nepean Agnew (Perseverance) Totals


Bulk 834S

+2.2 -2.2

72.1
+1.3

+92.2

k tonnes S To smelter Produced matte To refinery Produced Ni-Co sulphides Cu sulphide (NH4hS04 57.2 8.9" 14.9 0.9 0.8 24.5 72.1 Loss of 37 k tonnes of + 1.3%oS
a h

35.1

Assumes all nickel is pentlandite. Matte contains 20% S, 73% Ni. Ni produced = 32.5k tonnes.

Table 8.43 Sulphur lost to environment in treatment of copper, lead, zinc, and nickel ores
834S (A) (%0)

Quantities (B) (k tonnes)

(A) x (B) (%0 x k tonnes)

Lead ores Mt Isa Copper ores Total smelted Nickel ores Calculated losses Totals Average 834S (%0)

+14.5 + 13.0 +1.3 229.3 +11.3

21.0 171.3 37.0 2585.6

304.5 223.3 48.1

Case Studies and Potential Applications

399

8.9 SULPHUR ISOTOPE STUDIES OF ATMOSPHERIC S AND THE CORROSION OF MONUMENTS IN PRAGUE, CZECHOSLOVAKIA * 8.9.1 Introduction The aim of these studies was to differentiate possible sources of rainwater SOi-, atmospheric S, and gypsum crusts on monuments in Prague, Czechoslovakia. Prague is an industrial city surrounded by thermal and electrical power plants. A large conglomerate is located about 100 km to the north-west (Figure 8.38). Sulphur isotope abundances were determined for rainwater SO/-, flue gas, coal, dust particles, S02 and aerosol, and gypsum crusts on stone monuments. The flue gas was sampled at the Chvaletice power plant located to the east. Coal is transported there from West Bohemia. Local emissions were estimated at 50 400 tonnes yr-1 S02 plus 1650 tonnes from mobile sources. The mean concentration of rainwater SOi- was found to be 16 mg -1. The S02 concentrations fluctuated between 20 J-Lg m-3 in summer and a maximum of 200 J-Lg m-3 in winter. The dry and wet depositions were about 3.4 and 8.6 tonnes km-2yr-1 respectively (Novak, 1982). Pyrite, organic sulphur, and sulphate in coal were examined previously (Hokr et at., 1972). The pyrite content ranged from 2 to 20%, average 6%.
360'

270'

. .Chvaletice

.
~
180'

50 km

Figure 8.38 Locations of power plants around Prague, Czechoslovakia . F. Busek, 1. Cerny, and J. Sramek.

-""----

400 Table 8.44 Isotopic compositionof sulphur in brown coal, Czechoslovakia(after Hokr et al., 1972)
Form Sulphate Pyrite Organic sulphur Average
S34S (%0) --

Stable Isotopes

-3.2 -0.6 +1.2

- -1

Data from the mine that supplies the Chvaletice plant are summarized in Table 8.44. 8.9.2 Analytical The flue gas was sampled by a probe in the stack of the Chvaletice power plant at a temperature of about 200C. The S02 and S03 were collected simultaneously, oxidized in H202 solution, and precipitated as BaS04' Rainwater was collected with a manual sampler of 1 m2 area, which was covered during dry periods. The sampler was placed in a heavy polluted area of the inner city near a meteorological station. Dissolved sulphate was concentrated by ion exchange (Nehring et al., 1977) and precipitated as BaS04' S02 and aerosol were trapped using a Sartorius EM 100 high-volume sampler and Whatman 40 filters impregnated with KOH-triethanolamine solution. Sulphate aerosol was collected on a paper prefilter. Between 40 and 100 m3 of air were sampled. Dry deposition was sampled to compare with rainwater SO/-. Crust sulphates on monuments were dissolved in acid solution and precipitated as BaS04. Approximately 20 mg of BaS04 were converted into S02 by reaction with V20s (Haur et al., 1973). Isotope analyses were carried out on either an M86 or a Mat 251 ratio mass spectrometer. 8.9.3 Results and discussion 8.9.3.1 Atmospheric studies The flue gas in the power plant had 534S values (Table 8.45) close to that of the coal being consumed (534S ==-1%0). In comparison, the 534S values of 34 S02 samples ranged from -1 to +6, with an average near +2%0. The

Case Studies and Potential Applications

401

Table 8.45 Isotopic composition of sulphur in flue gas of the Chvaletice coal fired power plant. Date 1981 11.6 12.6 12.6 12.6 Time (h) 1505-2045 1035-1140 1215-1315 1415-1515 [SOz] (fLgm-3) 2079 4359 4344 3678 034S (%0) -1.4 +0.1 -0.8

SDz~
aerosol

monuments

~
dustfall ---4IJJ--E <l!//!J-

rain sulphates M ~

B <1!/!JP=

-==-5 0

power plant coal "'5


834S (0/00)

...10

Figure 8.39 Summary of environmental sulphur isotope data, Prague, 1981. B, M, and E designate the beginning, middle and end of storm events

Table 8.46 Isotopic composition of sulphate in dustfall Date, 1982 Exposure (h)

SO/(mg)

Mass flux (mg SOlm-zh-I 0.52 0.77 0.21 0.62 0.28 0.28 0.45

034S (%0)

29.1- 1.2 16.2-17.2 26.2- 1.3 24.3-27.3 27.3-30.3 1.5- 3.5 Mean

77.50 30.75 66.75 79.50 74.25 37.75

40.1 23.7 13.8 49.6 20.6 10.7

+9.0 +4.9 +6.7 +5.2 +5.1 +3.6 +5.8

402

Stable Isotopes

aerosol displayed the same variation, but was about 1%0 heavier on average than the S02 (Figure 8.39). The data for dustfall (Table 8.46) and rainwater SOi- were similar, with 034S values averaging +6%0. The average concentration of SOi- near the beginning of storms was about twice that found in the middle and end (Figure 8.40). There were no significant differences in the isotopic composition of S042- collected near the beginning and during mid-storm. The 034S values of SO i- collected near the end of storms appear to be higher on average than those collected near the beginning or during midstorm (Figure 8.39). Further, if the data are divided according to wind direction, there is no statistical difference in the 034S values of SO/associated with the beginning and end of storms coming from the northwest. In contrast, the 034Sshift during storms coming from other directions averaged +2%0. If the above shifts are meaningful, at least two sources of sulphur are involved. The negligible average shift when the wind was from the northwest suggests that the power plant complex in that direction constitutes a dominant source of atmospheric sulphur. Because of its dominance, a temporal trend was not observed during the storms. However, the average 034Svalue of +6%0 is somewhat higher than that of flue gases measured at the Chvaletice plant in the easterly direction. When the wind was southerly or south-westerly, the 034Svalues near the beginning of the storm also averaged +6%0, which is close to that of the dominant source. Therefore, the increase in 034S towards the end of the storm must indicate a second source. One interpretation is that the wind

_l
A&
0 20
2-

B I 40 (mgl-1)

54 Figure 8.40 Histograms of SO/- concentrations during the beginning, middle, and end of storms, Prague, 1981

Case Studies and Potential Applications

403

and rain scavenged the dominant pollutants from the atmosphere during the early part of the storm. The SOi- present in rain during the latter stages of the storm may have been carried by long-range transport. 8.9.3.2 Stone monuments Corrosion of calcareous stone (marble, limestone, some types of sandstone, etc.) is one negative aspect of industrialization. The damage to the surfaces is irreversible and can be described by the general reaction: CaC03 + (SOz, SO i- , NOn 03, H2O, H+, metals) ~ CaS04"2H2O

10 11 00 el

e e6 2
12 0

e9

13 0

14 0

15 0 016

18 0

400

Figure 8.41 Monument samplinglocations in the centre of Prague in relation to the Vltava River Valley. The black and white circles correspond to areas of higher and
.

lower sulphur pollution respectively

404

Stable Isotopes

Table 8.47 Sulphur isotope compositionof crusts on selected historical monuments in Prague
Location" 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. Description St. Nicolaus Church Old Town Hall Charles Bridge Towers Charles Bridge Arches Clam-Gallas Palais Tyn Church St. Clement Church Old Jewish Cemetery Henry's Tower St. Vitus Cathedral St. George Basilica Fortification Wall New Town Hall St. Ludmila Church Leopold's Gate St. Martin Rotunda St. Peter and Paul Church St. Philip Church
O:14S(%0)

+4.2 +4.5 +4.3 +4.2 +3.6 +4.0 +4.5 +3.2 +4.4 +3.1 +2.9 +1.9 +2.9 +1.8 +2.9 +2.3 +2.4 +2.1

" number corresponds to Fig. 8.41.

Table 8.48 Laboratory studies of reaction between building materials and S02 (2000 ppm S02, 80% relative humidity, 20C) Sample
O:14S (0/00)

Surface area (m2 g-l) 0.5 0.3--0.4 1.0-1.2

Reaction ability

Remarks

S02 Gaize Marble Sandstone

+12.5 +13.1 +13.5 +13.9

high medium low

Fe promoter pure stone binding medium

The gypsum crusts are more soluble than the original calcium carbonate and consequently the crusts fall off (Gauri and Holdren, 1981). Samples were collected from 18 of Prague's historic monuments (Figure 8.41). The locations are divided into two sets designated by black and white circles corresponding to areas of higher and lower pollution. In the former, the S02 and NOx concentrations in winter were both about 200 fLgm, -3, which was roughly five times the values recorded in the less polluted areas.

Case Studies and Potential Applications

405

Sulphur isotope data were obtained for samples carefully selected to exclude direct contact with rain and snow. Therefore the crusts were typical atmospheric sulphur reaction products. The average 334Svalues for samples from the more and less polluted areas were +4.1 and +2.5%0 respectively, with no overlap (Table 8.47). However, the ranges for S02 and aerosols extend well beyond that of the monument data so it is difficult to differentiate between aerosol and S02 as the prime reactant. However, in laboratory studies, the sulphate crust was found to be about 1%0enriched in 34S as compared to reactant S02 (Table 8.48). If the same isotope fractionation is relevant to the in situ monuments, then hydrolysis-oxidation reactions involving atmospheric S02 would seem to be the likely mechanism, as previously discussed by Longinelli and Bartelloni (1978) for monuments in Venice.

8.10

DETERMINATION OF TOTAL SULPHUR &34SVALUES IN ORE DEPOSITS OF THE PRECAMBRIAN OF INDIA*

8.10.1 Overview Sulphur isotope abundances and trace element concentrations were determined in mineral separates from sulphide ore horizons at Ingaldhal, Kalyadi, Thinthini, and Hutti in Karnataka State, India. These data have been interpreted using chemical solution models to provide an estimate of the isotopic composition of total sulphur in sulphide mineralization in the Archaean greenstone belts of the Dharwar Craton. 8.10.2 Introduction In the estimation of lithospheric components of sulphur in the global sulphur cycle, it is important to be able to estimate the isotopic composition of total sulphur in ore deposits. In situations where the ore deposit consists predominantly of a single mineral, the evaluation of 334Sis relatively simple. The adopted convention is to compute the average of measured 334Svalues of different samples of the mineral. The situation is more complicated in multimineralic ore deposits where ore-forming solutions have been influenced by a variety of parameters during transport and deposition. The relative contribution of these parameters to determining the distribution of sulphur isotopes in the different precipitating mineral phases must be evaluated (Ohmoto, 1972; Rye and Ohmoto, 1974). Our recent chemical modelling of trace element distribution in coexisting sulphide phases (precipitating from hydrothermal brines) has been used to evaluate the solution chemical
* A.G. Menon, G.V.A. Iyer and V.S. Venkatasubramanian

406

Stable Isotopes

parameters, and this information in turn has been used to obtain 834Stotal from fsz-foz plots.
8.10.3 Sulphur isotope data and estimation of ore solution parameters The 834S measurements were carried out on 27 sulphide samples from ore deposits at Ingaldhal (Chitradurga belt), Kalyadi, Thinthini, and Hutti, from the granite-greenstone terrain of Dharwar Craton, South India (Radhakrishna, 1976). The use of solution equilibria for interpreting sulphur isotopic abundance variations onfoz -pH plots was reported by Sakai (1968) and Ohmoto (1972). Working along these lines, three solution parameters, fsz, foz, and pH, were found to be suitable for explaining the data. The basic expression for making calculations of total sulphur isotopic composition is given below (Menon, 1979):
mtotal

= (fS/fOZ)1(CI + CzJbz)

where mtotal is the concentration of sulphur in solution, fsz and foz are the fugacities of sulphur and oxygen respectively, and CI and Cz are functions of equilibrium constants of solution equilibria, pH, and activities of various species in solution. This expression, when combined with the sulphur isotopic balance equation, enables the estimation of the total sulphur isotopic composition, o34Stotabfor the sulphide minerals that constitute the deposit. Sulphur isotopic data for the areas investigated are summarized in Tables 8.49 and 8.50 (Menon et al., 1981a, 1981b; Venkatasubramanian et al., 1981a, 1981b). At Ingaldhal, a mixed (seawater and magmatic) source has been 'suggested for the sulphur which has an estimated original sulphur isotopic composition

Table 8.49 Mineralization parameters for ore deposits, Dharwar Craton, South India Location Parameters mT (M) 034S (%) T (C) pH log fsz log foz log feoz Ingaldhal 0.01 0 to + 10 30350 5 to 7 -8 to -10 -27 to -31 Kalyadi 0.01 0 to + 5 20250 5 12 to -14 -39.5 Thinthini 0.01 0 200 5 -11 -33 Hutti 0.01 0 350 -13.6 -31 +2

Case Studies and Potential Applications Table 8.50 Summary of 334Svalues for ore deposits of India Deposit Andhra Pradesh Mangampeta Mineral 334S (0/00)
334STotal ('1'00)Reference

407

ba py bedded py gn -

+41.8 to +45.5 -1.1 to +2.5 +18.7 -0.1 to 2.6

Karunakaran (1976)

Rajasthan Rajpura-Dariba

Kolihan and Madhan Kudhan copper deposit Karnataka Ingaldhal

+10 to +23

Deb and Kumar (1982) Sofonov et at. (1980) * 10 Seccombe et at. (1981) 0 3 to 5 Sofonov et at. (1980) *

cp po sp gn cp gn sp py cp py cp po py-cp cp py asp po

-2.1 +5.4 +7.7 +11.1 -0.4

to +8.3 to +10.1 to +8.9 to + 11.7 to +0.1 -0.9 -0.2 +0.4 -1.5 to +1.5

Kalyadi

+2.4 to +6.7 +1.5 to +5.0 +5.4 -1.7 to -4.9 +0.8 to +3.5 +5.4 +0.9 to +1.6 +0.8 to +2.2

Thinthini Hutti

3 0

Sofonov et al. (1980) * *

034Stotal of about +10%0. The estimated log!s2 is -8 to -10; log!02' -31; and temperature about 350C. At Kalyadi, ore is believed to have been deposited from an solution with a sulphur isotopic equilibrium temperature of 200C estimated 034Stotal of +3 to +5%0. The estimated pH is [S2' -12 to -14; and log !02' -39.5.

-27 to igneous and an 5; log

408

Stable Isotopes

For Thinthini, the 834S values of the minerals plot in the SOl- stable region of the Isz -Ioz diagram. The estimated pH is 5; log Isz, -11; and log loz, - 33, and temperatures above 200C. At Hutti, the 534Stotal is estimated to be very near 0%0,suggesting a mantle source for the sulphur. The estimated log Isz is -13.6; log loz, -31; and log leoz, +2. Table 8.50 also summarizes 534Stotaldata for sulphide minerals in other deposits in India. 8.10.4 Trace elements in the evaluation of &34Stotal The estimated parameters of Table 8.49 may be further confirmed from the trace element analyses. The basis for this approach is the considerable variation in the trace element concentrations that has been observed in different samples of the same mineral in a given deposit. For example, the concentration of Pb in pyrite varies between 44 and 2643 ppm as compared to between 67 and 1384 ppm in chalcopyrite. Other trace elements, Co, Ni, Zn, As, and Bi, in the sulphide matrices also show similar variations (Table 8.51). Whereas it is recognized that trace element incorporation in sulphide matrices is governed by the lattice structure, there should also be chemical influences on their incorporation in the precipitating mineral phase. On the bases of concentrations and thermodynamic data (Barnes, 1979) for trace element complexes in solution, we undertook a theoretical study of this behaviour in terms of solution chemical parameters. We assumed that the trace element content in a mineral phase is not so much a function of its concentration in solution but the 'willingness' of the solution to release it into the precipitating mineral. Dpb, the distribution coefficient for lead, is defined as the ratio of the concentration of lead in the solid (mineral) phase and the liquid (mineralization solution) phase, i.e.
Table 8.51 Concentration of some trace elements in pyrites and cha1copyrites of Kalyadi Concentration ppm Mineral py cp Sample (number) Kl-p K2-p K8-p K2-c K4-c K8-c Pb 44 2643 103 245 1384 67 Co 4334 521 3730 40 95 55 Ni 385 744 304 133 173 65 Zn 309 815 624 1282 561 As 4 76 66 13 391 20 Bi 7 228 52 15 8 12

Case Studies and Potential Applications


Dpb

409
(6)

= mpb(S)/mpb

Further, assume, for simplicity, that the trace element is incorporated as a solid solution in the host matrix. Then
Dpb

mpb

(S)/"iimi

where mi corresponds to the concentration of various complex ions of lead in the solution. These include, in our present calculations, PbCI+, PbCho, PbCl3 - , PbCI/- , Pb(HSh - , Pb(HShO, PbS(H2Sh- The relevant equilibrium reactions when combined with the equilibria of various sulphide species in solution result in an expression of Dpb as
Dpb

= !(fsz,foz, pH, T, KiomCl-)

where !sz and !oz are the fugacities of sulphur and oxygen respectively, Tis the temperature, K are the equilibrium constants for the complex ion equilibria, and mCl- is the Cl- concentration in the mineralizing solution. The result of one such calculation is shown in Figure 8.42, with the shaded area indicating solution conditions prevailing at the time of mineral precipitation at Kalyadi. 8.10.5 Conclusions The complementary approach of sulphur isotope data and trace element concentration provides an effective cross-check on the physicochemical solution parameters prevailing at the time of sulphide mineralization. This approach is not only useful for computation of o34Stotalbut initial metal concentrations in the ore-forming solution may also be estimated. This information, in turn, should prove useful for isotope mass balance calculations for emissions to the atmosphere as these deposits are commercially developed. 8.11 POTENTIAL RESEARCH AREAS FOR SULPHUR ISOTOPE GEOCHEMISTRY IN NIGERIA*

8.11.1 Introduction The application of isotope analyses to the solution of environmental problems in Nigeria is in a rudimentary stage. Anthropogenic sulphur is abundant.

B.C.E. Egboka, M.M. Irogbenachi, and C.A. Eligwe.

410
0

Stable Isotopes

// // /
~!L'?~
CuSFeS4 -------10
~!L,p')

CuFe~

~~-FeS

-20
C\J
(f) 0> 0

( ) = pH

[ ]

-40 -50

-40 log f02

Figure 8.42 Distribution coeffJ.cient Dpb of lead plotted on the fs - fo diagram for MTs = 0.01 M, T = 250C, for the sulphide deposit at Kalya& (hatched area)

Research difficulties include low funding, lack of equipment, insufficient library facilities, and lack of international communication. Despite these problems, isotope geochemical research has an excellent future. Even though there are numerous geologic occurrences of S, there is little information on their isotopic compositions. This means that a realistic estimate for flux of sulphur in the Nigerian environment is not feasible at the present time.

Case Studies and Potential Applications

411

Research in this area must be encouraged and supported in order to make sulphur isotope data available to fit into models of the global sulphur budget. It must be emphasized that the United Nation's agencies and many international organizations have been playing positive roles in encouraging both regional and local research activities. The developing nations badly need such moral and financial encouragement to carryon research. Nigeria has a fairly large number of trained professionals in the Earth Sciences. Many of these indigenous experts have international training and research experience. They are suited to the task since they know the environment best and their devotion and patriotism would be significant driving factors. Nigeria has many geographic and geological features (Reyment, 1965; Figure 8.43). There are extensive coastal regions and geographical and geological similarities to parts of South America. Unique tropical and environments similar to the Sub-Sahel region extend far beyond her borders. It will be exciting to know whether environmental research carried out in other parts of the world can complement programmes in Nigeria.

8.11.2 Research areas in Nigeria Isotope geochemical research has yet to take off fully in Nigeria despite the fact that potential research areas abound. Figure 8.43 illustrates the extent to which the geology is krtown. Some isolated studies using 180, 2H, 3H, 14C, and BC have been conducted (United Nations Development Project, 1972; Oteze, 1981; Egboka, 1981a, 1981b). So far, there are no known records of where sulphur isotopes have been used.

8.11.3 Stable sulphur isotopes in basic geochemical research 8.11.3.1 Basement complex rocks Igneous and metamorphic rocks, 80% of which are Precambrian, cover about one-half of the landmass of Nigeria (Reyment, 1965). These rocks are highly fractured to form joints and faults. They have been subjected to intense weathering. The created sediments and soils are used for agricultural practices. The unconsolidated sediments are often water-bearing units or unconfined and confined aquifers that are tapped for domestic water supply and agriculture. Stable isotope analyses with emphasis on sulphur impoverishment, weathering processes, and soil pedogenesis would be particularly beneficial. The study of the hydrogeochemistry and biogeochemistry of many local ecosystems will be difficult without isotopic measurements.

~ ...... tV

\ \
~.A8EOKUTA

GULF OF GUINEA Figure 8.43 The geologic map of Nigeria (adapted from Reyment, 1965)

v, i:; <::J-

"

.g '" '"

c::;c

Case Studies and Potential Applications

413

8.11. 3.2 Lead-zinc mineralization and salt deposits As outlined by Kogbe and Obialo (1975), there are extensive deposits of lead and zinc sulphides in the Cretaceous sediments. Base metal deposits of economic value occur in the Abakaliki and Ogoja areas. Lead and zinc are mined in the towns of Enyigba, Ameri, and Ameka in the Abakaliki area. Potentially, sulphur isotope studies can (a) delineate the paragenesis of ores in these deposits and (b) trace the fate of SOz released during the smelting of these ores in the environment. Evaporite deposits occur in parts of Southeastern Nigeria within the Albian and Turonian marine (and continental) sedimentary formation of the Benue Trough (Uzuakpunwa, 1981). Their presence is marked by isolated salt ponds where salt-bearing horizons are faulted and contact predominantly shale formations. The evaporite zone which stretches from Abakaliki in Anambra State north-eastwards, to the Zurak area in Gongola State, is adjacent and nearly subparallel to Pb--Zn occurrences in eastern Nigeria. Early researchers have therefore considered the Pb--Zn mineralization and evaporite formation to be genetically related (Offodile, 1976; Orajaka, 1972). Isotopic data would bear directly on this interpretation. Sulphur occurs in most parts of Nigeria. Minor sulphide deposits occur in metagabbroic rocks exposed in the Egbe area of Western Nigeria (Bafor, 1981). The ore mineral consists of pyrrhotite, ilmenite, pentlandite, chalcopyrite, and pyrite. This Sudbury-type association, occurring also in similar host rocks, is being reported for the first time from the Nigerian basement complex. The paragenesis of these minerals can be evaluated using isotopic techniques. 8.11.3.3 Agulu-Nanka gully complex sulphur deposits The Agulu-Nanka gully complex began as an erosion feature around 1850. Over a time span of 135 years, the gully has expanded to an area of about 1100 kmz, greatly destroying ancestral homes and agricultural lands. The influence of geology and geotechnical factors (Egboka and Nwankwor, 1982) and iron and sulphur geochemistry have been described for this expansion. Gypsum and pyrite are found on the walls of the gully. At the bottom of the gully, seepage ground water flows as small streams, and a strong smell of hydrogen sulphide exists. Deposits of amorphous sulphur and pyrite are observed in the shale/mudstone layers. Isotopic studies should provide clearer explanations of the biogeochemistry of these deposits. 8.11.4 Fossil fuel deposits Sulphur and its compounds are found in coal and lignite, petroleum and natural gas, oil shales and tar sands, all of which occur extensively in Nigeria.

414

Stable Isotopes

8.11.4.1 Coal and mine drainage Sulphur in coal is an undesirable impurity because it affects the metallurgical usage of coal and is a major source of pollutant sulphur in the environment. The main sulphur compound in Nigerian seams is pyrite. Minor amounts of chalcopyrite (CuFeSz), pyrrhotite (FeS1.l)' chalcocite (CuzS), cubanite (CuFeS3)' sphalerite (ZnS), and galena (PbS) are known to be present. Nigeria has extensive deposits of coal and lignites (Simpson, 1954). The major coal deposits include the Tertiary hard brown coals of Afuze in the Niger Delta. It is lignitic, has high volatile content, is non-coking, and contains 1.5% sulphur. The Maestrichtian subbituminous to high volatile coals of Enugu in Anambra State and Okaba in Benue State are also noncoking and have S contents just less than 1%. The medium to high volatile coals of the Obi deposit in the Middle Benue Valley have proved to be moderately coking and contain 2-4% sulphur (Bello and Okeudo, 1981). The characteristics of some Nigerian coals are shown in Table 8.52. The exploitation of coals has resulted in severe acid mine drainage problems. About 1.36 x 104 m3 of acidic water is continuously pumped out of the Onyeama Mine into the Ekulu River daily. The hydrogeochemical parameters of the mine waters are given in Table 8.53. The mine water has a pH of 3.70 and a total iron content of 5.6 mg -1. Water from the nearby Iva Mine has a pH of 2.55, total iron content of 56.0 mg -1, and sulphate concentration of 428 mg -1. The source(s) of the ground water is still conjectural. It either originates from the overlying Ajali sandstone aquifers or the fractured zone and sandstone members of the Mamu formation aquifers. Stable and radioactive isotope studies of the waters of the two

Table 8.52 Composition of Nigerian coals (%) FORMA nON MAMU Orkupa Achimodo River to 18 to 36 to 8 to 42 0.5 1.4 to 2.0 14 to 15 14 34 6 40 Ezimo Lyinzu Stream 10 to 14 40 to 42 6 to 12 38 to 41 0.7 to 1.0 1.8 to 1.9 13 to 15 Nsukka 6 to 8 29 to 36 7 to 29 32 to 50 1.1 to 5.9

Deposit Moisture Volatiles Ash Fixed C* Sulphur


Nitrogen * *

Ogbakili 3 to 7 25 to 40 6 to 42 25 to 48 0.6 to 1.4 1.2 to 1.7 9 to 18

Oxygen

.* Ultimate analyses, ash-free, dry

Proximate analyses, air dried

Case Studies and Potential Applications Table 8.53 Hydrogeochemistry of Enugu coal mine groundwaters Location Odour Appearance (Hazen units) pH Dissolved Oz (mg 1-1) K+ (mg 1-1) Na+ (mg 1-1) 0- (mg 1-1) SOi- (mg 1-1) Total hardness (mg I-I) CaC03) Total Fe (mg 1-1) SiOz (mg 1-1) Onyeama mine Odourless 5 3.70 6 1.38 6.38 4.0 30 5.6 20 Ribadu mine Odourless 15 4.05 4 3.00 8.20 10.2 12 6.7 25

415

Iva Valley mine Odourless 5 2.55 4 9.88 3.50 20.0 428 80 56.0 60

formations should solve the groundwater source problem. Similar problems also exist in the other hydrogeological provinces. Sulphur isotope geochemical studies of the coal and the acid mine drainage problems would provide answers to the following problems: (a) The origin of the various forms of sulphur in the Nigerian coals (b) The source(s) of the sulphate acidity in the Enugu coal mine aquifers (c) The variability of sulphur isotopes in coal of one lithologic unit, of different lithologies but of the same age, and of different lithologies of different ages (d) The behaviour and fate of pollutant SOz released by these coals and lignite when they are combusted (e) The source materials, origin, and genesis and maturation of the coal deposits

8.11. 4.2 Petroleum and natural gas Nigeria has major deposits of petroleum and natural gas, the more economically oil-producing strata occurring in the Miocene. However, oil and gas seepages have been reported in older deposits such as the Owelli sandstone of Maestrichtian age. The Nigerian oil is light crude with a low sulphur content. Gas is currently wasted by flaring, but plans for its utilization are under discussion. Petroleum and natural gas production is a major source of environmental pollution. Oil spills on land, rivers, lakes, streams, marshes, and vegetation by the oil companies are a common feature. Studies in other parts of the

416

Stable Isotopes

world suggest that the isotopic composition of the sulphur can be used to tag the oil and subsequently follow its dispersal in the environment. Sulphur and other stable isotope studies can potentially provide information on the following: (a) Differentiation of reservoirs of different ages and formations (b) Trace oil maturation and migration from source rock to reservoir rock (c) Other environmental impacts of pollution caused by the production of crude oil and natural gas 8.11.4.3 Oil shales and tar sands Large-scale deposits of oil shale and tar sands have been discovered in areas of Western Nigeria. The areal extent, oil reserve, and geochemistry are being established. Isotopic studies can assist in ascertaining their origin and the potential environmental impacts of their utilization. 8.11.5 Emission of snlphur In Nigeria, anthropogenic sources of sulphur include the use and burning of coal, the iron and steel industry as found in Ajaokuta and Aladja, oil refineries at Warri, Port Harcourt, and Kaduna, gas flaring at Bonny, Warri, and Port Harcourt areas, and fertilizer applications. No attempt has thus far been made to apply sulphur isotope techniques in studying these various aspects of sulphur pollution in Nigeria. 8.11.6 ()ther research areas In addition to identifiable natural and anthropogenic sources of S compounds in Nigeria, there are other research areas where isotopic studies should prove useful. These include: (a) The interplay of the Saharan and Atlantic air masses (Garnier, 1967) (b) The release of hydrogen sulphide and dimethyl sulphide from tropical coastal waters, lakes, and wetlands (c) The release and uptake of sulphur by tropical vegetation (d) The dynamics of sulphur in the Savannah grasslands (e) Acidity of tropical rains and their impacts on tropical flora and fauna (f) Long-range transport of sulphur into and out of the country Numerous other areas of research can be cited but the list is adequate to emphasize the potential of stable isotope techniques in the study of biogeochemical phenomena in unique tropical ecosystems.

Case Studies and Potential Applications

417

REFERENCES
Abusada, S.M. (1981). Interpretation of environmental isotope data of Kuwait groundwater. J. Gulf and Arabian Penin. Studies, 2, 80-92. 'Australian Energy Statistics 1982' (1982). Department of National Development and Energy, Canberra. Ayres, D.E., Burns, M.S., and Smith, J.W. (1979). Sulphur-isotope study of the massive sulphide ore body at Woodlawn, New South Wales. J. Geolog. Soc. Austral., 26, 197-202. Bafor, B.E. (1981). The occurrence of sulphide mineralization in the Egbe area of Southwestern Nigeria. J. Min. Geo!., 18, 175-9. Barnes, H.L. (1979). Geochemistry of Hydrothermal Ore Deposits, 2nd edn, Wiley-Interscience, New York. Bauld, J., Chambers, L.A., and Skyring, G.W. (1979). Primary productivity, sulfate reduction and sulfur isotope fractionation in algal mats and sediments of Hamelin Pool, Shark Bay, W.A. Austral. J. Marine and Freshwater Res., 30, 753-64. Bello, A.O., and Okeudo, O.E. (1981). Coal selection blends and blast furnace coke quality production in Nigerian steel industry. J. Min. Geol., 18, 180-5. 'Black Coal in Australia 1981-82' (1982). Joint Coal Board, Sydney, Australia. Both, R.A., and Smith, J.W. (1975). A sulphur isotope study of base-metal mineralization in the Willyama Complex, Western New South Wales, Australia. Econ. Geol., 70, 308-18. Builov, V.V., and Builova, LV. (1976). Biogeochemical cycles of sulphur isotopes in soils. In: Proceedings of 1974 va Plenum of Scope in Moscow, Nauka, Moscow, pp. 258-64. Butler, G.P., Krouse, H.R., and Mitchell, R. (1973). Sulphur-isotope geochemistry of an arid, supratidal evaporite environment, Trucial Coast. In: Purser, B.H. (Ed.) The Persian Gulf: Holocene Carbonate Sedimentation and Diagenesis in a Shallow Epicontinental Sea, Springer-Verlag, New York, pp. 453-462. Cartwright, B., Merry, R.H., and Tiller, K.G. (1976). Heavy metal contamination of soils around a lead smelter at Port Pirie, South Australia. Austral. J. Soil Res., IS, 69-81. Chambers, L.A. (1982). Sulfur isotope study of a modern intertidal environment and the interpretation of ancient sulfides. Geochim. Cosmochim. Acta, 46, 721-8. Chambers, L.A., and Trudinger, P.A. (1979). Microbiological fractionation of stable sulfur isotopes: a review and critique. Geomicrobiol. J., 1, 249-93. Chapman, R.W. (1971). Climatic changes and the evolution of landforms in the eastern province of Saudi Arabia. Geol. Soc. Am. Bull., 88, 2713-28. Clarke, D.K. (1975). Regional Bore Survey, Southern Pirie Torrens Basin and Neighbouring Areas. In: Report Book 75/121, Geological Survey Engineering Division, Department of Mines, South Australia, 48 pp. Collins, P.L.F. (1981). The geology and genesis of the Cleveland tin deposit, western Tasmania: fluid inclusion and stable isotope studies. Econ. Geol., 76, 365-92. Deb, M., and Kumar, R. (1982). The volcano-sedimentary environment of RajpuraDariba polymetallic ore deposit, Udaipur District, Rajasthan. In: Proceedings of Symposium on Metallogeny of the Precambrian, Geological Society of India, Bangalore. Donnelly, T.H., Lambert, LB., Oehler, D.Z., Hallberg, J.A., Hudson, D.R., Smith, J.W., Bavinton, O.A., and Galding, L. (1978). A reconnaissance study of stable isotope ratios in Archaean rocks from the Yilgarn Block, Western Australia. J.

Ce%g. Soc. Austral., 24, 409-20.

418

Stable Isotopes

Driessen, A. (1980). Sulphur. In: Australian Mineral Industry Annual Review 1980, pp. 247-50. Eadington, P.J., Smith, J.W., and Wilkins, R.W.T. (1974). Fluid inclusion and sulphur isotope research, Mt. Morgan, Queensland. Australasian Institute of Mining and Metallurgy, Conference, Southern and Central Queensland, July 1974, pp. 441-4. Egboka, B.C.E. (1981a). Distribution patterns of bomb tritium, chloride, sulphate, oxygen-18 and deuterium in two shallow sand aquifers. Geochem. J., 15,305-14. Egboka, B.C.E. (1981b). Multitechnique approach using environmental tracers in hydrogeological studies. J. Min. Geol., 18, 96-102. Egboka, B.C.E., and Nwankwor, G.I. (1982). The Agulu-Nanka gully: an explanation for its origin. Q. J. Engng. Geol., Lond. Eriksson, E. (1963). The yearly circulation of sulphur in nature. J. Geophys. Res., 68, 4001-8. Ferguson, J., Burne, R.V., and Chambers, L.A. (1982). Lithification of peritidal carbonates by continental brines at Fisherman Bay, South Australia, to form a megapolygon/spelean limestone association. J. Sediment. Petrol., 52, 1127-47. Fleischer, E., Goldberg, M., Gat, J.R., and Magaritz, M. (1977). Isotopic composition of formation waters from deep drillings in southern Israel. Geochim. Cosmochim. Acta, 41, 511-25. Folk, R.L. (1973). Evidence for peritidal deposition of Devonian Caballos Novaculite, Marathon basin, Texas. Am. Assoc. Petrol. Geol. Bull., 57, 702-25. Forrest, J., and Newman, L. (1973). Sampling and analysis of atmospheric sulfur compounds for isotope ratio studies. Atmos. Environ., 7, 561-73. Forrest, J., and Newman, L. (1977). Further studies on the oxidation of sulfur dioxide and coal-fired power plant plumes. Atmos. Environ., 11, 465-74. Forrest, J., Klein, J.H., and Newman, L. (1973). Sulfur isotope ratios of some power plant flue gases: a method for collecting the sulfur-oxide. 1. Appl. Chem. Biotechnol., 23, 855-63. Garnier, B.J. (1967). Weather conditions in Nigeria. Climatological Research Series 2, McGill University, Montreal, Canada. Gat, J.R., and Issar, A. (1974). Desert isotope hydrology and water sources of the Sinai Desert. Geochim. Cosmochim. Acta, 38, 1117-31. Gauri, K. Lal, and Holdren, G.O. Jr (1981)., Pollutant effect on stone monuments. Environ. Sci. Tech., 15 (4), 386-90. Gibson, D.L. (1981). Oil shale in Australia-its occurrence and resources. AMI Q., 33 (3), 105-13. Green, G.R., Solomon, M., and Walshe, J.L. (1981). The formation of the volcanichosted massive sulfide ore deposit at Rosebery, Tasmania. Econ. Geol., 76, 304-38. Grinenko, V.A., and Grinenko, L.N. (1974). Geochemistry of Sulphur Isotopes, Nauka, Moscow, 272 pp. Harris, D.C., and Nickel, E.H. (1972). Pentlandite compositions and associations in some mineral deposits. Can. Mineral., II, 861-78. Haur, A., Hladikova, J., and Smejkal, V. (1973). Procedure of direct conversion of sulphate into S02 for mass spectrometer analysis of sulphur. Isotopenpraxis, 9, 329 pp. Hitchon, B., and Krouse, H.R. (1972). Hydrogeochemistry of the surface waters of the Mackenzie River drainage basin, Canada: III. Stable isotopes of oxygen, carbon and sulphur. Geochim. Cosmochim. Acta, 36, 1337-57. Hokr, Z., Kubant, J., and Hladikova, J. (1972). /)34Sof coal from NW Bohemia

Case Studies and Potential Applications

419

mines (in Czech). Research report, Geofond, Prague. Howarth, R.W., and Marino, R. (1984). Sulfate reduction in salt marshes, with some comparisons to sulfate reduction in microbial mats. In: Cohen, Y., Castenholtz, R.W., and Halvorsen, H.O. (Eds.) Microbiological Mats: Stromatolites, Alan R. Liss Inc., New York, pp. 245-63. Hulston, J.R., Taylor, CB., Lyon, G.L., Stewart, M.K., and Cox, M. (1981). Environmental isotopes in New Zealand hydrology. 2-Standards, measurement techniques, and reporting of measurements for oxygen -18, deuterium, and tritium in water. N.Z. J. Sri., 24, 313-22. Hyne, J.B. (1981). Recent developments in sulfur production from hydrogen sulfide containing gases. Presented at the 181st ACS Conference held in Atlanta, Georgia, 29 March-3 April 1981. Hyne, J.B. (1983). Hydrogen-sulphide-the basis of the Canadian sulphur industry. Alberta Sulphur Res. Ltd Q. Bul/.XX (1), April-June, 13-25. Jorgensen, B.B., and Cohen, Y. (1977). Solar Lake (Sinai). 5. The sulphur cycle of the benthic cyanobacterial mats. Limnol. Oceanogr., 22, 657-66. Karunakaran, C (1976). Sulphur isotopic composition of barite and pyrite from Mangampeta, Cuddapah District, Andhra Pradesh. J. Geol. Soc. India, 17, 181 pp. Knutson, J., Donnelly, T.H., and Tonkin, D.G. (1983). Geochemical constraints on the genesis of copper mineralization in the Mount Gunson Area, South Australia. Econ. Geol., 78, 250-74. Kogbe, CA., and Obialo, A.U. (1975). Statistics of mineral production in Nigeria (1946 to 1974) and the contribution of the mineral industry to the Nigerian Economy. In: Kogbe, C.A. (Ed.) Geology of Nigeria, Elizabethan Publishing Co., Lagos, Nigeria, 436 pp. Krouse, H.R. (1976). Sulphur isotope variations in thermal and mineral waters. Proceedings of International Symposium on Water-Rock Interaction, IAGC, Prague, 1974, pp. 340-7. Krouse, H.R. (1977). Sulphur isotope abundances elucidate uptake of atmospheric sulphur emissions by vegetation. Nature, 265, 45-6. Krouse, H.R., and Case, J.W. (1981). Sulphur isotope ratios in water, air, soil, and vegetation near Teepee Creek Gas Plant, Alberta. Water, Air, Soil Pol/ut., 15, 11-28. Krouse, H.R., and Case, J.W. (1983). Sulphur isotope abundances in the environment and their relation to long-term sour gas flaring, near Valleyview, Alberta. Prepared for Alberta Environment, Research Management Division and Pollution Control Division, Air Quality Control Branch by Department of Physics, University of Calgary. RMD Report 83/18, 110 pp. Krouse, H.R., Legge, A.H., and Brown, H.M. (1984). Sulfur gas emissions in the boreal forest: the West Whitecourt case study, V. Stable sulfur isotopes. Water, Air, Soil Pol/ut., 22, 321-47. Krouse, H.R., Viau, CA., Eliuk, L.S., Ueda, A., and Halas, S. (1987). Chemical and isotopic evidence of thermochemical sulphate reduction by light hydrocarbon gases in deep carbonate reservoirs. Nature, 333, 415-19. Kusakabe, M., Rafter, T.A., Stout, J.D., and Collie, T.W. (1976). Isotopic ratios of sulphur extracted from some plants, soils, and related materials. N.Z. J. Sci., 19, 433-40. Longinelli, A., and Bartelloni, M. (1978). Atmospheric pollution in Venice, Italy, as indicated by isotopic analyses. Water, Air, Soil Pol/ut., 10, 335 pp. Lowe, L.E., Sasaki, A., and Krouse, H.R. (1971). Variation of 534/532in soil

420

Stable Isotopes

fractions in Western Canada. Can. J. Soil Sci., 51, 129-31. Lyons, W.B., Hines, M.E., and Gaudette, H.E. (1984). Major and minor element pore water geochemistry of modern marine sabkhas: the influence of cyanobacterial mats. In: Cohen, Y., Castenholtz, R.W., and Halvorsen, H.O. (Eds.) Microbiological Mats: Stromatolites, Alan R. Liss Inc., New York, pp. 411-23. McKenzie, J.A, Hsu, K.J., and Schneider, J.F. (1980). Movement of subsurface waters under the sabkha, Abu Dhabi, UAE, and its relation to evaporative dolomite genesis. In: Zenger, D.H., Dunham, J.B., and Ethington, R.L. (Eds.) Concepts of Models of Dolomitization, SEPM Special Publication, 28 pp. Manuell, R.W. (1980). Sulfur content of Australian oil and petroleum products. In: Freney, J.R. and Nicholson, AJ. (Eds.) Sulfur in Australia, Australian Academy of Science, Canberra, pp. 118-32. Menon, AG. (1979). Sulphur isotope and trace element studies on sulphide deposits of Dharwar Craton. Unpublished Ph.D. thesis, Indian Institute of Science, Bangalore, India. Menon, A.G., Venkatasubramanian, V.S., and Anantha Iyer, G.V. (1981a). Sulphur isotope abundance variations in sulphides of the Dharwar Craton-part I: Kalyadi. J. Geol. Soc. India, 22, 391. Menon, AG., Venkatasubramanian, V.S., Vasudev, V.N., and Anantha Iyer, G.V. (1981b). Sulphur isotope abundance variations in sulphides of the Dharwar Cration-part III: Hutti. J. Geol. Soc. India, 22, 448. Mock, CM., and Roarty, M.J. (1980a). Lead. In: Australian Mineral Industry Annual Review, 1980, pp. 162-76. Mock, CM., and Roarty, M.J. (1980b). Zinc. In: Australian Mineral Industry Annual Review 1980, pp. 292-304. Nedwell, D.B., and Abram, J.W. (1978). Bacterial sulfate reduction in relation to sulfur geochemistry in two contrasting areas of saltmarsh sediment. Est. Coast. Marine Sci., 6, 341-51. Nehring, N.L., Bowen, P.A, and Truesdell, A.H. (1977). Technique for the conversion to carbon dioxide of oxygen from dissolved sulfate in thermal waters. Geothermics, 5, 63 pp. Newman, L., Smith, M., Forrest, J., Tucker, W., and Manowitz, B. (1971). A tracer method using stable isotopes of sulfur. In: Vogt, J.R. (Ed.) Proceedings of Nuclear Society Topical Meeting on Nuclear Methods in Environmental Research, Columbia, pp. 16-25. Newman, L., Forrest, J., and Manowitz, B. (1975a). The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulfur dioxide in the plume from an oil-fired power plant. Atmos. Environ., 9, 959-68. Newman, L., Forrest, J., and Manowitz, B. (1975b). The application of an isotopic ratio technique to a study of the atmospheric oxidation of sulfur dioxide in the plume from a coal-fired power plant. Atmos. Environ., 9, 969-77. Novak, I. (1982). Air pollution in Czechoslovakian emission regions (in Czech). Ochr. Ouzdusi, 14 (10), 145-7. Offodile, M.E. (1976). A hydrogeochemical interpretation of the Middle Benue and Abakaliki brine fields. J. Min. Geol., 13, 79 pp. O'Heare, J. (1980). Sulfur dioxide emission from power stations in Victoria. In: Freney, J.R. and Nicholson, A.J. (Eds.) Sulfur in Australia, Australian Academy of Science, Canberra, pp. 33-41. Ohmoto, H. (1972). Systematics of sulphur and carbon isotopes in hydrothermal ore deposits. Econ. Geol., 67, 551. Orajaka, S.O. (1972). Salt water resources of East Central State of Nigeria. Unpublished report. Late Professor of Geology, Department of Geology, University of Nigeria, Nsukka.

Case Studies and Potential Applications

421

Oteze, G.E. (1981). Preliminary report on the environmental isotope hydrology of the Main Rima aquifer waters. 17th Annual Conference of the Nigerian Mining and Geosciences Society, Calabar, Nigeria. Perkin, D.J. (1980). Copper. In: Australian Mineral Industry Annual Review 1980, pp. 103-10. Pratt, R. (1980). Nickel. In: Australian Mineral Industry Annual Review 1980, pp. 193-200. Radhakrishna, B.P. (1976). Mineralization episodes in the Dharwar Craton of Peninsular India. J. Geol. Soc. India, 17, 79 pp. Reyment, R.A. (1965). Aspects of Geology of Nigeria, Ibadan University Press, 145 pp. Robinson, B.W. (1976). The origin of mineralization at the Tui Mine, Te Aroha, New Zealand, in the light of stable isotope studies. Econ. Geol., 69, 910-25. Robinson, E., and Robbins, R.C. (1970). Gaseous sulfur pollutants from urban and natural sources. J. Air Pollut. Control Assoc., 20, 233-5. Romano, A.J., Klein, J.H., and Newman, L. (1975). An air sampling system to measure power plant effluents using a lightweight aircraft. J. Environ. Systems, 5, 271-89. Rouse, J.E., and Sherif, N. (1980). Major evaporite deposition from groundwater remobilized salts. Nature, 285, 470-2. Rye, R.O., and Ohmoto, H. (1974). Sulphur and carbon isotopes. A review. Econ. Geot., 69, 826 pp. Safonov, Yu. G., Radhakrishna, B.P., Krishna, Rao B., Vasudev, V.N., Krishnam, Raju K., Nosik, L.P., and Pashkov, Y.N. (1980). Mineralogical and geochemical features of endogene gold and copper deposits of South India. J. Geol. Soc. India, 21, 365 pp. Sakai, H. (1968). Isotopic properties of sulphur compounds in hydrothermal processes. Geochem. J., 2, 29. Scott, K.M., Smith, J.W., Sun, S.S., and Taylor, G.F. (1985). Proterozoic copper deposits in Northwest Queensland, Australia: Sulphur isotopic data. Mineralium Deposita, 20, 307-309. Seccombe, P.K., Groves, D.l., Binns, R.A., and Smith, J.W. (1978). A sulphur isotope study to test a genetic model for Fe-Ni sulphide mineralisation at Mt. Windarra, Western Australia. In: Robinson, B.W. (Ed.) Stable Isotopes in the Earth Sciences, DSIR Bulletin 220, pp. 287-300. Seccombe, P.K., Groves, D.l., Marston, R.J., and Barrett, F.M. (1981a). Sulfide paragenesis and sulfur mobility in Fe-Ni-Cu sulphide ores at Lunnon and Juan Main Shoots, Kambalda: textural and sulfur isotopic evidence. Econ. Geol., 76, 1675-85. Seccombe, P.K., Subbarao, K.V., and Pawar, J.N. (1981b). Sulphur isotopic composition of Ingaldhal sulphides, Karnataka State, India. J. Geol. Soc. India, 22, 326. Shampine, W.J., Dincer, T., and Noory, M. (1978). An evaluation of isotope concentrations in the groundwater of Saudi Arabia. In: Proceedings of the International Symposium on Isotope Hydrology, Vol. 2, IAEA, Proc. Ser. STI/ PUB/493, IAEA-SM228/23, pp. 443-63. Simpson, A. (1954). The Nigerian coalfield. The geology of parts of Onitsha, Owerri, and Benue Provinces. Geological Survey of Nigeria Bull., 24. Skyring, G.W., Chambers, L.A., and Bauld, J. (1983). Sulfate reduction in sediments colonized by cyanobacteria, Spencer Gulf, South Australia. Austral. 1. Marine and Freshwater Res., 34, 359-74. Smith, G .E. (1976). Sabkha and tidal-flat facies control of stratiform copper deposits in North Texas. In: Wolf, K.H. (Ed.) Handbook of Strata-Bound and Stratiform

422

Stable Isotopes

Ore Deposits, Vol. II, Regional Studies and Specific Deposits (Vol. 6, Ch. 8), Elsevier Scientific Publishing Company, Amsterdam, pp. 407--46. Smith, J.W., and Batts, B.D. (1974). The distribution and isotopic composition of sulfur in coal. Geochim. Cosmochim. Acta, 38, 121-31. Solomon, P.J. (1965). Investigations into sulfide mineralisation at Mount Isa, Queensland. Part 1: textural studies of the ores. Econ. Geol., 60, 737-65. Solomon, M., Rafter, T.A., and Jensen, M.L. (1969). Isotope studies on the Rosebery, Mount Farrell and Mount Lyell ores, Tasmania. Mineralium Deposita, 4, 172-99. Stanton, R.L., and Rafter, T.A. (1966). The isotopic composition of sulphur in some stratiform lead-zinc sulphide ores. Mineralium Deposita, 1, 16--19. Sun, S.S. (1983). Implications of Sand Pb isotope data to genesis of massive sulphide in the Cobar area. 6th Australian Geological Convention, Canberra, Geological Society of Australia. Abstracts 9, 307-309. Tanaka, S., and Hashimoto, Y. (1977). Studies on the behaviour of sulfur compounds (sulfur dioxide, sulfur trioxide, and hydrogen sulfide) in atmospheric air (in Japanese). Nippon Kagaku Kaishi, 5, 712-15. Tucker, W.D. (Ed.) (1969). The atmospheric diagnostics program at Brookhaven National Laboratory: Second Status Report, BNL 50206 (T-553). United Nations Development Project (1972). Hydrogeological Interpretation from Environmental Isotope Data in the Chad Basin, UNDP Special Fund Project AFR.RE. 79, Chad, IAEA, Vienna. Uzuakpunwa, A.B. (1981). The geochemistry and origin of the evaporite deposits in the Southern half of the Benue Trough. Earth Evo/. Sci., 1, 136--8. Venkatasubramanian, V.S., Setty, S.R., Anantha Iyer, G.V., and Menon, A.G. (1981a). Sulphur isotope abundance variations in sulphides of Dharwar Cratonpart II: Ingaldhal. J. Geo/. Soc. India, 22, 395. Venkatasubramanian, V.S., Setty, S.R., Anantha Iyer, G.V., and Menon, A.G. (1981b). Sulphur isotopic environment and trace element distribution in sulphur deposits of the Dharwar Craton. Presented at the National Symposium on Mass Spectrometry, Bombay. Walshe, J.L., and Solomon, M. (1981). An investigation into the environment of formation of the volcanic-hosted Mt. Lyell copper deposits using geology, mineralogy, stable isotopes and a six-component chlorite solid solution model. Econ. Geol., 76, 246--84. Zehnder, A.J.B., and Zinder, S.H. (1980). The sulphur cycle. In: Hutzinger, O. (Ed.) The Natural Environment and the Biogeochemical Cycles (The Handbook of Environmental Chemistry, Vol. 1, part A, Springer-Verlag, New York, pp. 105-45.

Appendix

A. The 834Sdistributions for sulphate in Broughton River* and bore waters, Spencer Gulf Location

834SS0/(%0) +20.1 +19.0 +16.9 + 16.4 +19.0 +19.1 +19.7 +18.2 +17.8 +20.8 +18.2 +15.5 +15.6 +9.0 +11.5 +10.7 +14.0

8180H20

(%0) -4.4 -6.2 -5.5 -5.9

8D (%0) -34.3 -31.3 -37.7 -39.0

Na+/CI(meq) 0.83 1.03 2.85 2.29 0.87 1.97 0.61 0.68 0.87 0.82 0.84 0.75 0.89 0.76 0.71 0.93 0.82

CI-/SO/(meq) 6.8 4.4 2.7 2.6 12.0 5.2 4.3 6.8 9.6 7.1 10.1 4.2 13.0 10.7 4.4 5.4

SO/(meq -1) 1 6 2 2 5 3 1 2 2 4 5 10 5 1 1 8 25

Mambray Creek

Baroota

Port Pirie

-5.2 -4.3

-32.0 -29.2

*Wood Point

-3.6 -3.6

-25.0 -26.0

Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko @ 1991 SCOPE, Published by John Wiley & Sons Ltd

424

Stable Isotopes

B. The 834Sdistributions for sulphate in peritidal ground waters, Spencer Gulf 834S (%0)
81XO

8D
(0/00)

Location

(%0)

Na+/Cl(meq)

Cl-/SO/(meq)

SO/(meq f-)

Red Cliff Point

Mambray Creek

Wood Point

Fisherman Bay

+18.6 +20.7 +19.3 +19.5 +18.7 +18.1 +18.3 + 16.4 +18.1 +20.5 +19.9 +20.5 +20.6 +16.0 +16.1 +19.1 +15.0 +15.0 +13.5 +13.5 +13.5 +16.8 +20.6 +20.0 +15.9 +15.9 +15.0 +15.5 +15.3 +15.7 +19.6 +17.7

-4.2 -4.4 -2.9 -4.4

-31.7 -37.4 -18.2 -20.7

+4.7 +5.5

+12.0 +15.6

+3.12

+2.3 -17.5

-0.11 -2.5 -20.4

0.93 0.98 1.03 0.95 0.94 0.84 0.79 0.75 0.84 0.83 0.85 0.84 0.86 0.78 0.80 0.83 0.82 0.78 0.75 0.77 0.78 0.83 0.84 0.84 0.81 0.83 1.04 0.82 0.81 0.79 0.85 0.84

8.3 6.7 8.5 6.8 7.9 8.6 17.4 23.5 11.7 9.8 11.6 10.6 9.7 8.5 8.2 14.7 8.9 10.7 10.7 10.7 10.9 14.8 10.7 9.7 7.6 7.8 9.2 9.1 9.2 9.7 13.0 15.8

127 718 45 110 35 79 173 153 178 319 237 80 95 242 240 204 236 335 349 275 336 224 373 377 56 82 386 226 67 140 266 63

Appendix

425

C. The o34Sdistributions for sulphate in peritidal ground waters, Shark Bay oD (%0) +24 +27 -17 +7 +11 -24 -4 -5 -8 -11 -6 0 -4 -3 -1 -23 -44 Na+/CI(meq) 0.92 0.87 0.87 0.91 0.89 0.98 0.87 0.85 0.87 0.88 0.97 0.95 0.91 0.92 0.88 0.87 0.95 0.98 1.15 SO/(meq -1)

Location Solar Ponds

034SS0/ (%0) +21.8 +22.0 +21.6 +20.8 +21.4 +21.4 +18.5 +20.3 +20.5 +20.0 +18.9 +18.9 +20.5 +19.9 +20.0 +20.9 +23.4 +17.9 +13.6

Ol80H20

(%0) +7.4 +8.1 +0.7 +3.3 +3.2 -1.1 +0.4 +0.8 +0.9 +2.2 +4.0 +3.8 +4.1 +2.9 +3.2 -1.5 -4.8

CI-/SO/(meq) 10.8 9.6 11.2 11.2 9.5 8.8 11.7 10.9 9.8 9.8 10.4 12.1 13.5 13.7 13.9 13.1 9.3 10.0 5.6

252 223 247 101 136 109 46 62 71 79 116 144 154 163 243 272 98 67 14

Nilemah Embayment

Playford Transect Flagpole Landing

CHAPTER 9

Summary
H.R. KROUSE AND V.A. GRINENKO

Summary statements have been made throughout the book, both in chapters devoted to the different 'spheres' and in individual case studies. There can be no doubt that stable sulphur and related oxygen isotope techniques are generally successful in investigations of anthropogenic sulphur in the environment. The isotope technique is best used for identifying the presence of pollutant sulphur in environmental receptors. Examples were cited where proportions of industrial sulphur in living trees and animals as well as surface layers of inanimate objects (e.g. monuments) have been successfully determined with isotope data. Isotope data are best utilized for long-term investigations. A short overexposure to H2S could prove lethal but there would be no discernable isotopic record of the cause because of the much greater sulphur reservoir in the victim. There are instances where short-term sampling is desirable, such as determining the dependence of 334S values of atmospheric sulphur compounds on wind direction. In such cases, 'yet to be developed', continuous sulphur isotope monitoring techniques would be ideal. Isotope techniques work best if the 334S value of anthropogenic sulphur is quite different from that of environmental receptors. Such is the case for emissions from sour gas processing in Alberta, Canada, where the isotopic differences (~334Svalues) between emitters and preindustrial receptors range from 20 to 50%0 (Section 8.2). However, examples are also given where successful interpretation seems possible with ~334S values of only 2-3%0. Such situations require a much larger data base for meaningful statistical analyses. The desirability of examining as many receptors as possible in an ecosystem cannot be overstressed. If one can have only a limited database because of fiscal constraints, lichens might be the first choice because they provide an integrated record of atmospheric sulphur. Next, leaves or needles might be examined since they have acquired sulphur from both the air and soil. If their 334Svalues are close to those of lichens, the system is either experiencing negligible pollutant sulphur or it has become dominated by industrial sulphur.
Stable Isotopes in the Assessment of Natural and Anthropogenic Sulphur in the Environment Edited by H.R. Krouse and V.A. Grinenko <9 1991 SCOPE, Published by John Wiley & Sons Ltd

428

Stable Isotopes

Sulphur content, normally available from sample processing, can be used to distinguish these extremes. If the 034S values of the leaves and lichens differ, the next logical step is an isotopic investigation of the soil. For a larger regional study, an alternative might be to analyse highest orders of the food chain. The hair of a grazing or browsing animal would provide an average value of isotopic composition for a wide area of grass and bushes. With such an approach, one has no information on the foliar S content since the S content of hair is relatively constant. Sulphur isotope data are best understood when used in combination with other techniques. Conversely, other techniques such as concentration data may prove fallible if not checked with isotope data. Increased sulphur contents in vegetation may not be due to the greater influence of emissions of a nearby industry (Section 7.2.5). Sulphur isotope data are necessary to resolve such anomalies. The supportive techniques may be isotopic as well as 'conventional'. In particular, many cases were cited where oxygen isotope abundances in sulphate proved informative. Their relationship with the 0180 values of local meteoric water is diagnostic of recent oxidation of lower valence state sulphur. Examples have also been given where the oxygen and hydrogen isotope compositions of water provide information on mixing, etc., in hydrospheric investigations of sulphur (e.g. Section 8.6). Many phenomena are classified as 'secondary effects'. The 'secondary disaster' of lake acidification due to the oxidation of fallout volcanic ash was cited in Section 8.5.3. There are questions as to the extent that natural fluxes are altered by anthropogenic additions. Emissions of gaseous sulphur compounds by plants may be enhanced by increased sulphur in the soil. Sulphate reduction in the uppermost layers of lake sediments may be activated by sudden SO/- additions (Sections 6.4.2.1). Secondary effects may extend beyond the sulphur cycle to influence the cycling of other elements. In turn anthropogenic inputs of other elements might influence sulphur cycling. The understanding of secondary effects requires many analytical approaches including isotope investigations. There are definite desirable avenues for future developments. Continuous isotopic monitoring of atmospheric or hydrospheric sulphur compounds would be ideal, but is unfortunately far from reality. Techniques have been developed for extracting trace sulphur from materials, e.g. teeth, and solid source mass spectrometry can be used to analyse submicrogram quantities (Section 3.1). The availability of 34S, 33S, and 36S-enriched compounds at reasonable costs would promote many relevant labelling experiments such as tracing long-range transport of atmospheric sulphur. Whereas isotopic data can be used to document the presence of anthropogenic sulphur in an environmental receptor, the influence (good or bad) of the pollutant on the receptor must be evaluated by biological

Summary

429

techniques such as species coverage. To this point, the relating of 'cause and effect' by isotope data or by any single technique has not been possible. The presence of industrial sulphur in foliage under stress does not necessarily mean that sulphur was the cause. The relation of effect to cause must embrace isotope data in combination with biological, biochemical, and physiological data. In summary, isotope data have proved extremely useful in many studies where anthropogenic sulphur has implications for ecosystems, health, corrosion of equipment, damage of cultural edifaces, and other diverse receptors. Successful applications have usually involved an isotope specialist in concert with participants from many disciplines. The team approach maximizes the return of knowledge for time invested.

Index

Acid rain (see Sulphate precipitation) Actinia, 186 Activated complex, 5 Adirondack mountains, USA, 21, 215 Aerodynamic sizing, 140 Aerosol, 36, 144, 399 Africa, 69, 88, 228, 409-14 African rivers, 228 Akansk-Achinsk, USSR, 155 Alberta, Canada, 140-1, 144, 154-5, 204, 247-8, 274-5, 278, 280, 284 (Fig.), 303, 309-31 Algae, 285, 289 Algal mats, 370, 382 Algoma, Canada, 199 Algonquin Provincial Park, Canada, 214 Altai, 88 Amazon river, 224 Amino acids, 268 Amu-Daria syncline, 111, 244 Analyses aerodynamic sizing, 140, 317 Kiba reagent, 184 O-isotopes, 55-9 sample pretreatment, 43-8 scanning electron microscopy, 141, 142 (Fig.), 316 sulphur atmosphere, 137-41, 312, 344-5 coal, 47 H2S in water, 344-5 petroleum, 47 soil, 268-9 trace-S, 184 volcanic fluxes, 117 Animals, 185-8, 299-303 Antarctica, 152

Aquatic organisms, 183-9 Arabia, 99 Archean, 14 Arctic, 151, 167,300 Argonne, Illinois, 34 Armavir, West Caucasus, 103 Asia, 69 Asse salt dome, Germany, 237 Astrakhan gas-condensate deposits, 112 Athabasca oil sands, 102 Atlantic Ocean, 152 Atmosphere, 34-7, 133-76, 383-97, 312-18, 331-43 aerodynamic sizing, 140 aerosols, 36, 144 anthropogenic-S, 147-9, 358-60, 383-97 biogenic sulphur, 144-7, 163-7, 242-51,344-7 carbon disulphide, CS2, 133-4, 165 carbonyl sulphide, COS, 133-4, 165 dimethyl sulphide, DMS, 17, 133-4, 165 dimethyl disulphide, DMDS, 133-4 dispersion, 168-71 dustfall, 400-1 electron microscopy, 142 (Fig.) emissions, power plant, 331-339, 400-2 geothermal areas, 246-8 gypsum dust, 68 high volume sampling arrays, 140 hydrogen sulphide, 17, 133-5, 154-5, 165, 344-7 methyl mercaptan, 133-4 monitoring with vegetation, 142 O-isotopes sulphate, 34-7, 157-58

--

432 Atmosphere O-isotopes (cont' d) sulphur dioxide, 37 water, 158 particle size, 140 particulates, 315-17 rain and snow, 150-5, 162 regional balance 159-62 relation to concentrations, 149, 153-7, 314 sampling of S-compounds, 44, 137-41 sea spray, 17 springs, 248-50 sulphate, 16, 34-7, 334-6 sulphur chemistry, 134-7 sulphur compounds, 133-8 sulphur dioxide, 37, 133-7, 154-5, 313-16, 333-8, 400 sulphur dust,135, 290 incorporation by plants, 290-6 incorporation by soil, 290-6 S-isotopes biogenic-S, 17, 163-7 H2S, 154-5, 344-7 H2SO4, 143 particulates, 141,316-17, 334-7 rain, snow, 150-5, 162, 352-5 relation to concentration, 149, 153-7, 314 relation to wind direction, 315 S02, 139, 141, 154-5, 312-16, 332-8 sulphur trioxide, 332 sulphuric acid, 143, 299 transport, 168-71 volcanic emissions, 9, 16, 116-25 Australia, 69, 88, 99, 271, 303, 385, 371-98 Azov Sea, 219 Baikal, USSR, 200 Balkhash, USSR, 200 Baltic Sea, 222 Balzac, Canada, 155 Barite conversion to AG2S, 44 conversion to S02, 51 O-isotopes, 38, 86, 249 radioactive sinter deposits, 38, 249 reduction with graphite, 56 S-isotopes, 67 (Fig.), 86, 249 Bark, entrapped sulphur, 325 Basic sills, S-isotopes, 8 (Fig.) Batch process, 6 Beggiatoa, 289 Benthic biomass, 189 Big Horn basin, USA, 244 Biogenic emissions ocean, 146-7, 164 (Fig.) springs, 147, 242-51 Biological sulphur cycle, 9 Biosphere, 282-311 Black Sea, 14, 185, 203, 217 Blood, 300 Bogs, 95 Boling, 92 Boraitotto, Italy, 91 Brahmaputra, 228 Brisbane, Australia, 303 Broken Hill, Australia, 90 Buggingen deposits, 86

Index

Calgary, Canada, 154-5 California, 274, 284 (Fig.) Cambrian, 82, 101 Canada, 82, 147, 154, 248, 274, 286 (see also Alberta, Ontario) Canberra, Australia, 303 Canon Diablo, 65 Cape Reykjanis, Iceland, 123 Carbon/sulphur ratio, plants, 282 Carbon-bonded S, 268 Carboniferous, 102, 243 Carbonyl sulphide, COS, 133-4, 146, 165 Cartagena, Spain, 90 Caspian Sea, 220 Casteliano, 88 Caucasus, 88, 151 Cenozic oils, 105 Central Europe, 88 Cephalopoda, 184 Chemosynthetic autotrophic organisms, 12, 301 Chesapeake Bay, USA, 147 China, 91 Chlamis latiaurata, 184 Chlorobium thiosulphatophilum, 13 Chlorophyta, 188 Chromatium vinosum, 13 Chromatium sp., 12 Chvaletice, Czechoslovakia, 147, 155, 399

Index Cladophora, 185 Claus process, 311 Closed system, 7 Coal extraction of sulphur, 47 low sulphur, 99 S-content, 399 S-isotopes, 15 (Fig.), 18, 96-9, 99, (Fig.), 100, 327, 332, 335, 383-7, 400 S-isotopes, emissions, 147, 154-5, 326, 335, 401 Codium vermila, 185 Coenzyme A, 187 Colorado River, 227-8 Connecticut, USA, 199, 203 Continental seas, 217-24 Copper-molybdenum deposits, 89 (Fig.) Copper-nickel deposits, 88 (Fig.), 392-4, 398 Corrosion of monuments, 402-4 Coverage, 296 Cretaceous, 82, 245, 274, 318 Crossfield, Canada, 141, 148 Crude oil, (see petroleum) Crust-mantle, 65 Crustal contamination, 66 Cucumis sativus, 297 Cyanobacterial mat, 370, 382 Cysteine, 187 Cystine, 187, 282, 299-300 Cystine kidney stones, 300 Czechoslovakia, 66, 150, 248, 399ff Deep sea hydrothermal vents, 301 Desulphotomaculum, 144 Desulphovibrio desulphuricans, 10, 144 Determination of 8180 of water, 58 Devonian, 82, 102, 109, 242 Dimethyl disulphide, DMDS, 133-4, 146 Dimethyl sulphide, DMS, 17, 133-4, 146, 165 Directional array sampling, 315 Dissolved organic sulphur, 204, 205 (Fig.), 318 Dustfall, 400-1 East Transbaikal, 89 East Pacific, 198

433 East Siberian platform, 100 Edmonton, Canada, 154, 312 Edwards aquifer, Texas, 239 El-Chichon, Mexico, 117, 125 Elemental-S atmosphere, 135, 290, 309, 325 deposits, 91-5, 411 entrapped in bark, 325 incorporation by plants, 290-6 incorporation by soil, 290-6 manufacture, 309-11, 325 oxidation, 279, 309-11 S-isotope fractionation during oxidation, 279 S-isotopes, 67 (Fig.), 91-5, 411 springs, 248 Ellesmere Island, Canada, 38, 199, 203 Enriched S-isotopes, potential for atmospheric studies, 168 Enteromorpha, 185 Epiphytic lichens, 142 Equilibrium isotope effects, 2 Eschka, 47 Etna, 118 Europe, 69 Evaporites, isotopic composition, 68-87 age curves, 82-6, 237 (Fig.) marine, 15, 38, 68-86, 237 (Fig.) non-marine, 86 reservoir estimate, 68 Feathers, 301-3 Filter packs for atmospheric sampling, 44 Finland, 88 Fisherman Bay, Australia, 379, 423 Flue gas, 335, 399 Fly ash, 328 Flybye Springs, NWT, Canada, 250 Flysch waters, 249 Food webs, 299-303 Forest soils, 146 Formation waters, 110 (Fig.), 242-5 Fossil fuels, 15-16, 15 (Fig.), 95-116, 412 (see also Coal, Petroleum, Sour gas) Fumaroles, 66, 117, 121-4 Gaurdak deposit, USSR, 94 Gazli, Middle Asia, 103

434 Geothermal areas, 245-7 Glacial tills, 233 Glutathione, 187 Gracefield, New Zealand, 157 Granite, S-concentration, 65 Great Salt Lake, 147, 154 Great Lakes, 151, 161 (Fig.), 199, 204, 212-15 Great Slave Lake, Canada, 247 Green Lake, Fayetteville, USA, 199 Greenhow, England, 90 Ground waters, 38, 177, 229-42, 277, 422-3 Groundwater, 229-42, 360-71 sulphate O-isotope composition, 232-42, 366-71 S-isotope composition, 232-42, 366-71 Gulf Coast oils, 108 Gypsum, 68, 149, 379 Hair, 300 Harz Mountains, 95 Helmstedt, Germany, 96 Heron Island, Great Barrier Reef, 300, 303 HI-reducible S, 268-9 High-volume sampler, 140, 337 Higher animals, 299-303 Huang He, 228 Hubbard Brook, New Hampshire, USA, 147 Humic acids, lake sediments, 213-14 Hydrogen sulphide abiogenic, in sediments, 198 atmosphere, 17, 133-4, 137, 344-7 biogenic, 15, 17, 144-7, 163-5,250, 344-7 Black sea, 218 emission rates, 17, 345 geothermal, 245-6 oxidation, 134-5, 143, 347 sour gas deposits Astrakhan, 113 Orenburg, 111, 113 Western Canada, 112 (Fig.) S-isotopes, 154-5 biogenic, 15, 17, 144-7, 163-5 geothermal, 245-7 hydrocarbon deposits, 109-16

Index volcanic, 122-4 tidal flats, 344-7 volcanic, 122-4 Hydrology (see Hydrosphere) Hydrosphere, 37, 177-266 (see also Continental seas, Groundwater, Lakes, Rivers) bogs, 95 dissolved organic-S, 318 formation waters, 110 (Fig.), 242-5 geothermal waters, 245-7 Kuwait, 360-71 Hydrothermal vents, 301 Iceland, 123 Igneous rock S-isotopes, 8 (Fig.) India, 247, 405-8 Indian Ocean, 185 Indus, 228 Iran, 99 Iraq, 91-2, 99 Isotope effects (see also Oxygen-isotope fractionation, Sulphur-isotope fractionation) equilibrium, 2 equilibrium constant, 3 fractionation factor, 6 kinetic, 5 partition functions, 3 Israel, 150, 161 Italy, 91 Japan, 88, 162,246, 287, 343-61 Jarosite, 250, 299 Jordan Rift Valley, 150 Jurassic, 82 Karelia, 88 Karst lakes, 20 Kazakhstan, 89, 151 Kazalehstan steppe, USSR, 20 Keystone, USA, 155 Kiba reagent, 46 Kidney stones, 300 Kilauea, 118, 120, 245 Kinetic isotope effects, 5 Kirgizia mountains, USSR, 151 Kola Peninsula, USSR, 151 Krafla, 245 Krakatoa, 125 Kunashir Island, 123

Index Kurume, 150 Kuwait, 99, 360-71 Labadie, USA, 155 Lacq, France, 110 Ladoga, USSR, 200 Lagoons, 146 Lake Tanganyika, 200 Lake Yanda, Antarctica, 38, 203 Lake Chernyi Kichiyer, USSR, 203 Lake Sakovo, USSR, 203 Lake Bolshoi Kichiyer, USSR, 203 Lake Kuznechikha, USSR, 203 Lake Kinneret, 200 Lake Creteil, 200 Lake Kononyer, USSR, 203 Lakes, 20, 38, 198-217 eutrophic, 205-8 humic acid, 213-15 karst, 20 meromictic, 208-11 oligotrophic, 204-5 oxygen isotope composition of sulphate, 205 sediments, 211-17 water column, 198, 203 Lead sulphide ores, (seeOre deposits) Lena River, USSR, 283 Li-AI-H4, 46 Lichens, 320, 323 Lignite, 95 Linsley Pond, USA, 199 Lithosphere, 38, 65-132 Lone Pine Sanctuary, Queensland, Australia, 300 Long Island, USA, 155 Lublin, Poland, 150, 157 Mackenzie River, 20, 246, 317 Madison formation, 111 Mafic rocks S-concentration, 65 Mambray Creek, South Australia, 190, 379 Mangrove, 185 Marine sulphate, (seeEvaporites, Ocean) Marine animals, 185-8 Marine sedimentary cycle, 14 Marsh sediments, 180-93 Mass spectrometry memory effects, 54

435
oxygen isotopes, 59 sulphur isotopes, 54 Meggen, 88 Mekong, 228 Mendeleev volcano, 123 Meteorites, 65 Methionine, 280, 297 Methyl mercaptan, CH3SH, 133-4, 146 Mexico, 91 Microbial biomass, 269 Middle East, 69, 100 Mikawa Bay, Japan, 147 Miocene, 82 Mishrak, Iraq, 92 Mississippi River, 228 Mississippian, 244 Miziiskaya platform, 102 Modelling of uptake of industrial S, 320 Mollusc, 184 Mongolia, 89 Monkeoppi formation, 82 Montana, 102 Monuments, corrosion, 402-4 Moscow, 151 Moss Bluff, USA, 91 Mosses, 142 Mt Ontake, 347 Mt Tom Pond, USA, 199, 203 Mt Agung, 147 Mt St Helens, USA, 118, 121 Nagoya, 150, 353-7 Nails, 300 Nairne, Australia, 88 Native sulphur, 18, 91-5 (see also Elemental-S) Natural oil seeps, 108 New Zealand, 151, 245, 279, 284 (Fig.), 285-6 New Haven, USA, 154 New York City, USA, 155 Newport, California, USA, 190 Nigeria, 410-16 Nitrogen/sulphur ratio, plants, 282 Norlinger Ries, Germany, 96 Norman Range, Canada, 38 North America, 69, 91 (seealso Canada, United States of America) North Dakota, 102 Northport, USA, 155

436 Northwest Territories, Canada, 286 Novgorod, 151 Novodimitrovskaya, USSR, 91

Index Parr bomb, 45, 47 Particulate sulphur, 138, (see also Elemental sulphur, Atmosphere) Peat, 16, 95 (see also Coal) Pedosphere, (see Soil) Pennsylvania, USA, 332 Pennsylvanian, 102, 244 Perch Lake basin, 232 Permian, 82, 102, 243 Persian Gulf, 378 Peru, 89 Petroleum contamination in sediments, 108 extraction of sulphur, 47 mean S-isotope composition, 107 S-isotopes, 15 (Fig.), 18, 100-108, 332-43, 387-91 Phaeophyta, 189 Phanerozoic, 68 Phosphoria formation, 111 Photosynthetic organisms, 12 Phytobenthos, 188 Phytoplankton, 146, 188 Pisa, 150, 160 Plants (seeVegetation) Plastic Lake watershed, 232 Pleistocene, 82 Poland, 91, 150 Polar bears, 187, 300 Polychaeta, 184 Polysaccharide sulphate, 45 Polysiphonia subilifera, 185 Pore water sulphate, 193, 203 Power plants Chwaletice, Czechoslovakia, 399 coal, 147, 155, 332, 342, 401 emissions, S-isotopes, 400-1 Kansk-Achinsk, USSR, 326-8 Northeastern USA, 331-43 Prague, 154 Proterozoic, 100 Prudhoe Bay, Alaska, USA, 108 Pyrite abiogenic, 198 coal, 399-400 lake sediments, 254 ocean sediments, 197-8 oxidation, 12-13, 33 S-isotopes, 67 (Fig.), 88 (Fig.) volcanic ash, 247 Queechy Lake, USA, 199, 203

Ocean (seealso Evaporites, Sulphate)


aerosols, 144-5 atmosphere, 144-147, 152 (Fig.) biogenic sulphide, 145-6, 164 (Fig.) carbonate, isotope composition, 251 emissions, 17 littorals, 17, 146 precipitation, 152 sediments, modern, 189-98 spray, 17, 152 sulphate, 14,37, 178-183 Ochoan, 82 Oder, 228 Oil shales, 391, 416 Oil, (see Petroleum) Oligocene, 82 Onega, USSR, 200 Ontake, Japan, 120 Ontario, Canada, 21, 102, 143, 155, 198-9, 202, 212-15, 274, 320 Open system, 7 Ordovician, 82 Ore deposits, 18, 87-91 Australia, 80, 391-7 India, 405-8 S-isotopes, 67 (Fig.) Cu-Mo, 89 (Fig.), 392-4 Native-S, 91-5, 411 Pb-Zn, 15, 90 (Fig.), 392-6 Orenburg gas-condensate deposit, 111 Organic-S, S-isotopes, 67 (Fig.) Oxygen-isotope fractionation exchange during sulphate reduction, 239-42 exchange reactions, 28-32 Soz-water, 28-32 sulphate crystallization, 33 sulphate reduction, 33 sulphate-water, 28 sulphide oxidation, 32 Oxygen-isotopes (see also Sulphate) Oz, atmosphere water, 34-5, 365-9 urine, 301 Pacific Ocean, 152 Paige Mountain, NWT, Canada, 147, 154, 248

Index
Radioactive-S, 190 Rain, 18-19, 365, 399 Ram River, Canada, 320 Rammelsberg, 88 Rayleigh distillation equation, 210 Reactions first order, 12 multi-step, 6, 12 unidirectional, 12 Red Sea, 198, 223 Red Deer, Canada, 154 Regional balances of atmospheric sulphur, 159-62 Rhine, 228 Rhodophyta, 189 Rhodopseudomonas sp., 13 Rice-fields, 146 River Niger, 227 River basins, 228 Rivers, 20, 224-9 a-isotopes, 25 sulphate concentration, 227-9 S-isotope variations, 38, 225-9, (Fig.), 283 Rock-water interactions, 230 Rot event, 85 Rotleigend salt deposits, 86 Russian platform, 100 Sabkhas, 369 Sakhalin, 150 Salmon, 186 Salt marshes, 146 Salt crystals on leaves, 298 Salt Lake City, Utah, USA, 158 Salt domes, 92 Santa Lucia, 88 Satsuma Iwo-Jima group, 122 Scanning electron microscopy, 142 (Fig.), 316 Sea spray, 144-6 Seals, 187, 300 Sediments distribution of sulphur compounds, 193 lake anthropogenic influence, 211-17 S-isotope composition, 211-17 marsh, 190 ocean, modern, 189-8 pore water sulphate, 193, 203

437
sulphide mineral formation, 197-8 sulphur-isotope balance, 195-7 sulphur-mass balance, 196-7 sulphur-isotope composition, 193 Selenium/sulphur ratio, 296, 302, 325 Serchio River, Italy, 38 Sernoye (reservoir), USSR, 200 Shark Bay, Western Australia, 372, 423 Shatt al Arab, 228 Shikote Alin, 65 Shimodava (Japan), 88 Showashinzan volcano, Hokkaido, 122 Shrimp, 187 Siberia, 150 Siberian platform, 89 Sicily, 93 Silurian, 102 Skyreholme, England, 90 Slotted cascade impactor, 140 Soil, 267-82 horizons, 323-4 a-isotopes, 277 sources of S, 271-2 S-forms, 267-71 C-bonded, 268 HI-reducible, 268 inorganic, 267 organic, 268-71 transformations of S, 271 transfer of S to, 280-2 sulphate, 267 S-isotopes, 272-82, 291 (Fig.) 318-24, 326-71 Solar Lake, 200, 203 Solfatara, 117, 121-4 Sour gas Alberta, 16, 122 (Fig.), 309-11 Astrakhan, 112-13 emissions, 312-26 Orenburg, 111, 113 processing, 309-11 S-isotopes, 109-16, 309-11 Souris River formation, Canada, 82 South America, 69, 91 Soviet Union (seeUSSR) Soxhlet extraction, 46 Spencer Gulf, South Australia, 372, 422 Spindleton, 92 Springs, 146, 247-50, 289 (Fig.), 298 barite sinter, 249

438 Springs (cont' d) biogenic emissions, 250 cave development, 248 jarosite depositing, 249 sulphate O-isotopes, 248-9 S-isotopes, 247-9 sulphide S-isotopes, 248-9 travertine depositing, 248 Sulphate assimilation, 10, 285-8 concentration formation waters, 293-6 lakes, 199-211 rivers, 225-9 seas, 217-25 sediments, 193-5 crystallization, 33 geothermal, 245-7 marine, 14-15, 68-86 ocean (see marine) O-isotopes, 37-8, 225 aerosol, 3fT-7 evaporites, 84 (Fig.) groundwater, 232-42 meromictic lakes, 209 rain, snow, 34-7, 150-9, 36fT-71 soil, 277 springs, 38, 247-9 urine, 301 pore water, 193, 203 precipitation, 8, 150-67, 352-61, 366, 400-2 reduction algal mats, 370 assimilatory (see sulphate assimilation) bacterial, 10-11, 192 chemical laboratory, 5, 32, 50-1 dissimilatory, 11 intensity in marches, 189-93 intensity in sediments, 189 light hydrocarbon gases, 113 mixture, 44 O-isotope fractionation, 33, 239-42 S-isotope fractionation, 5, 10-11, 239-42 thermochemical, 112-13 S-isotopes, 14, 18-20 aerosol

Index eutrophic lakes, 205-8 evaporites, 5, 68-86, 83, (Fig.), 85 (Fig.), 367 formation waters, 242-5 geothermal, 245-7 groundwater, 231-42, 365-6 lakes, 199-217 meromictic lakes, 208-11 minerals, 367 oligotrophic lakes, 203-4 rain, snow, 8, 150-67, 352-61, 365, 400-2 rivers, 38, 225-9, 282 seas, 217-25 springs, 247-9 terrestrial, 236 (Fig.) volcanic ash, 348 Stratosphere, 147, 154 Streams (see rivers) Sudbury, Canada, 21, 143, 155, 198, 202, 215 Sulphide (see also Hydrogen sulphide) abiogenic, in sediments, 198 mineral formation, sediments, 197-8 oxidation, 12-13, 33, 309, 348 S-isotopes (seealso Ore deposits) in lignite, 96 ocean sediments, 193-6 uptake by plants, 289 Sulphide ores (see Ore deposits) Sulphur isotope exchange reactions, 4 Sulphur dioxide, S02 atmosphere, 133-41, 154-5, 312-16, 332-8, 403 exposure threshold for plants, 290 geothermal, 245 heterogeneous oxidation, 136 homogeneous oxidation, 136 oxygen isotope composition, 37 preparation for sulphur isotope analyses, 48-53 purification, 52 S-isotopes atmosphere, 139, 141, 154-5, 312-16, 332-8 dependence on wind, 138-40, 315-16 sour gas processing, 309-11 Sulphur dust, (see Elemental-S, atmosphere) Sulphur hexafluoride, SF(" 48, 179

Index preparation for sulphur isotope analyses, 53-4 Sulphur-isotope abundances, 8-21 atmosphere, 16,312-17 basic sills, 8 (Fig.) biogenic H2S, 17, 163-7 coal, 15 (Fig.), 383-5 Cu-Ni deposits, 89 (Fig.), 392-4, 398 Cu-Mo deposits, 89 (Fig.) Cu-sandstone deposits, 90 (Fig.) emissions, power plant, 400-1 evaporites, 5, 68-86, 83 (Fig.), 85 (Fig.) fly ash, 327 feathers, 301-3 fossil fuels, 15-16, 15 (Fig.), 95-116 geothermal areas, 245-7 H2S in hydrocarbon deposits, 109-16 hair, 300-3 humans, 302-3 igneous rocks, 8 (Fig.) lakes, 20, 199-218 native S, 18, 91-5 ocean, 14 oil shale, 391 organic-S, 67 (Fig.) Pb-Zn deposits, 90 (Fig.), 373, 392-8 petroleum, 8 (Fig.), 15 (Fig.), 99-108, 331-43, 387-90 pyrite-polymetallic deposits, 88 (Fig.), 283, 392 rain, 8 (Fig.), 18-19, 150-2, 399 relation to metal pollution, 302 rivers, 20, 38, 225-9, 283 sea spray, 17 seas, 218-25 sedimentary sulphides, 8 (Fig.) snow, 8 (Fig.) (see also rain) sour gas processing emissions, 160 (Fig.) springs, 247-9 stratosphere, 147 sulphide ores, 87-91, 90 (Fig.) sulphur dioxide, 138-40, 400 terrestial sulphate, 236 (fig.) topographical effects, 295 urine, 301 volcanogenic, 8 (Fig.), 9, 93, 119-25 Sulphur-isotope fractionation, during atmospheric transformations, 143 exchange reactions, 4, 143

439 oxidation elemental S, 279 sulphide, 12-14 sulphate reduction assimilatory, 10, 285 chemical, 6, 32, 50-1 dissimilatory, 11 35S-labelled sulphate, 190 Sulphur trioxide, 32,49, 137, 332 Sulphuric acid from volcanic ash, 347 Sweden, 88, 240 Switzerland, 212 Tadzhik depression, USSR, 105 Tadzhikistan mountains, 151 Tar sands, 413 Tarnobzheg, Poland, 92 Tatarian, 82 Teepee Creek, Canada, 140 Tensleep formation, 111 Texas, 102 Thiobacilli, 12 Thiobacillus denitrificans, 235 Thiobacillus concretivorous, 13 Thompson, Canada, 274 Tidal flats, 344-7 Timano-Pechora, 100 Tokyo, 150, 353-7 Topographical factors, 295 Transition state theory, 5 Triassic, 102, 244 Troilite, 65 Tunisia, 284 (Fig.), 286 Tuscany, Italy, 38 Typha latifolia, 293 Uinta Basin, Utah, 105 Unidirectional processes, 5, 12 United States of America, 88, 91, 102, 105, 108, 121, 146-7, 154-5, 158, 199-203, 215, 244-5, 274, 284, 332 Upper Silurian, 82 Urals, 88, 151 Urine, 300 Usnea scabrata, 297 USSR, 20, 81, 89, 91, 94, 105, 109, 111-13, 155, 200, 203, 217-24, 222, 273, 282-5, 284 (Fig.), 326

Valleyview,Canada, 275, 280, 320

440 Vegetation, 20, 142, 282-99 algae, 185-9 atmospheric monitoring, 142 benthic weeds, 185 CIS ratios, 282 emission of reduced-S, 297 marine, 183 salt crystals on leaves, 298 sulphide assimilation, 184, 188 sulphur dioxide exposure threshold, 290 uptake, 288, 290-5 sulphur dust incorporation, 325 sulphur-forms, 282-3 sulphur isotopes, 283-99, 318-19 biological factors, 296 fractionation during emissions of reduced-S, 297 salt crystals on leaves, 298 topographical influence, 295 Zostera, 185, 189 Venezuela, 14 Venice, 150, 157, 160 Vistula, 228 Vladivostok, 150 Volcanoes activity classification, 121 ash, 120-1, 348 emissions, 16, 66 gases, 121-4 S-flux to atmosphere, 116-19 S-isotopes, 8-9, 119-25 total volcanic sulphur emitted to atmosphere< 124-5 Volcanogenic deposits, 93 Volga, 243 Volga-Urals, 110

Index

Wawa, Canada, 212, 273 Whiskers, 300 Whitecourt, Canada, 140, 155, 204, 319 Wind River Basin, USA, 102, 110, 244 Wood Point, Australia, 379 Wind direction, effect on S-isotopes of S02' 139 (Fig.) Yakutia, 150 Yellowstone, 245 Yudomski Souns event, 85 Zechstein, 82 Zhosaly Sopka, USSR, 283 Zinc sulphide ores (see Ore deposits) Zooplankton, 146 Zostera, 185, 189

S-ar putea să vă placă și