UNITED NATIONS ENVIRONMENT PROGRAMME INTERNATIONAL LABOUR ORGANISATION WORLD HEALTH ORGANIZATION INTERNATIONAL PROGRAMME ON CHEMICAL SAFETY ENVIRONMENTAL HEALTH CRITERIA 202 SELECTED NON-HETEROCYCLIC POLICYCLIC AROMATIC HYDROCARBONS This report contains the collective views of an international group of experts and does not necessarily represent the decisions or the stated policy of the United Nations Environment Programme, the International Labour Organisation, or the World Health Organization. First and second drafts prepared by staff members at the Fraunhofer Institute of Toxicology and Aerosol Research, Hanover, Germany, under the coordination of Dr R.F. Hertel, Dr G. Rosner, and Dr J. Kielhorn, in cooperation with Dr E. Menichini, Italy, Dr P.L. Grover, United Kingdom, and Dr J. Blok, Netherlands. Dr P. Muller, Canada, and Dr R. Schoeny and Dr T.L. Mumford, USA, prepared and revised the drafts of Appendix I. Published under the joint sponsorship of the United Nations Environment Programme, the International Labour Organisation, and the World Health Organization and produced within the framework of the Inter-Organization Programme for the Sound Management of Chemicals World Health Organization Geneva, 1998 The International Programme on Chemical Safety (IPCS), established in 1980, is a joint venture of the United Nations Environment Programme (UNEP), the International Labour Organisation (ILO), and the World Health Organization (WHO). The overall objectives of the IPCS are to establish the scientific basis for assessment of the risk to human health and the environment from exposure to chemicals, through international peer-review processes, as a prerequisite for the promotion of chemical safety, and to provide technical assistance in strengthening national capacities for the sound management of chemicals. The Inter-Organization Programme for the Sound Management of Chemicals (IOMC) was established in 1995 by UNEP, ILO, the Food and Agriculture Organization of the United Nations, WHO, the United Nations Industrial Development Organization, and the Organisation for Economic Co-operation and Development (Participating Organizations), following recommendations made by the 1992 United Nations Conference on Environment and Development, to strengthen cooperation and increase coordination in the field of chemical safety. The purpose of the IOMC is to promote coordination of the policies and activities pursued by the Participating Organizations, jointly or separately, to achieve the sound management of chemicals in relation to human health and the environment. WHO Library Cataloguing in Publication Data Selected non-heterocyclic polycyclic aromatic hydrocarbons. (Environmental health criteria ; 202) 1. Polycyclic hydrocarbons, Aromatic 2.Environmental exposure 3.Occupational exposure 4.Risk assessment - methods I.INternational Programme on Chemical Safety II.Series ISBN 92 4 157202 7 (NLM Classification: QD 341.H9) ISSN 0250-863X The World Health Organization welcomes requests for permission to reproduce or translate its publications, in part or in full. Applications and enquiries should be addressed to the Office of Publications, World Health Organization, Geneva, Switzerland, which will be glad to provide the latest information on any changes made to the text, plans for new editions, and reprints and translations already available. (c) World Health Organization 1998 Publications of the World Health Organization enjoy copyright protection in accordance with the provisions of Protocol 2 of the Universal Copyright Convention. All rights reserved. The designations employed and the presentation of the material in this publication do not imply the expression of any opinion whatsoever on the part of the Secretariat of the World Health Organization concerning the legal status of any country, territory, city or area or of its authorities, or concerning the delimitation of its frontiers or boundaries. The mention of specific companies or of certain manufacturers' products does not imply that they are endorsed or recommended by the World Health Organization in preference to others of a similar nature that are not mentioned. Errors and omissions excepted, the names of proprietary products are distinguished by initial capital letters. CONTENTS NOTE TO READERS OF THE CRITERIA MONOGRAPHS PREAMBLE WHO TASK GROUP ON ENVIRONMENTAL HEALTH CRITERIA FOR SELECTED NON-HETEROCYCLIC POLYCYCLIC AROMATIC HYDROCARBONS ENVIRONMENTAL HEALTH CRITERIA FOR SELECTED NON-HETEROCYCLIC POLYCYCLIC AROMATIC HYDROCARBONS 1. SUMMARY 1.1. Selection of compounds for this monograph 1.2. Identity, physical and chemical properties, and analytical methods 1.3. Sources of human and environmental exposure 1.4. Environmental transport, distribution, and transformation 1.5. Environmental levels and human exposure 1.5.1. Air 1.5.2. Surface water and precipitation 1.5.3. Sediment 1.5.4. Soil 1.5.5. Food 1.5.6. Aquatic organisms 1.5.7. Terrestrial organisms 1.5.8. General population 1.5.9. Occupational exposure 1.6. Kinetics and metabolism 1.7. Effects on laboratory mammals and in vitro 1.8. Effects on humans 1.9. Effects on other organisms in the laboratory and the field 2. IDENTITY, PHYSICAL AND CHEMICAL PROPERTIES, AND ANALYTICAL METHODS 2.1. Identity 2.1.1. Technical products 2.2. Physical and chemical properties 2.3. Conversion factors 2.4. Analytical methods 2.4.1. Sampling 2.4.1.1 Ambient air 2.4.1.2 Workplace air 2.4.1.3 Combustion effluents 2.4.1.4 Water 2.4.1.5 Solid samples 2.4.2. Preparation 2.4.3. Analysis 2.4.3.1 Gas chromatography 2.4.3.2 High-performance liquid chromatography 2.4.3.3 Thin-layer chromatography 2.4.3.4 Other techniques 2.4.4. Choice of PAH to be quantified 3. SOURCES OF HUMAN AND ENVIRONMENTAL EXPOSURE 3.1. Natural occurrence 3.2. Anthropogenic sources 3.2.1. PAH in coal and petroleum products 3.2.2. Production levels and processes 3.2.3. Uses of individual PAH 3.2.4. Emissions during production and processing of PAH 3.2.4.1 Emissions to the atmosphere 3.2.4.2 Emissions to the hydrosphere 3.2.5. Emissions during use of individual PAH 3.2.6. Emissions of PAH during processing and use of coal and petroleum products 3.2.6.1 Emissions to the atmosphere 3.2.6.2 Emissions to the hydrosphere 3.2.6.3 Emissions to the geosphere 3.2.6.4 Emissions to the biosphere 3.2.7. Emissions of PAH caused by incomplete combustion 3.2.7.1 Industrial point sources 3.2.7.2 Other diffuse sources 4. ENVIRONMENTAL TRANSPORT, DISTRIBUTION, AND TRANSFORMATION 4.1. Transport and distribution between media 4.1.1. Physicochemical parameters that dtermine environmental transport and distribution 4.1.2. Distribution and transport in the gaseous phase 4.1.3. Volatilization 4.1.4. Adsorption onto soils and sediments 4.1.5. Bioaccumulation 4.1.5.1 Aquatic organisms 4.1.5.2 Terrestrial organisms 4.1.6. Biomagnification 4.2. Transformation 4.2.1. Biotic transformation 4.2.1.1 Biodegradation 4.2.1.2 Biotransformation 4.2.2. Abiotic degradation 4.2.2.1 Photodegradation in the environment 4.2.2.2 Hydrolysis 4.2.3. Ultimate fate after use 5. ENVIRONMENTAL LEVELS AND HUMAN EXPOSURE 5.1. Environmental levels 5.1.1. Atmosphere 5.1.1.1 Source identification 5.1.1.2 Background and rural levels 5.1.1.3 Industrial sources 5.1.1.4 Diffuse sources 5.1.2. Hydrosphere 5.1.2.1 Surface and coastal waters 5.1.2.2 Groundwater 5.1.2.3 Drinking-water and water supplies 5.1.2.4 Precipitation 5.1.3. Sediment 5.1.3.1 River sediment 5.1.3.2 Lake sediment 5.1.3.3 Marine sediment 5.1.3.4 Estuarine sediment 5.1.3.5 Harbour sediment 5.1.3.6 Time trends of PAH in sediment 5.1.4. Soil 5.1.4.1 Background values 5.1.4.2 Industrial sources 5.1.4.3 Diffuse sources 5.1.4.4 Time trends of PAH in soil 5.1.5. Food 5.1.5.1 Meat and meat products 5.1.5.2 Fish and marine foods 5.1.5.3 Dairy products: cheese, butter, cream milk, and related products 5.1.5.4 Vegetables 5.1.5.5 Fruits and confectionery 5.1.5.6 Cereals and dried food products 5.1.5.7 Beverages 5.1.5.8 Vegetable and animal fats and oils 5.1.6. Biota 5.1.7. Animals 5.1.7.1 Aquatic organisms 5.1.7.2 Terrestrial organisms 5.2. Exposure of the general population 5.2.1. Indoor air 5.2.2. Food 5.2.3. Other sources 5.2.4. Intake of PAH by inhalation 5.2.5. Intake of PAH from food and drinking-water 5.3. Occupational exposure 5.3.1. Occupational exposure during processing and use of coal and petroleum products 5.3.1.1 Coal coking 5.3.1.2 Coal gasification and coal liquefaction 5.3.1.3 Pteroleum refining 5.3.1.4 Road paving 5.3.1.5 Roofing 5.3.1.6 Impregnation of wood with creosotes 5.3.1.7 Other exposures 5.3.2. Occupational exposure resulting from incomplete combustion of mineral oil, coal, and their products 5.3.2.1 Aluminium production 5.3.2.2 Foundries 5.3.2.3 Other workplaces 6. KINETICS AND METABOLISM IN LABORATORY MAMMALS AND HUMANS 6.1. Absorption 6.1.1. Absorption by inhalation 6.1.2. Absorption in the gastrointestinal tract 6.1.3. Absorption through the skin 6.2. Distribution 6.3. Metabolic transformation 6.3.1. Cytochromes P450 and PAH metabolism 6.3.1.1 Individual cytochrome P450 enzymes that metabolize PAH 6.3.1.2 Regulation of cytochrome P450 enzymes that metabolize PAH 6.3.2. Metabolism of benzo [a]pyrene 6.4. Elimination and excretion 6.5. Retention and turnover 6.5.1. Human body burdens of PAH 6.6. Reactions with tissue components 6.6.1. Reactions with proteins 6.6.2. Reactions with nucleic acids 6.7. Analytical methods 7. EFFECTS ON LABORATORY MAMMALS AND IN VITRO 7.1. Toxicity after a single exposure 7.1.1. Benzo [a]pyrene 7.1.2. Chrysene 7.1.3. Dibenz [a,h]anthracene 7.1.4. Fluoranthene 7.1.5. Naphthalene 7.1.6. Phenanthrene 7.1.7. Pyrene 7.2. Short-term toxicity 7.2.1. Subacute toxicity 7.2.1.1 Acenaphthene 7.2.1.2 Acenaphthylene 7.2.1.3 Anthracene 7.2.1.4 Benzo [a]pyrene 7.2.1.5 Benz [a]anthracene 7.2.1.6 Dibenz [a,h]anthracene 7.2.1.7 Fluoranthene 7.2.1.8 Naphthalene 7.2.1.9 Phenanthrene 7.2.1.10 Pyrene 7.2.2. Subchronic toxicity 7.2.2.1 Acenaphthene 7.2.2.2 Anthracene 7.2.2.3 Benzo [a]pyrene 7.2.2.4 Fluorene 7.2.2.5 Fluoranthene 7.2.2.6 Naphthalene 7.2.2.7 Pyrene 7.3. Long-term toxicity 7.3.1. Anthracene 7.3.2. Benz [a]anthracene 7.3.3. Dibenz [a,h]anthracene 7.4. Dermal and ocular irritation and dermal sensitization 7.4.1. Anthracene 7.4.2. Benzo [a]pyrene 7.4.3. Naphthalene 7.4.4. Phenanthrene 7.5. Reproductive effects, embryotoxicity, and teratogenicity 7.5.1. Benzo [a]pyrene 7.5.1.1 Teratogenicity in mice of different genotypes 7.5.1.2 Reproductive toxicity 7.5.1.3 Effects on postnatal development 7.5.1.4 Immunological effects in pregnant rats and mice 7.5.2. Naphthalene 7.5.2.1 Embryotoxicity 7.5.2.2 Toxicity in cultured embryos 7.6. Mutagenicity and related end-points 7.7. Carcinogenicity 7.7.1. Single substances 7.7.1.1 Benzo [a]pyrene 7.7.1.2 Benzo [e]pyrene 7.7.2. Comparative studies 7.7.2.1 Carcinogenicity 7.7.2.2 Further evidence 7.7.3. PAH in complex mixtures 7.7.4. Transplacental carcinogenicity 7.7.4.1 Benzo [a]pyrene 7.7.4.2 Pyrene 7.8. Special studies 7.8.1. Phototoxicity 7.8.1.1 Anthracene 7.8.1.2 Benzo [a]pyrene 7.8.1.3 Pyrene 7.8.1.4 Comparisons of individual PAH 7.8.2. Immunotoxicity 7.8.2.1 Benzo [a]pyrene 7.8.2.2 Dibenz [a,h]anthracene 7.8.2.3 Fluoranthene 7.8.2.4 Naphthalene 7.8.2.5 Comparisons of individual PAH 7.8.2.6 Exposure in utero 7.8.2.7 Mechanisms of the immunotoxicity of PAH 7.8.3. Hepatotoxicity 7.8.3.1 Benzo [a]pyrene 7.8.3.2 Comparisons of individual PAH 7.8.4. Renal toxicity 7.8.5. Ocular toxicity of naphthalene 7.8.6. Percutaneous absorption 7.8.7. Other studies 7.8.7.1 Benzo [k]fluoranthene 7.8.7.2 Benzo [a]pyrene 7.8.7.3 Phenanthrene 7.8.7.4 Comparisons of individual PAH 7.9. Toxicity of metabolites 7.9.1. Benzo [a]pyrene 7.9.2. 5-Methylchrysene 7.9.3. 1-Methylphenanthrene 7.10. Mechanisms of carcinogenicity 7.10.1. History 7.10.2. Current theories 7.10.3. Theories under discussion 7.10.3.1 Acenaphthene and acenaphthylene 7.10.3.2 Anthracene 7.10.3.3 Benzo [a]pyrene 7.10.3.4 Benz [a]anthracene 7.10.3.5 Benzo [c]phenanthrene 7.10.3.6 Chrysene 7.10.3.7 Cyclopenta [c,d]pyrene 7.10.3.8 Fluorene 7.10.3.9 Indeno[1,2,3- cd]pyrene 7.10.3.10 5-Methylchrysene 7.10.3.11 1-Methylphenanthrene 7.10.3.12 Naphthalene 7.10.3.13 Phenanthrene 7.10.3.14 Investigations of groups of PAH 8. EFFECTS ON HUMANS 8.1. Exposure of the general population 8.1.1. Naphthalene 8.1.1.1 Poisoning incidents 8.1.1.2 Controlled studies 8.1.2. Mixtures of PAH 8.1.2.1 PAH in unvented coal combustion in homes 8.1.2.2 PAH in cigarette smoke 8.1.2.3 PAH in coal-tar shampoo 8.2. Occupational exposure 8.3. Biomarkers of exposure to PAH 8.3.1. Urinary metabolites in general 8.3.2. 1-Hydroxypyrene 8.3.2.1 Method of determination 8.3.2.2 Concentrations 8.3.2.3 Time course of elimination 8.3.2.4 Suitability as a biomarker 8.3.3. Mutagenicity in urine 8.3.4. Genotoxicity in lymphocytes 8.3.5. DNA adducts 8.3.5.1 Method of determination 8.3.5.2 Concentrations 8.3.5.3 Suitability as a biomarker 8.3.6. Antibodies to DNA adducts 8.3.7. Protein adducts 8.3.8. Activity of cytochrome P450 8.3.9. Cell surface differentiation antigens in lung cancer 8.3.10. Oncogene proteins 9. EFFECTS ON OTHER ORGANISMS IN THE LABORATORY AND THE FIELD 9.1. Laboratory experiments 9.1.1. Microorganisms 9.1.1.1 Water 9.1.1.2 Soil 9.1.2. Aquatic organisms 9.1.2.1 Plants 9.1.2.2 Invertebrates 9.1.2.3 Vertebrates 9.1.2.4 Sediment-dwelling organisms 9.1.2.5 Toxicity of combinations of PAH 9.1.3. Terrestrial organisms 9.1.3.1 Plants 9.1.3.2 Invertebrates 9.1.3.3 Vertebrates 9.2. Field observations 9.2.1. Microorganisms 9.2.1.1 Water 9.2.1.2 Soil 9.2.2. Aquatic organisms 9.2.2.1 Plants 9.2.2.2 Invertebrates 9.2.2.3 Vertebrates 9.2.3. Terrestrial organisms 9.2.3.1 Plants 9.2.3.2 Invertebrates 9.2.3.3 Vertebrates 10 EVALUATION OF RISKS TO HUMAN HEALTH AND EFFECTS ON THE ENVIRONMENT 10.1. Human health 10.1.1. Exposure 10.1.1.1 General population 10.1.1.2 Occupational exposure 10.1..2 Toxic effects 10.1.2.1 Bioavailability 10.1.2.2 Acute toxicity 10.1.2.3 Irritation and allergic sensitization 10.1.2.4 Medium-term toxicity 10.1.2.5 Carcinogenicity 10.1.2.6 Reproductive toxicity 10.1.2.7 Immunotoxicity 10.1.2.8 Genotoxicity 10.2. Environment 10.2.1. Environmental levels and fate 10.2.2. Ecotoxic effects 10.2.2.1 Terrestrial organisms 10.2.2.2 Aquatic organisms 11 RECOMMENDATIONS FOR PROTECTION OF HUMAN HEALTH AND THE ENVIRONMENT 11.1. General recommendations 11.2. Protection of human health 11.3. Recommendations for further research 11.3.1. General 11.3.2. Protection of human health 11.3.3. Environmental protection 11.3.4. Risk assessment 12 PREVIOUS EVALUATIONS BY INTERNATIONAL BODIES 12.1. International Agency for Research on Cancer 12.2. WHO Water Quality Guidelines 12.3. FAO/WHO Joint Expert Committee on Food Additives 12.4. WHO Regional Office for Europe Air Quality Guidelines APPENDIX I. SOME APPROACHES TO RISK ASSESSMENT FOR POLYCYCLIC AROMATIC HYDROCARBONS I.1 Introduction I.2 Approaches to risk assessment I.2.1 Toxicity equivalence factors and related approaches I.2.1.1 Principle I.2.1.2 Development and validation I.2.1.2.1 Derivation of the potency of benzo [a]pyrene I2.1.2.2 Derivation of the relative potency of PAH other than benzo [a]pyrene I.2.1.3 Application I.2.2 Comparative potency approach I.2.2.1 Principle I.2.2.2 Development and validation I.2.2.3 Key implicit and explicit assumptions I.2.2.4 Application I.2.3 Benzo [a]pyrene as a surrogate for the PAH fraction of complex mixtures I.2.3.1 Principle I.2.3.2 Development and validation I.2.3.3 PAH profiles of complex mixtures I.2.3.4 Potency of complex mixtures I.2.3.5 Key implicit and explicit assumptions I.2.3.6 Application I.3 Comparison of the three procedures I.3.1 Individual PAH approach I.3.2 Comparative potency approach I.3.3 Benzo [a]pyrene surrogate approach APPENDIX II; SOME LIMIT VALUES II.1 Exposure of the consumer II.2 Occupational exposure II.3 Classification II.3.1 European Union II.3.2 USA REFERENCES RESUME RESUMEN Environmental Health Criteria PREAMBLE Objectives The WHO Environmental Health Criteria Programme was initiated in 1973, with the following objectives: (i) to assess information on the relationship between exposure to environmental pollutants and human health and to provide guidelines for setting exposure limits; (ii) to identify new or potential pollutants; (iii) to identify gaps in knowledge concerning the health effects of pollutants; (iv) to promote the harmonization of toxicological and epidemiological methods in order to have internationally comparable results. The first Environmental Health Criteria (EHC) monograph, on mercury, was published in 1976; numerous assessments of chemicals and of physical effects have since been produced. Many EHC monographs have been devoted to toxicological methods, e.g. for genetic, neurotoxic, teratogenic, and nephrotoxic effects. Other publications have been concerned with e.g. epidemiological guidelines, evaluation of short-term tests for carcinogens, biomarkers, and effects on the elderly. Since the time of its inauguration, the EHC Programme has widened its scope, and the importance of environmental effects has been increasingly emphasized in the total evaluation of chemicals, in addition to their health effects. The original impetus for the Programme came from resolutions of the World Health Assembly and the recommendations of the 1972 United Nations Conference on the Human Environment. Subsequently, the work became an integral part of the International Programme on Chemical Safety (IPCS), a cooperative programme of UNEP, ILO, and WHO. In this manner, with the strong support of the new partners, the importance of occupational health and environmental effects was fully recognized. The EHC monographs have become widely established, used, and recognized throughout the world. The recommendations of the 1992 United Nations Conference on Environment and Development and the subsequent establishment of the Intergovernmental Forum on Chemical Safety, with priorities for action in the six programme areas of Chapter 19, Agenda 21, lend further weight to the need for EHC assessments of the risks of chemicals. Scope The Criteria monographs are intended to provide critical reviews of the effect on human health and the environment of chemicals, combinations of chemicals, and physical and biological agents. They include reviews of studies that are of direct relevance for the evaluation and do not describe every study that has been carried out. Data obtained worldwide are used, and results are quoted from original studies, not from abstracts or reviews. Both published and unpublished reports are considered, and the authors are responsible for assessing all of the articles cited; however, preference is always given to published data, and unpublished data are used only when relevant published data are absent or when the unpublished data are pivotal to the risk assessment. A detailed policy statement is available that describes the procedures used for citing unpublished proprietary data, so that this information can be used in the evaluation without compromising its confidential nature (WHO, 1990). In the evaluation of human health risks, sound data on humans, whenever available, are preferred to data on experimental animals. Studies of animals and in-vitro systems provide support and are used mainly to supply evidence missing from human studies. It is mandatory that research on human subjects be conducted in full accord with ethical principles, including the provisions of the Helsinki Declaration. The EHC monographs are intended to assist national and international authorities in making risk assessments and subsequent risk management decisions. They represent a thorough evaluation of risks and are not in any sense recommendations for regulation or setting standards. The latter are the exclusive purview of national and regional governments. Content The layout of EHC monographs for chemicals is outlined below. * Summary: a review of the salient facts and the risk evaluation of the chemical * Identity: physical and chemical properties, analytical methods * Sources of exposure * Environmental transport, distribution, and transformation * Environmental levels and human exposure * Kinetics and metabolism in laboratory animals and humans * Effects on laboratory mammals and in-vitro test systems * Effects on humans * Effects on other organisms in the laboratory and the field * Evaluation of human health risks and effects on the environment * Conclusions and recommendations for protection of human health and the environment * Further research * Previous evaluations by international bodies, e.g. the International Agency for Research on Cancer, the Joint FAO/WHO Expert Committee on Food Additives, and the Joint FAO/WHO Meeting on Pesticide Residues Selection of chemicals Since the inception of the EHC Programme, the IPCS has organized meetings of scientists to establish lists of chemicals that are of priority for subsequent evaluation. Such meetings have been held in Ispra, Italy (1980); Oxford, United Kingdom (1984); Berlin, Germany (1987); and North Carolina, United States of America (1995). The selection of chemicals is based on the following criteria: the existence of scientific evidence that the substance presents a hazard to human health and/or the environment; the existence of evidence that the possible use, persistence, accumulation, or degradation of the substance involves significant human or environmental exposure; the existence of evidence that the populations at risk (both human and other species) and the risks for the environment are of a significant size and nature; there is international concern, i.e. the substance is of major interest to several countries; adequate data are available on the hazards. If it is proposed to write an EHC monograph on a chemical that is not on the list of priorities, the IPCS Secretariat first consults with the cooperating organizations and the participating institutions. Procedures The order of procedures that result in the publication of an EHC monograph is shown in the following flow chart. A designated staff member of IPCS, responsible for the scientific quality of the document, serves as Responsible Officer (RO). The IPCS Editor is responsible for the layout and language. The first draft, prepared by consultants or, more usually, staff at an IPCS participating institution is based initially on data provided from the International Register of Potentially Toxic Chemicals and reference data bases such as Medline and Toxline. The draft document, when received by the RO, may require an initial review by a small panel of experts to determine its scientific quality and objectivity. Once the RO finds the first draft acceptable, it is distributed in its unedited form to over 150 EHC contact points throughout the world for comment on its completeness and accuracy and, where necessary, to provide additional material. The contact points, usually designated by governments, may be participating institutions, IPCS focal points, or individual scientists known for their particular expertise. Generally, about four months are allowed before the comments are considered by the RO and author(s). A second draft incorporating the comments received and approved by the Director, IPCS, is then distributed to Task Group members, who carry out a peer review at least six weeks before their meeting. The Task Group members serve as individual scientists, not as representatives of any organization, government, or industry. Their function is to evaluate the accuracy, significance, and relevance of the information in the document and to assess the risks to health and the environment from exposure to the chemical. A summary and recommendations for further research and improved safety are also drawn up. The composition of the Task Group is dictated by the range of expertise required for the subject of the meeting and by the need for a balanced geographical distribution. The three cooperating organizations of the IPCS recognize the important role played by nongovernmental organizations, so that representatives from relevant national and international associations may be invited to join the Task Group as observers. While observers may provide valuable contributions to the process, they can speak only at the invitation of the Chairperson. Observers do not participate in the final evaluation of the chemical, which is the sole responsibility of the Task Group members. The Task Group may meet in camera when it considers that to be appropriate. All individuals who participate in the preparation of an EHC monograph as authors, consultants, or advisers must, in addition to serving in their personal capacity as scientists, inform the RO if at any time a conflict of interest, whether actual or potential, could be perceived in their work. They are required to sign a statement to that effect. This procedure ensures the transparency and probity of the process. When the Task Group has completed its review and the RO is satisfied as to the scientific correctness and completeness of the document, it is edited for language, the references are checked, and camera-ready copy is prepared. After approval by the Director, IPCS, the monograph is submitted to the WHO Office of Publications for printing. At this time, a copy of the final draft is also sent to the Chairperson and Rapporteur of the Task Group to check for any errors. It is accepted that the following criteria should initiate the updating of an EHC monograph: new data are available that would substantially change the evaluation; there is public concern about health or environmental effects of the agent because of greater exposure; an appreciable time has elapsed since the last evaluation. All participating institutions are informed, through the EHC progress report, of the authors and institutions proposed for the drafting of the documents. A comprehensive file of all comments received on drafts of each EHC monograph is maintained and is available on request. The chairpersons of task groups are briefed before each meeting on their role and responsibility in ensuring that these rules are followed. WHO TASK GROUP ON ENVIRONMENTAL HEALTH CRITERIA FOR SELECTED NON-HETEROCYCLIC POLYCYCLIC AROMATIC HYDROCARBONS Hanover, Germany, 25-29 September 1995 Members Dr P.E.T. Douben, Her Majesty's Inspectorate of Pollution, London, United Kingdom (Chairman) Dr P.L. Grover, Institute for Cancer Research, Sutton, United Kingdom Dr R.F. Hertel, Bundesgesinstitut für gesundheitlichen Verbraucherschutz und Veterinarmedizin, Berlin, Germany Professor J. Jacob, Biochemisches Institut für Umweltcarcinogene, Grosshausdorf, Germany Dr J Kielhorn, Fraunhofer Institute for Toxicology and Aerosol Research, Hanover, Germany Dr R.W. Luebke, National Health and Ecology Effects Laboratory, US Environmental Protection Agency, Research Triangle Park, NC, USA (Joint Rapporteur) Mr H. Malcolm, Institute of Terrestrial Ecology, Monks Wood, Huntingdon, Cambridgeshire, United Kingdom (Joint Rapporteur) Dr I. Mangelsdorf, Fraunhofer Institute for Toxicology and Aerosol Research, Hanover, Germany Dr E. Menichini, Istituto Superiore di Sanita, Rome, Italy Dr P. Muller, Ministry of Environment and Energy, Toronto, Ontario, Canada Dr J.L. Mumford, National Health and Environmental Effects Research Laboratory, US Environmental Protection Agency, Research Triangle Park, NC, USA Dr G. Rosner, Freiburg, Germany Dr R. Schoeny, National Center for Environmental Assessment, US Environmental Protection Agency, Cincinnati, OH, USA Dr T. Sorahan, Institute of Occupational Health, University of Birmingham, Birmingham, United Kingdom Dr Kimber L. White, Jr, Medical College of Virginia, Virginia Commonwealth University, Richmond, VA, USA (Vice-Chairman) Secretariat Dr E. Smith, International Programme on Chemical Safety, World Health Organization, Geneva, Switzerland Dr. M. Castegnaro, International Agency for Research on Cancer, Lyon, France Assisting the Secretariat Dr S. Artelt, Fraunhofer Institute for Toxicology and Aerosol Research, Hanover, Germany Dr A. Boehncke, Fraunhofer Institute for Toxicology and Aerosol Research, Hanover, Germany Dr O. Creutzenburg, Fraunhofer Institute for Toxicology and Aerosol Research, Hanover, Germany 1. SUMMARY 1.1 Selection of compounds for this monograph Polycyclic aromatic hydrocarbons (PAH) constitute a large class of compounds, and hundreds of individual substances may be released during incomplete combustion or pyrolysis of organic matter, an important source of human exposure. Studies of various environmentally relevant matrices, such as coal combustion effluents, motor vehicle exhaust, used motor lubricating oil, and tobacco smoke, have shown that the PAH in these mixtures are mainly responsible for their carcinogenic potential. PAH occur almost always in mixtures. Because the composition of such mixtures is complex and varies with the generating process, all mixtures containing PAH could not possible be covered in detail in this monograph. Thus, 33 individual compounds (31 parent PAH and two alkyl derivatives) were selected for evaluation on the basis of the availability of relevant data on toxicological end-points and/or exposure (Table 1). Since epidemiological studies, which are essential for risk assessment, were available only for mixtures, however, Sections 8 and 10 present the results of studies of mixtures of PAH, in contrast to the rest of the monograph. Numerous papers and reviews have been published on the occurrence, distribution, and transformation of PAH in the environment and on their ecotoxicological and toxicological effects. Only references from the last 10-15 years are cited in this monograph, unless no other information was available; reviews are cited for older studies and for further information. 1.2 Identity, physical and chemical properties, and analytical methods The term 'polycyclic aromatic hydrocarbons' commonly refers to a large class of organic compounds containing two or more fused aromatic rings made up of carbon and hydrogen atoms. At ambient temperatures, PAH are solids. The general characteristics common to the class are high melting- and boiling-points, low vapour pressure, and very low water solubility which tends to decrease with increasing molecular mass. PAH are soluble in many organic solvents and are highly lipophilic. They are chemically rather inert. Reactions that are of interest with respect to their environmental fate and possible sources of loss during atmospheric sampling are photodecomposition and reactions with nitrogen oxides, nitric acid, sulfur oxides, sulfuric acid, ozone, and hydroxyl radicals. Ambient air is sampled by collecting suspended particulate matter on glass-fibre, polytetrafluoroethylene, or quartz-fibre filters by means of high-volume or passive samplers. Vapour-phase PAH, which might volatilize from filters during sampling, are commonly trapped by adsorption on polyurethane foam. The sampling step is by far the most important source of variability in results. Table 1. Polycyclic aromatic hydrocarbons evaluated in this monograph Common name CAS name Synonyma CAS Registry No. Acenaphthylene Acenaphthylene91-20-3 Acenaphthene Acenaphthylene, 1,2-dihydro-208-96-8 Anthanthrene Dibenzo[def,mno]chrysene191-26-4 Anthracene Anthracene120-12-7 Benz[a]anthracene Benz[a]anthracene 1,2-Benzanthracene,56-55-3 tetraphene Benzo[a]fluorene 11 H-Benzo[a]fluorene 1,2-Benzofluorene238-84-6 Benzo[b]fluorene 11 H-Benzo[b]fluorene 2,3-Benzofluorene243-17-4 Benzo[b]fluoranthene Benz[e]acephenanthrylene 3,4-Benzofluoranthene205-99-2 Benzo[ghi]fluoranthene Benzo[ghi]fluoranthene 2,13-Benzofluoranthene203-12-3 Benzo[j]fluoranthene Benzo[j]fluoranthene 10,11-Benzofluoranthene205-82-3 Benzo[k]fluoranthene Benzo[k]fluoranthene 11,12-Benzofluoranthene207-08-9 Benzo[ghi]perylene Benzo[ghi]perylene 1,12-Benzoperylene191-24-2 Benzo[c]phenanthrene Benzo[c]phenanthrene 3,4-Benzophenanthrene195-19-7 Benzo[a]pyrene Benzo[a]pyrene 3,4-Benzopyreneb50-32-8 Benzo[e]pyrene Benzo[e]pyrene 1,2-Benzopyrene192-97-2 Chrysene Chrysene 1,2-Benzophenanthrene218-01-9 Coronene Coronene Hexabenzobenzene191-07-1 Cyclopenta[cd]pyrene Cyclopenta[cd]pyrene Cyclopenteno[cd]pyrene27208-37-3 Dibenz[a,h]anthracene Dibenz[a,h]anthracene 1,2:5,6-Dibenzanthracene53-70-3 Dibenzo[a,e]pyrene Naphtho[1,2,3,4-def]chrysene 1,2:4,5-Dibenzopyrene192-65-4 Dibenzo[a,h]pyrene Dibenzo[b,def]chrysene 3,4:8,9-Dibenzopyrene189-64-0 Dibenzo[a,i]pyrene Benzo[rst]pentaphene 3,4:9,10-Dibenzopyrene189-55-9 Dibenzo[a,l]pyrene Dibenzo[def,p]chrysene 1,2:3,4-Dibenzopyrene191-30-0 Fluoranthene Fluoranthene206-44-0 Fluorene 9H-Fluorene86-73-7 Indeno[1,2,3-cd]pyrene Indeno[1,2,3-cd]-pyrene 2,3-o-Phenylenpyrene193-39-5 5-Methylchrysene Chrysene, 5-methyl-3697-24-3 1-Methylphenanthrene Phenanthrene, 1-methyl-832-69-9 Table 1. (continued) Common name CAS name Synonyma CAS Registry No. Naphthalene Naphthalene91-20-3 Perylene Perylene peri-Dinaphthalene198-55-0 Phenanthrene Phenanthrene85-01-8 Pyrene Pyrene Benzo[def]phenanthrene129-00-0 Triphenylene Triphenylene 9,10-Benzophenanthrene217-59-4 Extensive lists of synonyms have been imported by the IARC (1983) and Loening & Merritt (1990). a Common synonym appearing in the literature b Also reported as benzo[def]chrysene Air is sampled at the workplace at low flow rates; particles are collected on glass-fibre or polytetrafluoroethylene filters and vapours on Amberlite XAD-2 resin. Devices for sampling stack gases are composed of a glass-fibre or quartz-fibre filter in front of a cooler to collect condensable matter and an adsorbent (generally XAD-2) cartridge. Vehicle exhausts are sampled under laboratory conditions, with standard driving cycles simulating on-road conditions. Emissions are collected either undiluted or after dilution with filtered cold air. Many extraction and purification techniques have been described. Depending on the matrix, PAH are extracted from samples with a Soxhlet apparatus, ultrasonically, by liquid-liquid partition, or, after sample dissolution or alkaline digestion, with a selective solvent. Supercritical fluid extraction from various environmental solids has also been used. The efficiency of extraction depends heavily on the solvent used, and many of the solvents commonly used in the past were not appropriate. Extracted samples are usually purified by column chromatography, particularly on alumina, silica gel, or Sephadex LH-20 but also by thin-layer chromatography. Identification and quantification are routinely performed by gas chromatography with flame ionization detection or by high-performance liquid chromatography (HPLC) with ultraviolet and fluorescence detection, generally in series. In gas chromatography, fused silica capillary columns are used, with polysiloxanes (SE-54 and SE-52) as stationary phases; silica-C18 columns are commonly used in HPLC. A mass spectrometric detector is often coupled to a gas chromatograph in order to confirm the identity of peaks. The choice of PAH to be determined depends on the purpose of the measurement, e.g. for health-orientated or ecotoxicological studies or to investigate sources. Testing for different sets of compounds may be required or recommended at national and international levels. 1.3 Sources of human and environmental exposure Little information is available on the production and processing of PAH, but it is probable that only small amounts of PAH are released as a direct result of these activities. The PAH found principally are used as intermediates in the production of polyvinylchloride and plasticizers (naphthalene), pigments (acenaphthene, pyrene), dyes (anthracene, fluoranthene), and pesticides (phenanthrene). The largest emissions of PAH result from incomplete combustion of organic materials during industrial processes and other human activities, including: - processing of coal, crude oil, and natural gas, including coal coking, coal conversion, petroleum refining, and production of carbon blacks, creosote, coal-tar, and bitumen; - aluminium, iron and steel production in plants and foundries; - heating in power plants and residences and cooking; - combustion of refuse; - motor vehicle traffic; and - environmental tobacco smoke. PAH, especially these of higher molecular mass, entering the environment via the atmosphere are adsorbed onto particulate matter. The hydrosphere and geosphere are affected secondarily by wet and dry deposition. Creosote-preserved wood is another source of release of PAH into the hydrosphere, and deposition of contaminated refuse, like sewage sludge and fly ash, contributes to emissions of PAH into the geosphere. Little information is available about the passage of PAH into the biosphere. PAH occur naturally in peat, lignite, coal, and crude oil. Most of the PAH in hard coals are tightly bound within the coal structure and cannot be leached out. The release of PAH into the environment has been determined by identification of a characteristic PAH concentration profile, but this has been possible in only a few cases. Benzo [a]pyrene has frequently been used as an indicator of PAH, especially in older studies. Generally, emissions of PAH are only estimates based on more or less reliable data and give only a rough idea of exposure. The most important sources of PAH are as follows: Coal coking: Airborne emissions of PAH from coal coking in Germany have decreased significantly over the last 10 years as a result of technical improvements to existing plants, closure of old plants, and reduced coke production. Similar situations are assumed to exist in western Europe, Japan, and the USA, but no data were available. Production of aluminium (mainly special coal anodes), iron, and steel and the binding agents used in moulding sand in foundries: Little information is available. Domestic and residential heating: Phenanthrene, fluoranthene, pyrene, and chrysene are emitted as major components. The emissions from wood stoves are 25-1000 times higher than those from charcoal-fired stoves, and in areas where wood burning predominates for domestic heating the major portion of airborne PAH may be derived from this source, especially in winter. The release of PAH during residential heating is thus assumed to be an important source in developing countries where biomass is often burnt in relatively simple stoves. Cooking: PAH may be emitted during incomplete combustion of fuels, from cooking oil, and from food being cooked. Motor vehicle traffic: The main compounds released from petrol-fuelled vehicles are fluoranthene and pyrene, while naphthalene and acenaphthene are abundant in the exhaust of diesel-fuelled vehicles. Although cyclopenta [cd]-pyrene is emitted at a high rate from petrol-fuelled engines, its concentration in diesel exhaust is only just above the limit of detection. The emission rates, which depend on the substance, the type of vehicle, its engine conditions, and the test conditions, range from a few nanograms per kilometre to > 1000 mg/km. PAH emissions from vehicle engines are dramatically reduced by fitting catalytic converter devices. Forest fires: In countries with large forest areas, fires can make an imprtant contribution to PAH emissions. Coal-fired power plants: PAH released into the atmosphere from such plants consist mainly of two- and three-ring compounds. In contaminated areas, the PAH levels in ambient air may be higher than those in the stack gases. Incineration of refuse: The PAH emissions in stack gases from this souce in a number of countries were < 10 mg/m3. 1.4 Environmental transport, distribution, and transformation Several distribution and transformation processes determine the fate of both individual PAH and mixtures. Partitioning between water and air, between water and sediment, and between water and biota are the most important of the distribution processes. As PAH are hydrophobic with low solubility in water, their affinity for the aquatic phase is very low; however, in spite of the fact that most PAH are released into the environment via the atmosphere, considerable concentrations are also found in the hydrosphere because of their low Henry's law constants. As the affinity of PAH for organic phases is greater than that for water, their partition coefficients between organic solvents, such as octanol, and water are high. Their affinity for organic fractions in sediment, soil, and biota is also high, and PAH thus accumulate in organisms in water and sediments and in their food. The relative importance of uptake from food and from water is not clear. In Daphnia and molluscs, accumulation of PAH from water is positively correlated with the octanol:water partition coefficient ( Kow). In fish and algae that can metabolize PAH, however, the internal concentrations of different PAH are not correlated with the Kow. Biomagnification - the increase in the concentration of a substance in animals in successive trophic levels of food chains - of PAH has not been observed in aquatic systems and would not be expected to occur, because most organisms have a high biotransformation potential for PAH. Organisms at higher trophic levels in food chains show the highest potential biotransformation. PAH are degraded by photodegradation, biodegradation by microorganisms, and metabolism in higher biota. Although the last route of transformation is of minor importance for the overall fate of PAH in the environment, it is an important pathway for the biota, since carcinogenic metabolites may be formed. As PAH are chemically stable, with no reactive groups, hydrolysis plays no role in their degradation. Few standard tests for the biodegradation of PAH are available In general, they are biodegraded under aerobic conditions, the biodegradation rate decreasing drastically with the number of aromatic rings. Under anaerobic conditions, degradation is much slower. PAH are photooxidized in air and water in the presence of sensitizing radicals like OH, NO3, and O3. Under laboratory conditions, the half-life of the reaction with airborne OH radicals is about one day, whereas reactions with NO3 and O3 usually have much lower velocity constants. The adsorption of high-molecular-mass PAH onto carbonaceous particles in the environment should stabilize the reaction with OH radicals. The reaction of two- to four-ring PAH, which occur mainly in the vapour phase, with NO3 leads to nitro-PAH, which are known mutagens. The photooxidation of some PAH in water seems to be more rapid than in air. Calculations based on physicochemical and degradation parameters indicate that PAH with four or more aromatic rings persist in the environment. 1.5 Environmental levels and human exposure PAH are ubiquitous in the environment, and various individual PAH have been detected in different compartments in numerous studies. 1.5.1 Air The levels of individual PAH tend to be higher in winter than in summer by at least one order of magnitude. The predominant source during winter is residential heating, while that during summer is urban motor vehicle traffic. Average concentrations of 1-30 ng/m3 of individual PAH were detected in the ambient air of various urban areas. In large cities with heavy motor vehicle traffic and extensive use of biomass fuel, such as Calcutta, levels of up to 200 ng/m3 of individual PAH were found. Concentrations of 1-50 ng/m3 were detected in road tunnels. Cyclopenta [cd]pyrene and pyrene were present at concentrations up to 100 ng/m3. In a subway station, PAH concentrations of up to 20 ng/m3 were measured. Near industrial sources, the average concentrations of individual PAH ranged from 1 to 10 ng/m3. Phenanthrene was present at up to a maximum of about 310 ng/m3. The background values of PAH are at least one or two orders of magnitude lower than those near sources like motor vehicle traffic. For example, the levels at 1100 m ranged from 0.004 to 0.03 ng/m3. 1.5.2 Surface water and precipitation Most of the PAH in water are believed to result from urban runoff, from atmospheric fallout (smaller particles), and from asphalt abrasion (larger particles). The major source of PAH varies, however, in a given body of water. In general, most samples of surface water contain individual PAH at levels of up to 50 ng/litre, but highly polluted rivers had concentrations of up to 6000 ng/litre. The PAH levels in groundwater are within the range 0.02-1.8 ng/litre, and drinking-water samples contain concentrations of the same order of magnitude. Major sources of PAH in drinking-water are asphalt-lined storage tanks and delivery pipes. The levels of individual PAH in rainwater ranged from 10 to 200 ng/litre, whereas levels of up to 1000 ng/litre have been detected in snow and fog. 1.5.3 Sediment The concentrations of individual PAH in sediment were generally one order of magnitude higher than those in precipitation. 1.5.4 Soil The main sources of PAH in soil are atmospheric deposition, carbonization of plant material, and deposition from sewage and particulate waste. The extent of pollution of soil depends on factors such as its cultivation, its porosity, and its content of humic substances. Near industrial sources, individual PAH levels of up to 1 g/kg soil have been found. The concentrations in soil from other sources, such as automobile exhaust, are in the range 2-5 mg/kg. In unpolluted areas, the PAH levels were 5-100 µg/kg soil. 1.5.5 Food Raw food does not normally contain high levels of PAH, but they are formed by processing, roasting, baking, or frying. Vegetables may be contaminated by the deposition of airborne particles or by growth in contaminated soil. The levels of individual PAH in meat, fish, dairy products, vegetables and fruits, cereals and their products, sweets, beverages, and animal and vegetable fats and oils were within the range 0.01-10 µg/kg. Concentrations of over 100 µg/kg have been detected in smoked meat and up to 86 µg/kg in smoked fish; smoked cereals contained up to 160 µg/kg. Coconut oil contained up to 460 µg/kg. The levels in human breast milk were 0.003-0.03 µg/kg. 1.5.6 Aquatic organisms Marine organisms are known to adsorb and accumulate PAH from water. The degree of contamination is related to the extent of industrial and urban development and shipping movements. PAH concentrations of up to 7 mg/kg have been detected in aquatic organisms living near industrial effluents, and the average levels of PAH in aquatic animals sampled at contaminated sites were 10-500 µg/kg, although levels of up to 5 mg/kg were also detected. The average levels of PAH in aquatic animals sampled at various sites with unspecified sources of PAH were 1-100 µg/kg, but concentrations of up to 1 mg/kg were found, for example, in lobsters in Canada. 1.5.7 Terrestrial organisms The concentrations of PAH in insects ranged from 730 to 5500 µg/kg. The PAH content of earthworm faeces depends significantly on the location: those in a highly industrialized region in eastern Germany contained benzo [a]pyrene at concentrations up to 2 mg/kg. 1.5.8 General population The main sources of nonoccupational exposure are: polluted ambient air, smoke from open fireplaces and cooking, environmental tobacco smoke, contaminated food and drinking-water, and the use of PAH-contaminated products. PAH can be found in indoor air as a result of residential heating and environmental tobacco smoke at average concentrations of 1-100 ng/m3, with a maximum of 2300 ng/m3. The intake of individual PAH from food has been estimated to be 0.10-10 µg/day per person. The total daily intake of benzo [a]pyrene from drinking-water was estimated to be 0.0002 µg/person. Cereals and cereal products are the main contributors to the intake of PAH from food because they are a major component of the total diet. 1.5.9 Occupational exposure Near a coke-oven battery, the levels of benzo [a]pyrene ranged from < 0.1 to 100-200 µg/m3, with a maximum of about 400 µg/m3. In modern coal gasification systems, the concentration of PAH is usually < 1 µg/m3 with a maximum of 30 µg/m3. Personal samples taken from operators of petroleum refinery equipment showed exposure to 2.6-470 µg/m3. In samples of air taken near bitumen processing plants at refineries, the total PAH levels were 0.004-50 µg/m3. Near road paving operations, the total PAH concentrations in personal air samples were up to 190 µg/m3, and the mean value in area air samples was 0.13 µg/m3. The PAH levels in personal air samples taken at an aluminium smelter were 0.05-9.6 µg/m3, but urine samples of workers at an aluminium plant contained very low levels. Area air samples contained PAH concentrations of up to 5 µg/m3 in one German foundry, 3-40 µg/m3 at iron mines and 4-530 µg/m3 at copper mines. The concentrations of PAH in cooking fumes in a food factory ranged from 0.07 to 26 µg/m3. 1.6 Kinetics and metabolism PAH are absorbed through the pulmonary tract, the gastrointestinal tract, and the skin. The rate of absorption from the lungs depends on the type of PAH, the size of the particles on which they are absorbed, and the composition of the adsorbent. PAH adsorbed onto particulate matter are cleared from the lungs more slowly than free hydrocarbons. Absorption from the gastrointestinal tract occurs rapidly in rodents, but metabolites return to the intestine via biliary excretion. Studies with 32P-postlabelling of percutaneous absorption of mixtures of PAH in rodents showed that components of the mixtures reach the lungs, where they become bound to DNA. The rate of percutaneous absorption in mice according to the compound. PAH are widely distributed throughout the organism after administration by any route and are found in almost all internal organs, but particularly those rich in lipids. Intravenously injected PAH are cleared rapidly from the bloodstream of rodents but can cross the placental barrier and have been detected in fetal tissues. The metabolism of PAH is complex. In general, parent compounds are converted via intermediate epoxides to phenols, diols, and tetrols, which can themselves be conjugated with sulfuric or glucuronic acids or with glutathione. Most metabolism results in detoxification, but some PAH are activated to DNA-binding species, principally diol epoxides, which can initiate tumours. PAH metabolites and their conjugates are excreted via the urine and faeces, but conjugates excreted in the bile can be hydrolysed by enzymes of the gut flora and reabsorbed. It can be inferred from the available information on the total human body burden that PAH do not persist in the body and that turnover is rapid. This inference excludes those PAH moieties that become covalently bound to tissue constituents, in particular nucleic acids, and are not removed by repair. 1.7 Effects on laboratory mammals and in vitro The acute toxicity of PAH appears to be moderate to low. The well-characterized PAH, naphthalene, showed oral and intravenous LD50 values of 100-500 mg/kg body weight (bw) in mice and a mean oral LD50 of 2700 mg/kg bw in rats. The values for other PAH are similar. Single high doses of naphthalene induced bronchiolar necrosis in mice, rats, and hamsters. Short-term studies showed adverse haematological effects, expressed as myelotoxicity with benzo [a]pyrene, haemolymphatic changes with dibenz [a,h]-anthracene, and anaemia with naphthalene; however, in a seven-day study by oral and intraperitoneal administration in mice, tolerance to the effect of naphthalene was observed. Systemic effects caused by long-term treatment with PAH have been described only rarely, because the end-point of most studies has been carcinogenicity. Significant toxic effects are manifested at doses at which carcinogenic responses are also triggered. In studies of adverse effects on the skin after dermal application, non- or weakly carcinogenic PAH such as perylene, benzo [e]pyrene, phenanthrene, pyrene, anthracene, acenaphthalene, fluorene, and fluoranthene were inactive, whereas carcinogenic compounds such as benz [a]anthracene, dibenz [a,h]-anthracene, and benzo [a]pyrene caused hyperkeratosis. Anthracene and naphthalene vapours caused mild eye irritation. Benzo [a]pyrene induced contact hypersensitivity in guinea-pigs and mice. Benz [a]anthracene, benzo [a]pyrene, dibenz [a,h]anthracene, and naphthalene were embrotoxic to mice and rats. Benzo [a]pyrene also had teratogenic and reproductive effects. Intensive efforts have been made to elucidate the genetic basis of the embryotoxic effect of benzo [a]pyrene. Fetal death and malformations are observed only if the cytochrome P450 monooxy-genase system is inducible, either in the mother (with placental permigration) or in the embryo. Not all of the effects observed can be explained by genetic predisposition, however: in mice and rabbits, benzo [a]pyrene had transplacental carcinogenic activity, resulting in pulmonary adenomas and skin papillomas in the progeny. Reduced fertility and oocyte destruction were also observed. PAH have also been studied extensively in assays for genotoxicity and cell transformation; most of the 33 PAH covered in this monograph are genotoxic or probably genotoxic. The only compounds for which negative results were found in all assays were anthracene, fluorene, and naphthalene. Owing to inconsistent results, phenanthrene and pyrene could not be reliably classified for genotoxicity. Comprehensive work on the carcinogenicity of PAH shows that 17 of the 33 studied are, or are suspected of being, carcinogenic (Table 2). The best-characterized PAH is benzo [a]pyrene, which has been studied by all current methods in seven species. PAH that have been the subject of 12 or more studies are anthanthrene, anthracene, benz [a]anthracene, chrysene, dibenz [a,h]-anthracene, dibenzo [a,i]pyrene, 5-methylchrysene, phenanthrene, and pyrene. Special studies of the phototoxicity, immunotoxicity, and hepatotoxicity of PAH are supplemented by reports on the ocular toxicity of naphthalene. Anthracene, benzo [a]pyrene, and some other PAH were phototoxic to mammalian skin and in cell cultures in vitro when applied with ultraviolet radiation. PAH have generally been reported to have immunosuppressive effects. After intraperitoneal treatment of mice with benzo [a]pyrene, immunological parameters were strongly suppressed in the progeny for up to 18 months. Increased liver regeneration and an increase in liver weight have also been observed. The effect of naphthalene in inducing formation of cataracts in the rodent eye has been attributed to the inducibility of the cytochrome P450 system in studies in which genetically different mouse strains were used. Theoretical models to predict the carcinogenic potency of PAH from their structures, based on a large amount of experimental work, were presented as early as the 1930s. The first model was based on the high chemical reactivity of certain double bonds (the K-region theory). A later systematic approach was based on the chemical synthesis of possible metabolites and their mutagenic activity. This 'bay region' theory proposes that epoxides adjacent to a bay region yield highly stabilized carbonium ions. Other theoretical approaches are the 'di-region theory' and the 'radical cation potential theory'. Many individual PAH are carcinogenic to animals and may be carcinogenic to humans, and exposure to several PAH-containing mixtures has been shown to increase the incidence of cancer in human populations. There is concern that those PAH found to be carcinogenic in experimental animals are likely to be carcinogenic in humans. PAH produce tumours both at the site of contact and at distant sites. The carcinogenic potency of PAH may vary with the route of exposure. Various approaches to assessing the risk associated with exposure to PAH, singly and in mixtures, have been proposed. No one approach is endorsed in this monograph; however, the data requirements, assumptions, applicability, and other features of three quantitative risk assessment processes that have been validated to some degree are described. 1.8 Effects on humans Because of the complex profile of PAH in the environment and in workplaces, human exposure to pure, individual PAH has been limited to scientific experiments with volunteers, except in the case of naphthalene which is used as a moth-repellant for clothing. After dermal application, anthracene, fluoranthene, and phenanthrene induced specific skin reactions, and benzo [a]pyrene induced reversible, regressive verrucae which were classified as neoplastic proliferations. The systemic effects of naphthalene are known from numerous cases of accidental intake, particularly by children. The lethal oral dose is 5000-15 000 mg for adults and 2000 mg taken over two days for a child. The typical effect after dermal or oral exposure is acute haemolytic anaemia, which can also affect fetuses transplacentally. Table 2. Summary of results of tests for genotoxicity and carcinogenicity for the 33 polycyclic aromatic hydrocarbons studies Compound Genotoxicity Carcinogenicity Acenaphthene (?) (?) Acenaphthylene (?) No studies Anthanthrene (+) + Anthracene - - Benz[a]anthracene + + Benzo[b]fluoranthene + + Benzo[j]fluoranthene + + Benzo[ghi]fluoranthene (+) (-) Benzo[k]fluoranthene + + Benzo[a]fluorene (?) (?) Benzo[b]fluorene (?) (?) Benzo[ghi]perylene + - Benzo[c]phenanthrene (+) + Benzo[a]pyrene + + Benzo[e]pyrene + ? Chrysene + + Coronene (+) (?) Cyclopenta[cd]pyrene + + Dibenz[a,h]anthracene + + Dibenzo[a,e]pyrene + + Dibenzo[a,h]pyrene (+) + Dibenzo[a,i]pyrene + + Dibenzo[a,l]pyrene (+) + Fluoranthene + (+) Fluorene - - Indeno[1,2,3-cd]pyrene + + 5-Methylchrysene + + 1-Methylphenanthrene + (-) Naphthalene - (?) Perylene + (-) Phenanthrene (?) (?) Pyrene (?) (?) Triphenylene + (-) +, positive; -, negative; ?, questionable Parentheses, result derived from small database Tobacco smoking is the most important single factor in the induction of lung tumours and also for increased incidences of tumours of the urinary bladder, renal pelvis, mouth, pharynx, larynx, and oesophagus. The contribution of PAH in the diet to the development of human cancer is not considered to be high. In highly industrialized areas, increased body burdens of PAH due to polluted ambient air were detected. Psoriasis patients treated with coal-tar are also exposed to PAH. Occupational exposure to soot as a cause of scrotal cancer was noted for the first time in 1775. Later, occupational exposure to tars and paraffins was reported to induce skin cancer. The lung is now the main site of PAH-induced cancer, whereas skin tumours have become more rare because of better personal hygiene. Epidemiological studies have been conducted of workers exposed at coke ovens during coal coking and coal gasification, at asphalt works, foundries, and aluminium smelters, and to diesel exhaust. Increased lung tumour rates due to exposure to PAH have been found in coke-oven workers, asphalt workers, and workers in Söderberg potrooms of aluminium reduction plants. The highest risk was found for coke-oven workers, with a standardized mortality ratio of 195. Dose-response relationships were found in several studies. In aluminium plants, not only urinary bladder cancer but also asthma-like symptoms, lung function abnormalities, and chronic bronchitis have been observed. Coke-oven workers were found to have decreased serum immunoglobulin levels and decreased immune function. Occupational exposure to naphthalene for five years was reported to have caused cataract. Several methods have been developed to assess internal exposure to PAH. In most of the studies, PAH metabolites such as urinary thioethers, 1-naphthol, b-naphthylamine, hydroxyphenanthrenes, and 1-hydroxypyrene were measured in urine. The latter has been used widely as a biological index of exposure. The genotoxic effects of PAH have been determined by testing for mutagenicity in urine and faeces and for the presence of micronuclei, chromosomal aberrations, and sister chromatid exchange in peripheral blood lymphocytes. In addition, adducts of benzo [a]pyrene with DNA in peripheral lymphocytes and other tissues and with proteins like albumin as well as antibodies to DNA adducts have been measured. 1-Hydroxypyrene in urine and DNA adducts in lymphocytes have been investigated as markers in several studies. 1-Hydroxpyrene can be measured more easily than DNA adducts, there is less variation between individuals, and lower levels of exposure can be detected. Both markers have been used to assess human exposure in various environments. Increased 1-hydroxpyrene excretion or DNA adducts were found at various workplaces in coke plants, aluminum manufacturing, wood impregnation plants, foundries, and asphalt works. The highest exposures were those of coke-oven workers and workers impregnating wood with creosote, who took up 95% of total of PAH through the skin, in contrast to the general population in whom uptake via food and tobacco smoking predominate. Estimates of the risk associated with exposure to PAH and PAH mixtures are based on estimates of exposure and the results of epidemiological studies. Data for coke-oven workers resulted in a relative risk for lung cancer of 15.7. On this basis, the risk of the general population for developing lung cancer over a lifetime has been calculated to be 10-4 to 10-5 per ng of benzo [a]pyrene per m3 air. In other words, about one person in 10 000 or 100 000 would be expected to develop lung cancer in his or her lifetime as a result of exposure to benzo [a]pyrene in air. 1.9 Effects on other organisms in the laboratory and the field PAH are acutely toxic to fish and Daphnia magna in combination with absorption of ultraviolet radiation and visible light. Metabolism and degradation alter the toxicity of PAH. At low concentrations, PAH can stimulate the growth of microorganisms and algae. The most toxic PAH for algae are benz [a]anthracene (four-ring), the concentration at which given life parameters are reduced by 50% (EC50) being 1-29 µg/litre, and benzo [a]pyrene (five-ring), with an EC50 of 5-15 µg/litre. The EC50 values for algae for most three-ring PAH are 240-940 µg/litre. Naphthalene (two-ring) is the least toxic, with EC50 values of 2800-34 000 µg/litre. No clear difference in sensitivity was found between different taxonomic groups of invertebrates like crustaceans, insects, molluscs, polychaetes, and echinoderms. Naphthalene is the least toxic, with 96-h LC50 values of 100-2300 µg/litre. The 96-h LC50 values for three-ring PAH range between < 1 and 3000 µg/litre. Anthracene may be more toxic than the other three-ring PAH, with 24-h LC50 values between < 1 and 260 µg/litre. The 96-h LC50 values for four-, five-, and six-ring PAH are 0.2-1200 µg/litre. Acute toxicity (LC50) in fish was seen at concentrations of 110 to > 10 000 µg/litre of naphthalene, 30-4000 µg/litre of three-ring PAH (anthracene, 2.8-360 µg/litre), and 0.7-26 µg/litre for four- or five-ring PAH. Contamination of sediments with PAH at concentrations of 250 mg/kg was associated with hepatic tumours in free-living fish. Tumours have also been induced in fish exposed in the laboratory. Exposure of fish to certain PAH can also cause physiological changes and affect their growth, reproduction, swimming performance, and respiration. 2. IDENTITY, PHYSICAL AND CHEMICAL PROPERTIES, AND ANALYTICAL METHODS 2.1 Identity The name 'polycyclic aromatic hydrocarbons' (PAH) commonly refers to a large class of organic compounds containing two or more fused aromatic rings, even though in a broad sense non-fused ring systems should be included. In particular, the term 'PAH' refers to compounds containing only carbon and hydrogen atoms (i.e. unsubstituted parent PAH and their alkyl-substituted derivatives), whereas the more general term 'polycyclic aromatic compounds' also includes the functional derivatives (e.g. nitro- and hydroxy-PAH) and the heterocyclic analogues, which contain one or more hetero atoms in the aromatic structure (aza-, oxa-, and thia-arenes). Some authors refer to polycyclic aromatic compounds as 'polycyclic organic matter', and the term 'polynuclear' is frequently used for 'polycyclic', as in 'polynuclear aromatic compounds'. More than 100 PAH have been identified in atmospheric particulate matter (Lao et al., 1973; Lee et al., 1976a) and in emissions from coal-fired residential furnaces (Grimmer et al., 1985), and about 200 have been found in tobacco smoke (Lee et al., 1976b, 1981). The selection of PAH evaluated in this monograph is discussed in Section 1. The nomenclature, common names, synonyms, and abbreviations used are given in Table 1 in that section. The structural formulae are shown in Figure 1. Molecular formulae, relative molecular masses, and CAS Registry numbers are given in Table 3. 2.1.1 Technical products Technical-grade naphthalene, also known as naphthalin and tar camphor, has a minimum purity of 95%. The impurities reported are benzo [b]thiophene (thianaphthene) when naphthalene is obtained from coal-tar and methylindenes when it is derived from petroleum (Society of German Chemists, 1989). Commercially available anthracene, also known by the trade name Tetra Olive N2G (IARC, 1983), has a purity of 90-95% (Hawley, 1987). The impurities reported are phenanthrene, chrysene, carbazole (Hawley, 1987), tetracene, naphthacene (Budavari et al., 1989), and pyridine at a maximum of 0.2% (IARC, 1983). The following purities were reported for other technical-grade products: acenaphthene, 95-99%; fluoranthene, > 95% (Griesbaum et al., 1989); fluorene, about 95%; phenanthrene, 90%; and pyrene, about 95% (Franck & Stadelhofer, 1987). The other compounds are generally produced as chemical intermediates and for research purposes (see also sections 3.2.2 and 3.2.3). Reference materials certified to be of geater than 99% purity are available for 22 of the PAH considered (Community Bureau of Reference, 1992); the remaining compounds are commercially available as chemical standards, with a purity of 99% or more. Table 3. Identity of polycyclic aromatic hydrocarbons covered in this volume ranked according to molecular mass Compound Molecular Relative CAS formula molecular Registry mass No. Naphthalene C10H8 128.291-20-3 Acenaphthylene C12H8 152.2208-96-8 Acenaphthene C12H10 154.283-32-9 Fluorene C13H10 166.286-73-7 Anthracene C14H10 178.2120-12-7 Phenanthrene C14H10 178.285-01-8 1-Methylphenanthrene C15H12 192.3832-69-9 Fluoranthene C16H10 202.3206-44-0 Pyrene C16H10 202.3129-00-0 Benzo[a]fluorene C17H12 216.3238-84-6 Benzo[b]fluorene C17H12 216.3243-17-4 Benzo[ghi]fluoranthene C18H10 226.3203-12-3 Cyclopenta[cd]pyrene C18H10 226.3 2720837-3 Benz[a]anthracene C18H12 228.356-55-3 Benzo[c]phenanthrene C18H12 228.3195-19-7 Chrysene C18H12 228.3218-01-9 Triiphenylene C18H12 228.3217-59-4 5-Methylchrysene C19H14 242.33697-24-3 Benzo[b]fluoranthene C20H12 252.3205-99-2 Benzo[j]fluoranthene C20H12 252.3205-82-3 Benzo[k]fluoranthene C20H12 252.3207-08-9 Benzo[a]pyrene C20H12 252.350-32-8 Benzo[e]pyrene C20H12 252.3192-97-2 Perylene C20H12 252.3198-55-0 Anthanthrene C22H12 276.3191-26-4 Benzo[ghi]perylene C22H12 276.3191-24-2 Indeno[1,2,3-cd]pyrene C22H12 276.3193-39-5 Dibenz[a,h]anthracene C22H14 278.453-70-3 Coronene C24H14 300.4191-07-1 Dibenzo[a,e]pyrene C24H14 302.4192-65-4 Dibenzo[a,h]pyrene C24H14 302.4189-64-0 Dibenzo[a,i]pyrene C24H14 302.4189-55-9 Dibenzo[a,l]pyrene C24H14 302.4191-30-0 PAH considered (Community Bureau of Reference, 1992); the remaining compounds are commercially available as chemical standards, with a purity of 99% or more. 2.2 Physical and chemical properties Physical and chemical properties relevant to the toxicological and ecotoxicological evaluation of the PAH are summarized in Table 4. It should be kept in mind that the values for any one parameter may be derived from different sources, with different methods of measurement or calculation, so that individual values cannot be compared directly unless the original sources are consulted. In particular, the vapour pressures reported in the literature for the same PAH vary by up to several orders of magnitude (Mackay & Shiu, 1981; Lane, 1989). Variations are also seen in the reported solubility in water of various PAH, although the values are generally within one order of magnitude (National Research Council Canada, 1983). Flash-points were available only for three compounds with high molecular mass (for naphthalene, 78.9°C by the open-cup method and 87.8°C by the closed-cup method; anthracene, 121°C by the closed-cup method; and phenanthrene, 171°C by the open-cup method). Explosion limits were available only for naphthalene (0.9-5.9 vol %) and ananthrene (0.6 vol %) (Lewis, 1992). Vapour density (air = 1) was 4.42 for naphthalene (IARC, 1973), 5.32 for acenaphthene, 6.15 for anthracene (Lewis, 1992), 6.15 for phenanthrene, and 8.7 for benzo[a]pyrene (National Institute for Occupational Safety and Health and Occupational Safety and Health Administration, 1981). The physical and chemical properties are largely determined by the conjugated alpha-electron systems, which vary fairly regularly with the number of rings and molecular mass, giving rise to a more or less wide range of values for each parameter within the whole class. At room temperature, all PAH are solids. The general characteristics common to the class are high melting- and boiling-points, low vapour pressure, and very low solubility in water. PAH are soluble in many organic solvents (IARC, 1983; Agency for Toxic Substances and Disease Registry, 1990; Lide, 1991) and are highly lipophilic. Vapour pressure tends to decrease with increasing molecular mass, varying by more than 10 orders of magnitude. This characteristic affects the adsorption of individual PAH onto particulate matter in the atmosphere and their retention on particulate matter during sampling on filters (Thrane & Mikalsen, 1981). Vapour pressure increases markedly with ambient temperature (Murray et al., 1974), which additionally affects the distribution coefficients between gaseous and particulate phases (Lane, 1989). Solubility in water tends to decreases with increasing molecular mass. For additional information, refer to section 4.1. PAH are chemically inert compounds (see also section 4.4). When they react, they undergo two types of reaction: electrophilic substitution and addition. As the latter destroys the aromatic character of the benzene ring that is affected, PAH tend to form derivatives by the former reaction; addition is often followed by elimination, resulting in net substitution. The chemical and photochemical reactions of PAH in the atmosphere have been reviewed Table 4. Physical and chemical properties of polycyclic aromatic compounds covered in this monograph, ranked by molecular mass Compound Colour Melting- Boiling- Vapour Densityc n-Octanol: Solubility in Henry's law pointa point pressure water water at 25°C constant at (°C) (°C) (Pa at 25°C) partition (µg/litre)d 25°C (kPa) coefficient (log Kow) Naphthalene Whiteb 81 217.9c 10.4g 1.15425 h 3.4j 3.17 x 104 4.89 x 10-2 k Acenaphthylene 92-93 8.9 x 10-1 g 0.89916/2 h 4.07f 114 x 10-3 l Acenaphthene Whiteb 95 279h 2.9 x 10-1 g 1.02490/4 h 3.92f 3.93 x 103 1.48 x 10-2 k Fluorene Whitee 115-116 295e 9.0 x 10-2 g 1.2030/4 h 4.18m 1.98 x 103 1.01 x 10-2 n Anthracene Colourlesso 216.4 342e 8.0 x 10-4 g 1.28325/4 h 4.5j 73 7.3 x 10-2 n Phenanthrene Colourlessp 100.5 340h 1.6 x 10-2 g 0.9804 h 4.6j 1.29 x 103 3.98 x 10-3 k 1-Methylphenanthrene 123 354-355y 5.07s 255 (24°C)t Fluoranthene Pale yellowh 108.8 375h 1.2 x 10-3 g 1.2520/4 h 5.22u 260 6.5 x 10-4 (20 °C)w Pyrene Colourlesse 150.4 393h 6.0 x 10-4 g 1.27123/4 h 5.18j 135 1.1 x 10-3 n Benzo[a]fluorene Colourlessx 189-190h 399-400y 5.32z 45 Benzo[b]fluorene Colourlessx 213.5 401-402y 1.226aa 5.75z 2.0 Benzo[ghi]fluoranthene Yellowbb 128.4 432cc 1.34523 dd Cyclopenta[cd]pyrene Orangex 170 439ee Benz[a]anthracene Colourlessb 160.7 400b 2.8 x 10-5 g 1.226aa 5.61f 14 Benzo[c]phenanthrene Colourlessx 66.1 1.265ff Chrysene Colourless 253.8 448h 8.4 x 10-5 1.27420/4 e 5.91u 2.0 with blue (20°C)gg fluoresenceb Triphenylene Colourlessx 199 425bb 1.3p 5.45hh 43 5-Methylchrysene Colourlessx 117.1 458ii 62 (27°C)jj Benzo[b]fluoranthene Colourlessi 168.3 481kk 6.7 x 10-5 6.12f 1.2ll 5.1 x 10-5 (20°C)gg (20°C)w Benzo[j]fluoranthene Yellowb 165.4 480ee 2.0 x 10-6 l 6.12mm 2.5nn Table 4. (continued) Compound Colour Melting- Boiling- Vapour Densityc n-Octanol: Solubility in Henry's law pointa point pressure water water at 25°C constant at (°C) (°C) (Pa at 25°C) partition (µg/litre)d 25°C (kPa) coefficient (log Kow) Renzo[k]fluoranthene Pale yellowh 215.7 480h 1.3 x 10-8 6.84m 0.76f 4.4 x 10-5 (20°C)oo (20°C)w Benzo[a]pyrene Yellowishe 178.1 496kk 7.3 x 10-7oo 1.351pp 6.50u 3.8 3.4 x 10-5 (20°C) Benzo[e]pyrene Pale yellowx 178.7 493kk 7.4 x 10-7qq 6.44rr 5.07 (23°C)tt Perylene Yellow to 277.5 503ss 1.35v 5.3uu 0.4 colourlessc Anthanthrene Golden yellowbb 264 547yy 1.39v Benzo[ghi]perylene Pale yellow- 278.3 545ii 1.4 x 10-8 ww 1.32920 xx 7.10u 0.26 2.7 x 10-5 greenbb (20°C)w Indeno[1,2,3-cd]pyrene Yellowi 163.6 536yy 1.3 x 10-8 6.58f 62f 2.9 x 10-5 (20°C)gg (20°C)w Dibenz[a,h]anthracene Colourlessi 266.6 524yy 1.3 x 10-8 1.282i 6.50zz 0.5 (27°C)jj 7 x 10-6 l (20°C) Coronene Yellowh 439 525aaa 2.0 x 10-10 qq 1.37b 5.4uu 0.14 Dibenzo[a,e]pyrene Pale yellowh 244.4 592vv Dibenzo[a,h]pyrene Golden yellowi 317 596vv Dibenzo[a,i]pyrene Greenish-yellowishi 282 594vv 3.2 x 10-10 mm 7.30hh 0.17l 4.31 x 10-6 l Dibenzo[a,l]pyrene Pale yellowi 162.4 595vv a From Karcheret al. (1985); Karcher (1988) b From Lewis (1992) c When two temperatures are given as superscripts, they indicate the specific gravity, i.e. the density of the substance at the first reported temperature relative to the density of water at the second reported temperature. When there is no value, or only one, for temperature, the datum is in grains per millilitre, at the indicated temperature, if any. Table 4 (continued) d From Mackay & Shiu (1977), except where noted e From Budavari (1989) f From National Toxicology Program (1993) g From Sonnefeld et al. (1983) h From Lide (1991) i From IARC (1977) j From Karickhoff et al. (1979) k From Mackay et al. (1979) l Calculated by Syracuse Research Center; from National Toxicology Program (1993) m Calculated as per Leo et al. (1971); from US Environmental Protection Agency (1980) n From Mackay & Shiu (1981) o When pure, colourless with violet fluorescence; from Budavari (1989) p From Hawley (1987) q From National Institute for Occupational Safety and Health and Occupational Safety and Health Administration (1981) r From Kruber & Marx (1938) s Calculated by Karcher et al. (1991) t From May et al. (1978) u From Bruggeman et al. (1982) v At ambient temperature; from Inokuchi & Nakagaki (1959) w From Ten Hulscher et al. (1992) x Personal observation by J. Jacob, Germany, on high-purity, certified reference materials y From Kruber (1937) z Calculated by Miller et al. (1985) aa From Schuyer et al. (1953) bb From IARC (983) cc From Kruber & Grigoleit (1954) dd From Ehrlich & Beevers (1956) ee Reported by Grimmer (1983a) ff From Beilstein Institute for Organic Chemistry (1993) gg Reported by Sims & Overcash (1983) hh Calculated by Yalkowsky & Valvani (1979) ii Calculated by White (1986) jj From Davis et al. (1942) kk From review by Bjorseth (1983); original references cited by White (1986) ll Temperature not given; reported by Sims & Overcash (1983) mm Calculated by National Toxicology Program (1993) nn Temperature not given; unpublished result cited by Wise et al. (1981) oo From US Environmental Protection Agency (1980) Table 4 (continued) pp From Kronberger & Weiss (1944) qq From review of Santodonato et al. (1981) rr Calculated by Ruepert et al. (1985) ss From Verschueren (1983) tt From Schwarz (1977) uu From Brooke et al. (1986) vv From Agency for Toxic Substances and Disease Registry (1990) xx From White (1948) yy Estimated from gas chromatographis retention time; from Grimmer (1983a) zz From Means et al. (1980) aaa From Von Boente (1955) (Valerio et al., 1984; Lane, 1989). After photodecomposition in the presence of air and sunlight, a number of oxidative products are formed, including quinones and endoperoxides. PAH have been shown experimentally to react with nitrogen oxides and nitric acid to form the nitro derivatives of PAH, and to react with sulfur oxides and sulfuric acid (in solution) to form sulfinic and sulfonic acids. PAH may also be attacked by ozone and hydroxyl radicals present in the atmosphere. The formation of nitro-PAH is particularly important owing to their biological impact and mutagenic activity (IARC, 1984a, 1989a). In general, the above reactions are of interest with regard to the environmental fate of PAH, but the results of experimental studies are difficult to interpret because of the complexity of interactions occurring in environmental mixtures and the difficulty in eliminating artefacts during analytical determinations. These reactions are also considered to be responsible for possible losses of PAH during ambient atmospheric sampling (see section 2.4.1.1). 2.3 Conversion factors Atmospheric concentrations of PAH are usually expressed as micrograms or nanograms per cubic meter. At 25°C and 101.3 kPa, the conversion factors for a compound of given relative molecular mass are obtained as follows: ppb = µg/m3 × 24.45/relative molecular mass µg/m3 = ppb × relative molecular mass/24.45. For example, for benzo [a]pyrene, 1 ppb = 10.3 µg/m3 and 1 µg/m3 = 0.0969 ppb. 2.4 Analytical methods Tables 5 and 6 present as examples a limited number of methods that are applied to 'real' samples of different matrices. The methods and sources were selected, as far as possible, according to the following criteria: accessibility of the bibliographic source, completeness of the description of the procedure, practicability with common equipment for this type of analysis (even if experienced personnel are required), recency, and whether it is an official, validated, or recommended method. 2.4.1 Sampling 2.4.1.1 Ambient air The physical state of PAH in the atmosphere must be considered when selecting the sampling apparatus. Compounds with five or more rings are almost exclusively adsorbed on suspended particulate matter, whereas lower-molecular-mass PAH are partially or totally present in the vapour phase (Coutant et al., 1988). When ambient air is monitored, it is common practice to monitor only particle-bound PAH Table 5. Analytical methods for polycyclic aromatic hydrocarbons in air Matrix Sampling, extraction Clean-up Analysis Limit of Reference detectiona Ambient air Sampling on GF+PUF, at 45 m3/h; Liquid-liquid partition GC/MS Yamasaki et al. Soxhlet extraction with cyclohexane with cyclohexane: (1982) H2O:DMSO, then CC with SiO2 Sampling on GF+PUF, at 30 m3/h; CC with Al2O3 + HPLC/FL 0.01-0.7 Keller & Soxhlet extraction with petroleum ether SiO2 ng/m3 Bidleman (1984) (GF) and DCM (PUF) Sampling on GF (particle diameter TLC with SiO2 HPLC/UV 0.01-0.3 Greenberg et al. < 15 µm), at 68 m3/h; Soxhlet extraction + FL ng/m3 (1985) with cyclohexane, DCM, and acetone Sampling on GF at 83 m3/h; sonication TLC with SiO2 GC/FID Valerio et al. (cyclohexane) (1992) Emissions Sampling by glass wool, condenser, Liquid-liquid partition GC/FID 10 ng/m3 Colmsjo et al. (municipal and XAD-2; extraction with acetone with DMF (1986a) incinerator) (glass-wool and XAD-2, by Soxhlet) Vehicle Sampling by GF and condenser; liquid- CC with SiO2 and GC/FID 2.5-20 ng Grimmer et al. exhaust liquid partition with acetone:H2O: Selphadex LH-20 per test (1979) cyclohexane and DMF:H,O:cyclohexane Sampling in dilution tunnel by Liquid-liquid partition GC/FID or Westerholm et PTFE-coated GF and condenser; Soxhlet with cyclohexane: GC/MS al. (1988) extraction of filter (DCM) and H2O:DMF condensate (acetone); remaining aqueous phase extracted with DCM Table 5. (continued) Matrix Sampling, extraction Clean-up Analysis Limit of Reference detectiona Indoor air Sampling on GF (particle diameter TLC with acetyloxylated Spectrofluorescence Lioy at al. (1988) < 10 µm) at 10 l/min; sonication cellulose (benzo[a]pyrene only) (cyclohexane) Sampling on quartz-fibre filtre and GC/MS Chuang at al. XAD-4 at 226 l/min; Soxhlet extraction (1991) with DCM Sampling on PTFE-coated GF at filtration; then CC HPLC/FL 0.02-0.12 Daisey & Gundel 20 l/minfor 24 h; Soxhlet extraction SiO2 cartridge), ng/m3 b (1993) with DCM optional Sampling on GF and PUF, at 20 litres/min GC/FID, GC/MS US Environmental for 24 h; Soxhlet extraction (10% ether: or HPLC/UV + FL Protection Agency n-hexane) (1990) Workplace air Sampling on PTFE filter and XAD-2 GC/FID 0.3-0.5 µg NIOSH (1994a,b) at 2 l/min; sonication or Soxhlet per sample extraction of filterc, extraction of HPLC/UV 0.05-0.8 µg XAD-2 with toluene (for GC) or + FL per sample acetonitrile (for HPLC) Workplace air Sampling on filter (GF, quartz fibre, CC (XAD-2) GC/FID approx 0.5 German PTFE or silver membrane) at 2 litres/min; µg/m3 Research sonication or Soxhlet extraction with Commission cyclohexane or toluene (1991) Tobacco Sampling by acetone trap; solvent CC (SiO2 + Sephadex GC/MS + ng/cigarette Lee at al. (1976b) smoke partition scheme (acids/bases/neutral LH-20); then NMR compounds/PAH) HPLC/UV Table 5 (continued) GC glass fibre; PUF, polyurethane foam; DMSO, dimethyl sulfoxide; CC, column chromatography; GC, gas chromatography; MS, mass spectrometry; DCM, dichloromethane; HPLC, high-performance liquid chromatography; FL, fluorescence detection; TLC, thin-layer chromatography; UV, ultraviolet detection; FID, flame-ionization detection; DMF, N-dimethylformamide; PTFE, polytetrafluoroethylene; NMR, nuclear magnetic resonance a Various PAH b The following PAH can be determined: fluoranthene, pyrene, chrysene, benzo[e]pyrene, benzo[b]fluoranthene, benzo[k]fluoranthene, benzo[a]pyrene, benzo[ghi]perylene, indeno[1,2,3-cd]pyrene. c Appropriate solvent must be determined by recovery tests on specific samples. Table 6. Analytical methods for polycyclic aromatic hydrocarbons in matrices other than air Matrix Extraction Clean-up Analysis Limit of Reference detectiona Tap-water Preconcentration on PUF; Liquid-liquid partition GC/FID or TLC 0.1 ng/litre Basu & Saxena extraction (with acetone and with cyclohexane: (Al2O3: acetyl (1978a) cyclohexane) H2O:methanol and celluose) with FL cyclohexane: H2O: detector DMSO; then CC OWN) Groundwater Liquid-liquid partition with CC (SiO2), if needed GC/FID µg/litre level US Environmental DCM GC/MS 10 µg/litre Protection Agency HPLC/UV + FL 0-01-2 µg/litre (1986a) Wastewater Liquid-liquid partition with CC (SiO2), if needed GC/FID or 0.01 -0.2 µg/litre US Environmental DCM HPLC/UV+FL (by HPLC) Protection Agency (1984a) Seawater Liquid-liquid partition with CC (SiO2 + Al2O2) GC/FID or Desideri at al. n-hexane or CCl4 HPLC/UV (1984) Soil Sonication with DCM CC (Al2O2); then GC/MS 1 µg/kg Vogt at liquid-liquid partition al. (1987) (n-hexane:H2O:DMSO) Soxhlet extraction with DCM CC (Florisil cartridge) HPLC/UV + FL 1 µg/kg Jones et a[. (1989a) Sediment Soxhlet extraction with DCM CC (SiO2 + Sephadex HPLC/DAD/MS Quilliam & Sim LH20) (1988) Sonication with acetone: CC (Florisil) HPLC/UV + FL 1-160 µg/kg Marcus et al. n-hexane (1988) Table 6. (continued) Matrix Extraction Clean-up Analysis Limit of Reference detectiona Meat and fish (I) digestion (alcoholic KOH), Liquid-liquid partition GC/FID 2.5-20 ng/ Grimmer & products (I), then liquid-liquid partition with cyclohexane: sample Bohnke (1979b) vegetable oils (methanol: H2O:cyclohexane) H2O:DMF); then CC (II), and sewage (II) dissolution in cyclohexane (SiO2 + Sephadex sludge (III) (III) refluxing with acetone LH20) Food (total Refluxing with alcoholic KOH, Liquid-liquid partition HPLC/FL 0.002-0.7 µg/kg Dennis et al. diet) extraction with isooctane (isooctane:H2O:DMF); (1983) then CC (SiO2 cartridge) Saponification with alcoholic CC (SiO2) HPLC/FL 0.03-2 µg/kg de Vos et al. KOH, extraction with (1990) cyclohexane Saponikation wit ahoholic CC (Florisil); then TLC/UV+FL 0.02 µg/kg Howard (1979); KOH, extraction with liquid-liquid partition (benzo[a]pyrene) Fazio (1990) isooctane isooctane:H2O:DMSO) Seafood Digestion with alcoholic KOH, CC (Al2O3 + SiO2 + HPLC/FL 0.01-0.6 µg/kg Perfetti et al. extraction with TCTFE C18 cartridge) (1992) Smoked food Digestion with alcoholic KOH, CC (Al2O3 + SiO2); HPLC/UV+FL 0.03-0.4 Joe et al. (1984) extraction with TCTFE liquid-liquid partition µg/kg (cyclohexane:H2O:DMSO) Refluxing with cyclohexane or Liquid-liquid partition TLC/FLb (only 0 0.5 ng/kg IUPAC (1987) TCTFE, extraction with with cyclohexane:H2O: benzo[alpyrene) methanol:H2O DMF); then CC (SiO2) Solid waste Soxhlet extraction with DCM CC (SiO2), if needed GC/FID µg/kg level US Environmental or sonication with GC/MS 1-200 mg/kg Protection Agency DGM:acetone HPLC/UV + FL µg/kg level (1986b) Table 6. (continued) Matrix Extraction Clean-up Analysis Limit of Reference detectiona Mineral oil and Liquid-liquid partition with CC (SiO2 + Sephadex GC/FID 100 ng/kg Grimmer & fuel cyclohexane:H2O:DMF) LH20) Bohnke (1979a) Medicinal oil Liquid-liquid partition CC (SiO2 + Sephadex HPLC/FL + 0.2-200 ng/kg Geahchan at al. (cyclohexane: H2O:DMF) LH20) GC/FID (1991) Plants Sonication (acetonitrile), CC (SiO2) GC/FID Coates et al. extraction with pentane (1986) Urine Adjusted to pH3, extraction CC (SiO2 cartridge) HPLC/FLc Becher & Bjorseth in C18 cartridge, metabolites (1983) reduced with hydriodic acid Urine and Addition of HCl, refluxing CC (SiO2) + Sephadex GC/MSd Jacob at al. (1989) faeces with toluene, addition of LH20 methanol and diazomethanol in ether (faeces saponified before acidification) Tissue Homogenization (benzene: CC (Florisil) GC/MS 5-50 µg/kg Liao et al. (1988) n-hexane) Skine Sonication of exposure pads HPLC/FL 6 ng/cm2 Jongeneelen et al. with DCM, centrifugation (1988a) Table 6. (continued) PUF, polyurethane foam; DMSO, dimethyl sulfoxide; CC, column chromatography; GC, gas chromatography; FID, flame ionization detection; FL, fluorescence detection; DCM, dichloromethane; MS, mass spectrometry; UV, ultraviolet detection; DAD, diode-array detector; DMF, N-dimethylformamide; TLC, thin-layer chromatography; TCTFE, 1,1,2-trichlorotrifluoroethane a Various PAH b Benzo[a]pyrene content estimated to be > 0.6 µg/kg (screening method) c Determination of unmetabolized and metabolized PAH d Determination of pyrene and 1-hydroxypyrene e Measurement of skin contamination with soft polypropylene exposure pads mounted on skin sites (Menichini, 1992a), probably because of the increased work involved in trapping volatile compounds, both in assembling the sampling unit and in analysing samples, and also because lighter compounds are of lesser toxicological interest. Of the PAH that are classified as 'probably' and 'possibly' carcinogenic to humans (IARC, 1987), only benz [a]anthracene is found at significant levels in the vapour phase (Van Vaeck et al., 1984; Coutant et al., 1988; Baek et al., 1992). Sampling is generally performed by collecting total suspended particulate matter for 24 h on glass-fibre filters by means of high-volume samplers. Other filters that have been used are quartz fibres (Hawthorne et al., 1992), polytetrafluoroethylene (PTFE) membranes (Benner et al., 1989; Baek et al., 1992), and, in comparisons, PTFE-coated glass fibres (Lindskog et al., 1987; De Raat et al., 1990). The effects of these materials on the decomposition of PAH during sampling have been compared (see section 2.2). Some studies indicated that higher recoveries are obtained with PTFE and PTFE-coated filters (Lee et al., 1980a; Grosjean, 1983); however, more recent investigations did not confirm this finding (Lindskog et al., 1987; Ligocki & Pankow, 1989; De Raat et al., 1990). Moreover, when cellulose acetate membrane filters were compared with glass-fibre filters, they had similar efficiency for collecting heavier PAH, but the former had greater efficiency for collecting three- and four-ring compounds (Spitzer & Dannecker, 1983). The most widely used method for trapping vapour-phase PAH is adsorption on plugs of polyurethane foam located behind the filter (Keller & Bidleman, 1984; Chuang et al., 1987; De Raat et al., 1987a; Benner et al., 1989; Hawthorne et al., 1992). This method is widely accepted, probably because of the low pressure drop, the low blanks, the low cost, and ease of handling. Among the other sorbents tested (see also reviews by Leinster & Evans, 1986; Davis et al., 1987), further polymeric materials have received particular attention, including Amberlite XAD-2 resin, which is a valid alternative to polyurethane foam (Chuang et al., 1987), Porapak PS, which has been successfully tested in combination with a silanized glass-fibre filter at a flow rate of 2 m3/h (Jacob et al., 1990a), and Tenax(R) (Baek et al., 1992). The trapped vapours contain both the PAH that were initially present in the vapour phase and those already collected on the filter and volatilized during sampling (the 'blowing-off' effect) (Van Vaeck et al., 1984; Coutant et al., 1988). The amount of PAH found in the vapour phase increases with ambient temperature (Yamasaki et al., 1982). Samplers incorporating an annular denuder, as well as a filter and back-up trap, have been used to investigate phase distribution and artefact formation (Coutant et al., 1988, 1992). Sampling times are restricted to 24 h in order to avoid sample degradation and losses. Grimmer et al. (1982) proposed a useful method for controlling losses due to chemical degradation and volatilization from filters which is based on the invariability of PAH profiles (i.e. the ratio of all PAH to one another) at different collection times. The adsorption of gas-phase PAH onto a quartz-fibre filter has been investigated as a possible sampling artefact (Hart & Pankow, 1994); the results suggested that overestimation of particle-associated PAH can be avoided by replacing quartz-fibre filters with a PTFE membrane filters, or can be corrected by using back-up quartz-fibre filters. Elutriators and cascade impactors have been used to achieve particle size-selective sampling of PAH (Menichini, 1992a). Instruments designed as additions to high-volume samplers are available, including 'PM10' inlets, which allow collection of airborne particles with a 50% cutoff at the aerodynamic diameter of 10 m (US Environmental Protection Agency, 1987a; Lioy et al., 1988; Hawthorne et al., 1992), and cascade impactors (Van Vaeck et al., 1984; Catoggio et al., 1989). When PAH are collected in indoor air, samplers operating at 20 or 200 litre/min are commonly used. The filter and sorbent materials are those used for outdoor air (Wilson et al., 1991; see also Table 5). The sampling step is by far the most important source of variability in the results of atmospheric PAH determination. Most investigations are difficult to compare because of differences in factors such as season, meteorological conditions, time of day, number and characteristics of sampling sites, and sampling parameters (Menichini, 1992a). Passive biological sampling has been investigated as an approach to long-term sampling of atmospheric PAH (Jacob & Grimmer, 1992), and preliminary correlation factors have been determined by comparing the PAH profiles in biological (plants, particularly) and air samples. Of the matrices tested, spruce sprouts were found to be the most suitable. 2.4.1.2 Workplace air The general considerations described for ambient air are also valid for the working environment. Less volatile PAH may be retained than in ambient air because of the high temperatures that are often found at the workplace. In the potroom of an aluminium plant where Sderberg electrodes were used, 42% of benz [a]anthracene was found in the vapour phase (Andersson et al., 1983), and in an iron foundry at a site where the temperature of the PAH source was 600-700°C, four- to seven-ring PAH represented about 70% of the total in the vapour phase (Knecht et al., 1986). Glass-fibre or PTFE filters are usually used to collect particle-bound PAH. A number of back-up systems can be used to efficiently trap volatile PAH, including liquid impingers and solid sorbents such as Tenax(R)-GC, Chromosorb, and XAD-2 (Bjorseth & Becher, 1986; Davis et al., 1987). The latter seems to be the most practical. The US National Institute for Occupational Safety and Health (1994a,b) recommended use of a PTFE-laminated membrane followed by a tube containing two sections of XAD-2. For sampling in bright sunlight, opaque or foil-wrapped filter cassettes can be used to prevent degradation. The exposure of workers is estimated by taking air samples at various locations in the workplace or by personal sampling, in which workplace air is pumped through a filter attached to clothing close to the breathing zone for a specified time. Both procedures provide an estimate and not a precise measurement of an individual's exposure. 2.4.1.3 Combustion effluents The validity of a collected sample, i.e. the degree to which it reflects the 'true' composition of the emission, is a crucial factor in the determination of PAH in emissions. The problems associated with efficient collection of volatile PAH are enhanced when sampling combustion effluents, such as stack gases and vehicle exhausts, because of the elevated temperatures at sampling positions. A sampling device for stack gases is constituted by a glass- or quartz-fibre filter, followed by a special unit which generally consists in a cooler for collecting condensable matter and an adsorbent cartridge (Colmsjö et al., 1986a; Funcke et al., 1988). Tenax(R) has been used as an adsorbent (Jones et al., 1976), but XAD-2 seems to be more suitable (Warman, 1985) and is generally preferred. Two sampling procedures have been described in detail by the US Environmental Protection Agency (1986c). In the first ('Modified method 5 sampling train'), the unit basically includes a glass- or quartz-fibre filter kept at around 120°C, a condenser coil that conditions the gas at a maximum of 20°C, and a bed of XAD-2 jacketed to maintain the internal gas temperature at about 17°C. The second ('Source assessment sampling system') is often used for stationary investigations (Warman, 1985). The apparatus consists of a stainless-steel probe, which enters an oven containing the filter, preceded by three cyclone separators in series, with cutoff diameters of 10, 3, and 1 m; the volatile organic compounds are cooled and trapped on XAD-2. The sorbent is followed by a condensate collection trap and an impinger train. Motor vehicle exhausts are sampled under laboratory conditions, by chassis or engine dynamometer testing. Standard driving cycles are employed to simulate on-road conditions (Stenberg, 1985; see also section 3.2.7.2). Two basic techniques have been used to collect, sample, and analyse exhaust (Levsen, 1988; IARC, 1989a). In the first-raw gas sampling-the exhaust pipe is connected directly to the sampling apparatus; undiluted emissions are cooled in a condenser and then allowed to pass through a filter for collection of particulates (Grimmer et al., 1979, 1988a; Society of German Engineers, 1989). A second technique-dilution tube sampling-is now often used, in which hot exhaust is diluted with filtered cold air in a tunnel, from which samples are collected isokinetically. This technique simulates the process of dilution that occurs under real conditions on the road (US Environmental Protection Agency, 1992a). Particles are almost always collected on glass-fibre, glass-fibre with PTFE binder, quartz-fibre filters, or PTFE membranes; the latter have been reported to be particularly efficient and chemical inert (Lee & Schuetzle, 1983). Glass-fibre filters impregnated with liquid paraffin are also used (Grimmer et al., 1979; Society of German Engineers, 1989). Vapour-phase PAH (Stenberg, 1985) may be collected by cryo-condensation (Stenberg et al., 1983) or on an adsorbent trap with a polymeric material such as XAD-2 (Lee & Schuetzle, 1983). Artefacts may be introduced during collection on filters as a result of chemical conversion of PAH, particularly into nitro-PAH and oxidation products (Lee & Schuetzle, 1983; Schuetzle, 1983; IARC, 1989a). These effects have not been fully evaluated. 2.4.1.4 Water The concentrations of PAH in uncontaminated groundwater supplies and in drinking-water are generally very low, at 0.1 and 1 ng/litre (see sections 5.1.2.1 and 5.1.2.2). This implies that serious errors arising from adsorption losses and contamination occur during collection and storage of samples or that a preconcentration step may be needed to enrich the sample. It is recommended that sampling be performed on-site, directly in the extraction vessel (Smith et al., 1981). Various solid sorbents have been successfully used for preconcentration (Smith et al., 1981), including Tenax(R)-GC, prefiltered if necessary (Leoni et al., 1975); XAD resins (Griest & Caton, 1983); open-pore polyurethane foam (Basu et al., 1987); and prepacked disposable cartridges of bonded-phase silica gel (Chladek & Marano, 1984; Van Noort & Wondergem, 1985a). Solid sorbents have limitations when the sample contains suspended material, since adsorbed PAH may be lost by filtration (Van Noort & Wondergem, 1985a). 2.4.1.5 Solid samples Some foodstuffs (Liem et al., 1992), soil, sediment, tissues, and plants usually require homogenization before a sample is extracted. 2.4.2 Preparation As most environmental samples contain only small amounts of PAH, sophisticated techniques are required for their detection and quantification. Therefore, efficient extraction from the sample matrix is usually followed by one or more purification steps, so that the sample to be analysed is as free as possible from impurities and interference. Many extraction and purification techniques and combinations ('isolation schemes') have been described, validated, and recommended, but no single scheme is commonly recognized as 'the best' for a given matrix. The isolation schemes have been classified according to groups of matrices (Jacob & Grimmer, 1979; Grimmer, 1983a), as summarized briefly below. PAH are extracted from a sample (Lee et al., 1981; Santodonato et al., 1981; Grimmer, 1983a; Griest & Caton, 1983) with: - a Soxhlet apparatus, from filters loaded with particulate matter, vehicle exhausts, or sediments; - directly by liquid-liquid partition, for water samples; or - after complete dissolution (e.g. fats and vegetable and mineral oils) or alkaline digestion of samples (e.g. meat products) by a selective solvent such as N,N-dimethylformamide (Natusch & Tomkins, 1978) or dimethyl sulfoxide. Complete extraction of PAH from samples such as soot emitted by diesel engines, carbon blacks, and other carbonaceous materials is particularly difficult. Extraction of PAH from soil, sediment, sewage sludge, and vehicle exhaust particulates by refluxing with various solvents has been investigated. In all cases, toluene was found to be the most efficient solvent, especially for vehicle exhaust (Jacob et al., 1994). As an alternative to Soxhlet extraction, ultrasonic extraction (Griest & Caton, 1983) has advantages in terms of reduced time of extraction (minutes versus hours) and superior recovery efficiency and reproducibility, particularly for solid samples and filters loaded with particulate matter. Comparisons of techniques depend, however, on the matrix, solvent, and experimental conditions. Supercritical fluid extraction (Langenfeld et al., 1993) has gained attention as a rapid alternative to conventional liquid extraction from polyurethane foam sorbents (Hawthorne et al., 1989a), soil (Wenclawiak et al., 1992), and other environmental solids such as urban dust, fly ash, and sediment (Hawthorne & Miller, 1987). This technique can also be directly coupled with on-column gas chromatography (see section 2.4.3.1); the extract is quantitatively transferred onto the gas chromatographic column for a rapid (< 1 h) analysis with maximal sensitivity. This technique has been used for urban dust samples (Hawthorne et al., 1989b). Extracted samples are usually purified from interfering substances by adsorption column chromatography. The classical sorbents, alumina and silica gel, are widely used. In addition, the hydrophobic Sephadex LH-20 has been found to be suitable for isolating PAH from nonaromatic, nonpolar compounds, which is important if the sample is analysed by gas chromatography (Grimmer & Böhnke, 1979a); It has also been used in partition chromatography as a carrier of the stationary phase, to separate PAH from alkyl derivatives (Grimmer & Böhnke, 1979b). Chromatography on silica gel and Sephadex is often combined (Jacob & Grimmer, 1979; Grimmer, 1983a). Clean-up has also been achieved by eluting extracted samples through XAD-2 (soil samples: Spitzer & Kuwatsuka, 1986), XAD-2 and Sephadex LH-20 in series (foods: Vaessen et al., 1988), or Florisil (food, water, and sediment samples: references given in Table 6). Conventional chromatographic columns may be substituted by prepacked commercial cartridges, which have advantages in terms of time and solvents consumed and reproducibility. For example, silica cartridges have been used to purify foodstuffs (Dennis et al., 1983), urine (Becher & Bjorseth, 1983), vehicle emissions (Benner et al., 1989), mineral oil mist (Menichini et al., 1990), and atmospheric samples (Baek et al., 1992); soil samples have been cleaned up on Florisil cartridges (Jones et al., 1989a). Preparative thin-layer chromatography is also used for, e.g. air particulates (see Table 5) and vegetable oils (Menichini et al., 1991a). Handling of samples in the absence of ultraviolet radiation is recommended at all stages in order to avoid photodecomposition of PAH (Society of German Engineers, 1989; US Environmental Protection Agency, 1990; US National Institute for Occupational Safety and Health, 1994a,b). It is also generally recommended that possible sources of interference and contamination be controlled, particularly from solvents (US Environmental Protection Agency, 1984a, 1986b, 1990), and that samples be refrigerated until extraction (US Environmental Protection Agency, 1984a; US National Institute for Occupational Safety and Health, 1994a,b). 2.4.3 Analysis PAH are now routinely identified and quantified by gas chromatography or high-performance liquid chromatography (HPLC). Each technique has a number of relative advantages. Both are rather expensive, particularly HPLC, and require qualified operating personnel; nevertheless, they are considered necessary in order to analyse 'real' samples for a large number of PAH with accuracy and precision. 2.4.3.1 Gas chromatography Excellent separation (< 3000 plates per meter) is obtained by the use of commercially available fused silica capillary columns, making it possible to analyse very complex mixtures containing more than 100 PAH. The most widely used stationary phases are the methylpolylsiloxanes: especially SE-54 (5% phenyl-, 1% vinyl-substituted) and SE-52 (5% phenyl-substituted), but SE-30 and OV-101 (unsubstituted), OV-17 (50% phenyl-substituted), Dexsil 300 (carborane-substituted) and their equivalent phases are also used. Chemically bonded phases are used increasingly because they can be rinsed to restore column performance and undergo little 'bleeding' at the high temperatures of analysis (about 300°C) that are required for determining high-boiling-point compounds. Nematic liquid crystal phases (Bartle, 1985) have also been used to separate some isomeric compounds that are poorly resolved by siloxane phases, such as chrysene and triphenylene on N,N'-bis (para-methoxy-benzylidene)-a,a'-bi- para-toluidine (Janini et al., 1975) and N,N'-bis (para-phenylbenzylidene)-a,a'-bi- para-toluidine (Janini et al., 1976). Splitless or on-column injection is necessary to gain sensitivity in trace analysis, the latter being preferred as it allows better reproducibility. Flame ionization detectors are almost always used because of the excellent linearity, sensitivity, and reliability of their response. Since the signal is related linearly to the carbon mass of the compound, PAH are recorded in proportion to their quantities, and the chromatogram shows the quantitative composition of the sample directly. Because flame ionization detectors are non-selective, samples for gas chromatography must be highly purified. Peak identification, which is done routinely from data on retention, must be confirmed by analysing samples on a different gas chromatographic column, by an independent technique, such as HPLC, or by directly coupling a mass spectrometric detector to the gas chromatograph (Lee et al., 1981; Olufsen & Bjorseth, 1983; Bartle, 1985; Hites, 1989). Mass spectrometers have gained wide acceptance. They are powerful tools for identifying compounds, especially when commercially available libraries of reference spectra are used to match the spectra obtained and to control the purity of a compound. As isomeric compounds often have indistinguishable spectra, however, the final assignment must also be based on retention. On-line coupling of liquid chromatography, capillary gas chromatography, and quadrupole mass spectrometry has been used to determine PAH in vegetable oils (Vreuls et al., 1991). 2.4.3.2 High-performance liquid chromatography The packing material considered most suitable for separating PAH consists of silica particles chemically bonded to linear C18 hydrocarbon chains; selection of the appropriate phase has been discussed in detail by Wise et al. (1993). Typically, 25-cm columns packed with 5-m particles are used in the gradient elution technique, and the mobile phase consists of mixtures of acetonitrile and water or methanol and water ('reversed-phase HPLC'). As the efficiency of separation that can be achieved with HPLC columns is much lower than that with capillary gas chromatography, HPLC is generally less suitable for separating samples containing complex PAH mixtures. The advantages of HPLC derive from the capabilities of the detectors with which it is used. Those most widely used for PAH are ultraviolet and fluorescence detectors, generally arranged in series, with flow-cell photometers or spectrophotometers. Both, but especially the latter, are highly specific and sensitive: the detection limits with fluorescence are at least one order of magnitude lower than those with ultraviolet detection. The specificity of fluorescence detectors allows the determination of individual PAH in the presence of other nonfluorescing substances. In addition, since different PAH have different absorptivity or different fluorescence spectral characteristics at given wavelengths, the detectors can be optimized for maximal response to specific compounds. This may prove advantageous in the identification of unresolved components. In particular, wavelength-programmed fluorescence detection, to measure changes in excitation and emission wavelengths during a chromatographic run (Hansen et al., 1991a), is being used for the analysis of environmental samples (Wise et al., 1993). HPLC is suitable to a limited degree for lower-molecular-mass compounds like naphthalene, acenaphthene, and acenaphthylene, for which the detection limits are relatively high (US Environmental Protection Agency, 1984a). Owing to the selectivity of packing materials, various isomers that cannot be separated efficiently on the usual capillary gas chromatographic columns can be resolved at baseline and identified by HPLC. Such isomers include the pairs chrysene-triphenylene and benzo [b]fluoranthene-benzo [k]fluorathene (Wise et al., 1980). Coupling of a mass spectrometer to HPLC has also been used in detecting PAH (e.g., Quilliam & Sim, 1988). As much information on isomeric structure can be obtained from spectra seen during the elution of chromatographic peaks, an ultraviolet diode-array detector has been used to confirm peaks (Dong & Greenberg, 1988; Kicinski et al., 1989). For applications of HPLC to determination of PAH, reference should be made to published reviews (Lee et al., 1981; Wise, 1983, 1985). 2.4.3.3 Thin-layer chromatography Thin-layer chromatography is commonly used only for identifying individual compounds, such as benzo [a]pyrene, during screening (IUPAC, 1987) or for identifying selected PAH, such as the six PAH that WHO (1971) recommended be determined in drinking-water (Borneff & Kunte, 1979). It is an inexpensive, quick analytical technique but has low separation efficiency. The last parameter is improved by two-dimensional processes (see, e.g. Borneff & Kunte, 1979). Quantification may be done by spectrophotometric or spectrofluorimetric methods in solution after the scrubbed substance spot has been extracted (Howard, 1979; Fazio, 1990) or in situ by scanning spectrofluorimetry (Borneff & Kunte, 1979). Acetylated cellulose is the adsorbent that has been used most widely for one-step separation of PAH fractions, and mixed aluminium oxide and acetylated cellulose have been used for two-dimensional development (Daisey, 1983). 2.4.3.4 Other techniques A number of unconventional instruments and techniques based on spectro-scopic principles have been developed as possible alternatives to the chromatographic methods for PAH. Most of them are, however, expensive, require skilled personnel, and are not yet considered useful for the practising analyst (Wehry, 1983; Vo-Dinh, 1989). Low-temperature luminescence in frozen solutions ('Shpol'skii effect') has been used for various environmental samples, particularly to identify methylated PAH isomers (Garrigues & Ewald, 1987; Saber et al., 1987). This technique was used widely in the countries of former Soviet Union (Dikun, 1967). Synchronous luminescence and room temperature phosphorimetry have been reported to be simple, cost-effective techniques for screening PAH (Vo-Dinh et al., 1984; Abbott et al., 1986). Infrared analysis, particularly Fourier transform infrared spectroscopy coupled to gas chromatography (Stout & Mamantov, 1989), and capillary supercritical fluid chromatography (Wright & Smith, 1989) have also been used. Various environmental samples have been analysed by packed column supercritical fluid chromatography, with rapid separation of PAH (Heaton et al., 1994). 2.4.4 Choice of PAH to be quantified The choice of PAH depends on the purpose of the measurement. For example, carcinogenic PAH are of interest in studies of human health, but other, more abundant PAH may be of interest in ecotoxicological studies. Quantification of a number of PAH is advantageous when the profiles are to be correlated with sources and/or effects. Table 7 lists the PAH that are required or recommended to be determined at national or international levels. According to an EEC (1980) Directive, which followed a WHO (1971) recommendation, the concentrations of six reference compounds (also known as 'Borneff PAH') must be measured in drinking-water in order to check its compliance with the cumulative limit value for the PAH class. The choice of these six PAH by WHO was not based on toxicological considerations but on the fact that analytical investigations were then largely confined to these relatively easily detected compounds (WHO, 1984). Table 7. Some polycyclic aromatic hydrocarbons required or recommended for determination by various authorities Compound WHO/EECa US EPAb European Italyd Norwaye (drinking- (waste Aluminium (air) water) water) Associationc Health Environment Acenaphthene X Acenapthylene X Anthracene X X X Anthanthrene X X Benz[a]anthracene X X X X X Benzo[a]fluorene X Benzo[a]pyrene X X X X X Benzo[b]fluoranthene X X X X X X Benzo[b]fluorene X Benzo[c]phenanthrene X X Benzo[e]pyrene X Benzo[ghi]perylene X X X X Benzo[j]fluoranthene X X X X Benzo[k]fluoranthene X X X X X X Chrysene X X X X Cyclopenta[a]pyrene X X Dibenzo[a,e]pyrene X X X Dibenz[a,h]anthracene X X X X X Dibenzo[a,h]pyrene X X X Dibenzo[a,i]pyrene X X X Dibenzo[a,l]pyrene X X Fluoranthene X X X X Fluorene X Indeno[1,2,3-cd]pyrene X X X X X X Naphthalene X X Phenanthrene X X X Pyrene X X X Triphenylene X Table 7 (continued) a Recommended by WHO (1971) and required by an EEC (1980) Directive b Required by the US Environmental Protection Agency (1984a) for the analysis of municipal and industrial wastewater c Recommended by the European Aluminium Association, Environmental Health and Safety Secretariat (1990) d Recommended by the Italian National Advisory Toxicological Committee for health-related studies (Menichini, 1992b) e Recommended at the International Workshop on polycyclic aromatic hydrocarbons (State Pollution Control Authority and Norwegian Food Control Authority, 1992) for studies of health and on the environment The method required by the US Environmental Protection Agency (1984a) for the analysis of municipal and industrial wastewater covers the determination of 16 'priority pollutant PAH' considered to be representative of the class. Outside the USA, this list of compounds is often taken as a reference list for the analysis of various environmental matrices. 3. SOURCES OF HUMAN AND ENVIRONMENTAL EXPOSURE Appraisal Coal and crude oils contain polycyclic aromatic hydrocarbons (PAH) in considerable concentrations owing to diagenetic formation in fossil fuels. The main PAH produced commercially are naphthalene, acenaphthene, anthracene, phenanthrene, fluoranthene, and pyrene. The release of PAH during production and processing, predominantly of plasticizers, dyes, and pigments, is of only minor importance. Most PAH enter the environment via the atmosphere from incomplete combustion processes, such as: - processing of coal and crude oil: e.g. refining, coal gasification, and coking; - heating: power plants and residential heating with wood, coal, and mineral oil; - fires: e.g. forest, straw, agriculture, and cooking; - vehicle traffic; and - tobacco smoking. Industrial processes such as coal coking, aluminium, iron and steel production, and foundries make important contributions to the levels of PAH in the environment. An important indoor source of exposure to airborne PAH, especially in developing countries, is cooking fumes (see section 5.2). The hydrosphere and the geosphere are affected secondarily by wet and dry deposition. PAH are released directly into the hydrosphere, for example during wood preservation with creosotes. Deposition of contaminated refuse like sewage sludge and fly ash may cause further emissions into the geosphere. It is very difficult to identify a source on the basis of the ratio of the measured concentrations of different individual PAH, and such studies are in most cases inconclusive. 3.1 Natural occurrence In some geographical areas, forest fires and volcanoes are the main natural sources of PAH in the environment (Baek et al., 1991). In Canada, about 2000 tonnes of airborne PAH per year are attributed to natural forest fires (Environment Canada, 1994). On the basis of samples from volcanoes, Ilnitsky et al. (1977) estimated that the worldwide release of benzo [a]pyrene from this source was 1.2-14 t/year; no estimate was given of total PAH emissions from this source. Coal is generally considered to be an aromatic material. Most of the PAH in coal are tightly bound in the structure and cannot be leached out, and the total PAH concentrations tend to be higher in hard coal than in soft coals, like lignite and brown coal. Hydroaromatic structures represent 15-25% of the carbon in coal. The PAH identified include benz [a]anthracene, benzo [a]pyrene, benzo [e]pyrene, perylene, and phenanthrene (Neff, 1979; Anderson et al., 1986). Table 8 shows the typical contents of PAH in different crude oils, such as those derived from coal conversion or from shale. Table 8. Polycyclic aromatic hydrocarbon content of crude oils from various sources Compound PAH content (mg/kg) in crude oil from Coala Petroleum Shale Acenaphthene 1700/1800 147-348 147-903 Anthracene 4100 204-321 231-986 Anthanthrene Trace/< 800 NR 0.3 Benz[a]anthracene Trace/< 2200 1-7 1 Benzo[a]fluorene 2100/2500 11-22 53 Benzo[a]pyrene < 500/< 1200 0.1-4 3-192 Benzo[b]fluorene < 1500/3400 < 13 140 Benzo[c]phenanthrene < 600/< 2200 NR NR Benzo[e]pyrene < 1200/1300 0.5-29 1-19 Benzofluorenesb < 500/< 1300 23 NR Benzo[ghi]fluoranthene 3200 NR NR Benzo[ghi]perylene 4300/6600 ND-8 1-5 ND-5 Chrysene < 1500/2500 7-26 3-52 Coronene NR 0.2 NR Dibenz[a,h]anthracene NR 0.4-0.7 1-5 Fluoranthene < 1900/< 3700 2-326 6-400 Fluorene 5300/9900 106-220 104-381 1-Methylphenanthrene < 1200/< 5100 > 21 NR Naphthalene 100/2800 402-900 203-1390 Perylene Trace/< 600 6-31 0.3-68 Phenanthrene 12 000/20 400 > 129-322 221-842 Pyrene 14 200/35 000 2-216 18-421 Triphenylene NR 3/13 0.5 From Guerin at al. (1978), Weaver & Gibson (1979), Grimmer at al. (1983a), Sporstol et al. (1983), IARC (1985, 1989b) Ranges represent at least three values; NR, not reported; ND, not detected a Two crude oils from coal conversion; single measurements b Isomers not specified Two rare PAH minerals have been described: the greenish-yellow, fluorescent curtisite from surface vents of hot springs at Skagg Springs, California, USA, and the bituminous mercury ore idrialite from Idria, Yugoslavia, the two main components of which are chrysene and dibenz [a,h]-anthracene. These minerals are assumed to have been formed by the pyrolysis of organic material at depths below that at which petroleum id generated (West et al., 1986). 3.2 Anthropogenic sources 3.2.1 PAH in coal and petroleum products Commercial processing of coal leads first to coal-tars, which are further processed to yield pitch, asphalt, impregnating oils (creosotes for the preservation of wood), and residue oils such as anthracene oil (IARC, 1985). The concentration of PAH in coal-tars is generally ¾ 1%; naphthalene and phenanthrene are by far the most abundant compounds, occurring at concentrations of 5-10%. Comparable levels were detected in high-temperature coal-tar pitches. The PAH content of soots is about one order of magnitude lower, and that of carbon and furnace blacks ranges from about 1 to 500 mg/kg, pyrene being present at the highest concentration (IARC, 1984a; Nishioka et al., 1986). The PAH contents of some impregnating oils, bitumens, asphalts, and roof paints are shown in Table 9. In bitumens, PAH constitute only a minor part of the total content of polyaromatic compounds. Table 9. Polycyclic aromatic hydrocarbon content of impregnating oils, bitumens, asphalts, and roof paints Compound Concentration (mg/kg) Impregnating Bitumens Road tar (asphalt, Roof oils (oil-derived) coal-derived) paint Anthracene 1600-22 500 0.01-0.32 4170-14 400 2380 Anthranthrene NR Trace-1.8 NR NR Benz[a]anthracene 169-11 700 0.14-35 6820-24 100 6640 Benzo[a]pyrene 45-3490 0.1-27 5110-10 400 5950 Benzo[b]fluoranthene 42-3630 5 4490-10 900 5420 Benzo[e]pyrene 65-2020 0.03-52 3300-6750 3820 Benzo[gh]perylene 57-570 Trace-15 2390-2730 3270 Benzo[k]fluoranthene 24-2610 0.024-0.19 3170-7650 4470 Chrysene NR 0.04-34 NR Chrysene + 779-12 900 NR 6820-26 100 7700 triphenylene Coronene NR 0.2-2.8 NR NR Fluoranthene 703-85 900 0.15-5 23 500-61 900 12 100 Fluorene 8040-58 400 NR 6310-15 500 2220 Indeno[1,2,3-cd]pyrene 57-273 Trace 3100-3530 3320 Perylene 66-744 0.08-39 1550-2300 1730 Phenanthrene 7070-159 300 0.32-7.3 20 300-52 500 8180 Pyrene 604-46 400 0.08-38 15 100-42 500 8960 Triphenylene NR 0.3-7.6 NR NR From IARC (1985), Lehmann et al. (1986), Knecht & Woitowitz (1990); NR, not reported; ranges represent at least three values The concentrations of PAH in petrol and diesel fuels for vehicles and in heating oils are several parts per million. Almost all compounds are present at < 1 mg/kg; only phenanthrene, anthracene, and fluoranthene are sometimes found at > 10 mg/kg (Herlan, 1982). The PAH levels in unused engine lubricating oils are of the same order of magnitude. During the use of petrol-fuelled engine oils, the PAH content rises dramatically, by 30-500 times; in comparison, the total PAH levels in used diesel-fuelled engine oils were only 1.4-6.1 times greater than that in an unused sample. The major constituents of used oils are pyrene and fluoranthene, although benzo [b]fluoranthene, benzo [j]-fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, and dibenz [a,h]anthracene were also detected at considerable concentrations (IARC, 1984a; Carmichael et al., 1990). PAH have also been found in machine lubricating and cutting oils, which is of interest for the estimation of exposure in the workplace. The concentrations were < 7 mg/kg, although phenanthrene may have been present at a higher level (Grimmer et al., 1981a; Rimatori et al., 1983; Menichini et al., 1990; Paschke et al., 1992). PAH were detected in coloured printing oils, the concentrations of individual compounds varying between < 0.0001 and 63 mg/kg (Tetzen, 1989). By far the most abundant compounds were fluoranthene and pyrene (> 1 mg/kg); benzo [ghi]fluoranthene, cyclopenta [cd]pyrene, benz [a]anthracene, benzo [c]-phenanthrene, chrysene, triphenylene, benzo [b+j+k]fluoranthenes, benzo [a]pyrene, benzo [e]pyrene, anthanthrene, benzo [ghi]perylene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, and coronene were found at concentrations of < 0.5 mg/kg. 3.2.2 Production levels and processes Most of the PAH considered in this monograph are formed unintentionally during combustion and other processes. Only a few are produced commercially, including naphthalene, acenaphthene, fluorene, anthracene, phenanthrene, fluoranthene, and pyrene (Franck & Stadelhofer, 1987). The most important industrial product is naphthalene (see section 3.2.3). In 1987, about 220 kt of this compound were produced in western Europe, 190 kt in eastern Europe, 170 kt in Japan, and 110 kt in the USA (Fox et al., 1988); in 1986, > 1 kt was produced in Canada (Environment Canada, 1994). In 1985, about 2.5 kt of acenaphthene and 20 kt of anthracene were produced worldwide (Franck & Stadelhofer, 1987). In 1986, 0.1-1 t anthracene and 1 t fluorene were produced in Canada (Environment Canada, 1994). In 1993, a major producer in Germany produced < 5000 t anthracene, < 1000 t acenaphthene, < 500 t pyrene, < 50 t phenanthrene, and < 50 t fluoranthene (personal communication, Rütgers-VfT AG, 1994). The substances are not synthesized chemically for industrial purposes but are isolated from products of coal processing, mainly hard coal-tar. The raw material is concentrated and the product purified by subsequent distillation and crystallization. Only naphthalene is sometimes isolated from pyrolysis residue oils, olefin fractions, and petroleum-derived fractions; it is also obtained by distillation and crystallization (Collin & Höke, 1985; Franck & Stadelhofer, 1987; Griesbaum et al., 1989; Collin & Höke, 1991). In the USA in 1970, the distribution of capacity was about 60% coal-tar- and 40% petroleum-derived naphthalene (Gaydos, 1981); more detailed data were not available. The purity of the technical-grade products is 90-99% (Collin & Höke, 1985; Franck & Stadelhofer, 1987; Griesbaum et al., 1989; Collin & Höke, 1991; see also Section 2). 3.2.3 Uses of individual PAH The uses of commercially produced PAH are as follows (Collin & Höke, 1985; Franck & Stadelhofer, 1987; Griesbaum et al., 1989; Collin & Höke, 1991): - naphthalene: main use: production of phthalic anhydride (intermediate for polyvinyl chloride plasticizers); also, production of azo dyes, surfactants and dispersants, tanning agents, carbaryl (insecticide), alkylnaphthalene solvents (for carbonless copy paper), and use without processing as a fumigant (moth repellent) (see Figure 2); - acenaphthene: main use, production of naphthalic anhydride (intermediate for pigments); also, for acenaphthylene (intermediate for resins); - fluorene: production of fluorenone (mild oxidizing agent); - anthracene: main use, production of anthraquinone (intermediate for dyes); also, use without processing as a scintillant (for detection of high-energy radiation); - phenanthrene: main use, production of phenanthrenequinone (intermediate for pesticides); also, for diphenic acid (intermediate for resins) - fluoranthene: production of fluorescent and vat dyes; - pyrene: production of dyes (perinon pigments). 3.2.4 Emissions during production and processing of PAH The emissions of PAH during industrial production and processing in developed countries are not thought to be important in comparison with the release of PAH from incomplete combustion processes, since closed systems and recycling procedures are usually used. Few data were available. 3.2.4.1 Emissions to the atmosphere No data were available. 3.2.4.2 Emissions to the hydrosphere During the refining of aromatic hydrocarbons, and especially hard coal-tar, 80-190 t/year were estimated to be released to the hydrosphere in western Germany until 1987. This quantity was reduced to 8-19 t/year by the installation of new adsorption devices (sand filtration and adsorbent resin) by the two German hard coal-tar refineries in 1989 and 1991 (Klassert, 1993). 3.2.5 Emissions during the use of individual PAH Only naphthalene is used directly (as a moth repellent) without further processing. On the assumption that all naphthalene-containing moth repellent is emitted into the atmosphere, the emissions would have been about 15 000 t/year in western Europe in 1986, about 4400 t/year in Japan in 1987, and about 5500 t/year in the USA in 1987 (Fox et al., 1988). 3.2.6 Emissions of PAH during processing and use of coal and petroleum products Coal coking, coal conversion by gasification and liquefaction, petroleum refining, and the production and use of carbon blacks, creosote, coal-tar, and bitumen from fossil fuels may produce significant quantities of PAH (Anderson et al., 1986). A great deal of information on emissions of PAH is available in the literature; this monograph gives an overview of the most reliable values. The emission profile depends on the source, and specific emission profiles are detectable only in the direct vicinity of the corresponding source. Generally, emissions are estimated on the basis of more or less reliable databases, which are not identified in most publications. The values reported give only a rough idea of the situation. 3.2.6.1 Emissions to the atmosphere (a) Coal coking During coal coking, PAH are released into the ambient air mainly when an oven is loaded through the charging holes and new coal is suddenly brought into contact with the hot oven, and from leaks around oven doors and battery-top lids (Bjorseth & Ramdahl, 1985; Slooff et al., 1989). The specific emission factor for both benzo [a]pyrene and benzo [e]pyrene during coal coking was 0.2 mg/kg coal charged (Ahland et al., 1985). The emission factor for total PAH was estimated to about 15 mg/kg coal charged (Bjorseth & Ramdahl, 1985). Stack gases were measured about 8 m away from the aperture through which coke was discharged at a Belgian coking battery. Although the effluent may have been slightly diluted with ambient air, the following PAH concentrations were detected: benz [a]anthracene plus chrysene, 580 ng/m3; benzo [k]fluoranthene, 500 ng/m3; benzo [a]pyrene plus benzo [e]pyrene, 470 ng/m3; fluoranthene, 330 ng/m3; pyrene, 180 ng/m3 benzo [ghi]perylene, 140 ng/m3; anthracene plus phenanthrene, 130 ng/m3; and perylene, 44 ng/m3 (Broddin et al., 1977). The release of total PAH in 1985 was estimated to about 630 t/year in the USA, 18 t/year in Sweden, and 5.1 t/year in Norway (Bjorseth & Ramdahl, 1985). The authors emphasized that their data are subject to uncertainty and should be used only as an indication of the order of magnitude. In 1990, the total PAH emission in Canada was estimated to be 13 t/year (Environment Canada, 1994). Further estimates of total annual emissions of individual PAH compounds during the coking of coal are shown in Table 10. Table 10, Estimated annual emissions of polycyclic aromatic hydrocarbons during coal coking in the Netherlands and western Germany Compound Annual Year Reference emission (t/year) Netherlands Anthanthrene 0.5 Before 1989 Slooff at al. (1989) Benz[a]anthracene 0.3 1988 Slooff at al. (1989) Benzo[a]pyrene 0.1 Before 1989 Slooff at al. (1989) Benzo[ghi]perylene 0.2 1988 Slooff et al. (1989) Benzo[k]fluoranthene 0.1 1988 Slooff at al. (1989) Chrysene 0.2 1988 Slooff at al. (1989) Fluoranthene 1.1 1988 Slooff at al. (1989) lndeno[1,2,3-cd]pyrene 0.1 1988 Slooff et al. (1989) Naphthalene 1.3 1987 Slooff et al. (1988) 2.0 Before 1989 Slooff et al. (1989) Phenanthrene 2.1 1988 Slooff et al. (1989) Western Germany Benzo[a]pyrene 1.1 1990 Ministers for the Environment (1992); 1.7 Zimmermeyer et al. (1991) Naphthalene 10.0 1987 Society of German Chemists (1989) The emission factors for benzo [a]pyrene in the coking industry in the North-Rhine Westphalia area of Germany have been assumed to have been reduced to an average of about 60 mg/t coke. The newest plants have emission factors of 40 mg/t coke (Eisenhut et al., 1990). The reduction in PAH discharge was brought about by technical improvements to existing plants, closure of old plants and their partial replacement by new plants, and a reduction in coke production (Zimmermeyer et al., 1991). Decreasing trends in the annual emissions of airborne PAH during coke production are also assumed to have occurred in other industrialized countries (western Europe, Japan, and the USA), but no data were available. (b) Coal conversion PAH emission factors measured in the USA during gasification of coal at the end of the 1970s ranged from about 1 µg/g burnt coal for chrysene and 1500 µg/g burnt coal for naphthalene. Three qualities of coal were analysed for naphthalene, acenaphthylene, fluorene, anthracene, phenanthrene, pyrene, benz [a]anthracene, chrysene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, indeno[1,2,3- cd]pyrene, and dibenzo [a,h]pyrene (Nichols et al., 1981). In 1981, the stack gas of one US pilot coal gasification plant with an outdoor filter contained 0.2 and 2.1 µg/m3 naphthalene at two sampling times and 6.8 µg/m3 phenanthrene (Osborn et al., 1984). Acenaphthylene was detected at concentrations of 0.11-0.12 µg/m3 in the stack gases of two Canadian pilot coal liquefaction plants (Leach et al., 1987). (c) Petroleum refining The average profile of PAH compounds in petroleum refineries indicates that at least 85% of the total concentration is made up of two-ring compounds (naphthalene and its derivatives) and 94% of two- and three-ring compounds. Compounds with five rings or more contributed less than 0.1% at the catalytic cracking unit. In turn-round operations on reaction and fractionation towers, naphthalene and its methyl derivatives accounted for more than 99% of the total PAH (IARC, 1989b). Little information is available on the concentrations of PAH in stack gases. The levels in one French (Masclet et al., 1984) and two US petroleum refining plants (Karlesky et al., 1987) are available (Table 11); no information was given about the sampling site in the French facility, but sampling in the US plants was at the distillation device and below the cracking tower. The results depended on which fuel was burnt and the positioning and type of sampling device in the stack. Table 11. Polycyclic aromatic hydrocarbon concentrations in the stack gases of petroleum refinery plants in France and the USA Compound Concentration (µg/m3) France USA Acenaphthene NR 0.018-0.035 Acenaphthylene NR 0.013/0.019 Anthracene 3.9 0.003-0.041 Benz[a]anthracene 1.6 0.051-0.801 Benzo[a]pyrene 0.4 0.261-3.17 Benzo[b]fluoranthene 1.3 0.323-0.616a Benzo[e]pyrene 2.8 NR Benzo[ghi]perylene 0.7 0.23/0.382 Benzo[k]fluoranthene 0.5 NR Chrysene 1.7 0.021-0.252 Coronene 1.0 NR Dibenzo[a,h]anthracene NR 0.177 Fluoranthene 2.3 0.030-0.577 Fluorene 2.4 0.041-2.48 Indeno[1,2,3-cd]pyrene 1.2 0.25/0.538 Naphthalene NR 0.052-0.113 Perylene ND ND Phenanthrene 7.9 0.040-9.13 Pyrene 4.3 0.016-3.56 From Masclet et aL (1984) and Karlesky et al. (1987) NR, not reported; ND, not detected, limit of detection not stated; /, single measurements a Plus benzo[k]fluoranthene Few data are available on the total release of PAH into the atmosphere during petroleum refining. In western Germany, the emissions of naphthalene during petroleum refining, including hard coal-tar processing, were estimated to be 11 t/year (year not given; Society of German Chemists, 1989). In the Netherlands, the release of total PAH in 1988 was estimated to be about 7 t/year; the burning of pitch contributed 6.6 t/year, regeneration of catalyst, 0.4 t/year, and refining, < 0.01-0.1 t/year (Slooff et al., 1989). In Canada, about 0.1 t total PAH were emitted into the atmosphere in 1990 (Environment Canada, 1994). (d) Other processes In a US oil-furnace carbon black plant, the following mean emission factors per kg carbon black produced were found for individual PAH in three runs in the main vent gas: acenaphthylene, 800 g; pyrene, 500 g; anthracene plus phenanthrene, 70 g; fluoranthene, 60 g; benzo [ghi]fluoranthene, 40 g; benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, 30 g; benzo [a]pyrene plus benzo [e]pyrene plus perylene, 30 g; benzo [ghi]perylene plus anthanthrene, 23 g; chrysene plus benz [a]anthracene, 9 g; indeno[1,2,3- cd]pyrene, < 2 g; and benzo [c]phenanthrene, < 2 g. The release of PAH into ambient air cannot be estimated from these emission factors, however, as an additional combustion device is fitted in most US carbon-black plants in which the process vent gases are burnt (Serth & Hughes, 1980). Compounds with five or more rings (e.g. benzo [a]pyrene) contributed about 0.3% to the total PAH released from the bitumen processing unit of a refinery (IARC, 1989b). The emissions of PAH from batch asphalt mixers are assumed to be low and to occur mainly in combustion gases (IARC, 1984a), although no experimental data were available. Few estimates have been made of the annual emissions of PAH from processes in which coal and coal products are used. The total release of PAH to the atmosphere during asphalt production in 1985 was estimated to be about 4 t in the USA, 0.1 t in Norway, and 0.3 t in Sweden (Bjorseth & Ramdahl, 1985). In Canada, the amount emitted in 1990 was estimated to be about 2.5 t (Environment Canada, 1994). The amount released during carbon-black production and processing in 1985 was estimated to be about 3 t in the USA and < 0.1 t in Sweden (Bjorseth & Ramdahl, 1985). In the Netherlands in 1988, about 3.3 t of total PAH were emitted during the storage and transport of anthracene oil, an intermediate in the processing of hard coal-tar (Slooff et al., 1989). (e) Use of impregnating oils (creosotes) in wood preservation Estimates of the total input of PAH into the atmosphere from wood preservation with creosotes were available only for the Netherlands for unspecified years, at about 320 t/year (Slooff et al., 1989) and 840 t/year (Berbee, 1992). In 1988, the PAH input during storage of preserved material was estimated by the same authors to be about 200 t naphthalene, 110 t phenanthrene, 30 t fluoranthene, 5 t anthracene, 1.1 t benz [a]anthracene, and 0.02 t benzo [k]fluoranthene. 3.2.6.2 Emissions to the hydrosphere (a) Coal coking The concentrations of PAH reported in wastewater effluents are shown in Table 12. The removal of PAH by biological oxidation in two US coal coking plants was 93 to > 99%. Higher-molecular-mass PAH, benzo [a]pyrene, dibenz [a,h]anthracene, and benzo [ghi]perylene, comprised a greater fraction (about 60%) of the total PAH content in the effluent than in the input stream (Walters & Luthy, 1984). The total concentration of PAH discharged into the aqueous environment from a Norwegian coking plant was estimated to be about 23 kg/d (Berglind, 1982). On the basis of Dutch emission factors, the release in western Europe in 1985 of fluoranthene was calculated to be about 5 t and that of benzo [a]pyrene about 0.7 t (Berbee, 1992). The total annual input of PAH into the aqueous environment of the Netherlands was estimated to be about 1.7 t (year not given; Slooff et al., 1989). (b) Coal conversion The PAH content of wastewater from coal and shale conversion was < 0.5 mg/litre (Guerin, 1977). In raw, untreated wastewaters from a US pilot coal liquefaction plant, numerous PAH were found to emanate from the liquefaction section, the untreated hydrogenation section, and the still bottoms processing device when two kinds of coal were tested; for example, benzo [a]pyrene was found at a concentration of 0.3-52 µg/litre (Robbins et al., 1981). Numerous PAH were found in raw wastewater samples from two US pilot coal gasification plants (Walters & Luthy, 1981; Abbott et al., 1986), the maximum level of benzo [a]pyrene being 5.0 µg/litre. No information was available about total PAH emissions into the aqueous environment from commercial coal conversion plants. In groundwater near a US in-situ coal gasification site, naphthalene was found at a concentration of 2.7 µg/litre and acenaphthene and fluorene at < 0.1 µg/litre (Pellizzari et al., 1979). Until 1988, the final effluent from the two hard coal-tar refineries in western Germany contained an average of 50 mg/litre naphthalene, with a maximum of 120 mg/litre. The annual emission of this compound was thus calculated to be about 80 t. By 1991, the estimated release of naphthalene had been reduced to about 8 t/year by the addition of adsorption devices (Klassert, 1993). (c) Petroleum refining and offshore oil-well drilling PAH concentrations in wastewater effluents from these sources are summarized in Table 13. A refinery-activated sludge unit with a dual-media filter removed about 95% of the five-ring PAH and 99% of the four-ring PAH from the effluent of a petroleum refinery (Pancirov et al., 1980). A similar elimination efficiency was found for dissolved air flotation treatment of refinery wastewater and subsequent removal by activated sludge. Air stripping of the compounds in the sewage plant seemed to be of minor importance (Snider & Manning, 1982). The concentrations of PAH with more than three rings were found to be < 0.05 µg/litre even in the input to a sewage device and < 0.02 µg/litre in the final effluent (German Society for Mineral-oil and Coal Chemistry, 1984). The authors stated that these levels were of the same order of magnitude as the background concentrations in surface waters. The discharge of total PAH from a Norwegian petroleum refinery was about 0.26 kg/day (Berglind, 1982). The total concentration of PAH released into the North Sea from offshore oil-well drilling activities was about 2.5 t/year in 1987, comprising 2 t/year from drill rinsing and 0.2 t/year from shipping (Slooff et al., 1989). (d) Use of impregnating oils (creosotes) in wood preservation PAH were detected at levels of milligrams per litre in groundwater under a former wood preserving facility in Florida, USA. The concentrations of lower-molecular-mass creosote constituents were smaller in the groundwater than in an unweathered standard, probably because of greater mobility and biodegradability (Mueller & Lantz, 1993; Middaugh et al., 1994). Model experiments with fresh and seawater were carried out to determine the release of PAH from marine pilings made from southern pine and preserved with creosote (Ingram et al., 1982). The PAH levels per litre fresh water in the leachate at 20°C after immersion for three days were: naphthalene, 200-350 g; acenaphthene, 190-230 g; phenanthrene, 190-230 g; fluorene, 120-150 g; acenaphthylene, 51-88 g; anthracene, 48-76 g; fluoranthene, 27-30 g; pyrene, 12 g; and benz [a]anthracene, 11-19 g. The concentrations in seawater were three to four times lower. The amounts of PAH leached increased with increasing temperature. The concentrations in leachates from pilings that had been in seawater for 12 years were of the same order of magnitude. In contrast, rapidly decreasing PAH concentrations were found three months after the start of the experiment in runoff rainwater from spruce and pine pilings impregnated with hard coal-tar (van Dongen, 1987). The total PAH emissions into water and soil in the Netherlands from commercial wood preservation were about 28 t/year (year not given). The release of 10 PAH into water during the storage of creosote-preserved wood was about 16 t/year; the PAH measured were naphthalene, anthracene, phenanthrene, fluoranthene, benz [a]anthracene, benzo [a]pyrene, benzo [ghi]-perylene, and indeno[1,2,3- cd]pyrene) (Slooff et al., 1989). In Canada, the maximum release of PAH into water and soil from creosote-treated wood products was estimated to be 2000 t/year, on the basis of the PAH content of creosote, the volume of treated wood, the retention rates of the substances for different wood species, and an estimated 20% loss of PAH during the time the wood was in service, i.e. 40 years for pilings and 50 years for railroad ties (Environment Canada, 1994). Table 12. Polycyclic aromatic hydrocarbon concentrations (µg/litre) in wastewater effluents from coal coking plants Compound [1] [2] [3]a [4] [5] Acenapthene NR NR NR 0.009-2.5 NR Acenaphthylene NR NR NR NR NR Anthracene 0.31 NR NR 0.0-2.0 0.1 Anthanthrene ND NR 0.040/0.600 NR NR Benzo[j+k]fluoranthene NR NR NR NR NR Benz[a]anthracene 2.0 11.1 0.504/4.9 NR NR Benzo[a]fluoranthene 0.8 NR NR NR NR Benzo[a]pyrene NR 3.8 0.622/4.841 4.7-25 NR Benzo[b]fluoranthene NR NR NR NR NR Benzo[a]fluorene 0.81 NR NR NR NR Benzo[c]phenanthrene ND NR 0.042/0.699 NR NR Benzo[e]pyrene NR NR 0.323/2.928 NR NR Benzofluoranthenesb NR 6.9 1.010/8.741 NR NR Benzo[ghi]fluoranthene ND NR 0.042/0.663 NR NR Benzo[ghi]perylene 2.0 NR 0.445/2.835 0-9.0 NR Chrysene NR 7.2 0.732/6.440 1.8-42 NR Dibenz[a,h]anthracene NR NR NR 0.06-3.0 NR Fluoranthene 2.8 11.2 NR 1.3-10 NR Fluorene NR NR NR 0.0-1.0 NR Indeno[1,2,3-cd]pyrene NR NR 0.371/3.051 NR NR 1-Methylphenanthrene ND NR NR NR NR Naphthalene NR NR NR 0-4.1 NR Perylene ND NR 0.117/1.348 NR NR Phenanthrene 0.4 NR NR 0.45-2.3 0.5 Pyrene 4.0 12.9 NR NR 0.38-60 [1] Effluent channel water from one US coking plant (Griest, 1980); [2] Effluent channel water from one US coking plant (Griest at al., 1981); [3] Raw wastewater from two coking plants in western Germany (Grimmer at al., 1981 b); [4] Effluents from two US coking plants downstream of company-owned biological oxidation device (Walters & Luthy, 1984); [5] Final effluent after biological oxidation; no further information (Jockers at al., 1988) When the water samples were filtered through solid sorbents, the results may be underestimates of the actual content of polycyclic aromatic hydrocarbons (see section 2.4.1.4) ND, not detected, limit of detection not given; NR not reported a /, single measurements b Isomers not specified Table 13. Polycyclic aromatic hydrocarbons in effluents after wastewater treatment in petroleum refineries (µg/litre) Compound [1] [2] [3] [4] [5] Acenaphthene NR 4.0 < 0.1-6 NR NR Acenaphthylene NR 1.8 < 0.1-< 1 NR NR Anthracene NR 11 < 0.01-< 2 0.26 NR Benz[a]anthracene NR 0.6 < 0.02-< 1 NR NR Benzo[a]pyrene 0.57 0.1 0.1-2.9 0.11 NR Benzo[b]fluoranthene < 0.1 0.2 < 0.06 NR NR Benzo[c]phenanthrene NR 0.2 NR NR NR Benzo[e]pyrene 0.65 0.3 NR NR NR Benzo[ghi]fluoranthene < 0.4 NR NR NR NR Benzo[ghi]perylene 0.36 NR < 0.2-< 1 NR NR Benzo[j]fluoranthene < 0.2 NR NR NR NR Benzo[k]fluoranthene < 0.2 0.4a < 0.2 NR NR Chrysene < 0.03 1.4b < 0.02-1.4 NR NR Coronene < 0.01 NR NR NR NR Dibenz[a,h]anthracene NR NR < 0.3-< 1 NR NR Fluoranthene < 0.2 16.0 < 0.1-< 10 0.26 NR Fluorene NR 3.4 < 0.1-< 1 1.2 NR Indeno[1,2,3-cd]pyrene < 0.02 NR < 1 NR NR 1-Methylphenanthrene NR 4.2 NR NR NR Naphthalene NR 2.4 < 0.1-< 10 15 0.06-9 Perylene 0.14 NR NR NR NR Phenanthrene NR 111.0 < 0.2-< 0.5 7.1 0.02-1.2 Pyrene 0.07 16.1 < 0.1-7 NR NR Triphenylene < 0.03 NR NR NR NR [1] Final effluent from one US petroleum refinery (Pancirov et al., 1980); [2] Effluent from one Norwegian petroleum refinery after treatment in oil-separation devices, oil traps, and retention ponds (Berglind, 1982); [3] Average results for final effluent from 17 US petroleum refineries (Snider & Manning, 1982); [4] Final effluent from one Australian petroleum refinery (Symons & Crick, 1983); [5] Average results for the final effluent from six petroleum refineries in western Germany (German Society for Mineral-oil and Coal Chemistry, 1984) When water samples were filtered through solid sorbents, the results may be underestimates of the actual PAH content (see section 2.4.1.4). NR, not reported a With benzo[j]fluoranthene b With triphenylene (e) Other sources PAH may be released into the hydrosphere during leaching of stocks of coal by rain. In model leaching experiments, naphthalene, acenaphthene, fluorene, phenanthrene, anthracene, fluoranthene, pyrene, chrysene, benz [a]anthracene, benzo [k]fluoranthene, and benzo [a]pyrene were detected at concentrations in the low microgram per litre range, with a maximum of 100 µg/litre; for example, benzo [a]pyrene was found at 0.6 µg/litre (Stahl et al., 1984; Fendinger et al., 1989). PAH were also found in sludge from US coke processing plants in the following concentrations (average of five samples): naphthalene, 430 mg/kg; phenanthrene, 260 mg/kg; acenaphthene, 78 mg/kg; pyrene, 30 mg/kg; chrysene, 28 mg/kg; benzo [a]pyrene, 3.8 mg/kg; benzo [b]fluoranthene, 3.8 mg/kg; and benzo [ghi]perylene, 0.9 mg/kg (Tucci, 1988). PAH may also leach into drinking-water from coal-tar or asphalt coatings on storage tanks and water distribution pipes. Samples from a five-year-old coal-tar-coated water tank in the USA contained 0.21 µg/litre phenanthrene plus anthracene, 0.081 µg/litre fluoranthene, 0.071 µg/litre pyrene, 0.025 µg/litre naphthalene, and 0.021 µg/litre fluorene (Alben, 1980). Measurements in numerous US drinking-water systems showed that PAH accumulate in the water during transport in these pipes. The total concentration of fluoranthene, benzo [j]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, indeno[1,2,3- cd]-pyrene, and benzo [ghi]perylene after transport was in the low nanogram per litre range (Basu et al., 1987). In 1994, a PAH concentration of 2.7 µg/litre was measured in accordance with the German Directive on drinking-water (6.9 µg/litre measured in accordance with US regulations), which was due to transport through a tar-coated pipe in a central water reservoir; phenanthrene was present at a concentration of 2.8 µg/litre and pyrene at 1.2 µg/litre (State Chemical Analysis Institute, Freiburg, 1995). The release of PAH from this source cannot be estimated from the available data. During offshore oil and gas production, PAH-containing drilling muds are discharged directly into the sea. The PAH concentrations at some oil and gas platforms in the Gulf of Mexico and the North Sea were found to be 1900 µg/litre for naphthalene and < 0.01 µg/litre each for chrysene, benzo [b]fluo-ranthene, and dibenz [a,h]anthracene (van Hattum et al., 1993). The total PAH passing into the oceans from shipping have not been estimated. The worldwide discharge of PAH into the oceans from refineries, marine transportation, and industrial effluents of crude oil was estimated to be about 6 t/year in 1973 and 4.6 t/year in the early 1980s (Suess, 1976), but the basis for these estimates is unknown. 3.2.6.3 Emissions to the geosphere The average PAH concentrations in soil from more than 20 former coking sites in Germany were: naphthalene, 1000 mg/kg; phenanthrene, 500 mg/kg; fluoranthene, 200 mg/kg; pyrene, 200 mg/kg; anthracene, 50 mg/kg; and benzo [a]pyrene, 3-5 mg/kg. During vertical leaching, the compounds are distributed according to their mobility. PAH with high-boiling points and low water solubility are present at the highest concentrations at the surface, and more mobile compounds accumulate in deeper soil layers. Naphthalene is usually leached into groundwater, in which it is relatively soluble (Hoffmann, 1993). The sediment of an effluent channel at one US coking plant contained the following concentrations of PAH (dry weight basis): fluoranthene, 31 mg/kg; pyrene, 23 mg/kg; benzo [b+j+k]fluoranthenes, 23 mg/kg; benzopyrenes, 19 mg/kg; benz [a]anthracene, 15 mg/kg; chrysene plus triphenylene, 15 mg/kg; benzo [ghi]perylene, 7.3 mg/kg; benzo [a]fluorene, 7.2 mg/kg; anthracene, 6.7 mg/kg; perylene, 3.8 mg/kg; phenanthrene, 3.6 mg/kg; benzo [b]fluorene, 3.2 mg/kg; benzo [ghi]fluoranthene, 2.3 mg/kg; anthanthrene, 2.3 mg/kg; benzo [c]phenanthrene, 2.1 mg/kg; and 1-methylphenanthrene, 0.71 mg/kg. In the sediment of an effluent from one US petroleum tank farm, anthracene was detected at 3.4 mg/kg, benz [a]anthracene at 0.13 mg/kg, and benzo [a]pyrene at < 0.049 mg/kg (Griest, 1980). Oily sludge originating from a dissolved air flotation unit of the treatment system of a US petrochemical plant effluent was applied to sandy loam samples seven times during a 920-day active disposal period followed by a 360-day inactive 'closure' period, and the decreases in the concentrations of fluorene, phenanthrene, anthracene, fluoranthene, pyrene, benz [a]anthracene, chrysene, triphenylene, benzo [ghi]fluoranthene, benzo [b]fluoranthene, benzo [j]fluoran-thene, benzo [k]fluoranthene, perylene, benzo [a]pyrene, benzo [e]pyrene, and benzo [ghi]perylene in soil were determined. The initial PAH levels ranged from 0.9 mg/kg benzo [j]fluoranthene to 270 mg/kg phenanthrene (dry weight basis). After 1280 days, the three-ring compounds (fluorene, phenanthrene, anthracene) had almost completely disappeared, with 0.2-6.9% remaining, the four-ring substances (fluoranthene, benz [a]anthracene, chrysene) had been partly degraded, and the five-ring compounds remained at fairly high concentrations (Bossert et al., 1984). PAH may be released into soil from polluted industrial sludges and during commercial wood preservation; however, no estimates of the total PAH input into this compartment were available. 3.2.6.4 Emissions into the biosphere Use of anti-dandruff shampoos containing hard coal-tar may lead to increased body concentrations of PAH, as measured by urinary excretion of the PAH metabolite 1-hydroxypyrene. One shampoo had a total PAH content of 2800 mg/kg, including 290 mg/kg pyrene and 56 mg/kg benzo [a]pyrene (no further specification) (van Schooten et al., 1994). Application of a 2% crude coal-tar solution in petrolatum led to significantly increased PAH levels in the blood of five volunteers (Storer et al., 1984; see also Section 8). Measurements of hard coal-tar-containing shampoos in Germany showed concentrations of 7-61 mg/kg benzo [a]pyrene. In wood-tar-containing shampoos, benzo [a]pyrene was detected at concentrations in the low microgram per kilogram range, but 150 mg benzo [a]pyrene were found in one tar bath (State Chemical Analysis Institute, Freiburg, 1995). 3.2.7 Emissions of PAH due to incomplete combustion PAH not only pre-exist in fossil fuels but more are formed during pyrolysis by a radical mechanism (see Zander, 1980). The domestic activities that may result in significant emissions of PAH emissions are vehicle traffic, tobacco smoking, broiling and smoking of foods, and refuse burning. The industrial activities that result in PAH release are aluminium production with use of Söderberg electrodes, iron and steel production, foundries, tyre production, power plants, incinerators, and stubble burning (Anderson et al., 1986) 3.2.7.1 Industrial point sources (a) Emissions to the atmosphere (i) Power plants fired with coal, oil, and gas fossil fuels PAH emitted into the atmosphere from coal-fired power plants consist mainly (69-92%) of two- and three-ring compounds, i.e. naphthalene and phenanthrene and their mono- and dimethyl derivatives. Naphthalene is by far the major component of PAH fractions (31-35%), although high concentrations of phenanthrene and fluorene are also observed (Bonfanti et al., 1988). Specific emission factors of 0.02 g emitted per kg combusted were measured for benzo [a]pyrene and 0.03 µg/kg for benzo [e]pyrene (Ahland et al., 1985). The concentrations of PAH in stack gases from comparable coal- and oil-fired power plants are shown in Table 14. It is difficult to find a characteristic PAH profile for coal-fired plants. The concentrations were low during undisturbed combustion (Guggenberger et al., 1981; Warman, 1985). Low-molecular-mass PAH are found at higher concentrations than high-molecular-mass compounds in coal combustion effluents (Warman, 1985); the low-molecular-mass PAH phenanthrene, fluoranthene, and pyrene were detected at particularly high concentrations, whereas benzo [a]pyrene was found at a level typical of that in ambient air (Kanij, 1987). The specific emission factor for benzo [a]pyrene was 3.5-230 µg/t burnt coal (Ahland & Mertens, 1980). As the contribution of benzo [a]pyrene to the total release of PAH is small, it was considered not to be a suitable indicator for this source (Guggenberger et al., 1981). In contaminated areas, the PAH concentrations in ambient air may be higher than those in the stack gases, which result from after-burning (Guggenberger et al., 1981). Table 14. Concentrations of polycyclic aromatic hydrocarbons (ng/m3) in stack gases of coal- and oil-fired power plants Compound Fuel [1] [2] [3] [4] [5] [6]a Acenaphthene Coal NR NR NR NR NR ND-24 Anthracene Coal NR 0.5 < 10-1800 0.4-100 2-65 19-120 Anthanthrene Coal NR NR NR NR < 0.2-< 0.6 NR Benz[a]anthracene Coal NR 0.6 < 20-1400 NR 1-40 NR Benzo[a]pyrene Coal < 0.1-0.7b 1.3 0.5-790 0.1-120 0.1-1.9 NR < 0.5c Oil < 0.5-7 NR NR NR NR NR Benzo[b]fluoranthene Coal < 0.1-3b,d 2.0 30/40k NR 0.3-12 NR < 0.1-0.4c,d (1/880e) Oil < 0.1-39a NR NR NR NR NR Benzo[b]fluorene Coal NR NR NR NR < 2-< 6 NR Benzo[c]phenanthrene Coal NR NR 0.2 NR NR NR Benzo[e]pyrene Coal NR ND < 10-810 NR 3-< 18 NR Benzo[ghi]perylene Coal NR NR < 10-1400 NR NR NR Coal < 0.5-3b 1.2 < 10-< 100 3-22 < 2-< 6 NR < 0.5c Oil < 0.5-40 NR NR NR NR NR Benzo[j]fluoranthene Coal NR NR NR NR < 5-< 13 NR Benzo[k]fluoranthene Coal < 0.1-2b 0.9 20 NR 1.7-2.5 NR < 0.1-1.3c Oil < 0.1-29 NR NR NR NR NR Chrysene Coal NR 1.8 < 10-< 600 0.1-28 1-41 ND-56 < 10-310e 3.8g Coronene Coal 1-3b 0.9 < 100 NR NR NR < 2c Oil < 2-36 NR NR NR NR NR Dibenz[a,h]anthracene Coal < 0.5-2b NR < 100 NR NR NR < 0.5c Oil < 0.5-26 NR NR NR NR NR Table 14. (continued) Compound Fuel [1] [2] [3] [4] [5] [6]a Fluoranthene Coal NR 4.1 < 10-22 100 0.5-240 20-720 NR Fluorene Coal NR 1.9 NR NR NR 2-140 Indeno[1,2,3-cd]pyrene Coal NR 1.7 < 10-< 100 NR < 0.1-< 1.4 NR 1-Methylphenanthrene Coal NR NR < 20-90 NR NR NR Naphthalene Coal NR NR NR 10-1800 NR 420-2100 Perylene Coal < 0.1-0.2b NR NR NR NR NR < 0.1c Oil < 0.1-15 ND < 10-< 100 NR < 0.2-0.9 NR Phenanthrene Coal NR 5.2 < 20-33 200 26-640 32-2930 NR Pyrene Coal NR 1.3 9-5800 0.2-2850 5-335 ND-311 Triphenyene Coal NR NR NR NR 20-77 NR [1] Coal- and oil-fired power plants in the former FRG (Guggenberger et al., 1981); [2] One French coal-fired power plant (Masclet at al., 1984); [3] 10 Swedish coal-fired power plants (Warman, 1985); [4] One US coal-fired power plant (Junk at al., 1986); [5] One Dutch coal-fired power plant (Kanij, 1987); [6] One German coal-fired power plant with circulating fluid bed combustion (Wienecke at al., 1992) NR, not reported; ND, not detected, limit of detection not given a Various coal qualities b Hard coal c Brown coal d With benzo[e]pyrene e Isomers not specified f With triphenylene g With benz[a]anthracene The inputs of PAH into the atmosphere from power plants were: about 0.001 t benzo [a]pyrene in western Germany in 1981 (Ahland et al., 1985) and 0.1 t in 1983 (Grimmer, 1983a); about 1 t/year total PAH in the USA; 0.1 t in Norway and 6.6 t in Sweden in 1985 (Bjorseth & Ramdahl, 1985); about 2 t total PAH in the Netherlands in 1988 (Slooff et al., 1989); and about 11 t total PAH in Canada in 1990 (Environment Canada, 1994). These numbers may be subject to uncertainty and should be used only as an indication of the order of magnitude of e.g. the concentration in stack gases that is to be expected from experimental values. Actual information on PAH emissions from oil- and gas-fired power plants was not available. PAH emissions from coal-fired power plants have been claimed to be negligible in Germany due to the installation of appropriate filter systems, despite the vast amount of stack gases produced (Zimmermeyer et al., 1991; Ministers for the Environment, 1992). (ii) Incinerators Numerous PAH are formed under simulated incinerator conditions from plastics such as polystyrene, polyethylene, polyvinyl chloride, and their mixtures (Hawley-Fedder et al., 1984a,b,c, 1987). PAH were detected at the following concentrations in the stack gases from a British municipal incinerator: pyrene, 1.6 µg/m3; benz [a]anthracene plus chrysene, 0.72 µg/m3; fluorene, 0.58 µg/m3; benzo [ghi]perylene, 0.42 µg/m3; benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, 0.32 µg/m3; perylene, 0.18 µg/m3; indeno[1,2,3- cd]pyrene, 0.18 µg/m3; coronene, 0.04 µg/m3; and benzo [a]pyrene plus benzo [e]pyrene, 0.02 µg/m3 (Davies et al., 1976). When PAH were sampled at a height of about 10 m above the ground in the 110-m chimney of an incineration plant in Sweden, no measurable amounts of PAH, at a limit of detection of 10 ng/m3, were found during normal operating conditions or during start-up in the morning; however, inactivity over a weekend resulted in detectable concentrations of individual PAH, covering three orders of magnitude up to around 100 µg/m3 (Colmsjö et al., 1986a). Comparable results were obtained at a pilot incineration plant in Canada (Chiu et al., 1991). Only phenanthrene plus anthracene was found in measurable amounts in the stack gas (limit of detection not stated). The total release of PAH from this plant was estimated to be 80-100 ng/m3. The concentrations of PAH emitted in the stack gases from an Italian municipal solid waste incinerator were: 0.1-1.9 µg/m3 indeno[1,2,3- cd]pyrene, 0.63 µg/m3 acenaphthene, 0.57-2.5 µg/m3 phenanthrene, 0.36-4.4 µg/m3 perylene, 0.35-0.55 µg/m3 benzo [e]pyrene, 0.25-3.6 µg/m3 benz [a]anthracene, 0.23 µg/m3 benzo [k]fluoranthene, 0.22 µg/m3 dibenz [a,h]anthracene, 0.19 µg/m3 benzo [b]fluoranthene, 0.15-0.67 µg/m3 pyrene, 0.15-0.73 µg/m3 acenaphthylene, 0.11-0.23 µg/m3 chrysene, 0.08 µg/m3 anthracene, 0.069 µg/m3 fluorene, 0.068-1.3 µg/m3 fluoranthene, 0.05-1.1 µg/m3 benzo [a]pyrene, and 0.014-0.47 µg/m3 benzo [ghi]perylene, depending on the firing conditions and the composition of the waste (Morselli & Zappoli, 1988). The benzo [a]pyrene concentrations in stack gases from commercial waste incinerators in western Germany were estimated to be 1-6 µg/m3 (Johnke, 1992). Controlled incineration of automobile tyres for thermal and electric energy has been estimated to result in considerable release of PAH into the atmosphere. In laboratory experiments, the following concentrations were found in flue gas at an incineration temperature of 677°C (per kg rubber): 930 mg pyrene, 760 mg fluoranthene, 390 mg phenanthrene, 290 mg anthracene, 220 mg acenaphthylene, 120 mg chrysene, 84 mg benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, 66 mg benz [a]anthracene, 18 mg benzo [e]pyrene, 11 mg benzo [a]pyrene, 3.8 mg perylene, 3.3 mg benzo [ghi]fluoranthene, 2.0 mg dibenz [a,h]anthracene, 1.5 mg benzo [ghi]perylene, 1.2 mg naphthalene, and 0.5 mg indeno[1,2,3- cd]pyrene (Jacobs & Billings, 1985). On the basis of data from Hartung & Koch (1991) on the number of tyres incinerated in western Germany in 1987, the annual emissions from this source can be calculated as follows: 160 t pyrene, 130 t fluoranthene, 70 t phenanthrene, 50 t anthracene, 40 t acenaphthylene, 20 t chrysene, 14 t benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]-fluoranthene, 10 t benz [a]anthracene, 3 t benzo [e]pyrene, 2 t benzo [a]pyrene, 0.5 t benzo [ghi]fluoranthene, 0.3 t dibenz [a,h]anthracene, 0.3 t benzo [ghi]-perylene, 0.2 t naphthalene, and 0.1 t indeno[1,2,3- cd]pyrene. The total PAH levels in stack gases from incinerators in different countries were: Italy, 0.0075-0.21 mg/m3; Japan, 0.002-0.04 mg/m3; Sweden, 0.001 mg/m3; and Canada, 0.00002-0.02 mg/m3 (WHO, 1988). The results for traditional incinerators could not be compared with those for plants with additional abatement techniques on the basis of the available data. The total PAH emissions to the atmosphere resulting from incineration of refuse were about 0.001 t benzo [a]pyrene in western Germany in 1989 (Ministers for the Environment, 1992) and about 0.0003 t in 1991 (Johnke, 1992), about 50 t total PAH in the USA, 0.3 t in Norway and 2.2 t in Sweden in 1985 (Bjorseth & Ramdahl, 1985); and about 2.4 t total PAH in Canada in 1990 (Environment Canada, 1994). In Germany, the contribution of stack gases from commercial incinerators is estimated to be < 4% of the total stack gas volume from combustion processes. One of the main confounders of and contributors to stack gases from combustion is motor vehicle traffic (Johnke, 1992), indicating that PAH released from incinerators are probably of minor importance. (iii) Aluminium production The production of coal anodes, used in the electrolytic production of aluminium, from pitch and petroleum coke may still be an important source of PAH, but confirmatory data are not available. Estimates of PAH released during the production of aluminium in the Netherlands in 1988 ranged from about 0.3 t benzo [ghi]perylene to 24 t naphthalene (Slooff et al., 1989). The estimated total airborne PAH released in 1985 was about 1000 t in the USA, 160 t in Norway, and 35 t in Sweden (Bjorseth & Ramdahl, 1985). In 1990, the input of total PAH from this source into the atmosphere in Canada was 930 t (Environment Canada, 1994). In horizontal and vertical Söderberg aluminium production processes in Sweden, the emission factors per tonne of aluminium were 0.11 kg benzo [a]pyrene and 4.4 kg total PAH for the horizontal process and 0.01 kg benzo [a]pyrene and 0.7 kg total PAH for the vertical process (Alfheim & Wikström, 1984). In a Norwegian vertical Söderberg aluminium production plant, the emission factors were 0.005-0.015 kg/t aluminium for benzo [a]pyrene and 0.3-0.5 kg/t for total PAH (European Aluminium Association, 1990). (iv) Iron and steel production The total emissions of PAH resulting from iron and steel production with carbon electrodes containing tar and pitch in Norway was estimated to be about 34 t in 1985 (Bjorseth & Ramdahl (1985), but the database for this estimate is limited. The release of total PAH from metallurgical processes in Canada where similar electrodes were used, including ferro-alloy smelters but excluding aluminium production, was estimated to be 19 t in 1990 (Environment Canada, 1994). (v) Foundries PAH are formed during casting by thermal decomposition of carbonaceous ingredients in foundry moulding sand, and they partly vaporize under the extremely hot reducing conditions at the mould-metal interface. Thereafter, the compounds are adsorbed onto soot, fume, or sand particles. Organic binders, coal powder, and other carbonaceous additives are the predominant sources of PAH in iron and steel foundries (IARC, 1984b). In pyrolysis experiments with green-sand additives, the highest PAH levels were found in coal-tar pitch, with values per kilogram of additive of 3100 mg benzo [a]pyrene, 3000 mg benzo [b+j+k]fluoranthenes, 3000 mg pyrene, and 2900 mg fluoranthene; the lowest levels were found in vegetable product additives, such as maize starch: 26 mg pyrene, 16 mg fluoranthene, 3 mg benzo [b+j+k]fluoranthenes, and 2 mg benzo [a]pyrene (Novelli & Rinaldi, 1979). Less than 0.002 mg/kg benzo [a]pyrene was found in foundry moulding sand when petrol resin, polystyrol, or polyethylene was used as the carrier and 7.5 mg/kg when hard coal was used as the carrier. The PAH content was directly correlated with the amount of hydrocarbon carrier in the sand (Schimberg et al., 1981). The following levels of PAH were found in the stack gases of one French automobile foundry: fluoranthene, 980 ng/m3; benz [a]anthracene, 830 ng/m3; benzo [a]pyrene, 570 ng/m3; benzo [b]fluoranthene, 460 ng/m3; indeno[1,2,3- cd]pyrene, 370 ng/m3; anthracene, 250 ng/m3; benzo [k]fluoranthene, 220 ng/m3; perylene, 160 ng/m3; benzo [ghi]perylene, 130 ng/m3; chrysene, 110 ng/m3; coronene, 28 ng/m3; and pyrene, 15 ng/m3. No further information was given about the sampling site (Masclet et al., 1984). The total emission of PAH into the atmosphere from iron foundries in the Netherlands was estimated to be about 1.3 t in 1988 (Slooff et al., 1989). (vi) Other industrial sources The estimated release of 10 PAH into the atmosphere in the Netherlands in 1988 was about 1.3 t from sinter processes and 0.2 t/year from phosphorus production (Slooff et al., 1989). (b) Emissions to the hydrosphere (i) Aluminium production PAH levels in wastewater from aluminium production in Norwegian plants are shown in Table 15. At the beginning of the 1970s, the release of anthracene and phenanthrene into the aqueous environment from aluminium production in western Europe was estimated to be 180 t/year (Palmork et al., 1973). About 0.6 t/year are released into water by the aluminium producing industry in the Netherlands (Slooff et al., 1989). (ii) Other industrial sources No recent data were available on PAH emissions into the aqueous environment from coal- or oil-fired power plants. PAH were found in the final effluent from a British municipal incinerator at concentrations ranging from < 0.01 µg/litre each for coronene and indeno[1,2,3- cd]pyrene to 0.62 µg/litre fluoranthene. The calculated daily output of single compounds was in the low milligram range, with a maximum of 16 mg/d. Actual data were not available (Davies et al., 1976). Numerous PAH were detected in the final effluent from a Norwegian ferro-alloy smelter in which the wastewater from gas scrubbers was treated by chemical flocculation. The concentrations were 50 µg/litre phenanthrene, 45 µg/litre pyrene, 40 µg/litre fluoranthene, 39 µg/litre acenaphthylene, 27 µg/litre fluorene, 17 µg/litre acenaphthene, 13 µg/litre chrysene plus triphenylene, 11 µg/litre anthracene, 10 µg/litre naphthalene, 10 µg/litre benz [a]anthracene, 9 µg/litre benzo [b]fluoranthene, 6 µg/litre benzo [j]fluoranthene plus benzo [k]fluoranthene, 6 µg/litre benzo [e]pyrene, 6 µg/litre benzo [a]pyrene, 3 µg/litre benzo [c]phenanthrene, 3 µg/litre indeno[1,2,3- cd]pyrene, 3 µg/litre benzo [ghi]perylene, 2 µg/litre benzo [a]fluorene, 2 µg/litre benzo [b]fluorene, 2 µg/litre perylene, and 1 µg/litre dibenz [a,h]-anthracene. The PAH contents of wastewater from gas washers in one Norwegian steel production plant were of the same order of magnitude (Berglind, 1982). Table 15. Polycyclic aromatic hydrocarbon concentrations [µg/litre] in wastewater from aluminium production in Norway Compound [1] [2] [3] Acenaphthene NR NR 5 Acenaphthylene NR NR 1 Anthracene 1.1-2.8 0.9 10 Anthenthrene < 1-3.2 NR NR Benzo[b+k]fluoranthenes 6.8-38.1 NR NR Benzo[j+k]fluoranthenes NR 10.5 5 Benz[a]anthracene 2.5-5.6 14.6 11 Benzo[a]fluorene 1.5-3.4 8.2 13 Benzo[a]pyrene 1.3-7.4 13.5 4 Benzo[b]fluoranthene NR 21.2 9 Benzo[b]fluorene 1.3-3.0 7.2 2 Benzo[c]phenanthrene NR NR 3 Benzo[e]pyrene 2.6-16.4 17.0 5 Benzo[ghi]perylene NR 8.3 2 Chrysene and triphenylene 5.8-16.0 27.3 17 Coronene < 1-2.0 NR NR Dibenz[a,h]anthracene NR NR 1 Fluoranthene 12.4-20.8 7.5 124 Fluorene NR NR 3 Indeno[1,2,3-cd]pyrene NR 8.1 2 1-Methyphenanthrene NR 0.4 NR Naphthalene NR NR 1 Perylene NR 3.2 1 Phenanthrene 14.0-23.1 1.8 34 Pyrene 5.6-15.3 6.4 76 [1] Two samples of wastewater with two runs each from one aluminium production plant (Kadar at al., 1980); [2] Wastewater from one aluminiurn production plant; no further information (Olufsen, 1980); [3] Effluent from gas washers from one aluminium smelter (Berglind, 1982) When the water samples were filtered through solid sorbents, the results may be underestimates of the actual content (see section 2.4.1.4). NR, not reported The release of 10 PAH into water from different industries in the Netherlands was estimated to be 4 t/year (Slooff et al., 1989). (c) Emissions to the geosphere The levels of PAH in ash samples from various incinerators are shown in Table 16. The values given by Eiceman et al. (1979) were based on the gas chromatographic responses of pyrene and benzo [a]pyrene. The concentrations of PAH in ashes from coal-fired power plants were of the same magnitude as the background levels of these compounds in soil, but fly ash from municipal waste incinerators may contain significantly higher levels (Guerin, 1977; Kanij, 1987). The total PAH content of filter residues in incinerators was about 0.20-0.5 µg/g. The compounds are assumed to be tightly bound to particle surfaces and not mobile in an aqueous environment in the absence of organic solvents (WHO, 1988). In a comparison of 26 incineration plants, combustion conditions were shown to have a marked influence on PAH release (Wild et al., 1992). The material dredged from harbour areas may have a significant PAH content (see also sections 5.3.3 and 5.3.4). The annual load of naphthalene, anthracene, phenanthrene, fluoranthene, benz [a]anthracene, chrysene, benzo [k]-fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, and indeno[1,2,3- cd]pyrene in material dredged from Rotterdam harbour was about 12 t (year not given). The main PAH were fluoranthene and benz [a]anthracene (Slooff et al., 1989). 3.2.7.2 Other diffuse sources (a) Atmosphere (i) Mobile sources PAH are released into the atmosphere by motor vehicle traffic. The profile of the PAH released and the quantity of PAH in the exhaust are fairly similar, independently of the type of engine and the PAH content of the fuel, indicating that the emitted compounds are formed predominantly during combustion (Meyer & Grimmer, 1974; Janssen, 1980; Stenberg, 1985; Williams et al., 1989). PAH accumulate in used engine oil, but the importance of the PAH content of engine oil on emissions is still under discussion. Janssen (1980), Pischinger & Lepperhoff (1980), and Stenberg (1985) assumed that the PAH content of the oil played only a minor role, but Williams et al. (1989) showed in tests with diesel fuel that it may contribute considerably to the release of particulate PAH. There is also doubt about whether PAH emissions are indepen-dent of the aromaticity of the fuel. Janssen (1980) stated that release of PAH into the atmosphere is not increased if the aromaticity does not exceed a concentration of 50% volume (see also Schuetzle & Frazier, 1986). According to Stenberg (1985), the release of PAH by automobile traffic is dependent on the: Table 16. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in ash samples from coal-fired power plants and municipal waste and sewage sludge incinerators Compound Coal-fired Municipal waste incinerators Sewage power plants sludge incinerators Netherlands USA Canada Japan Netherlands Canada UK Italy (UK) [1] [2] [3] [3] [3] [4] [5] (mean) [6] [5] Acenaphthene + NR NR NR NR NR NR 1-258(7.8) 289-1022i NR fluoranthene Acenaphthylene NR NR NR NR NR 3.35 NR 5-1394 NR Anthracene < 0.14-0.5 NR 10/500 10/10 200 NR 1-62(2.3) 42-651 NR Anthanthrene < 0.24-< 0.5 NR NR NR NR NR NR NR Benz[a]anthracene < 0.6-< 1.2 NR NR NR NR NR 1-1646a(12) 280-1278 3a Benzo[a]fluoranthene NR 36.8 NR NR NR NR NR NR Benzo[a]pyrene < 0.29-< 1.8 NR ND/400 ND/ND ND NR 1-596(8.2) 1014-3470 3 Benzo[b]fluoranthene < 0.6-< 0.29 NR NR NR NR NR 1-873(5.7) 1818 6 Benzo[b]fluorene < 2.0-< 4 11.8 NR NR NR NR NR NR Benzo[e]pyrene < 2.9-< 6 NR NR NR NR NR NR 458-1786 NR Benzo[ghi]perylene < 1.6-1.7 NR NR NR NR NR 10-9507 (62.3) 700-2377 135 Benzo[j]fluoranthene < 4.5-< 9 NR ND/400b ND/NDb NDb NR NR NR Benzo[k]fluoranthene < 0.15-< 2.8 NR NR NR NR NR 1-276(1.5) 1535 NR Chrysene < 1.5-< 3 NR NR NR NR NR NR 570-1973 NR Coronene NR NR NR NR NR NR 3-238 (31.3) 36 Dibenz[a,h]anthracene < 4.2-< 8.2 NR NR NR NR NR 1-167(5.2) 57/69 1 Fluoranthene 1.1-5.2 < 13.4 2/500 3/ND 20 2.14-43.2 1-765 (8.6) 1684-10 890 1 Fluorene NR NR ND/10 ND/ND 60 2.57/4.41 NR 45-522 NR Indeno[1,2,3-cd]pyrene < 0.82-< 1.6 NR NR NR NR NR NR 478-1343 NR Naphthalene NR 8.3 NR NR NR NR 4/15(0.2) NR Perylene < 0.16-< 0.3 NR NR NR NR NR NR 259 NR Phenanthrene 4.0-43 17.6 NR NR NR 8.76-154c 2-5402 (36.5) 1616-7823 6 Pyrene 0.72-2.9 < 19.0 1/500 1/ND 10 2.47-19.6 1-3407 (45.3) 1863-8799 10 Triphenylene < 2.5-< 5.0 NR NR NR NR 12.7a NR NR Table 16 (continued) [1] Pulverized coal ash (Kanij, 1987); [2] Fly ash (Guerin, 1977); [3] Fly ash (Eiceman at al., 1979); [4] Fly ash (Chiu at al., 1991); [5] Fly ash 26 incinerators with different firing techniques (Wild et al., 1992); [6] Fly ash from electrostatic precipitator and scrubber (Morselli & Zappoli, 1988) NR, not reported; ND, not detected; /, single measurements a With chrysene b Isomers not specified c With anthracene i Only acenaphthene - aromaticity of the fuel; - starting temperature: Starting at -10°C results in threefold higher PAH emissions than a standardized cold start (+ 23°C); the emission factors measured by Larssen (1985) were significantly higher in winter than in summer. - ambient temperature: Low ambient temperatures (5-7°C) increase PAH emissions from petrol-fuelled vehicles by five to 10 times, depending on the engine used. - test conditions: Three standardized test cycles are in general use: a test developed by the Economic Commission for Europe of the United Nations (ECE) in Europe; the Federal Test Procedure (FTP) in the USA; and the Japanese test cycle in Japan. Emissions at cold start may be lower and those at hot start slightly higher in the FTP than in the ECE test, but overall agreement between the tests is good. - air:fuel ratio (l): Small variations around l = 1, representing stoichiometric levels of fuel and air, do not affect PAH emissions significantly; richer mixtures lead to increasing PAH emissions, and bad ignition at l = 0.8 causes a sharp increase in PAH emissions. - type of fuel: Emissions of the sum of phenanthrene, fluoranthene, pyrene, benzo [ghi]fluoranthene, cyclopenta [cd]pyrene, benz [a]-anthracene, chrysene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]-pyrene, indeno[1,2,3- cd]pyrene, benzo [ghi]-perylene, and coronene decreased in the FTP cycle as follows: diesel (total PAH; 960 µg/km) > petrol (170 µg/km) > petrol containing methanol or ethanol (43-110 µg/km) > methanol = liquefied petro-leum gas = catalyst-equipped petrol-fuelled vehicles (6-9 µg/km) (Stenberg, 1985). In comparable measurements, similar results were obtained but with a much lower average emission rate for diesel-fuelled vehicles: 186 µg/km for total PAH, including fluoranthene, pyrene, benz [a]anthracene, chrysene, benzo [b]-fluoranthene, benzo [e]-pyrene, benzo [a]pyrene, perylene, indeno[1,2,3- cd]pyrene, benzo [ghi]-perylene, and coronene. It was not stated whether the difference in the emission rates was due to the numbers of PAH chosen for analysis (Lies et al., 1986). PAH emissions in the exhaust from spark-ignition automobile engines can be reduced by operation with lean air:fuel ratios, smaller quenching distances in the combustion chamber, and increased cylinder wall temperatures in the engine (Pischinger & Lepperhoff, 1980; Lepperhoff, 1981). Diesel-fuelled engines with low emissions of total unburnt gaseous hydrocarbons have low rates of PAH emission. Control can therefore be achieved by using conventional techniques for reducing unburnt gaseous hydrocarbons (Williams et al., 1989). Fluoranthene and pyrene constitute 70-80% of total PAH emissions from vehicles (Lies et al., 1986; Volkswagen AG, 1989; see also Table 17), whereas the emissions from one diesel-fuelled truck consisted mainly of naphthalene and acenaphthene (Nelson, 1989). Although cyclopenta [cd]pyrene is emitted at a high rate from petrol-fuelled engines, its concentration in diesel exhaust is just above the limit of detection, probably because the oxidizing conditions in diesel-fuelled engines decompose this relatively reactive compound (Lies et al., 1986). The amounts of PAH released from vehicles with three-way catalytic converters are much lower than those from vehicles without catalysts (Table 18). The total amount of PAH was increased by a factor of about 40 between new and used catalytic converters (Hagemann et al., 1982). PAH emissions from diesel-fuelled vehicles can be reduced by > 90% by a combination of a catalytic converter and a particulate trap, as shown by experiments with a heavy-duty diesel-fuelled truck (Westerholm et al., 1989). Westerholm et al. (1991) found benz [a]anthracene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [e]-pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, benzo [ghi]perylene, fluoranthene, pyrene, anthracene, and coronene in much lower amounts than other investigators, while some other PAH that were not measured by other investigators, especially phenanthrene and 1-methylphenanthrene, were detected at quite high concentrations. These differences are possibly due to the driving cycle used. Measurements made on particulate matter in the exhausts of light- and heavy-duty diesel-fuelled vehicles with different fuel qualities showed concentrations of 1 mg/kg each of benz [a]anthracene, benzo [b]fluoranthene plus benzo [j]fluoranthene, benzo [a]pyrene plus benzo [e]pyrene, and benzo [ghi]perylene and 290 mg/kg pyrene. The results were strongly dependent on the driving cycle and individual engine conditions (CONCAWE, 1992). The PAH concentrations measured in the exhaust gases of different vehicles are shown in Table 19. The differences in PAH emissions from petrol- and diesel-fuelled vehicles are still under discussion. When the data of Behn et al. (1985) are compared with those of Klingenberg et al. (1992), diesel-fuelled vehicles emitted larger amounts of PAH than petrol-fuelled vehicles. Benzo [a]pyrene was emitted at a rate of 6 µg/km from a petrol-fuelled vehicle without a catalyst and at 5 µg/km from a diesel-fuelled vehicle (Gibson, 1982). When the PAH emissions from 10 petrol- and 20 diesel-engined vehicles were measured under three urban cycles, the mean emission factors (µg/km) for benzo [a]pyrene were 12 with petrol and 0.56 with diesel in a cold, low-speed cycle, 0.50 with petrol and 0.37 with diesel in a hot, low-speed cycle, and 0.37 with petrol and 0.24 with diesel in a hot, free-flow cycle (Combet et al., 1993). Considerably higher emission rates were found from four petrol-fuelled passenger cars without catalysts, 11 with catalysts, and eight diesel-fuelled passenger cars, two of which had oxidation catalysts, on a chassis dynamometer at the USA FTP 75 cycle. The diesel-fuelled vehicles emitted about as much benzo [a]pyrene as the petrol-fuelled vehicles without catalysts (5-25 µg/km), while the petrol-fuelled vehicles with catalysts had emission rates significantly below 2 µg/km. The diesel-fuelled vehicles with oxidation catalysts had emissions of about 5 µg/km (Klingenberg et al., 1992). The following emission factors were given for motorcycles and two-stroke mopeds: 1000 µg/km naphthalene, < 32-650 µg/km phenanthrene, < 11-170 µg/km anthracene, < 5-110 µg/km fluoranthene, < 2-11 µg/km chrysene, < 2-11 µg/km indeno[1,2,3- cd]pyrene, < 1-1200 µg/km benz [a]anthracene, 0-63 µg/km benzo [ghi]perylene, 0-16 µg/km benzo [a]pyrene, and 0-11 µg/km benzo [k]fluoranthene (Slooff et al., 1989). Further PAH emissions may result from the abrasion of asphalt by vehicle traffic, so that PAH in asphalt and bitumens (see section 3.2.1) may contribute considerably to the total PAH emissions due to automobile traffic. The abrasion caused by spiked tyres in winter was estimated to be 20-50 mg/km (Lygren et al., 1984). Another source of PAH from motor vehicle traffic is clutch and break linings, which are subject to considerable thermal stress, sometimes resulting in pyrolytic decomposition of abraded particles. Numerous PAH were found in the abraded dust of brake and clutch linings in one study, but the values show large standard deviations, due, presumably, to the fact that the substances are adsorbed onto asbestos fibres from which they are difficult to separate (Knecht et al., 1987). Total PAH release from clutch and brake linings cannot be estimated from the available data. Rubber vehicle tyres contain highly aromatic oils as softeners. These oils, which can contain up to 20% PAH, are used at concentrations of 15-20% in rubber blends (Duus et al., 1994). In Sweden, it was considered that the input of PAH to the atmosphere from rubber particles was important (National Chemicals Inspectorate, 1994). According to estimates for Belgium, western Germany, and the Netherlands in 1985, the annual PAH input into the atmosphere from vehicle traffic ranges from < 10 t/year for benzo [ghi]fluoranthene, benz [a]anthracene, benzo [k]-fluoranthene, benzo [a]pyrene, and indeno[1,2,3- cd]pyrene, to < 10-20 t/year for anthracene, fluoranthene, and chrysene, to 10-70 t/year for phenanthrene, to about 100-1000 t/year for naphthalene (Slooff et al., 1989). Values of the same order of magnitude were reported for emissions of naphthalene in 1987 (Society of German Chemists, 1989) and benzo [a]pyrene in 1989 in western Germany (Ministers for the Environment, 1992) and for total PAH in 1985 in Norway and Sweden (Bjorseth & Ramdahl, 1985). The total annual PAH input from vehicle traffic in the USA in 1985 was about 2200 t/year (Bjorseth & Ramdahl, 1985). In Canada, the total PAH input was estimated to be about 200 t in 1990; 155 t were assumed to be due to diesel-fuelled and 45 t to petrol-fuelled vehicles (Environment Canada, 1994). Table 17. Polycyclic aromatic hydrocarbon emission factors (µg/km) for petrol-fuelled vehicles Compound [1] [2] [3] [4] [5] [6] [7] Anthracene NR 0.7/0.7a NR 2/99b NR 21-42 0.6 37/1988c Anthanthrene NR 0.2/1.3 NR NR NR NR NR Benzo[b+j+k]fluoranthene NR NR NR NR NR NR 7.6 Benzo[b+k]fluoranthene NR 3.9/7.0 NR NR 0.23-0.54/2.55-9.20 NR NR Benz[a]anthracene NR 5.7/5.9 3.5-9.0 NR 0.06-0.35/2.5-8.0 5-16 5.1 Benzo[a]pyrene NR 1.9/4.51 1.5-14.5 0.06-2/1-12b 0.06-0.62/1.30-10.4 2-11 3.7 Benzo[e]pyrene NR 2.6/6.2 NR 0.2/2-14b 0.08-0.54/2.54-9.20 NR 5.1 Benzo[ghi]fluoranthene NR 5.6/12 NR NR NR NR 8.8 Benzo[ghi]perylene NR 5.9/13 NR NR 0.19-0.75/1.45-17.5 5-21 18.9 Benzo[j]fluoranthene NR 1.1/0.9 NR NR NR NR NR Benzo[k]fluoranthene NR NR NR NR NR 0-5 NR Chrysene NR 6.7/8.7 NR NR 0.12-0.73/2.78-23.1 11-42 7.7 Coronene NR 6.5/12 1.5-20.0 NR NR NR 29.5 Cyclopenta[cd]pyrene NR 2.9/12 NR NR NR NR 16.5 Fluoranthene NR 14/20 NR 3/139-211b 2.7/43.3d 11-158 10.4 ND/186-280c Indeno[1,2,3-cd]pyrene NR 1.7/3.6 NR NR 0.06-0.43/0.83-6.67 5-21 4.2 Naphthalene 8100-8600a NR NR NR NR 2300f NR 210-2651 Perylene NR 0.3/0.5 NR NR 0.01-0.06/0.25-1.82 NR NR Phenanthrene NR 2.6/2.9 NR NR NR 84-210 1.8 Pyrene NR 28/31 43-184 4-16/12-268b 2.9/43.0b NR 19.2 ND/124-360c Table 17 (continued) NR, not reported; ND not detected (detection limit not stated); /, single measurements a Two driving distances b Only particulate phase considered c Only gaseous phase considered d Average e Depending on analytical conditions f With converter [1] From measurements in tunnel with converters (Hampton at al., 1983); [2] One vehicle without converter (Alsberg et al., 1985); [3] Various tests conducted mainly in the 1970s, some unstandardized, different numbers of vehicles, without converters (Stenberg, 1985); [4] FTP cycle only, number of vehicles not given; year of manufacture 1980-85 = petrol-engine vehicles with converter; 1973-81 = petrol-engine vehicles without converter (Schuetzle & Frazier, 1986); [5] Various standardized test procedures; four petrol-engine vehicles without, seven with three-way-converter for each test, all with four or five cylinders (Volkswagen AG, 1988); [6] No information about test cycle or number of cars tested; city roads, motorways and other roads tested; no distinction between vehicles with and without converter, unless otherwise stated (Slooff et al., 1988, 1989); [7] One petrol-engine vehicle without converter in USFTP test cycle (Strandell at al., 1994) Table 18. Polycyclic aromatic hydrocarbon emission factors (µg/km) for diesel-fuelled vehicles Compound [1] [2] [3] [4] [5] [6] [7] [8] Acenaphthene NR NR NR NR NR NR 41-128 NR Anthracene 17/63 65-273a 1.2/3.0 NR 21-73b 3.3 2.9-26 4.6 1305-5568c Benzo[b+j+k]fluoranthene NR NR NR NR NR NR 1.7-12d 5.0 Benzo[b+k]fluoranthene 2.6/47 NR 3.9/6.1 5.57-14.96 NR 0.29 NR NR Benz[a]anthracene 8/43a NR 4.0/7.0 2.73-3.91 11-21b 0.47 0.7-9.6 2.0 Benzo[a]fluorene NR NR NR NR NR 2.4 NR NR Benzo[a]pyrene < 1/20 0.6-34a 1.6/2.2 2.09-7.23 1-5 < 0.06 0.5-3.2 1.5 Benzo[e]pyrene 3/38 2-40a 2.5/4.1 2.40-52.8 NR 0.15 1.1-9.9 4.0 Benzo[ghi]fluoranthene NR NR 4.0/12 NR NR 1.5 NR 10.6 Benzo[ghi]perylene < 1/18 NR 1.9/3.1 2.84-26.3 9e < 0.13 0.5-3.7 2.0 Chrysene 14/67 NR 11/25 4.7-21.1 16-42b 2.8f 3.5-28 3.7 Coronene NR NR 0.3/20.7 NR NR < 0.01 NR NR Cyclopenta[cd]pyrene NR NR 3.6/3.9 NR NR 0.18 NR 4.0 Fluoranthene 58/200 139-580a 13/38 70g 21-105b 17 14-34 43.7 186-771c Fluorene NR NR NR NR NR NR 38-228 NR Indeno[1,2,3-cd]pyrene NR NR 1.5/2.3 0.89-7.52 9e < 0.04 NR 1.2 1-Methylphenanthrene NR NR NR NR NR 41 NR NR Naphthalene NR NR NR NR 2100-6302b NR 1030-1805 NR Perylene < 1/2 NR NR 0.23-1 NR < 0.01 NR NR Phenanthrene 295/524 NR 4.6/25 NR NR 2.9 79-308 54.8 Pyrene < 0-9/22 24-734a 20/104 66.9g NR 11 9-30 35.4 702-982c Table 18 (continued) NR, not reported; /, single measurements; [1] ECE test; two passenger cars with < 50 000 and > 100 000 km odometer readings (Scheepers & Bos, 1992); [2] FTP cycle; number of vehicles not given; year of manufacture, 1980-85 (Schuetzle & Frazier, 1986); [3] Chassis dynamometer; one heavy-duty vehicle (Westerholm et al., 1986); [4] Various standardized testing procedures; seven vehicles with four or five cylinders for each test (Volkswagen AG, 1988); [5] No information on test cycle or number of cars tested; three traffic situations (Slooff at al., 1989); [6] Bus cycle simulating public transport (duration 29 min; driving distance, 11.0 km; average speed, 22.9 km/h); one heavy-duty truck; measurement of particle phase (Westerholm at al., 1991); [7] Bus cycle (duration, about 10 min after warm-up, each ramp consisting of 10 s acceleration, 10 s constant speed of 12 km/h, 4.5 s deceleration, 7 s idling); three trucks and two buses without particle trap, two buses with particle trap (Lowenthal et al., 1994); [8] US FTP cycle; one passenger car (Strandell at al., 1994) a Particle phase b Automobiles and trucks c Gas phase d Isomers not specified e Trucks f With triphenylene g Average Table 19. Polycyclic aromatic hydrocarbon concentrations (µg/m3) in the exhaust gases of different vehicles Compound [1] [2] [3] Acenaphthene NR NR < 0.02-0.81 Acenaphthylene NR NR < 0.02-4.16 Anthracene NR NR < 0.02-6.45 Anthanthrene 0.02-0.07 0.11-0.12 NR Benz[a]anthracene 1.91-2.24 3.53-4.64 NR Benzo[a]pyrene 0.46-0.76 2.03-2.33 < 0.02-4.97 Benzo[b]fluoranthene 1.53-2.04a 7.37-8.58a 0.06-6.63 Benzo[b]fluorene NR NR 0.11-12.7 Benzo[e]pyrene 1,07-1.24 2.46-2.90 0.09-6.16 Benzo[ghi]fluoranthene 0.46-0.59 4.81-7.19 NR Benzo[ghi]perylene 0.76-1.04 3.42-4.41 0.22-1.81 Benzo[k]fluoranthene NR NR < 0.02-2.68 Chrysene 2.37-2.97b 7.37-8.58b 0.07-25.48 Coronene 0.26-0.30 1.82-2.32 < 0.02-1.80 Cyclopenta[cd]pyrene 1.86/2.26 5.80-6.09 NR Dibenz[a,h]anthracene 0.04-0.07 0.32-0.35 < 0.02-0.44 Fluoranthene 11.83-13.09 20.90-25.30 0.16-35.94 Fluorene NR NR 0.06-2.16 Indeno[1,2,3-cd]pyrene 0.30-0.41 2.89-4.06 < 0.02-0.80 Perylene 0.10-0.26 0.21-0.33 0.13-5.55 Phenanthrene NR NR < 0.02-4.16 Pyrene 6.86-8.96 12.20-15.20 0.06-21.31 NR, not reported; /, single measurements; [1] One vehicle with spark-ignition engine on chassis dynamometer at 75% of maximum engine performance (velocity, about 50 km/h) with varying test periods (Behn et al., 1985); [2] One turbo-charged diesel-fuelled vehicle on chassis dynamometer at 75% of maximum engine perfornance (velocity, about 50 km/h) and a test period of 0.5 h; three tests for each component (Behn at al., 1985); [3] Two diesel-fuelled truck engines at different engine speeds (Moriske at al., 1987) a With benzo[k]fluoranthene b With triphenylene Measurements of PAH concentrations in a Belgian highway tunnel in 1991 were used to calculate emission factors of 2 µg/km for indeno[1,2,3- cd]pyrene and coronene and 32 µg/km for benzo [ghi]perylene. The corresponding annual PAH emissions in Belgium were estimated to be 0.11 t/year for perylene and anthanthrene and 1.3 t/year for benzo [ghi]perylene; the combined release of pyrene, benz [a]anthracene, chrysene, benzo [b]fluoranthene, benzo [j]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [e]pyrene, perylene, anthanthrene, benzo [ghi]perylene, indeno[1,2,3- cd]pyrene, dibenzo [a,c]anthracene, dibenzo [a,h]anthracene, and coronene was 8.3 t/year (De Fré et al., 1994). The importance of PAH released by aircraft is also under discussion. While Bjorseth & Ramdahl (1985) classified the maximum emission in Norway and Sweden in 1985 of 0.1 t/year as small, Slooff et al. (1989) estimated that the release of naphthalene, anthracene, phenanthrene, fluoranthene, benz [a]-anthracene, chrysene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]-perylene, and indeno[1,2,3- cd]pyrene was 51 t/year in 1985. The following concentration ranges were measured in the exhaust gases from two US by-pass turbine engines at various power settings: naphthalene, 0.77-4.7 µg/m3; phenanthrene, 0.46-1.3 µg/m3; pyrene, 0.15-0.61 µg/m3; fluoranthene, 0.13-0.51 µg/m3; acenaphthene, 0.03-0.21 µg/m3; anthracene, 0.029-0.12 µg/m3; benzofluoranthenes (unspecified), 0.028-0.096 µg/m3 (isomers not specified); chrysene, 0.026-0.064 µg/m3; benzo [a]pyrene, 0.021-0.073 µg/m3; benz [a]anthracene, 0.019-0.16 µg/m3; acenaphthylene, 0.017-0.31 µg/m3; benzo [e]pyrene, 0.017-0.057 µg/m3; dibenz [a,h]anthracene, 0.011-0.064 µg/m3; indeno[1,2,3- cd]pyrene, 0.011-0.054 µg/m3; and benzo [ghi]perylene, 0.011-0.045 µg/m3. Cyclopenta [cd]pyrene was not detected (limit of detection not stated) (Spicer et al., 1992). (ii) Domestic residential heating The main PAH released by domestic slow-combustion furnaces and hard-coal and brown-coal coal stoves were fluoranthene, pyrene, and chrysene, which comprised 70-80% of the total PAH in model experiments (Ahland & Mertens, 1980). The specific emission factors for various fuels used in residential heating are shown in Table 20 for coal stoves and Table 21 for wood stoves (Bjorseth & Ramdahl, 1985). Few data are available on the release of PAH from oil stoves. Benzo [a]pyrene was detected at a concentration of < 0.05 µg/kg in one burner-boiler combination (Meyer et al., 1980), and 0.006 and 4 µg/kg benzo [a]pyrene and 0.02 and 15 µg/kg benzo [e]pyrene were found during testing of atomizer and vaporizer oil heating techniques, respectively (Ahland et al., 1985). PAH emissions from residential oil heating seem to be about one order of magnitude lower than those from coal stoves. Table 20. Specific polycyclic aromatic hydrocarbon emission factors (mg/kg) for residential coal stoves Compound [1] [2] [3] [4] [5] [6] Acenaphthene NR NR NR NR 65 NR Acenaphthylene NR NR NR 0.427 NR 7.74 Anthracene 0.0039 NR > 0.595 2.113 26a,b 1.49 Anthanthrene NR NR 0.03-0.08 0.665 NR NR Benz[a]anthracene NR NR 1.04-3.68 7.181 NR 0.61 Benzo[a]fluorene 0.0009 NR NR 1.366 NR NR Benzo[a]pyrene 0.0003 0.014-17.4 0.043-1.3 4.303 5c NR Benzo[b]fluoranthene 0.0002 NR 2.028d 6.102 NR NR Benzo[b]fluorene 0.0007 NR NR 0.874 NR NR Benzo[c]phenanthrene NR NR 1.462e 2.215 4 NR Benzo[e]pyrene 0.0005 0.09-16.2 0.40-1.70 3.994 NR NR Benzofluoranthenesf NR NR 0.90-3.20 NR 6 NR Benzo[ghi]fluoranthene NR NR NR 3.323 NR 0.67 Benzo[ghi]perylene 0.0001 NR 0.30-0.50 3.855 NR NR Benzo[j]fluoranthene NR NR NR 6.782 NR NR Benzo[k]fluoranthene NR NR 0.569 NA NR NR Chrysene 0.0016g NR 2.09 9.571 6h 0.68 1.39-5.60g Coronene NR NR 0.081 1.898 NR NR Cyclopenta[cd]pyrene NR NR 0.145 3.590 NR NR Dibenz[a,h]anthracene NR NR 0.113 NR 5 NR Fluoranthene 0.016 NR 3.30-17.0 28.4 9a 3.47 Fluorene NR NR < 0.065 1.05 44 1.64 Indeno[1,2,3-cd]pyrene 0.0002 0.20-0.60 4.60 NR 4 NR 1-Methylphenanthrene NR NR NR 2.217 NR NR Naphthalene NR NR NR NR 254 35.7 Perylene NR NR 0.20-0.50 1.134 NR NR Phenanthrene 0.046 NR > 3.69 3.984 NR 7.42 Pyrene 0.020 NR 2.98-12.0 26.589 8 3.38 Triphenylene NR NR 0.804 NR NR NR Table 20 (continued) NR, not reported; [1] One new residential stove fuelled with charcoal (Ramdahl et al., 1982); [2] Five coal types: hard-coal and brown-coal briquettes and anthracite (Ahland etal., 1985); [3] Burning of brown coal in different domestic stoves; single values refer to one slow-combustion stove; ranges refer to one slow-combustion stove and one permanent built-in combustion stove at medium load (Grimmer et al., 1983a); [4] One slow combustion stove fueled with hard-coal briquettes (Grimmer at al., 1985); [5] One warm-air furnace and one hot-water boiler fuelled with three different bituminous coals (Hughes & DeAngelis, 1982); [6] Samples from chimney of a detached family house with brown-coal heating in Leipzig, Germany (Engewald et al., 1993) a In particulate phase b With phenanthrene c With benzo[e]pyrene and perylene d With benzo[j]fluoranthene e With benzo[ghi]fluoranthene f Isomers not specified g With triphenylene h With benz[a]anthracene Table 21. Specific polycyclic aromatic hydrocarbon emission factors (mg/kg) for residential wood stoves Compound [1] [2] [3] Anthracene 0.119-1.859 10.4-146.3a 130/3600 Benz[a]anthracene 0.060-0.781 NR 55/740 Benzo[a]fluorene 0.018-0.845 NR NR Benzo[a]pyrene 0.046-0.617 1.1-11.6b NR Benzo[b]fluoranthene 0.108-1.016 NR NR Benzo[b]fluorene 0.011-0.393 NR NR Benzo[c]phenanthrene NR 0.2-2.3 NR Benzo[e]pyrene 0.035-0.350 NR NR Benzofluoranthenesc NR 1.5-15.9 NR Benzo[ghi]fluoranthene NR 0.4-6.7 NR Benzo[ghi]perylene 0.034-0.544 1.1-9.9 NR Chrysene 0.481-0.829d 1.3-37.1e 67/770d Cyclopenta[cd]pyrene 0.04-0.720 0.5-8.9 15/800 Fluoranthene 0.296-3.245 1.2-31.6 190/2300 Indeno[1,2,3-cd]pyrene 0.033-0.415 NR NR 1-Methylphenanthrene 0.141-2.213 NR NR Perylene 0.023-0.274 NR NR Phenanthrene 0.834-8.390 NR 480/7500 Pyrene 0.232-3.822 1.3-24.0 160/2100 NR, not reported; /, single measurements; [1] One small residential wood stove burning spruce and birch; normal and slow burning of each kind of wood (Ramdahl at al., 1982); [2] One zero-clearance fireplace with heat circulation and two airtight wood stoves (baffled and non-baffled) fuelled with red oak and yellow pine with different moisture contents (Peters at al., 1981); [3] One wood-burning stove with and without catalytic combustor (Tan et al., 1992) a With phenanthrene b With benzo[e]pyrene and perylene c Isomers not specified d With triphenylene e With benz[a]anthracene Numerous PAH, including acenaphthene, acenaphthylene, fluorene, phenanthrene, anthracene, 1-methylphenanthrene, fluoranthene, pyrene, benzo [a]fluorene, benzo [ghi]fluoranthene, benzo [c]phenanthrene, cyclo-penta [cd]pyrene, benz [a]anthracene, chrysene plus triphenylene, benzo [b]-fluoranthene, benzo [j]fluoranthene, benzo [j]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, perylene, indeno[1,2,3- cd]pyrene, benzo [ghi]perylene, and anthanthrene, were detected in atmospheric emissions from straw-burning residential stoves, at concentrations mainly in the range of 10 µg/kg to 19 mg/kg (Ramdahl & Muller, 1983). The total PAH content of barbecue briquettes was 2.5-13 µg/g sample. PAH were found in coal and charcoal briquettes but not in lava stones or pressed sawdust briquettes (Kushwaha et al., 1985). The PAH content of soot from domestic open fires was 3-240 mg/kg benzo [a]pyrene, 2-190 mg/kg chrysene, 2-100 mg/kg benz [a]anthracene, 1-77 mg/kg indeno[1,2,3- cd]pyrene, 2-39 mg/kg benzo [e]pyrene, 1-29 mg/kg benzo [ghi]perylene, 1-18 mg/kg coronene, 1-14 mg/kg perylene, and 1-12 mg/kg anthracene (Cretney et al., 1985). The amounts of PAH emitted from coal-fired domestic stoves seem to depend on the quality of the coal used and on the firing technique. Generally, hard coal has a higher energy content than other fuels; thus, less total PAH is emitted per unit of gained energy. The lowest specific emission factors for benzo [a]pyrene and benzo [e]pyrene were found with anthracite and the highest with gas coal and gas-flame coal (Ahland et al., 1985). Model experiments with a slow-combustion stove showed that pitch-bound hard-coal briquettes emitted about 10 times more PAH than bitumen-bound briquettes (Ratajczak et al., 1984). The use of pitch-bound hard-coal briquettes for domestic heating may thus be an important source of PAH in the atmosphere. Use of this fuel was restricted by law to permanent combustion stoves in western Germany in 1974, and since 1976 only bitumen-bound hard-coal briquettes have been produced there (Ratajczak et al., 1984). There is no comparable information for other countries. The levels of airborne PAH from a permanent combustion stove burning brown coal were two to four times higher than those from a slow-combustion stove with a medium load (Grimmer et al., 1983a). About 25-1000 times more PAH are produced from burning wood than from the same mass of charcoal. Since the yield of energy per unit mass is similar, burning wood also produces more PAH per unit of energy. Burning conditions are apparently the major determinant of emission and are much more important than the kind of wood (Ramdahl et al., 1982). In areas where domestic heating is predominantly by wood burning, most airborne PAH may come from this source, especially in winter (e.g.Cooper, 1980). Using benz [a]pyrene as an indicator in extensive measurements in New Jersey, USA, the amounts emitted were found to be more than 10 times higher during the heating period than in seasons when heating is not required. An assessment of combustion source also showed that residential combustion of wood was the decisive factor (Harkov & Greenberg, 1985). About 43-47% of the total PAH released in winter in Fairbanks, Alaska, came from residential wood stoves (Guenther et al., 1988). The PAH concentrations in gases in the chimney stacks of residential coal and oil furnaces are given in Table 22. The highest levels were found during the start of the burning process (Brockhaus & Tomingas, 1976). Measurements with five qualities of coal showed that Extrazit(R), a specially treated coal, emitted smaller quantities of smoke and the lowest PAH levels, and anthracite briquettes emitted the highest levels. Presumably, the high PAH emissions from anthracite briquettes are due to the binding agent, hard coal-tar, which has an especially high PAH content. Furnaces with atomizer oil burners seemed to emit less PAH than those with vaporizers. Measurements in a slow-combustion stove and a tiled stove showed that the highest concentrations of PAH were associated with dust of a particle size of < 2.1 m. As for residential heating with wood, in areas where the predominant form of domestic heating is coal burning, a major proportion of airborne PAH may come from this source, especially in winter (Moriske et al., 1987). Table 22. Polycyclic aromatic hydrocarbon concentrations (µg/m3) in stack gases from residential coal and oil stoves Compound Coal Oil Benz[a]anthracene 0.0157-2630 0.2-0.6 Benzo[a]pyrene 0.0016-1270 0.19-0.67 Benzo[b]fluoranthene 0.0188-3270 0.004-0.68 Benzo[e]pyrene 0.0261-3430 0.4-6.9 Benzo[ghi]perylene 0.010-1670 0.41-3.4 Benzo[k]fluoranthene 0.0044-1250 0.18-0.36 Chrysene 0.0142-2590 0.1-0.5 Coronene 0.003-710 0.15-0.47 Dibenz[a,h]anthracene 0.002-410 NR Fluoranthene 0.0393-6830 0.0134 Perylene 0.0015-2730 0.31-0.8 Pyrene 0.0066-1650 0.1-0.9 From Brockhaus & Tomingas (1976); one permanent combustion stove burning anthracite and brown-coal briquets and vaporizer and atomizer oil burners; NR, not reported Estimates of annual PAH emissions due to residential heating are available for a few countries: - In western Germany, the benzo [a]pyrene emissions were about 10 t in 1981 (Ahland et al., 1985), 7 t in 1985, and 2.5 t in 1988, mainly resulting from coal heating. The reduction in the release of PAH into the atmosphere due to domestic heating resulting from increasing use of oil and gas during the last 30-40 years was estimated to be 90-99% (Zimmermeyer et al., 1991). - In the Netherlands, the estimated release in 1985 was < 1 t/year each for benzo [k]fluoranthene and indeno[1,2,3- cd]pyrene, < 10 t/year each for anthracene, fluoranthene, benz [a]anthracene, chrysene, benzo [a]-pyrene, and benzo[ghi]perylene, and 48-70 t/year each for naphthalene and phenanthrene, mainly resulting from wood heating (Slooff et al., 1989). - The total PAH input, mainly from coal and wood heating, was about 63 t in Norway, 130 t in Sweden, and 720 t in the USA in 1985 (Bjorseth & Ramdahl, 1985). - In Canada in 1990, the total PAH released due to residential heating, mainly wood burning, was about 500 t (Environment Canada, 1994). (iii) Open burning PAH may be released to the atmosphere during forest and agricultural fires, burning of accidentally spilled oil, disposal of road vehicles and especially automobile tyres, open burning of coal refuse and domestic and municipal waste, and open fires. The release of PAH into the atmosphere from the burning of wastes, including road vehicles, in the open is decreasing in industrialized countries due to comprehensive regulations. Laboratory experiments with pine needles gave the following specific PAH emission factors (per kg pine needle): 980-20 000 g pyrene, 690-15 000 g fluoranthene, 580-12 000 g anthracene plus phenanthrene, 540-29 000 g chrysene plus benz [a]anthracene, 420-6200 g benzo [ghi]-perylene, 170-4300 g indeno[1,2,3- cd]pyrene, 140-8800 g benzo [c]-phenanthrene, 130-13 000 g benzofluoranthenes (isomers not specified), 61-800 g benzo [e]-pyrene, 38-3500 g benzo [a]pyrene, and 24-2100 g perylene, depending on the amount of needles, area, and type of fire. Fires moving with the wind and low fuel loading resulted in significantly smaller amounts of PAH than fires moving against the wind and high fuel loading (McMahon & Tsoukalas, 1978). The emission factor for acenaphthene was 230-1000 µg/kg dry straw (Ramdahl & Mller, 1983) and 660 µg/kg dry wood (Alfheim et al., 1984). In model experiments with crude oil spilled on water, numerous PAH were found, including acenaphthene, acenaphthylene, phenanthrene, anthracene, 1-methylphenanthrene, fluoranthene, pyrene, fluorene, benzo [a]fluorene, benzo [b]fluorene, benz [a]anthracene, chrysene plus triphenylene, benzo [b]-fluoranthene, benzo [ghi]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, perylene, indeno[1,2,3- cd]pyrene, benzo [ghi]perylene, and coronene, at concentrations of ¾ 1000 mg/kg individual substance in both the soot and the burn residue (Benner et al., 1990). Even though the open burning of oil spilled on water results in a lower PAH content than in crude oil (see Table 8), this source may be of local importance, e.g. near tanker accidents. Between the early and the mid-1970s, the total release of PAH (including nitrogen-containing analogues and quinone degradation products) into the atmosphere in the USA due to open burning was estimated to be about 4000 t/year (Agency for Toxic Substances and Disease Registry, 1990). The total PAH input from forest and agricultural fires in 1985 was estimated to be 13 t in Norway, 1.3 t in Sweden, and 1000 t in the USA, and that from open fires to be 0.4 t in Norway and 100 t in the USA (Bjorseth & Ramdahl, 1985). The release of all PAH into the atmosphere from the burning of scrap electrical cable in 1988 was about 17 t (Slooff et al., 1989). In Canada in 1990, the total PAH emissions from agricultural burning and open-air fires were estimated to be about 360 t and those from forest fires to be about 2000 t (Environment Canada, 1994). (iv) Other diffuse sources The total PAH released into the atmosphere in the Netherlands from roofing tar and asphalt in 1988 was estimated at 0.5 t/year (Slooff et al., 1989). (c) Emissions to the hydrosphere (i) Motor vehicle traffic The main source of PAH in the aqueous environment as a result of motor vehicle traffic is highway run-off, which contains asphalt and soot particles and is washed by rainfall and storm water or snow into surface waters and soil (see also 3.2.7.2 (a) (i)). The available data are summarized in Table 23. Higher PAH concentrations were found in highway run-off in winter than in summer; this was attributed to the increased abrasion of the road surface due to use of steel-studded tyres in winter (Berglind, 1982). It was estimated that an average of < 10 µg/km per vehicle per day of total PAH are transported via pavement runoff water. Most is transported to nearby surroundings as small particles of dust (see also section 3.2.7.2; Lygren et al., 1984). In contrast, storm water runoff near a US highway was of considerable importance for adjacent water bodies. In the test area, over 50% of the total PAH input into a nearby river came from highway runoff. The runoff loading factor was given as 24 mg/km per vehicle (Hoffman et al., 1985). Table 23. Polycyclic aromatic hydrocarbon concentrations (µg/litre) in highway runoff Compound [1] [2] [3] [4] [5] Acenaphthene 0.016/0.087 0.195/5.126 NR NR NR Acenaphthylene 0.045 0.557/16.804 NR NR NR Anthracene 0.042-0.214 0.486/8.917 0.379 0.165 0.246 Benzo[j+k]fluoranthene 0.089/0.277 NR NR NR 0.207 Benz[a]anthracene 0.031-0.139 0.341/0.863 0.677 0.228 NR Benzo[a]fluorene 0.018-0.170 0.587 NR 0.179 0.396 Benzo[a]pyrene 0.061-0.120 0.537/1.255 0.602 0.250 NR Benzo[b]fluoranthene 0.129/0.157 NR NR 0.799 1.501 Benzo[b]fluorene 0.033/0.097 0.356/0.366 NR NR 0.192 Benzo[c]phenanthrene NR 0.250 NR NR NR Benzo[e]pyrene 0.108/0.202 0.238/1.665 0.609 0.360 0.630 Benzofluoranthenesa 0.401/0.695 1.087/2.712 1.171 NR NR Benzo[e]perylene 0.100-0.299 NR 0.551 0.391 0.319 Chyrsene + triphenylene 0.194-0.433 1.472/2.752 1.147 0.665 1.070 Fluoranthene 0.321-1.573 4.065/15.322 2.665 1.820 3.143 Fluorene 0.0088-0.564 0.432/11.093 0.096 0.485 1.237 Indeno[1,2,3-cd]pyrene 0.061-0.154 0.344/0.666 NR NR NR 1-Methylphenanthrene 0.030-1.073 0.637/2.308 1.366 2.117 Naphthalene NR 2.59 NR 0.123 0.195 Perylene 0.048 NR NR NR NR Phenanthrene 0.068-2.668 3.297/38.10 1.385 4.055 6.787 Pyrene 0.363-1.449 3.026/12.094 2.002 1.886 3.066 NR, not reported; /, single measurements; [1] Run-off samples from a Norwegian highway north of Oslo in summer and winter 1980-82 (Berglind, 1982); [2] Snow 20 and 50 m from the same highway in February 1981 (Berglind, 1982); [3] Snow from a frozen Norwegian lake 50 m from a highway with high traffic density in winter 1981-82 (Gjessing at al., 1984); [4] Snow from a Norwegian highway south of Oslo with concrete pavement, February 1972 (Lygren at al., 1984); [5] Snow from a Norwegian highway south of Oslo with asphalt pavement, February 1972 (Lygren et al., 1984) When the water samples were filtered through solid sorbents, the results may be underestimates of the actual content (see section 2.4.1.4). a Isomers not specified (ii) Sewage treatment The concentrations of PAH in final effluents from municipal sewage treatment facilities are generally in the low microgram per litre range and are almost always < 0.1 µg/litre (Nicholls et al., 1979; Young et al., 1983; van Luin & van Starkenburg, 1984; Kröber & Häckl, 1989). Maximum values of 29 µg/litre naphthalene and 7 µg/litre acenaphthene were detected in one US sewage treatment plant, and 8 µg/litre benzo [a]pyrene were found in one German plant (Young et al., 1983; Kröber & Häckl, 1989), but no explanation was given for these unusually high concentrations. It was concluded that final effluents contain PAH at a background level (van Luin & van Starkenburg, 1984). Naphthalene was found at a concentration of 9.3 kg/year in the final effluent from one US municipal sewage plant (Hoffman et al., 1984). The annual emissions of naphthalene, anthracene, phenanthrene, fluoranthene, benz [a]anthracene, chrysene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, and indeno[1,2,3- cd]pyrene from Dutch sewage treatment plants into surface waters were estimated to be about 0.6 t. The amount of these PAH transported into the Netherlands from other European countries via the Rhine, Meuse, and Scheldt rivers was estimated to be 65 t/year (year and database not given). The main compounds were fluoranthene (18 t/year) and naphthalene (15 t/year) (Slooff et al., 1989). (iii) Other sources PAH have been found in wastewaters from power stations, from garages with car-wash devices, and from a German car-wash storage tank at the following concentrations: fluoranthene, 1.3-7.7 µg/litre; pyrene, 3.5-28 µg/litre; benz [a]anthracene, 0.49-1.9 µg/litre; chrysene, 1.2-6.0 µg/litre; benzo [e]pyrene, 4.7-16 µg/litre; benzo [a]pyrene, 0.40-8.8 µg/litre; benzo [b]fluoranthene, 1.2-3.6 µg/litre; and benzo [k]-fluoranthene, 0.51-0.72 µg/litre (Baumung et al., 1985). Wastewaters from power stations could be an important local source of PAH. Numerous PAH were detected in leachate plumes from refuse landfills in western Germany and the USA (Grimmer et al., 1981b; Götz, 1984; Reinhard et al., 1984). Concentrations < 0.1 µg/litre were detected of benzo [ghi]-fluoranthene, benz [a]anthracene, benzo [c]phenanthrene, chrysene, benzofluoranthenes (isomers not specified), benzo [a]pyrene, benzo [e]pyrene, perylene, anthanthrene, benzo [ghi]perylene, and indeno[1,2,3- cd]pyrene (Grimmer et al., 1981b). Naphthalene was found at a concentration > 100 µg/litre, and acenaphthene, fluorene, anthracene, phenanthrene, and pyrene were found at concentrations of 1-30 µg/litre (Götz, 1984; Reinhard et al., 1984). The importance of this source for groundwater pollution cannot be estimated from the available data. (c) Emissions to the geosphere (i) Motor vehicle traffic PAH were deposited within 100 m of a highway at a concentration of 100-200 µg/km per vehicle per day in winter as small particles of dust resulting from the abrasion of asphalt by steel-studded tyres (Lygren et al., 1984). Studies of adsorption on various soil types showed that most PAH in highway runoff is retained on the soil surface (Gjessing et al., 1984). (ii) Open burning Phenanthrene, fluoranthene, triphenylene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, indeno[1,2,3- cd]pyrene, and coronene were determined in the soil of burning sites in western Oregon, USA. Before burning, the PAH concentrations in the top 2 cm of the soil layer ranged from 0.8 ng/g dry weight for benzo [a]pyrene to 4.4 ng/g for fluoranthene and triphenylene. One week after burning, the concentrations ranged from 0.9 ng/g for benzo [k]fluoranthene to 19 ng/g for triphenylene. The finding that the PAH levels did not increase appreciably after burning indicates that the bulk of the PAH were retained within the litter rather than passing into the soil (Sullivan & Mix, 1983). (iii) Disposal of sewage sludge and fly ash from incineration When sewage sludge is applied to soils, adsorbed PAH are added to the geosphere. The PAH concentrations in municipal aerobic and anaerobic sewage sludge are given in Table 24. In a detailed survey of the PAH concentrations in soil to which anaerobic sludges had been applied between 1942 and 1961 in the United Kingdom, the total PAH content increased to over 125 mg/kg up to 1948 but had decreased to about 29 mg/kg by 1961. The authors attributed the declining levels to a decrease in atmospheric PAH contamination from smoke emissions (Wild et al., 1990). No seasonal variation in the content or profile of PAH was detected in western Germany by Grimmer et al. (1980), but Süss (1980) found the highest PAH load in sewage sludge in January-April and the lowest in July and October. Human faeces seemed to contribute little to the PAH content of sewage sludge (Grimmer et al., 1980). The most important emission sources could not be identified, but McIntyre et al. (1981) concluded that the PAH content of sewage sludge originating from British treatment works with significant flows of industrial effluent was higher than that in works dealing with predominantly domestic effluents. After application of compost over three years to an agricultural soil in Spain, no accumulation of PAH was observed (Gonzalez-Vila et al., 1988). It was shown, however, that the extent of accumulation is dependent on the duration, frequency, and concentration of application. After 10 years of sludge spreading, considerable quantities of PAH were detected in both a sandy loam and a clay soil Table 24. Polycyclic aromatic hydrocarbons concentrations (mg/kg dry weight) in municipal sewage sludge Compound [1] [2] [3] [4] [5] [6] [7] [8] Acenaphthene NR NR NR NR NR NR ND NR Anthracene NR NR NR 0.89-44 NR NR ND-10.0 NR Anthanthrene 0.00-2.10 0.03-1.8 NR NR NR NR NR NR Benz[a]anthracene 0.62-19.0 0.91-17.3 NR NR NR NR ND-2.1 NR Benzo[a]fluorene 0.28-9.00 0.56-7.9 NR NR NR NR NR NR Benzo[a]pyrene 0.54-13.3 0.41-14.3 0.12-9.14 NR NR 0.29-2.00 ND-0.64 NR Benzo[b]fluoranthene NR NR 0.06-9.14 NR < 1-1.3 0.29-1.80 ND-1.100 NR Benzo[e]pyrene 0.53-12.4 0.48-12.3 NR NR NR NR NR NR Benzofluoranthenesa 1.07-23.7 1.02-24.8 NR NR NR NR NR NR Benzo[ghi]perylene 0.40-8.70 0.34-10.9 0.06-9.14 NR NR < 0.1-3.41 ND-1.21 NR Benzo[k]fluoranthene NR NR 0.06-4.57 NR NR 0.15-1.00 ND-0.500 NR Chrysene 0.78-23.7 1.24-22.2 NR 0.25-13 NR NR NR NR Dibenz[a,h]anthracene NR NR NR 13 NR NR ND-0.25 NR Fluoranthene 0.61-51.6 4.10-28.2 0.34-11.45 0.35-7.1 < 1-10.4 0.54-7.67 0.216-5.14 5.2/5.6b Fluorene NR NR NR NR NR NR ND-2.9c 3.5/5.8 Indeno[1,2,3-cd]pyrene 0.30-7.40 0.28-9.4 0.06-6.68 NR NR 0.24-2.08 ND-0.640 NR Naphthalene NR NR NR 0.9-70 NR NR NR 4.5/8.6 Perylene 0.14-6.40 0.09-3.1 NR NR NR NR NR NR Phenanthrene NR NR NR 0.89-44 NR NR 0.30-40 15.2/18.6d Pyrene 0.90-47.2 3.20-25.3 NR 0.33-18N R NR ND-7.6 NR NR, not reported; /, single measurements; ND, not detected (limits of detection, 0.2-1 mg/kg); [1] Samples from 25 sewage treatment plants in western Germany 1976-78 (Grimmer et al., 1980); [2] Samples from three sewage treatment facilities in western Germany before 1979 (Suss, 1980); [3] Samples from 12 British sewage treatment works (McIntyre at al., 1981); [4] Samples from 20 US sewage treatment works (Naylor & Loehr, 1982); [5] Samples from six Dutch municipal sewage treatment plants (van Luin & van Starkenburg, 1984); [6] 31 sludge samples from different sewage treatment works in western Germany (Witte et al., 1988); [7] Anaerobic sludge samples from 13 sewage treatment plants in western Germany 1985-88 (Krober & Hackl, 1989); [8] Anaerobic sludge samples from one Spanish sewage treatment facility in spring 1985 and autumn 1986 (Gonzalez-Villa at al., 1988). a Isomers not specified b With pyrene c With acenaphthylene d With anthracene (Diercxsens & Tarradellas, 1987). The annual addition of PAH to soil from sewage sludge in the Netherlands was estimated as follows: 0.1 t naphthalene, 0.1 t anthracene, 1.5 t phenanthrene, 2.3 t fluoranthene, 0.6 t benzo [a]anthracene, 0.6 t chrysene, 0.4 t benzo [k]fluoranthene, 0.6 t benz [a]pyrene, 0.6 t benzo [ghi]-perylene, and 0.6 t indeno[1,2,3- cd]pyrene (year and database not given; Slooff et al., 1989). The annual contribution of PAH to landfill in the United Kingdom from fly ash from coal combustion (see also Table 16) exceeded that from municipal solid-waste incineration by a factor of about 10, with the exception of naphthalene, the level of which was about 20 000-fold higher in fly ash from coal combustion than in that from solid-waste incineration. The annual PAH loads from solid-waste incineration were about 0.01 kg naphthalene and 3.5 kg benzo [ghi]perylene, whereas those from coal combustion were about 15 kg each of anthracene, benzo [k]fluoranthene, and dibenz [a,h]anthracene and 1200 kg pyrene (Wild et al., 1992). (iv) Waste dumping Soil cores taken from a hazardous waste disposal site in Spain containing petroleum tar residues and lubricating oils as the major organic wastes contained 62 mg/kg 1-methylphenanthrene, 53 mg/kg naphthalene, 52 mg/kg benzo [a]fluorene, 30 mg/kg benzo [ghi]fluoranthene, 25 mg/kg benzo [c]-phenanthrene, 0.5-0.71 mg/kg acenaphthene, 0.2-48 mg/kg fluorene, 0.2-390 mg/kg phenanthrene, 0.110 mg/kg anthanthrene, 0.1-210 mg/kg pyrene, 0.1-200 mg/kg acenaphthylene, 0.1-140 mg/kg anthracene, 0.1-140 mg/kg benzo [e]pyrene, 0.1-145 mg/kg benzo [a]pyrene, 0.1-50 mg/kg benzo [b]fluorene, 0.08-130 mg/kg chrysene plus triphenylene, 0.08-90 mg/kg indeno[1,2,3- cd]pyrene, 0.06-130 mg/kg benz [a]anthracene, 0.05-290 mg/kg fluoranthene, 0.03-75 mg/kg benzo [ghi]perylene, 0.03-0.2 mg/kg perylene, and 0.01-0.4 mg/kg dibenz [a,h]anthracene (Navarro et al., 1991). There can be appreciable movement of PAH into soil from waste dumping, especially of hazardous refuse. The dumping conditions are decisive for the amount of PAH released. Annual emissions of PAH in the Netherlands in 1987 due to the spreading of contaminated composts onto soils were estimated to be 1 t benz [a]anthracene, 1 t chrysene, 1 t benzo [k]fluoranthene, 0.5 t benzo [ghi]-perylene, 0.5 t indeno[1,2,3-cd]pyrene, and 0.4 t benzo [a]pyrene (Slooff et al., 1989). (d) Biosphere Perch (Perca fluviatilis) were not significantly contaminated after an oil spill in Finland due to a tanker accident. The concentrations of acenaphthene, acenaphthylene, fluorene, phenanthrene, anthracene, 1-methylphenanthrene, fluoranthene, pyrene, benzo [a]fluorene, benzo [b]fluorene, chrysene, triphenyl-ene, and benzofluoranthenes in both contaminated and control groups were between < 0.1 and 0.2 µg/kg each in muscle and < 0.1 and 16 µg/kg in bile. The investigators concluded that the fish with the highest load would probably not have survived and others had moved to less contaminated areas. Additionally, the cold climate caused clumping of the spilled oil, which then drifted to the coast (Lindström-Seppä et al., 1989; see also sections 4.1.5.1 and 5.1.7.1). 4. ENVIRONMENTAL TRANSPORT, DISTRIBUTION, AND TRANSFORMATION Appraisal The transport and distribution of polycyclic aromatic hydrocarbons (PAH) in the environment depend on their physicochemical properties of very low solubility in water and low vapour pressure, and high partition coefficients for n-octanol:water (log Kow) and organic carbon:water (log Koc). PAH are stable towards hydrolysis as they have no reactive groups. In the gaseous phase, PAH and particularly those of higher molecular mass, are mainly adsorbed to particulate matter and reach the hydrosphere and geosphere by dry and wet deposition. Little is volatilized from water phases owing to their low Henry's law constants. The log Koc values indicate strong adsorption to the organic matter of soils, so that migration does not usually occur. The log Kow values indicate high bioaccumulation. Few experimental data are available on the biodegradation of PAH. In general, they are biodegradable under aerobic conditions, and the biodegradation rates decrease drastically with the number of aromatic rings. Under anaerobic conditions, biodegradation appears to be very slow. The bioconcentration factors measured in the water phase vary widely according to the technique used. High values are seen for some algae, crustaceans, and molluscs, but those for fish are much lower owing to rapid biotransformation. The bioaccumulation factors for aquatic and terrestrial organisms in sediment and soil are generally very low, probably because of the strong adsorption of PAH onto the organic matter of soils and sediments, resulting in low bioavailability. The photodegradation of PAH in air and water has been studied intensively. The most important degradation process in both media is indirect photolysis under the influence of radicals like OH, O3, and NO3. The measured degradation rate constants vary widely according to the technique used. Under laboratory conditions, the half-life of the reaction of PAH with airborne OH radicals is about one day. Adsorption of high-molecular-mass PAH onto carbonaceous particles in the environment has a stabilizing effect. Formation of nitro-PAH has been reported from two- to four-ring PAH in the vapour phase during photooxidation with NO3. For some PAH, photodegradation in water seems to be more rapid than in air. According to model calculations based on physicochemical and degradation parameters, PAH with four or more aromatic rings persist in the environment. 4.1 Transport and distribution between media 4.1.1 Physicochemical parameters that determine environmental transport and distribution The transport and distribution of PAH in the environment are the result of the following physicochemical parameters: - Aqueous solubility: PAH are hydrophobic compounds with very low solubility in water under environmental conditions: the maximum at room temperature is 32 mg/litre for naphthalene, and the minimum is 0.14 µg/litre for coronene (see Table 4). - Vapour pressure: The vapour pressure of PAH under environmental conditions is very low: the maximum at room temperature is 10.4 Pa for naphthalene, and the calculated minimum is 3 × 10-12 Pa for dibenzo [a,i]pyrene (see Table 4). - n-Octanol:water partition coefficient (log Kow): The affinity of PAH to organic phases is much higher than that for water. The log Kow values range from 3.4 for naphthalene to 7.3 for dibenzo [a,i]pyrene (see Table 4), indicating that the potential for bioaccumulation is high. - Organic carbon:water partition coefficient (log Koc): The sorption coefficients of PAH to the organic fraction of sediments and soils are summarized in Table 25. The high values indicate that PAH sorb strongly to these fractions. The wide variation in the results for individual compounds are due to the very long exposure necessary to reach steady-state or equilibrium conditions, which can lead to underestimation of sorption coefficients; furthermore, degradation in the overlying aqueous phase can lead to overestimates of the actual values. 4.1.2 Distribution and transport in the gaseous phase PAH are emitted mainly to the atmosphere (see Section 3), where they can be both transported in the vapour phase and adsorbed onto particulate matter. The distribution between air and particulate matter under normal atmospheric conditions depends on the lipophilicity, vapour pressure, and aqueous solubility of the substance. Generally, PAH with few (two to four) aromatic rings occur in the vapour phase and are adsorbed, whereas PAH consisting of more aromatic rings exist mainly in the adsorbed state (Hoff & Chan, 1987; McVeety & Hites, 1988; Baker & Eisenreich, 1990). PAH are usually adsorbed onto particles like fly ash and soot that are emitted during combustion. Table 25. Organic carbon normalized sorption coefficients (Koc) of polycyclic aromatic hydrocarbons Compound log Koc Comments Reference Acenaphthene 5.38 Average on sediments Kayal & Connell (1990) 3.79 RP-HPLC on CIHAC Szabo et al. (1990) 3.59 RP-HPLC on PIHAC Szabo at al. (1990) Acenaphthylene 3.83 RP-HPLC on CIHAC Szabo et al. (1990) 3.75 RP-HPLC on PIHAC Szabo et al. (1990) Anthracene 4.42 Average sorption isotherms on Karickhoff et al. (1979) sediment 3.74 Suspended particulates Herbes et al. (1980) 4.20 Soil, shake flask, UV Karickhoff (1981) 3.95/4.73 Lake Erie with 9.6 mg C/litre Landrum et al. (1984a) 4.87/5.70 Huron river with 7.8 mg C/litre Landrum et al. (1984a) 4.20 Soil, shake flask, LSC Nkedl-Kizza et al. (1985) 4.93 Fluorescence, quenching interaction Gauthier et al. (1986) with humic acid 4.38 HPLC Hodson & Williams (1988) 5.76 Average on sediments Kayal & Connell (1990) 4.41 RP-HPLC Pussemier et al. (1990) 4.53 RP-HPLC on CIHAC Szabo at al. (1990) 4.42 RP-HPLC on PIHAC Szabo at al. (1990) Benz[a]anthracene 4.52 Suspended particles Herbes et al. (1980) 6.30 Average on sediments Kayal & Connell (1990) 7.30 Specified particulate Bromen et al. (1990) Benzo[a]pyrene 6.66 LSC Eadie et al. (1990) 6.26 Average on sediments Kayal & Connell (1990) 8.3 Specified particulate Broman et al. (1990) 4.0 Predicted to be dissolved Broman et al. (1990) Benzo[e]pyrene 7.20 Specified particulate Broman at al. (1990) 4.00 Predicted to be dissolved Broman at al. (1990) Benzo[k]fluoranthene 5.99 Average on sediments Kayal & Connell (1990) 7.00 Specified particulate Broman et al. (1990) 4.00 Predicted to be dissolved Broman at al. (1990) Table 25. (continued) Compound log Koc Comments Reference Chrysene 6.27 Average on sediments Kayal & Connell (1990) 6.90 Specified particulate Broman et al. (1990) 4.0 Predicted to be dissolved Broman at al. (1990) Coronene 7.80 Specified particulate Broman et al. (1990) 5.0 Predicted to be dissolved Broman et al. (1990) Dibenz[a,h]anthracene 6.31 Average of 14 soil or sediment Means et al. (1980) samples, shake flask, LSC Fluoranthene 6.38 Average on sediments Kayal & Connell (1990) 4.74 RP-HPLC on CIHAC Szabo et al. (1990) 4.62 RP-HPLC on PIHAC Szabo et al. (1990) 6.30 Specified particulate Broman et al. (1990) 4.0 Predicted to be dissolved Broman et al. (1990) Fluorene 5.47 Average on sediments Kayal & Connell (1990) 3.76 RP-HPLC Pussemier et al. (1990) 4.15 RP-HPLC on CIHAC Szabo et al. (1990) 4.21 RP-HPLC on PIHAC Szabo et al. (1990) Naphthalene 3.11 Average sorption isotherms on Karickhoff at al. (1979) sediments 2.38 Suspended particulates Herbes et al. (1980) 2.94 Karickhoff (1981) 3.0 McCarthy & Jimenez (1985); McCarthy et al. (1985) 2.73-3.91 Aquifer materials Stauffer et al. (1989) 3.15/2.76 Podoll et al. (1989) 5.00 Average on sediments Kayal & Connell (1990) 2.66 Average on sediments Kishi et al. (1990) 3.11 Soil, RP-HPLC Szabo et al. (1990) 3.29 Sandy surface soil Wood et al. (1990) Table 25. (continued) Compound log Koc Comments Reference Phenanthrene 4.36 Average sorption isotherms on Karickhoff et al. (1979) sediments 4.28 Hodson & Williams (1988) 6.12 Average on sediments Kayal & Connell (1990) 4.22 RP-HPLC on CIHAC Szabo et al. (1990) 4.28 RP-HPLC on PIHAC Szabo et al. (1990) 4.42 Sandy surface soil Wood et al. (1990) Pyrene 4.92 Average isotherms on sediments Karickhoff et al. (1979) 4.90 Sediment, shake flask, sorption Karickhoff et al. (1979) isotherm 4.81 Average of soil and sediment Means et al. (1979) Shake flask, LSC, sorption isotherms 4.80 Average of 12 soils and sediments Means et al. (1980) Shake flask, LSC, sorption isotherms 4.78 Soil and sediment; calculated Kow Means at al. (1980) 4.83 Sorption isotherms Karickhoff (1981) 3.11/3.46 Sediment suspensions Karickhoff & Morris (1985) 4.80/5.13 Hodson & Williams (1988) 5.65 LSC Eadie et al. (1990) 5.29 Soil Jury at al. (1990) 6.51 Average on sediments Kayal & Connell (1990) 4.83 RP-HPLC Pussemier et al. (1990) 4.82 RP-HPLC on CIHAC Szabo et al. (1990) 4.77 RP-HPLC on PIHAC Szabo et al. (1990) 6.50 Specified particulate Broman et al. (1990) 4.0 Predicted particulate Broman et al. (1990) Triphenylene 6.90 Specified particulate Broman et al. (1990) 4.00 Predicted to be dissolved Broman et al. (1990) RP-HPLC, reversed-phase high-performance liquid chromatography; CIHAC, chemical-induced humic-acid column; PIHAC, physical-induced humic-acid column; UV, ultraviolet; C, carbon; LSC, liquid scintillation chromatography PAH are ubiquitous in the environment, probably because they are distributed for long distances without significant degradation (Lunde, 1976; De Wiest, 1978; Bjorseth & Sortland Olufsen, 1983; McVeety & Hites, 1988), e.g. from the United Kingdom and the European continent to Norway and Sweden during winter (Bjorseth & Lunde, 1979). Washout ratios calculated from measurements in rain and snow in the area of northern Lake Superior, during one year showed that airborne PAH adsorbed onto particulate matter result in effective wet deposition, while gaseous PAH are removed to only a minor degree (McVeety & Hites, 1988). 4.1.3 Volatilization Henry's law constant gives a rough estimate of the equilibrium distribution ratio of concentrations in air and water but cannot predict the rate at which chemicals are transported between water and air. The constants for PAH are very low, ranging from 49 Pa .m3/mol for naphthalene to 0.000449 Pa .m3/mol for dibenzo [a,i]pyrene (see Table 4). The rates of removal and volatilization of PAH (Table 26) are strongly dependent on environmental conditions such as the depth and flow rate of water and wind velocity. Although PAH are released into the environment mainly in air, considerably higher concentrations are found in aqueous samples because of the low vapour pressure and Henry's law constants of PAH. The volatilization half-life for naphthalene from a 22.5-m water body was found experimentally to be 6.3 h, whereas the calculated value was 2.1 h (Klöpffer et al., 1982). Calculations based on a measured air:water partition coefficient for river water 1 m deep with a water velocity of 0.5 m/s and a wind velocity of 1 m/s gave a volatilization half-life of 16 h for naphthalene (Southworth, 1979). The value calculated for evaporative loss of naphthalene from a 1-m water layer at 25°C was of the same order of magnitude (Mackay & Leinonen, 1975). Naphthalene was volatilized from soil at a rate of 30% after 48 h, with neglible loss of PAH with three or more rings (Park et al., 1990). 4.1.4 Adsorption onto soils and sediments PAH are adsorbed strongly to the organic fraction of soils and sediments (see section 4.1.1 and Table 25). Some PAH may be degraded biologically in the aerobic soil layer, but this process is slow, because sorption to the organic carbon fraction of the soil reduces the bioavailability. For the same reason, leaching of PAH from the soil surface layer to groundwater is assumed to be negligible, although detectable concentrations have been reported in groundwater (see section 5.1.2.2). Table 26. Rates of volatilization of polycyclic aromatic hydrocarbons Compound Rate constant Half-life (h)a Comments Reference Anthracene Removal rate constants (estimated) from Southworth (1977) water column At 25°C in midsummer sunlight: 0.002 h-1 347 - in deep, slow, somewhat turbid water 0.001 h-1 693 - in deep, slow, muddy water 0.002 h-1 347 - in deep, slow, clear water 0.042 h-1 17 - in shallow, fast, clear water 0.179 h-1 3.9 - in very shallow, fast, clear water 62 Calculated half-life for a river 1 m deep Southworth (1979) with water velocity of 0.5 m/s and wind velocity of 1 m/s Benz[a]anthracene 500 Calculated half-life for a river 1 m deep Southworth (1979) with water velocity of 0.5 m/s and wind velocity of 1 m/s Benzo[a]pyrene 1550 Calculated half-life for a river 1 m deep Southworth (1979) with water velocity of 0.5 m/s and wind velocity of 1 m/s <1 × 10-5 S-1 > 19 Sublimation rate constant from glass Cope & Kalkwarf surface at 24 °C at an airflow of 3 litre/min (1987) Naphthalene 1.675 × 10-9 Rate of evaporation estimated at 20 00 Guckel at al. (1973) mol.cm-2h-1 and air flow of 50 litre/h 7.15 Calculated half-life from 1 m depth of water Mackay & Leinonen (1975) 16 Half-life for surface waters Southworth (1979) 200 In a lake, considering current velocity and wind speed in combination with typical re-aeration rates Perylene <1 × 10-5 S-1 > 19 Sublimation rate constant from glass Cope & Kalkwarf surface at 24°C at an air flow of 3 litre/min (1987) Pyrene 1.1 × 10-4 S-1 1.8 Sublimation rate constant as loss from Cope & Kalkwarf glass surface at 24°C at an air flow of 3 litre/min (1987) Table 26 (continued) For comparison of results for which only rate constants are reported, half-lives have been estimated from the equation: t1/2 = In2 k where t1/2 is the half-life and k is the rate constant. The calculated values are reported in italics. 4.1.5 Bioaccumulation The ability of a substance to bioconcentrate in organisms in the aqueous phase is expressed as the bioconcentration factor. For substances like PAH, with high n-octanol:water partition coefficients, long exposures are necessary to achieve equilibrium conditions, so that results obtained under non-equilibrium conditions can result in underestimates of the bioconcentration factor. Bioaccumulation may also vary with the metabolic capacity of the organism (see section 4.2.1.2). Bioconcentration can also be calculated as the ratio between the rates of uptake (k1) and depuration (k2). This method has the advantage that relatively short exposures can be used. It is therefore preferred for PAH, as constant concentrations of compounds like benzo [a]pyrene are very difficult to maintain over a long period. 4.1.5.1 Aquatic organisms Aquatic organisms may accumulate PAH from water, sediments, and their food. In general, PAH dissolved in pore water are accumulated from sediment (McElroy & Sisson, 1989), and digestion of sediment may play an important role in the uptake of PAH by some species. Although organisms can accumulate PAH from food, the relative importance of uptake from food and water is not clear (Farrington, 1991). The bioconcentration factors of PAH in different species are shown in Table 27; this is not a comprehensive presentation of all of the available data but provides examples of the accumulation of some PAH in different groups of organisms. Species that metabolize PAH to little or no extent, like algae, oligochaetes, molluscs, and the more primitive invertebrates (protozoans, porifers, and cnidaria), accumulate high concentrations of PAH, as would be expected from their log Kowvalues, whereas organisms that metabolize PAH to a great extent, like fish and higher invertebrates such as arthropods, echinoderms, and annelids, accumulate little or no PAH (James, 1989). Remarkably high bioconcentration factors have been measured for phenanthrene, anthracene, pyrene, benzo [a]anthracene, and benzo [a]pyrene in the amphipod Pontoporeia hoyi, which has a 20-50% lipid content by wet weight and no capacity to biotransform PAH (Landrum, 1988). The ratio of the concentration of an individual PAH in a bottom-dwelling organism and in the sediment, the bioaccumulation factor, is usually < 1 when expressed as wet weight. In a coastal area, the bioaccumulation factors for 16 PAH in polychaete species varied from 4.9 to 21.8 on a dry-weight basis (Bayona et al., 1991). Measurements of the concentrations of PAH in P. hoyi and in the sediment at three sites with different organic carbon contents gave bioaccumulation factors close to 1 on a wet-weight basis, corrected for the 64-mm sieved fraction (Eadie et al., 1982). The lipid- and organic carbon-based bioaccumulation factors in clams (Macoma baltica) for naphthalene and chrysene added to sediment were 0.78 and 0.16, respectively (Foster et al., 1987). In a study in which clams were exposed for 28 days to six sediments contaminated with different concentrations of PAH (and other organic pollutants) and with an organic carbon content of 0.86-7.4%, the bioaccumulation factors (normalized with respect to lipid content and organic carbon content) ranged from 0.15 to 0.85 (Ferraro et al., 1990). Species that can biotransform PAH have internal concentrations well below the concentration in the sediment. The average bioaccumulation factors (normalized with respect to lipid content and organic carbon content) for eel, pike, and roach at two locations were 0.1 and 0.015. The lowest bioaccumulation factor was found at the site with the highest PAH concentration (128 mg/kg, organic carbon-based), probably due to the inductive capability of the fish to biotransform PAH. This was confirmed by the finding of increased hepatic metabolic activity for PAH in the fish (Van der Oost et al., 1991). 4.1.5.2 Terrestrial organisms Little information is available on the accumulation of PAH in terrestrial organisms. The bioaccumulation factors of 22 PAH in the earthworm Eisenia foetida at six sites varied from 0.23 to 0.6 on an ash-free dry-weight basis (Rhett et al., 1988). The half-life of labelled benzo [a]pyrene in crickets (Acheta domesticus) was 13 h; after 48 h, 36% of the injected dose was unchanged benzo [a]pyrene. After topical application of piperonyl butoxide, a known inhibitor of the mixed-function oxidase system, the level of polar metabolites in the excreta had decreased by approximately 75% within 8 h of injection of benzo [a]pyrene. After articular application of benzo [a]pyrene at 0.29 ng/µl in hexane, some of the dose accumulated internally; the highest level of polar metabolites was found after 24 h (Kumi et al., 1991). The concentration of PAH in vegetation is generally considerably lower than that in soil, the bioaccumulation factors ranging from 0.0001-0.33 for benzo [a]pyrene and from 0.001-0.18 for 17 other PAH tested. It was concluded that some terrestrial plants take up PAH through their roots and/or leaves and translocate them to various other parts (Edwards, 1983). When bush beans (Phaseolus vulgaris Pr.) were exposed to radiolabelled anthracene in a nutrient solution for 30 days during flowering and seed production, more than 90% of the compound was metabolized. Of the total 14C radiolabel, 60% was found in the roots, 3% in the stems, 3% in the leaves, 0.1% in the pods, and 17% in the nutrient solution; 16% was unaccounted for (Edwards, 1986). Table 27. Measured bioconcentration factors of polycyclic aromatic hydrocarbons in aquatic organisms Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Acenaphthene Fish Lepomis 14C S 8.94 28 d 387 Equi Barrows et al. macrochirus (1980) Anthracene Algae Chlorella fusca HPLC S 50 1 d 7 770a NS Geyer et al. (1984) Crustaceans Daphnia magna 14C, TLC S 35 1 d 511 k1/k2 McCarthy et al. (1985) Daphnia magna HPLC S 15 1 d 970 NS Newsted & Giesy (1987) Daphnia magna HPLC S 5.58 24 h 2699 NS Oris et al. (1990) Daphnia pulex Spect S 6 24 h 917 Southworth et al. (1978) Hyalella azteca 14C IF 0.0082 8 h/7 h 2089 k1/k2 Landrum & 14C,TLC 1 800 k1/k2 Scavia (1983) 14C IF 0.0066 8 h/7 h 10985 k1/k2 14C, TLC 9096 k1/k2 Pontoporeia hoyi 14C TLC F 4-17 8 W d 16857 k1/k2 Landrum (1982) Pontoporeia hoyi 14C TLC F 4.6-16.9 6 h/14 d 39727 k1/k2 Landrum (1988) Oligochaetes Stylodrilus -C, TLC F < 6 6 hS d 5051 k1/k2 Frank et al. heringianus (1986) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Fish Lepomis 14C S 0.7 4 h/60 h 900 k1/k2 Spacie et al. macrochirus 14C, TLC 675 (1983) Leuciscus idus UC S 50 3 d 910 NS Freitag A al. melanotus (1985) Oncorhynchus 14C HPLC R 12 18h 190 NS Linder & mykiss 14C HPLC R 12 18 h 270 NS Bergman (1984) Oncorhynchus 14C HIPLC R 50 72 h/144 h 9000 k1/k2 Linder et al. mykiss 9200 (1985) Pimephales HPLC S 6.61 24 In 1016 NS Oris et al. (1990) promelas Benz[a]anthracene Algae Chlorella fusca 14C S 50 1 d 3180 NS Freitag et al. (1985) Crustaceans Daphnis magna 14C TLC S 0.8 1 d 2920 k1/k2 McCarthy et al. (1985) Daphnis pulex Spect S 6 1 d 10109 Southworth et al. (1978) Daphnia pulex HPLC S 1.8 1 d 10226 NS Newsted & Giesy (1987) Pontoporaia hoyi 14C, TLC F 0.62-1.11 6 h/14 d 63000 k1/k2 Landrum (1988) Fish Leuciscus idus 14C S 50 3 d 350 NS Freitag et al. melanotus (1985) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Benzo[a]fluorene Crustaceans Daphnis magna HPLC S 4.8 1d 3668 NS Newsted & Glesy (1987) Benzo[b]fluorene Crustaceans Daphnia magna HPLC S W 1 d 7725 NS Newsted & Giesy (1987) Benzo[a]pyrene Algae Periphyton 14C F 1 1 d 9000 NS Leversee et al. (1981) Crustaceans Daphnis magna 14C S/F 1 6 h 2440 k1/k2 Leversee et al. (1981) Daphnia magna 14C 3050 NS Leversee et al, 14C HPLC 2837 k1/k2 (1981) Daphnia magna 14C TLC S 0.63 1 d 5770 k1/k2 McCarthy et al. (1985) Daphnia magna HPLC S 1.5 1 d 12761 NS Newsted & Giesy (1987) Daphnia pulex 14C S 1.20 24 h 458 NS Trucco et al. 14C S 0.47 24 h 745 NS (1983) 14C S 5.42 24 h 803 NS 14C S 3.21 24 h 1 106 NS 14C S 2.20 24 h 1 259 NS 14C S 1.50 24 h 2 720 NS Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Pontoporeia hoyi 14C, TLC S 0.002-2.6 6 h/14 d 73 000 k1/k2 Landrum (1988) Oligochaetes Stylodrilus 14C, TLC F < 0.03 6 h/8 d 7 048 k1/k2 Frank et al. heringianus (1986) Molluscs Mysis relicta 14C F - 6 h/10-26d 8 297 k1/k2 Evans & Landrum (1989) Ostrea edulis 14C, GLC S 65.7 3 d 58 NS Riley et al. Ostrea edulis 14C, GLC S 65.7 3 d 59 NS (1981) Ostrea edulis 14C, GLC S 65.7 3 d 62 NS Physa sp. 14C, GLC S 2.5 3 d 2 177 NS Lu et al. (1977) Rangia cuneata 14C S 30.5 24 h 236 NS Neff & Anderson 14C S 30.5 24 h 187 NS (1975) Insects Chironomus 14C S 1 8 h/48 h 970 k1/k2 Leversee et al. riparius 14C 600 NS (1981) 14C, HPLC 166 NS Culex pipiens 14C, GLC S 2.5 3 d 37 NS Lu et al. (1977) quinquefasciatus Hexagenia limbata 14C, TLC F - 6 h/14 d 5 870 k1/k2 Landrum & Poore (1988) Fish Lepomis 14C-extraction F 1 2 d/4 d 3 208 k1/k2 Jimenez et al. macrochirus (1987) Lepomis 14C S/F 1 4 h/4 h 4 700 k1/k2 Leversee et al. macrochirus 14C 4 h 120 NS (1981) 14C, HPLC 4 h 12.5 NS Lepomis 14C S 1 4 h/20 h 4 900 k1/k2 Spacie et al. macrochirus 14C, TLC 490 k1/k2 (1983) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Lepomis 14C, TLC S 0.5 5 h/100 h 2 657 k1/k2 McCarthy & macrochirus Jimenez (1985) Leuresthes tenuis Spect S 2 15 d 241 Equi Winkler et al. (1983) Oncorhynchus GC-HPLC F 0.4 10 d 920 NS Gerhart & mykiss Carlson (1978) Salmo salar 14C S 1 48 h/96 h 2 310 k1/k2 Johnsen et al. (1989) Benzo[e]pyrene Crustaceans Daphnis magna HPLC S 0.7 1 d 25 200 NS Newsted & Giesy (1987) Benzo[ghi]perylene Crustaceans Daphnia magna HPLC S 0.2 1 d 28 288 NS Newsted & Giesy (1987) Benzo[k]fluoranthene Crustaceans Daphnia magna HPLC S 1.4 1 d 13 225 NS Newsted & Giesy (1987) Chrysene Crustaceans Daphnia magna 14C S 48 48 h/40 h 5 500 NS Eastmond et al. (1984) Daphnia magna HPLC S 0.7 1 d 6 088 NS Newsted & Giesy (1987) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Dibenz[a,h]anthracene Algae Chlorella fusca 14C S 50 1 d 2 398 NS Freitag et al. (1985) Crustaceans Daphnia magnia HPLC S 0.4 1 d 50 119 NS Newsted & Giesy (1987) Fish Leuciscus idus 14C S 50 3 d 10 NS Freitag et al. melanotus (1985) Fluoranthene Crustaceans Crangon HPLC F 2.4 4 d/14 d 180 k1/k2 McLeese & septemspinosa Burridge (1987) Daphnia magna HPLC S 9 1 d 1 742 NS Newsted & Giesy (1987) Molluscs Mya arenaria HPLC F 2.4 4 d/14 d 4 120 k1/k2 McLeese & Burridge (1987) Mytilus edulis HPLC F 2.4 4 d/14 d 5 920 k1/k2 McLeese & Burridge (1987) Polychaetes Neiris virens HPLC F 2.4 4 d/14 d 720 k1/k2 McLeese & Burridge (1987) Fish Oncorhynchus GC-HPLC F 3.31 21 d 378 Equi Gerhart & mykiss Carlson (1978) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Fluorene Crustaceans Daphnia magna HPLC S 17 1 d 506 NS Newsted & Giesy (1987) Fish Lepomis - IF 20, 37 30 d 1 800 Equi Finger et al. macrochirus - IF 86 30 d 700 Equi (1985) - IF 175, 353 30 d 200 Equi Naphthalene Algae Selenastrum GC S 2,000 1 d 18 000b NS Casserly et al. capricornutum (1983) Chlorella fusca 14C S 50 1 d 130a NS Geyer et al. (1984) Insects Somatochlora Spect S 10 48 h 1 548 NS Correa & Coler cingulata Spect S 100 48 h 178 NS (1983) Crustaceans Daphnia magna 14C, HPLC S 1 000 1 d 19.3 k1/k2 McCarthy et al. (1985) Daphnia magna 14C S 1 800 48 h/40 h 50 NS Eastmond et al. (1984) Daphnia pulex Spect S 1 000 1 d 131 k1/k2 Southworth et al. (1978) Daphnia pulex 14C S 2 292 4 h 677 NS Trucco et al. 14C S 0.45 24 h 10 844 NS (1983) 14C S 2.742 4 h 2 337 NS Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Fish Fundulus 14C S 20 4 h 2.2 NS DiMichele & heteroclitus Taylor (1978) Lepomis 14C, HPLC S 1 000 24 h/36 h 310 k1/k2 McCarthy & macrochirus 14C, HPLC S 100 24 h/36 h 320 k1/k2 Jimenez (1985) Oncorhynchus 14C S 23 8 h/24 h 253 k1/k2 Melancon & Lech mykiss (1978) Perylene Algae Chlorella fusca 14C S 50 1 d 2 010 NS Freitag et al. (1985) Crustaceans Crangon HPLC F 0.4 4 d/14 d 175 k1/k2 McLeese & septemspinosa Burridge (1987) Daphnia magnia HPLC S 0.6 1 d 7 190 NS Newsted & Giesy (1987) Molluscs Mya arenaria HPLC F 0.4 4 d/14 d 100 000 k1/k2 McLeese & Burridge (1987) Mytilus edulis HPLC F 0.4 4 d/14 d 105 000 k/q McLeese & Burridge (1987) Polychaetes Neiris virens HPLC F 0.4 4 d/14 d 180 k1/k2 McLeese & Burridge (1987) Fish Leuciscus idus 14C S 50 3 d < 10 NS Freitag et al. melanotus (1985) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Phenanthrene Bacteria Mixed Spect S 30-300 2 h 6 300c NS Steen & Karickhoff (1981) Algae Selenastrum GC S 1000 1 d 36 970b NS Casserly et al. capricornutum (1983) Chlorella fusca 14C S 50 1 d 1 760a NS Geyer et al. (1984) Insects Hexagenia limbata 14C F - 6 h/14 d 1640 k1/k2 Landrum & Poore (1988) Crustaceans Crangon HPLC F 4.3 4 d/14 d 210 k1/k2 McLeese & septemspinosa Burridge (1987) Daphnia magna HPLC S 40.1 1 d 323 NS Newsted & Giesy (1987) Daphnia magna 14C S 60 48 h/40 h 600 NS Eastmond et al. (1984) Daphnia pulex 14C S 6.01 24 h 1 165 NS Trucco et al. 14C S 3.10 24 h 1 032 NS (1983) 14C S 3.45 24 h 1 424 NS Daphnia pulex Spect S 30 1 d 325 k1/k2 Southworth et al. (1978) Pontoporeia hoyi 14C-TLC F 0.7-7.1 6 h/14 d 28 145 k1/k2 Landrum (1988) Oligochaetes Stylodrilus 14C-TLC F < 200 6 h/8 d 5 055 k1/k2 Frank et al. heringianus (1986) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Molluscs Mya arenaria HPLC F 4.3 4 d/14 d 1 280 k1/k2 McLeese & Burridge (1987) Mytilus edulis HPLC F 4.3 4 d/14 d 1 240 k1/k2 McLeese & Burridge (1987) Polychaetes Neiris virens HPLC F 4.3 4 d/14 d 500 k1/k2 McLeese & Burridge (1987) Pyrene Bacteria Mixed Spect S 1-20 2 h 24 600c NS Steen & Karickhoff (1981) Algae Selenastrum GC S 500 1 d 55 800b NS Casserly et al. capricornutum (1983) Crustaceans Crangon HPLC F 1.7 4 d/14 d 225 k1/k2 McLeese & septemspinosa Burridge (1987) Daphnis magna HPLC S 5.7 24 h 2 702 NS Newsted & Giesy (1987) Daphnis pulex Sped S 50 24 h 2 702 k1/k2 Southworth et al. (1978) Pontoporeia hoyi 14C-TLC F 0.002-0.011 6 h/14 d 16 600 k1/k2 Landrum (1988) Molluscs Mya arenaria HPLC F 1.7 4 d/14 d 6 430 k1/k2 McLeese & Burridge (1987) Mytilus edulis HPLC F 1.7 4 d/14 d 4 430 k1/k2 McLeese & Burridge (1987) Table 27. (continued) Species Analysis Test Concentration Duration of Bioconcentration Type Reference system in water exposure factor (in (µg/litre) or uptake/ wet weight) depuration period Oligochaetes Stylodrilus 14C-TLC F < 26.4 6 h/8 d 6 588 k1/k2 Frank et al. heringianus (1986) Polychaetes Neiris virens HPLC F 1.7 4 d/14 d 700 k1/k2 McLeese & Burridge (1987) Fish Oncorhynchus GC-HPLC F 3.89 21 d 72.2 Equi Gerhart & mykiss Carlson (1978) Triphenylene Crustaceans Crangon HPLC F 0.5 4 d/14 d 270 k1/k2 McLeese & septemspinosa Burridge (1987) Daphnia magna HPLC S 1.7 1 d 9 066 NS Newsted & Giesy (1987) Molluscs Mya arenaria HPLC F 0.5 4 d/14 d 5 540 k1/k2 McLeese & Burridge (1987) Mytilus edulis HPLC F 0.5 4 d/14 d 11 390 k1/k2 McLeese & Burridge (1987) Polychaetes Neiris virens HPLC F 0.5 4 d/14 d 2 560 k1/k2 McLeese & Burridge (1987) Table 27 (continued) 14C, measurement of radioactivity in a liquid scintillation counter: as parent compounds cannot be differentiated from metabolites with this method, additional extraction is usually performed. S, static exposure system; Equi, at equilibrium Corg/Cw; HPLC, high-performance liquid chromatography; NS, not steady-state Corg/Cw; TLC, thin-layer chromatography; k1/k2, kinetics: uptake rate/depuration rate; Spect, spectroscopy; F, flow-through system; R, static renewal system; GLC, gas-liquid chromatography; GC, gas chromatography; IF, intermittent flow system a Based on dry weight (5 × wet weight) b Based on total suspended solids c Based on dry weight 4.1.6 Biomagnification Biomagnification, the increase in the concentration of a substance in animals in successive trophic levels of food chains, has been determined in a number of studies. When Daphnia pulex were exposed to water or algae contaminated with naphthalene, phenanthrene, benz [a]anthracene, or benzo [a]pyrene, naphthalene accumulated to the greatest extent from algal food, (bioconcentration factor, 11 000), whereas benz [a]anthracene and benzo [a]pyrene accumulated more from water (bioconcentration factors, 1100 and 2700, respectively). It must be emphasized that because of the short exposure (24 h), the last two compounds would not have reached equilibrium (Trucco et al., 1983). In a study of bioaccumulation and biomagnification in closed laboratory model ecosystems, green algae (Oedogonium cardiacum), D. magna, mosquito larvae (Culex pipiens quinquefasciatus), snails (Physa sp.), and mosquito fish (Gambusia affinis) were exposed for three days to 2 µg/litre of 14C-benzo [a]pyrene. Of the radiolabel accumulated, 88% was attached to parent compound in snails, 22% in mosquito larvae, and none in fish. The parent compound represented 46% of the total extractable radiolabel in mosquito larvae and 90% in Daphnia. The bioconcentration factors were 5300 for algae, 12 000 for mosquito larvae, 82 000 for snails, 140 000 for Daphnia, and 930 for fish. Despite the apparent absence of bioconcentration in fish, accumulation is assumed to be due to food-chain transfer, as no accumulation of benzo [a]pyrene was found in a study of uptake from water. Biomagnification was also studied in a terrestrial-aquatic system, by adding 14C-benzo [a]pyrene to Sorghum vulgare seedlings and allowing them to be eaten by fourth-instar salt-marsh caterpillar larvae (Estigmene acrea); the labelled products entered the terrestrial and aquatic phases as products such as faeces. The food-chain organisms were the same as in the model aquatic ecosystem. After a 33-day interaction period, the concentrations of benzo [a]pyrene were 0.01 µg/litre water and 36.1 µg/kg algae, with bioconcentration factors of 3600, 490, 2100, and 30, respectively. Most of the radiolabel was found on polar products or as unextractable radioactivity, which comprised 25% of the total in snails, 63% in fish, 67% in mosquito larvae, and 79% in algae (Lu et al., 1977). Trophic transfer of benzo [a]pyrene metabolites between benthic organisms was studied by feeding Nereis virens 14C-benzo [a]pyrene and harvesting them five days later. The worm homogenate contained 14% parent compound, 7.2% organic-soluble metabolites, 58% water-soluble metabolites, and 21% bound material. Flounder (Pseudiopleuronectes americanus) were then given doses of 4.8-19 g of either pure benzo [a]pyrene homogenized in unexposed Nereis or the worm-metabolite mixture by gavage and analysed after 24 h of incubation. On the basis of the radiolabel recovered from the fish tissues, assuming comparable accumulation efficiency, flounder appear to have at least a limited ability to accumulate polar, conjugated, and bound metabolic products of benzo [a]pyrene from the diet. The parent compound represented 5-15% of the radiolabel in liver and 6-7% in intestine; conjugated metabolites represented 40-60% of the label in liver and 60-70% in intestine; and bound metabolic products represented 30% in liver and 10-20% in intestine (McElroy & Sisson, 1989). 4.2 Transformation On the basis of model calculations, Mackay et al. (1992) classified some PAH according to their persistence in air, water, soil, and sediment (Table 28). Table 28. Suggested half-life classes of polycyclic aromatic hydrocarbons in various environmental compartments Class Half-life (h) Mean Range 1 17 10-30 2 55 30-100 3 170 100-300 4 550 300-1000 5 1 700 1000-3000 6 5 500 3000-10 000 7 17 000 10 000-30 000 8 55 000 > 30 000 Compound Air Water Soil Sediment Acenalphthylene 2 4 6 7 Anthracene 2 4 6 7 Benz[a]anthracene 3 5 7 8 Benzo[a]pyrene 3 5 7 8 Benzo[k]fluoranthene 3 5 7 8 Chrysene 3 5 7 8 Dibenz[a,h]anthracene 3 5 7 8 Fluoranthene 3 5 7 8 Fluorene 2 4 6 7 Naphthalene 1 3 5 6 Perylene 3 5 7 8 Phenanthrene 2 4 6 7 Pyrene 3 5 7 8 From Mackay et al. (1992) 4.2.1 Biotic transformation 4.2.1.1 Biodegradation Information on the biodegradation of PAH in water and soil under aerobic and anaerobic conditions is summarized in Table 29. The few results available from standard tests for biodegradation in water show that PAH with up to four aromatic rings are biodegradable under aerobic conditions but that the biodegradation rate of PAH with more aromatic rings is very low. Biodegradation under anaerobic conditions is slow for all components (Neff, 1979). The reactions normally proceed by the introduction of two hydroxyl groups into the aromatic nucleus, to form dihydrodiol intermediates. Bacterial degradation produces cis-dihydrodiols (from a dioxetane intermediate), whereas metabolism in fungal or mammalian systems produces trans-dihydrodiol intermediates (from an arene oxide intermediate). The differences in the metabolic pathways are due to the presence of the cytochrome P450 enzyme system in fungi and mammals. Algae have been reported to degrade benzo [a]pyrene to oxides, peroxides, and dihydroxydiols (see below). Owing to the high biotransformation rate (see also section 4.2.1.2), the concentrations of PAH in organisms and water are usually not in a steady state. Freely dissolved PAH may be rapidly degraded under natural conditions if sufficient biomass is available and the turnover rates are fairly high (see Table 29). Biodegradation is the major mechanism for removal of PAH from soil. PAH with fewer than four aromatic rings may also be removed by volatilization and photolysis (see also sections 4.1.4 and 4.2.2.1). The rate of biodegradation in soil depends on several factors, including the characteristics of the soil and its microbial population and the properties of the PAH present. Temperature, pH, oxygen content, soil type, nutrients, and the presence of other substances that can act as co-metabolites are also important (Sims & Overcash, 1983). Biodegradation is further affected by the bioavailability of the PAH. Sorption of PAH by soil organic matter may limit the biodegradation of compounds that would normally undergo rapid degradation (Manilal & Alexander, 1991); however, no significant difference was found in the biodegradation rate of anthracene in water with 10 and 1000 mg/litre suspended material (Leslie et al., 1987). In Kidman sandy loam, the biodegradation rates varied between 0.23 h-1 (or 5.5 d-1) for naphthalene and 0.0018 d-1 for fluoranthene (see Table 29). In a study with sandy loams, forest soil, and roadside soil partially loaded with sewage sludge from a municipal treatment plant, the following half-lives (in days) were found: 14-48 for naphthalene, 44-74 for acenaphthene plus fluorene, 83-193 for phenanthrene, 48-210 for anthracene, 110-184 for fluoranthene, 127-320 for pyrene, 106-313 for benz [a]anthracene plus chrysene, 113-282 for benzo [b]fluoranthene, 143-359 for benzo [k]fluoranthene, 120-258 for benzo [a]pyrene, 365-535 for benzo [ghi]perylene, and 603-2030 for coronene (Wild & Jones, 1993). Table 29. Biodegradation of polycyclic aromatic hydrocarbons (PAH) Compound Rate constant Half-life Comments Reference Acenaphthene 100% degradation Significant degradation with rapid adaptation; Tabak et al. after 7 d static flask screening; settled domestic waste (1981) as inoculum; experiments with 5 and 10 mg/litre PAH at 25°C; detection by GC 295-2448 h Aerobic half-life; aerobic soil column Kincannon & Lin (1985) 1180-9792 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation half-life (1991) 0% degradation Japanese Ministry of Trade and Industry test Japanese Ministry of after 7 d with 100 mg/litre PAH and 30 mg/litre sludge International Trade and Industry (1992) < 3.2 year Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Acenaphthylene 98% degradation Significant degradation with rapid adaptation; Tabak et al. after 7 d statis flask screening; settled domestic waste (1981) as inoculum; 5 or 10 mg/litre PAH at 25°C; detection by GC 1020-1440 h Aerobic half-life; soil column Kincannon & Lin (1985) 4080-5760 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 0% degradation Japanese Ministry of Trade and Industry test Japanese Ministry of after 4 weeks with 100 mg/litre PAH and 30 mg/litre sludge International Trade and Industry (1992) Table 29. (continued) Compound Rate constant Half-life Comments Reference Anthracene 0.061 h-1 10 h Microbial degradation in Third Creek water Southworth incubated 18 h at 25°C: (1977) Removal rate constants from water column at 25°C in midsummer sunlight: 0.060 h-1 12 h - in deep, slow, somewhat turbid water 0.030 h-1 23 h - in deep, slow, muddy water 0.061 h-1 11 h - in deep, slow, clear water 0.061 h-1 11 h - in shallow, fast, clear water 0.061 h-1 11 h - in very shallow, fast, clear water 0.035 h-1 20 h Microbial degradation rate constant Herbes et al. (1980) 51-92% degradation Significant degradation with gradual Tabak et al. after 7 d adaptation; static flask screening; settled (1981) domestic waste as inoculum; experiments with 5 and 10 mg/litre PAH at 25°C; detection by GC 1200-11 040 h Aerobic half-life; aerobic soil die-away Coover & Sims (1987) 20O g dry weight of soil at -0.33 bar Park et al. [33 kPa] soil moisture at 25°C: (1990) 0.0052 d-1 3200 h - Kidman sandy foam; initial test concentration, 210 mg/kg 0.0138 d-1 1200 h - McLaurin sandy loam; initial test concentration, 199 mg/kg 4800-44 160 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation half-life (1991) Table 29. (continued) Compound Rate constant Half-life Comments Reference 1.9% degradation Japanese Ministry of Trade and Industry test Japanese Ministry of after 2 weeks with 100 mg/litre PAH and 30 mg/litre sludge International Trade and Industry (1992) Anthracene 33% after 16 months Degradation in soil in co-metabolic closed Bossert & bottle with 1-phenyldecane as primary Bartha (1986) substrate; 20°C; initial test concentration, 1 mg/g; abiotic loss, 60% 5% after 56 d Batch test with river water; initial concentration, Fedorak et al. 20 mg/litre related to dissolved organic carbon; (1982) no mineralization during first 19 days; 20°C Serum bottle radiorespirometry in five soils Grosser et al. contaminated with hydrocarbons: (1995) 10-60% after 64 d - initial concentration, 31.3 ng/g - Inoculated with enriched culture of Mycobacteriarn sp. and initial test concentration of 37.7 ng/g; biodegradation rate without enriched culture, 18% after 64 d Static test in bioreactor in enriched mixed Walter et al. culture; anthracene oil (38 g/litre) which also (1990) contained 62 mg/g fluorene; 30°C: 100% after 3 d - under aerobic conditions 90% after 20 d - under anaerobic conditions 17-45 d Aerobic degradation in surface Donneybrook Bulman et al. sandy loam from Canadian pasture; initial test (1987) concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20 00 and water-holding capacity of 60% of the soil Table 29. (continued) Compound Rate constant Half-life Comments Reference 7.9 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Benz[a]anthracene 2448-16 320 h Aerobic soil die-away at 10-30°C Groenewegen & Stolp (1976); Coover & Sims (1987) 0% degradation No significant degradation under conditions of Tabak et al. (1981) after 7 d method; static flask sceening; settled domestic waste as inoculum; experiment with 5 and 10 mg/littre PAH at 25°C; detection by GC 0.0026 d-1 6400 h Kidman sandy loam Park et al. (1990) 9792-65 280 h Anaerobic half-life; estimated unacclimatized Howard et al. (1991) aqueous aerobic biodegradation 16% after Degradation in soil in co-metabolic closed Bossert & Bartha 16 months bottle with 1-phenyldecane as primary (1986) substrate; 20°C; initial test concentration, 1 mg/g; abiotic loss, 18% 0-40% after 64 d Serum bottle radiorespirometry in five soils Grosser et al. contaminated with hydrocarbons; initial (1995) concentration, 31.3 ng/g 130-240 d Aerobic degradation in surface samples of Bulman et al. Donneybrook sandy loam from Canadian (1987) pasture; initial test concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20°C and water-holding capacity of 60% of the soil Table 29. (continued) Compound Rate constant Half-life Comments Reference 8.1 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Benzo[a]pyrene 0.2-0.9 Aquatic fate rate for bacterial protein Barnsley (1975) µmol.h-1mg-1 3.5 × 10-5 h-1 19 800 h Estimated rate constant in soil and water Ryan & Cohen (1986) 1368-12 702 h Aerobic half-life at 10-30°C; soil die-away Coover & Sims (1987) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; 33 mg/kg at 25°C: 0.0022 d-1 7416 h - Kidman sandy loam 0.0030 d-1 5496 h - McLaurin sandy loam 5472-50 808 h Anaerobic half-life; estimated unacclimatized Coover & Sims aqueous aerobic biodegradation (1987) < 8% after 160 d Serum bottle radiorespirometry in five soils Grosser et al. contaminated with hydrocarbons; initial (1995) concentration, 105 ng/g 218-347 d Aerobic degradation in surface samples of Bulman et al. Donneybrook sandy loam from Canadian (1987) pasture; initial test concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20°C and water-holding capacity of 60% of the soil 8.2 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Table 29. (continued) Compound Rate constant Half-life Comments Reference Benzo[b]fluoranihene 8640-14 640 h Aerobic half-life; estimated unacclimatized Coover & Sims aqueous aerobic biodegradation (1987) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; initial test concentration, ± 38 mg/kg at 25°C: 0.0024 d-1 7056 h - Kidman sandy loam 0.0033 d-1 5064 h - McLaurin sandy loam 34 560-58 560 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 9 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Benzo[ghi]perylene 14 160-15 600 h Aerobic half-life; aerobic soil dieaway at Coover & Sims 10-30°C (1987) 56 640-62 400 h Anaerobic half-life; aerobic soil dieaway at Coover & Sims 10-30°C (1987) 9.1 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Benzo[k]fluoranthene 21 840-51 360 h Aerobic half-life; aerobic soil dieaway Coover & Sims (1987) 87 360-205 440 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) Table 29. (continued) Compound Rate constant Half-life Comments Reference 8.7 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Chrysene 59% degradation Significant degradation with gradual Tabak et al. (1981) after 7 d adaptation; static flask screening; settled domestic waste as inoculum; experiment with 5 mg/litre PAH at 25°C; detection by GC 38% degradation No significant degradation under conditions of Tabak et al. (1981) after 7 d method; static flask sceening; settled domestic waste as inoculum; experiment with 10 mg/litre PAH at 25°C; detection by GC 8904-24 000 h Aerobic half-life; aerobic soil dieaway Coover & Sims (1987) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; initial test concentration, ± 100 mg/kg at 25°C: 0.0019 d-1 8904 h - Kidman sandy loam 0.0018 d-1 9288 h - McLaurin sandy loam 35 616-96 000 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 11 % after 16 Degradation in soil in co-metabolic closed Bossert & Bartha months bottle with 1-phenyldecane as primary (1986) substrate; 20°C; initial test concentration, 1 mg/g; abiotic loss, 5% Table 29. (continued) Compound Rate constant Half-life Comments Reference 224-328 d Aerobic degradation in surface samples of Bulman et al. Donneybrook sandy loam from Canadian (1987) pasture; initial test concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20°C and water-holding capacity of 60% of the soil 8.1 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Coronene 16.5 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Dibenz[a,h]anthracene 8664-22 560 In Aerobic half-life; aerobic soil die-away Coover & Sims (1987); Park et al. (1990) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; initial test concentration, ± 13 mg/kg at 25°C: 0.0019 d-1 8664 h - Kidman sandy loam 0.0017 d-1 10 080 h - McLaurin sandy loam No degradation Degradation in soil in co-metabolic closed Bossert & Bartha after 16 months bottle with 1-phenyldecane as primary (1986) substrate; 20°C; initial test concentration, 1 mg/g Table 29. (continued) Compound Rate constant Half-life Comments Reference Fluoranthene 2.2 × 10-3 Aquatic fate rate with bacterial protein Barnsley (1975) µmol h-1mg-1 100% degradation Significant degradation with gradual adaptation; Tabak et al. (1981) after 7 d static flask screening; settled domestic waste as inoculum; experiment with 5 mg/litre PAH at 25°C; detection by GC 0% degradation No significant degradation under conditions of Tabak et al. (1981) after 7 d method; static flask screening; settled domestic waste as inoculum; experiment with 10 mg/litre PAH at 25°C; detection by GC 3360-10 560 h Aerobic half-life; aerobic soil dieaway Coover & Sims (1987) 0.19 h-1 3.6 h In atmosphere Dragoescu & Friedlander (1989) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; initial test concentration, 900 mg/kg at 25°C: 0.0018 d-1 9048 h - Kidman sandy loam 0.0026 d-1 6432 h - McLaurin sandy loam 13 440-42 240 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 34-39 d Aerobic degradation in surface samples of Bulman et al. Donneybrook sandy loam from Canadian (1987) pasture; initial test concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20°C and water-holding capacity of 60% of the soil Table 29. (continued) Compound Rate constant Half-life Comments Reference 7.8 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Fluorene 45-77% degradation Significant degnadation with gradual adaptation; Tabak et al. (1981) after 7 d static flask screening; settled domestic waste as inoculum; experiment with 5 and 10 mg/litre PAH at 25°C; detection by GC Degradation of 30 µg/litre in natural river water Lee & Ryan (1976) (Skidway River; salinity, 20%): 100% after 1000 d - Turnover time in June at incubation time of 48 h 0% after 72 h - February or May 30% after 1 week Degradation of non-autoclaved groundwater Lee et al. (1984) samples of ± 0.06 mg/litre by microbes 768-1440 h Aerobic half-life; aerobic soil diaway Coover & Sims (1987) 3072-5760 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 0% degradation Japanese Ministry of Trade and Industry test Japanese Ministry of after 4 weeks with 100 mg/litre PAH and 30 mg/litre sludge International Trade and Industry (1992) 100% after 36 h Batch test with enriched culture of Arthrobacter Grifoll et al. sp.; initial test concentration, 483 µmol/litre; (1992) 22°C Table 29. (continued) Compound Rate constant Half-life Comments Reference < 3.2 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Indeno[1,2,3-cd]pyrene 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; initial test concentration, ± 8 mg/kg at 25°C: 0.0024 d-1 6912 h - Kidman sandy loam 0.0024 d-1 6936 h - McLaurin sandy loam Naphthalene Degradation in natural river water (Skidway Lee & Ryan River; salinity, 20%): (1976) 500 d - Turnover time in February at incubation time of 48 h; test concentration, 40 µg/litre 46 d - Turnover time in May at incubation time of 24 h; test concentration, 40 µg/litre 79 d - Turnover time in May at incubation time of 8 h; test concentration, 40 µg/litre 30 d - Turnover time in May at incubation time of 24 h; test concentration, 130 µg/litre Degradation of 130 µg/litre in natural water Lee & Ryan 330 d offshore with salinity of 35%: turnover time (1976) in May at incubation time of 24 h 0.0403.3 × 10-6 At depth of 5-10 m in laboratory water basin Lee & Anderson g/litre per d (1977) 100% after 8 d In gas-oll-contaminated groundwater Kappeler & circulated through sand inoculated with Wuhrmann groundwater under aerobic conditions (1978) Table 29. (continued) Compound Rate constant Half-life Comments Reference 168 h In oil-polluted estuarine stream Lee (1977) 576 h In clean estuarine stream 1500 h In coastal waters 40 800 h In the Gulf Stream 12h Aerobic half-life; die-away in oil-polluted Walker & Colwell creek (1976) Anaerobic half-life: Hambrick et al. 600 h at pH 8 (1980) 6200 h at pH 5 24-216 h In deep, slowly moving, contaminated water Herbes (1981); Wakeham et al. (1983) 0.23 h-1 3.O h Microbial degradation rate constant Herbes et al. (1980) 100% degradation Significant degradation with rapid adaptation; Tabak et al. (1981) after 7 d static flask screening; settled domestic waste as inoculum; experiments with 6 and 10 mg/litre PAH at 25°C; detection by GC 100% degradation Degradation of non-autoclaved groundwater Lee et al. (1984) after 7 d samples of ± 0.04 mg/litre by microbes 0.024 d-1 693 h Groundwater with nutrients and acclimatized Vaishnav & Babeu microbes (1987) 0.013 d-1 1279 h River water with acclimatized microbes 0.018-1 924 h River water with nutrients and acclimatized microbes Table 29. (continued) Compound Rate constant Half-life Comments Reference 200 g dry weight of soil at -0.33 bar Park et al. (1990) [-0.0032 kPa] soil moisture; initial test concentration, 101 mg/kg at 25°C: 0.377 d-1 50 h - Kidman sandy loam 0.308 d-1 53 h - McLaurin sandy loam 2% degradation Japanese Ministry of Trade and Industry test Japanese Ministry of after 4 weeks with 30 mg/litre PAH and 100 mg/litre sAdge International Trade and Industry (1992) < 2.1 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH Perylene No degradation Degradation in soil in co-metabolic closed Bossert & Bartha after 16 months bottle with 1-phenyldecane as primary (1986) substrate; 20°C; initial test concentration, 1 mg/g Phenanthrene 100% degradation Significant degradation with rapid adaptation; Tabak et al. after 7 d static flask screening; settled domestic waste (1981) as inoculum; experiments with 5 and 10 mg/litre PAH at 25°C; detection by GC 383-4800 h Aerobic half-life; aerobic soil die-away Coover & Sims (1987) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [-0.0032 kPa] soil moisture; initial test concentration, 900 mg/kg at 25°C: 0.0447 d-1 384 h - Kidman sandy loam 0.0196 d-1 840 h - McLaurin sandy loam Table 29. (continued) Compound Rate constant Half-life Comments Reference 1536-19 200 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 96 h Inorganic solution Manilal & Alexander 264 h Kendaia soil (1991) 54% degradation Japanese Ministry of Trade and Industry test Japanese Ministry of after 4 weeks with 100 mg/litre PAH and 30 mg/litre sludge International Trade and Industry (1992) > 62 % after Degradation in soil in co-metabolic closed Bossert & Bartha 16 months bottle with 1-phenyldecane as primary (1986) substrate; 20°C; initial test concentration, 1 mg/g; abiotic loss significant Serum bottle radiorespirometry in five soils Grosser et al. (1995) contaminated with hydrocarbons: 38-55% after 64 d - initial concentration, 31.3 ng/g 80% after 32 d - inoculated with enriched culture of Mycobacterium sp. and an initial test concentration of 17.9 ng/g 9.7-14 d Aerobic degradation in surface samples of Bulman et al. (1987) Donneybrook sandy loam from Canadian pasture; initial test concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20°C and water-holding capacity of 60% of the soil 5.7 years Field tests of rural British soils amended with Wild et al. metal-enriched sewage sludges with (1991) 0.1-15.1 mg/kg PAH Table 29. (continued) Compound Rate constant Half-life Comments Reference Pyrene 100% degradation Significant degradation with rapid adaptation; Tabak et al. after 7 d static flask screening; settled domestic waste (1981) as inoculum; experiment with 5 mg/litre PAH at 25°C; detection by GC 0% degradation No significant degradation under conditions of Tabak et al. after 7 d method; static flask screening; settled domestic (1981) waste as inoculum; experiments with 5 and 10 mg/litre PAH at 25°C; detection by GC 5040-46 600 h Aerobic half-life at 10-30°C; aerobic soil Coover & Sims die-away (1987) 0.29 h-1 2.4 h In atmosphere Dragoescu & Friedlander (1989) 200 g dry weight of soil at -0.33 bar Park et al. (1990) [33 kPa] soil moisture; initial test concentration, ± 690 mg/kg at 25°C: 0.0027 d-1 6240 h - Kidman sandy loam 0.0035 d-1 4776 h - McLaurin sandy loam 20 160-182 400 h Anaerobic half-life; estimated unacclimatized Howard et al. aqueous aerobic biodegradation (1991) 70% after 16 months Degradation in soil in co-metabolic closed Bossert & Bartha bottle with 1-phenyldecane as primary (1986) substrate; 20°C; initial test concentration, 1 mg/g; abiotic loss, 27% Table 29. (continued) Compound Rate constant Half-life Comments Reference Serum bottle radiorespirometry in five soils Grosser et al. contaminated with hydrocarbons: (1995) 25-70% after 64 d - initial concentration, 8.5 ng/g 54% after 32 d - inoculated with enriched culture of Mycobacterium sp. and an initial test concentration of 7.7 ng/g 52.4% after 96 h Mineralization test with Mycobacterium sp.; Heitkamp et al. 24°C; initial test concentration, 0.5 mg/litre (1988) 48-58 d Aerobic degradation in surface Donneybrook Bulman et al. (1987) sandy loam from Canadian pasture; initial test concentrations, 5 and 50 mg/kg; up to 400 days' exposure at 20°C and water-holding capacity of 60% of the soil 8.5 years Field tests of rural British soils amended with Wild et al. (1991) metal-enriched sewage sludges with 0.1-15.1 mg/kg PAH GC, gas chromatography In order to compare numbers when only rate constants are reported, the half-lives were estimated from the formula: t1/2 = In2 k where t1/2 is the half-life and k is the rate constant. The calculated values are reported in italics. After biodegradation of pyrene by a Mycobacterium sp., cis- and trans-4,5-pyrene dihydrodiol and pyrenol were the initial ring oxidation products. The main metabolite was 4-phenathroic acid. The ring fission products were 4-hydroxyperinaphthenone and cinnamic and phthalic acids (Heitkamp et al., 1988). The pyrene-metabolizing Mycobacterium sp. can also use phenanthrene and fluoranthene as the sole source of carbon. Phenanthrene was degraded and 1-hydroxy-2-naphthoic acid, ortho-phthalate, and protocatechuate were detected as metabolites. 1-Hydroxy-2-naphthoic acid did not accumulate, indicating that it is further metabolized (Boldrin et al., 1993). A strain of Arthobacter sp. was isolated that was capable of metabolizing fluorene as a sole energy source: 483 nmol/ml were degraded completely within 36 h, and four major metabolites were detected: 9-fluorenol, 9 H-fluoren-9-one, 3,4-dihydrocoumarin, and an unidentified polar-substituted aromatic compound. Fluorenol was not degraded further, suggesting that it and fluorenone are products of a separate metabolic pathway from that which produces dihydrocoumarin, the polar compound, and the energy for cell growth. The bacteria could also degrade phenanthrene (Grifoll et al., 1992). The degradation of PAH was studied in a culture made from activated sludge, polychlorinated biphenyl-degrading bacteria, and chlorophenol-degrading mixed cultures, adapted to naphthalene. The metabolites of naphthalene were 2-hydroxybenzoic acid and 1-naphthalenol, those of phenanthrene were 1-phrenanthrenol and 1-hydroxy-2-naphthalenecarboxylic acid, and that of anthracene was 3-hydroxy-2-naphthalenecarboxylic acid. The authors concluded that the biotransformation pathway proceeds via initial hydroxylation to ring cleavage, to yield the ortho or meta cleavage intermediates, which are further metabolized via conventional metabolic pathways (Liu et al., 1992). The metabolism of PAH by fungi is similar to that by mammalian cells. For example, Cunninghamella elegans in culture metabolizes benzo [a]pyrene to the trans-7,8-diol, the trans-9,10-diol, 3,6-quinone, 9-hydroxybenzo [a]pyrene, 3-hydroxybenzo [a]pyrene, and 7,8-dihydro-7,8-dihydroxybenzo [a]pyrene (Cerniglia, 1984). In a further experiment, C. elegans metabolized about 69% of added fluorene after 24 h. The major ethyl acetate-soluble metabolites were 9-fluorenone (62%), 9-fluorenol, and 2-hydroxy-9-fluorenone (together, 7%). The degradation pathway was similar to that in bacteria, with oxidation at the C9 position of the five-member ring to form an alcohol and the corresponding ketone. 2-Hydroxy-9-fluorenone had not been found as a metabolite previously (Pothuluri et al., 1993). 4.2.1.2 Biotransformation Biotransformation is often advanced as an explanation for the differences in PAH profiles seen in aquatic organisms and in the medium to which they were exposed. Furthermore, all of the metabolites of PAH may not have been identified or quantified. This section addresses biotransformation in organisms other than bacteria and fungi, which is discussed in section 4.2.1.1, above. The uptake of naphthalene and benzo [a]pyrene was studied in three species of marine fish: the mudsucker or sand goby (Gillichthys mirabilis), the sculpin (Oligocottus maculosus), and the sand dab (Citharichthys stigmaeus). In all three species, biotransformation took place rapidly in the liver. The uptake of naphthalene was greater than that of benzo [a]pyrene. The major metabolite of benzo [a]pyrene appeared to be 7,8-dihydroxy-7,8-dihydroxy benzo [a]pyrene, while the major metabolite of naphthalene was 1,2-dihydro-1,2-dihydroxy-naphthalene. The gall-bladder was the major storage site for the PAH and their metabolites. Naphthalene and its metabolites were removed at a higher rate than benzo [a]pyrene and its metabolites (Lee et al., 1972). Transformation of naphthalene and benzo [a]pyrene in the bluegill sunfish Lepomis macrochirus took place very rapidly, benzo [a]pyrene having the highest rate (McCarthy & Jimenez, 1985). L. macrochirus were exposed in a flow-through system to 4 nmol/litre benzo [a]pyrene for 48 h, followed by a 96-h depuration period, at 13 or 23°C in the presence or absence of food. Both polar and nonpolar metabolites were found. After 48 h, the polar metabolites comprised 10% of the benzo [a]pyrene metabolites in fed fish at 13°C, 20% in unfed fish at 23°C, and 30% in fed fish at 23°C (Jimenez et al., 1987). In rainbow trout (Oncorhynchus mykiss) exposed to naphthalene at 0.5 mg/litre for 24 h, the bile contained 65-70% metabolites, the liver contained 5-10%, and muscle < 1% (Melancon & Lech, 1978). In L. macrochirus exposed to 8.9 ± 2.1 µg/litre acenaphthene for 28 days, the half-life for metabolism was less than one day. No information was given on metabolites (Barrows et al., 1980). The depuration of anthracene was investigated in O. mykiss during simulated day and night cycles of 16 and 8 h, respectively. After a 96-h clearance period, the metabolites contributed 2-3% of the depurated substance, half of which came from the bile. No specific metabolites were reported (Linder & Bergman, 1984). After L. macrochirus had been exposed to anthracene at 8.9 µg/litre or benzo [a]pyrene at 0.98 µg/litre for 4 h, the rates of biotransformation were 0.26 and 0.082 nmol/g per h, respectively, and 8% of the anthracene and 88% of the benzo [a]pyrene were metabolized (Spacie et al., 1983). Benzo [a]pyrene is transformed in the Japanese medaka (Oryzias latipes) and the guppy (Poecilia reticulata), the main metabolite being the 7,8-diol-9,10-epoxide (Hawkins et al., 1988). Two benthic organisms, the European fingernail clam (Sphaerium corneum) and larvae of the midge Chironomus riparius, both metabolized benzo [a]pyrene. In the larvae, the main metabolite appeared to be 3-hydroxybenzo [a]pyrene; a quinone isomer was also found. Only a very small amount of 3-hydroxy-benzo [a]pyrene was found in the clam. No diol metabolites were found in either species (Borchert & Westendorf, 1994). After exposure of the benthic oligochaete Stylodrilus heringianus to either anthracene and pyrene or phenanthrene and benzo [a]pyrene, 2% degradation of each PAH was reported within 24 h (Frank et al., 1986). The half-lives for metabolism in D. magna were 0.5 h for 1.8 mg/litre naphthalene, 9 h for 0.06 mg/litre phenanthrene, and 18 h for 0.023 mg/litre chrysene (Eastmond et al., 1984). In amphipod Hyalella azteca was exposed to 0.043 nmol/ml anthracene for 8 h, the rates of biotransformation were 2.2 ± 0.5 nmol/g dry weight per h with no substratum, 3.0 ± 0.8 in the presence of washed sand from a local lake, and 1.0 ± 0.15 in the presence of sediment from the lake (Landrum & Scavia, 1983). The amphipod Rhepoxynius abronius metabolizes benzo [a]pyrene (Plesha et al., 1988). When two marine amphipods were exposed to a sediment containing 5.1 ng/mg of this compound, R. abronius metabolized 49% and Eohaustorius washingtonianus metabolized 27% of the benzo [a]pyrene after one day. The main metabolites appeared to be 7,8-dihydro-7,8-dihydroxy-benzo [a]pyrene, 9,10-dihydro-9,10-dihydroxybenzo [a]pyrene, 3-hydroxy-benzo [a]pyrene, and 9-hydroxybenzo [a]pyrene. The ratio of 7,8-dihydro-7,8-dihydroxybenzo [a]pyrene to 9,10-dihydro-9,10-dihydroxybenzo [a]pyrene in normal-phase high-performance liquid chromatography was 1.2 for R. abronius and 0.7 for E. washingtonianus (Reichert et al., 1985). No biotransformation of benzo [a]pyrene or phenanthrene was found in mayflies (Hexagenia limbata) or in the amphipod Pontoreia hoyi (Landrum & Poore, 1988). In a study of the route of metabolism of benzo [a]pyrene in green algae (Selenastrum capricornutum) exposed to 1.2 µg/litre for four days, with simulated day and night periods, the major dihydrodiol metabolites identified were the cis-4,5-diol (< 1%), the cis-7,8-diol (13%), the 9,10-diol (36%), and the cis-11,12-diol (50%), demonstrating the presence of a dioxygenase enzyme for this type of algae (Lindquist & Warshawsky, 1985), as suggested by Cody et al. (1984). Payne (1977) reported, however, that aryl hydrocarbon hydroxylase was not present in Fucus and Ascophyllum sp. of marine algae. Benzo [a]pyrene was not biotransformed in periphyton after 0.25 or 4 h. In cladocerans (D. magna) exposed to 1.0 µg/litre benzo [a]pyrene, the biotransformation rate after exposure for 6 h was 1.07 ± 0.20 nmol/g dry weight per h. In midge larvae (C. riparius) exposed to 0.6-1.5 µg/litre, the biotrans-formation rate was 3.6 ± 0.7 nmol/g dry weight per h after exposure for 1 h and 2.7 ± 0.3 after 4 h. In L. macrochirus exposed to 1.0 µg/litre, the biotransformation rate was 0.20 ± 0.03 nmol/g dry weight per h after 1 h and 0.37 ± 0.04 after 4 h. In chironomids, 3-hydroxybenzo [a]pyrene was the major metabolite after 8 h, representing 4.4% of the total water activity; smaller amounts of 7-hydroxy-benzo [a]pyrene and the 9,10- and 7,8-dihydroxydiols of benzo [a]pyrene were also found (Leversee et al., 1981). After exposure of benthic species to benzo [a]pyrene for one to four weeks, the following percentages of metabolites were found: E. washingtonianus, 22% in the whole body; R. abronius, 74% in the whole body; clams (Macoma nasuta), < 5% in the body and < 5 in the hepatopancreas; shrimp (Pandalus platyceros), 94% in the hepatopancreas; and the English sole (Parophrys vetulus), 94% in the body, 99% in the liver and > 99% in the bile (Varanasi et al., 1985). Mosquito larvae (C. pipens quinquefasciatus) were exposed for three days to 0.002 mg/litre benzo [a]pyrene in the presence or absence of the mixed-function oxidase inhibitor piperonyl butoxide at 0.0025 mg/litre. Parent benzo [a]pyrene represented 22% of the excreted PAH in the absence of piperonyl butoxide and 86% in its presence. After three days' exposure of snails (Physa sp.) to the same concentration of benzo [a]pyrene with or without piperonyl butoxide at 0.0025 mg/litre, parent benzo [a]pyrene represented 88% in the absence of the inhibitor and 85% in its presence. The authors suggested that snails are deficient in microsomal oxidases. In mosquito fish (G. affinis) exposed similarly, no parent benzo [a]pyrene was found in the absence of piperonyl butoxide but 21% in its presence (Lu et al., 1977). In an aquatic ecosystem, plankton, green algae (Oedogonium cardiacum), D. magna, mosquito larvae (C. pipiens quinquefasciatus), snails (Physa sp.), and mosquito fish (G. affinis) were exposed to 0.002 mg/litre benzo [a]pyrene for three days. Parent benzo [a]pyrene represented 83, 90, 46, 70, and 55% in the four organisms, respectively. The substance was metabolized to unidentified hydroxylated polar compounds. The finding of 55% parent benzo [a]pyrene in the fish was attributed to food-chain transfer, as none was found after direct exposure. A terrestrial-aquatic ecosystem was also exposed to benzo [a]pyrene by applying 0.2 mg of radiolabelled compound to Sorghum vulgare seedlings to simulate atmospheric fall-out and allowing them to be consumed by fourth-instar salt-marsh caterpillar larvae (E. acrea). Faecal products then entered the terrestrial and aquatic ecosystem described above, which was left for 33 days. The maximum radiolabel (0.005 ppm) was detected in the aquatic phase after 14 days. Unmetabolized benzo [a]pyrene accounted for 7.1% of the total extractable radiolabel in fish, 19% in snails, 32% in algae, and 34% in mosquitoes. Addition of the mixed-function oxidase inhibitor, piperonyl butoxide, resulted in 12% parent benzo [a]pyrene in fish, 34% in snails, 48% in the algae, and no change in mosquitoes (Lu et al., 1977). The biotransformation of 19 PAH was studied in the food chain seston (plankton) -> blue mussel (Mytilus edulis L.) -> common eider duck (Somateria mollissima L.) in the open, northern Baltic Sea. The concentrations of the PAH in the eider duck showed the distribution gallbladder > adipose tissue > liver. There was a high flux of the PAH in the food chain, but the concentration did not increase with increasing trophic level, indicating that the PAH were biotransformed rapidly. There was little biotransformation in the plankton. The distribution of the PAH in blue mussels was different from that in plankton, perhaps due to metabolic activity in the mussel. Biotransformation of PAH with a relative molecular mass of 252 was rapid in the ducks (Broman et al., 1990). In beans (Phaseolus vulgaris L.) exposed to 15 g anthracene per plant, uptake via the roots was rapid, 90% being metabolized within 30 days (Edwards, 1986). These investigations are summarized in Table 30. As the rate of metabolism depends not only on the species but also on factors such as temperature, pH, and other experimental conditions, the results are difficult to compare. Some general conclusions can, however, be drawn: - The biotransformation potential of aquatic organisms depends on the activity of cytochrome P450-dependent mixed-function oxidases, which are important for oxidation, the first step in the metabolism of xenobiotics such as PAH (James, 1989). - The tissues in which biotransformation mainly takes place are liver, lung, kidney, placenta, intestinal tract, and skin (Cerniglia, 1984). - The initial transformation step in invertebrates usually occurs more slowly than in vertebrates (James, 1989). Monoxygenation of PAH is faster in higher invertebrates like arthropods, echinoderms, and annelids and slowest in more primitive invertebrates like protozoa, profina, cnidaria, and molluscs (Neff, 1979). - In general, invertebrates excrete PAH metabolites inefficiently (James, 1989). - In higher organisms and algae, metabolites are usually produced by monooxygenase activity, resulting in the formation of epoxides, phenols, diols, tetrols, quinones, and conjugates. - It is not clear whether molluscs have cytochrome P450 activity (Moore et al., 1989). Table 30. Biotransformation of polycyclic aromatic hydrocarbons by various organisms Species Compound Biotransformation rate Reference Fungi Cunninghamella elegans Benzo[a]pyrene No information Cerniglia (1984) Algae Selenastrum capticornutum Benzo[a]pyrene Relatively fast Lindquist & Warshawsky (1985) Oedogenium cardiacum Benzo[a]pyrene 15% after 3 d in Lu et al. (1977) aquatic ecosystem Fucus sp. Various None Payne(1977) Ascophyllum sp. Various None Molluscs Sphaerium corneum Benzo[a]pyrene Very fast (no carcinogenic Borchert & Westendorf (1994) metabolites) Physa sp. Benzo[a]pyrene 12% after 3 d Lu et al. (1977) Mytilus edulis L. Different No information Broman et al. (1990) Crustaceae Hyalella azteca Anthracene 2.2 nmol/g dw/h in water Landrum & Scavia (1983) Hyalella azteca Anthracene 3.0 nmol/g dw/h 5 water/ Landrum & Scavia (1983) sediment Daphnia magna Benzo[a]pyrene 1.07 nmol/g dw/h after 6 h Leversee et al. (1981) Daphnia magna Benzo[a]pyrene 10% after 3 d in aquatic Lu et al. (1977) ecosystem Pontoporeia hoyi Benzo[a]pyrene None Landrum & Poore (1988) Pontoporeia hoyi Benzo[a]pyrene None after 48 h Evans & Landrum (1989) Mysis relicta Benzo[a]pyrene No information Evans & Landrum (1989) Rhepoxynius abronius Benzo[a]pyrene No information Plesha et al. (1988) Rhepoxynius abronius Benzo[a]pyrene 74% after 1-4 weeks Varanasi et al. (1985) Rhepoxynius abronius Benzo[a]pyrene 49% after 1 d Reichert et al. (1985) Eohaustorius washingtonianus Benzo[a]pyrene 27% after 1 d Reichert et al. (1985) Eohaustorius washingtonianus Benzo[a]pyrene 22% after 1-4 weeks Varanasi et al. (1985) Table 30. (continued) Species Compound Biotransformation rate Reference Pandalus platyceros Benzo[a]pyrene < 5% after 1-4 weeks Varanasi et al. (1985) Parophrys vetulus Benzo[a]pyrene 94% after 1-4 weeks Varanasi et al. (1985) Daphnia magna Chrysene 50% after 18 h Eastmond et al. (1984) Daphnia magna Naphthalene 50% after 0.5 h Eastmond et al. (1984) Daphnia magna Phenanthrene 50% after 9 h Eastmond et al. (1984) Fish Lepomis macrochirus Acenaphthene Half-life, < 1 d Barrows et al. (1980) Lepomis macrochirus Anthracene 8% after 4 h Spacie et al. (1983) Oncorhynchus mykiss Anthracene 2-3% after 24 h Linder & Bergman (1984) Gillichthys mirabilis Benzo[a]pyrene Rapid in liver Lee et al. (1972) Oligocottus maculosus Benzo[a]pyrene Rapid in liver Lee et al. (1972) Citharichthys stigmaeus Benzo[a]pyrene Rapid in liver Lee et al. (1972) Lepomis macrochirus Benzo[a]pyrene Very fast McCarthy & Jimenez (1981) Lapomis macrochirus Benzo[a]pyrene 88% after 4h Spacie et al. (1983) Lepomis macrochirus Benzo[a]pyrene 0.20-0.37 nmol/g dry Leversee et al. (1981) weight per h Oryzias latipes Benzo[a]pyrene No information Hawkins (1988) Poecilia reticulata Benzo[a]pyrene No information Hawkins (1988) Rhepoxynius abronius Benzo[a]pyrene None Plesha et al. (1988) Gambusia affinis Benzo[a]pyrene 100% after 3 d in water Lu et al. (1977) 40% after 3 d in aquatic ecosystem Gillichthys mirabilis Naphthalene Rapid in liver Lee et al. (1972) Oligocottus maculosus Naphthalene Rapid in liver Lee et al. (1972) Citharichthys stigmaeus Naphthalene Rapid in liver Lee et al. (1972) Lepomis macrochirus Naphthalene Very fast McCarthy & Jimenez (1981) Worm Stylodrilus heringianus Various None Franck et al. (1986) Table 30. (continued) Species Compound Biotransformation rate Reference Insects Chironomus riparius Benzo[a]pyrene Very fast (no carcinogenic Bochert & Westendorf (1994) metabolites) Chironomus riparius Benzo[a]pyrene 2.7-3.6 nmol/g dry weight Leversee et al. (1981) per h Hexagenia limbata Benzo[a]pyrene None Landrum & Poore (1983) Culex pipiens Benzo[a]pyrene 78% after 3 d Lu et al. (1977) quinquefasciatus Somatochlora cingulata Naphthalene No information Correa & Coler (1990) Bird Somateria mollissima L. Various Fast for PAH with Broman et al. (1990) molecular mass > 252 Plant Phaseolus vulgaris L. Anthracene 90% after 30 d Edwards (1986) - In crustaceans, biotransformation differs greatly between species and for different PAH. Biotransformation of naphthalene, anthracene, phenanthrene, and chrysene appears to occur rapidly, while that of benzo [a]pyrene is generally slower. Only Reichert et al. (1985) reported significant degradation in R. abronius (49%) and E. washingtonianus (27%) within one day. - It is not clear how rapidly biotransformation occurs in insects. - Too little information was available on algae, plants, and fungi for conclusions to be drawn. 4.2.2 Abiotic degradation Abiotic processes may account for the removal of 2-20% of two- and three-ring PAH from soil (Park et al., 1990). In soils partly amended with PAH-containing sewage sludge, 24-100% was removed, and naphthalene was eliminated almost completely by volatilization and photodegradation (Wild & Jones, 1993). 4.2.2.1 Photodegradation in the environment PAH can be expected to be photodegraded in air and water but to a very low extent in soils and sediments, owing to low light intensity. In natural waters, photodegradation takes place only in the upper few centimetres of the aqueous phase. Information on the photodegradation of PAH in air and water is summarized in Table 31; however, as the testing conditions varied widely, general conclusions cannot be drawn. PAH are photodegraded in air and water by two processes: direct photolysis by light with a wavelength < 290 nm and indirect photolysis by least one oxidizing agent such as OH, O3, and NO3 in air and ROO radicals in water. In general, indirect photolysis - photooxidation - is the more important process. The reaction rates of PAH with airborne OH radicals measured under standard conditions are given in Table 32, which shows that most of the calculated half-lives are one day or less. Under environmental conditions, PAH of higher molecular mass, i.e. those with more aromatic rings, are almost completely adsorbed onto fine particles (see section 4.1.2); this reduces the degradation rate markedly. Degradation half-lives of 3.7-30 days were reported for the reaction with NOx of various PAH adsorbed onto soot. The degradation was much slower in the absence of sunlight. PAH did not react significantly with SO2 (Butler & Crossley, 1981). PAH in wood smoke and gasoline exhaust did not degrade significantly during winter in extreme northern and southern latitudes owing to low temperatures and the low angle of the sun (Kamens et al., 1986a). In summer, however, at a temperature of 20°C, the half-lives of individual PAH were in the range of 30-60 min (Kamens et al., 1986b). The degradation rate increased further with increasing humidity (Kamens et al., 1991). Table 31. Photodegradation of polycyclic aromatic hydrocarbons Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Acenaphthene Air, particles Determined in rotary photoreactor Behymer & with 25 µg/g on: Hites (1985) 2.0 - silica gel 2.2 - alumina 44 - fly ash Water 0.23 h-1 3.0 Rate constant in distilled water Fukuda et al. (1988) Acenaphthylene Air, particles Determined in rotary photoreactor Behymer & with 25 µg/g on: Hites (1985) 0.7 - silica gel 2.2 - alumina 44 - fly ash Anthracene Air, water 0.58 Measured in atmosphere and water Southworth from aqueous photolysis rate (1979) constant for midday summer sunlight at 35°N Air, particles Determined with 25 µg/g on: Behymer & 2.9 - silica gel Hites (1985) 0.5 - alumina 48 - fly ash Water Removal rate constants from water Southworth at 25°C in midsummer sunlight: (1979) 0.004 h-1 173 - in deep, slow, somewhat turbid water <0.001 h-1 > 700 - in deep, slow, muddy water 0.018 h-1 38 - in deep, slow, clear water 0.086 h-1 8 - in shallow, fast, clear water 0.238 h-1 3 - in very shallow, fast, clear water Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Water Half-lives calclulated from average Southworth light intensity over 24 h: (1977) 1.6 - in summer 4.8 - in winter Water Half-lives calculated for direct Zepp & sunlight at 40°N at midday in Schlotzhauer midsummer: (1979) 0.75 - near surface water 108 - inland water 125 - inland water with sediment partitioning 0.75 - direct photochemical transformation near water surface Water 0.66 h-1 1.0 In distilled water Fukuda et al. (1988) Benz[a]anthracene Air, particles First-order daytime decay rate Kamens et al. constants with soot particle loading of: (1988) 0.0125 min-1 0.9 - 1000-2000 ng/mg 0.0250 min-1 0.5 - 30-350 ng/mg Air, particles Determined with ± 25 µg/g on: Behymer & 4.0 - silica gel Hites (1985) 2.0 - alumina 38 - fly ash Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Water Calculated rate constant in pure Mill et al. water: (1981) 13.4 × 10-5s-1 1.4 - at 366 nm and in sunlight at 23-28°C, early March 2.28 × 1O-5s-1 8.4 - at 313 nm with 1% acetonitrile in filter-sterilized natural water 5 Early March Benzo[a]pyrene Air, particles Determined with 25 µg/g on: Behymer & 4.7 silica gel Hites (1985) 1.4 - alumina 31 - fly ash Air particles First-order daytime decay rate Kamens et al. constants with soot particle loading of: (1988) 0.0090 min-1 1.3 - 1000-2000 ng/mg 0.0211 min-1 0.54 - 30-350 ng/mg Air, particles < 6.1 × 10-4 m/s Ozonization rate constant measured Cope & at 24°C with O3 = 0.16 ppm and Kalkwarf light intensity of 1.3 kW/m3 (1987) Air 0.37-1.1 Estimated Lyman et al. (1982) Air 1 Sunlight in mid-December Mill & Mabey (1985) Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Air, water Calculated rate constants for Mill et al. direct photolysis: (1981) 3.86 × 10-4s-1 0.69 - in pure water at 366 nm and in sunlight at 23-28°C, late January 1.05 × 10-5s-1 1.1 - at 313 nm with 1-20% acetonitrile in filter-sterilized natural water, mid-December Water Computed near-surface half-life for Zepp & direct photochemical transformation Schlotzhauer of a natural water body: (1979) 0.54 - latitude 40°N, midday, midsummer 77 - no sedimentmater partitioning 312 - sediment; water partitioning in a 5-m deep inland water body Air > 1 Summer Valerio et al. Days Winter (1991) Methanol 2 Irradiated at 254 nm Lu et al. (1977) Benzo[b]fluoranthene Air, particles First-order daytime decay rate Kamens et al. constants with soot particle loading of: (1988) 0.0065 min-1 1.8 - 1000-2000 ng/mg 0.0090 min-1 1.3 - 30-350 ng/mg Air, water 8.7-720 Based on measured rate of Lane & Katz photolysis in heptane irradiated with (1977); Muel light at > 290 nm & Saguem (1985) Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Benzo[ghi]perylene Air, particles Determined with 25 µg/g on: Behymer & 7.0 - silica gel Hites (1985) 2.2 - alumina 29 - fly ash Air, particles First-order daytime photodegradation Kamens et al. rate constants for adsorption (1988) on wood soot particles in an outdoor Teflon chamber for soot loading of: 0.0077 min-1 1.5 - 1000-2000 ng/mg 0.0116 min-1 1.0 - 30-350 ng/mg Benzo[k]fluoranthene Air, particles First-order daytime decay constants Kamens et al. for soot loading of: (1988) 0.0047 min-1 2.5 - 1000-2000 ng/mg 0.0013 min-1 8.9 - 30-350 ng/mg Air, water 3.8-499 Based on measured rate of photolysis Muel & in heptane under November Saguem sunlight, adjusted by ratio of (1985) sunlight photolysis half-lives in water: heptane Chrysene Air, particles Determined with 25 µg/g on: Behymer & 100 - silica gel Hites (1985) 78 - alumina 38 - fly ash Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Air, particles First-order daytime decay constants Kamens et al. for soot loading of: (1988) 0.0056 min-1 2.1 - 1000-2000 ng/mg 0.0090 min-1 1.3 - 30-350 ng/mg Air, water 4.4 Calculated for direct photochemical Zepp & transformation near surface of Schlotzhauer a water body at 40°N at midday in (1979) midsummer Water 13 Estimated on basis of photolysis Lyman et al. in water in winter (1982) Dibenzo[a,h]anthracene Air, water 782 Based on measured rate of photolysis Muel & in heptane in November sun Saguem 6 After adjusting ratio of sunlight (1985) photolysis in water: heptane Fluoranthene Air, particles Determined with 25 µg/g on: Behymer & 74 - silica gel Hites (1985) 23 - alumina 44 - fly ash Air, water 63 Computed, adjusted for approximate Lyman et al. winter sunlight intensity (1982) Air, water Calculated photochemical transformation Zepp & near surface of water body: Schlotzhauer 21 - at 40°N, midday, midsummer (1979) 3800 - 5-m deep inland water body with no sediment:water partitioning 4800 - with sediment:water partitioning Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Water 3800 Summer sunlight in surface water Mill & Mabey (1985) Fluorene Air, particles Determined in rotary photoreactor Behymer & with 25 µg/g on: Hites (1985) 110 - silica gel 62 - alumina 37 - fly ash Naphthalene Water 13 200 Calculated, 5-m deep inland water Zepp & Schlotzhauer (1979) Water 0.028 h-1 25 Half-life in distilled water Fukuda et al. (1988) Perylene Air, particles Determined with 25 µg/g on: Behymer & 3.9 - silica gel Hites (1985) 1.2 - alumina 35 - fly ash Air, glass < 4.7 × 10-5 m/s Ozonization rate constant measured Cope & from glass surface at 24°C with 03 Kalkwarf - 0.16 ppm and light intensity of (1987) 1.3 kW/m2 Phenanthrene Air, particles Determined with 25 µg/g on: Behymer & 150 - silica gel Hites (1985) 45 - alumina 49 - fly ash Water 3 Based on measured aqueous photolysis Zepp & quantum yields, midday, mid-summer, Schlotzhauer 40°N (1979) Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Air, water 25 Adjusted for approximate winter Lyman et al. sunlight intensity (1982) Air, water Calculated, direct sunlight photolysis, Zepp & midday, midsummer, 40°N: Schlotzhauer 8.4 - near surface water (1970) 1400 - 5-m deep inland water body with no sediment:water partitioning 1650 - with sedimentmater partitioning Water 0.11 h-1 6.3 Half-life in distilled water Fukuda et al. (1988) Pyrene Air, particles Determined with 25 µg/ml on: Behymer & Hites 21 - on silica gel (1985) 31 - on alumina 46 - on fly ash Air, particles Adsorption on airborne particles Valerio et al. by sunlight: (1991) 1 - in summer Days - in winter Air, water 1.014 h-1 0.68 Based on measured aqueous photolysis Zepp & quantum yields, midday, Schlotzhauer summer, 40°N (1979) Air, water 2.04 Based on measured aqueous photolysis Lyman et al. quantum yields, adjusted for (1982) approximate winter sunlight intensity Air, glass < 1.05 × 10-4 m/s Ozonization rate on glass surface Cope & at 24°C with O3 = 0.16 ppm and Kalkwarf light intensity of 1.3 kW/m2 (1987) Table 31. (continued) Compound Compartment Photolysis Half-life Comments Reference rate constant (h) Water Calculated, direct sunlight photolysis, Zepp & midday, midsummer, 40°N: Schlotzhauer 0.58 - near surface water (1979) 100 - 5-m deep inland water body with no sediment:water partitioning 142 - with sediment:water partitioning Water 100 Summer sunlight photolysis in Mill & Mabey surface water (1985) In order to compare numbers reported only as rate constants, half-lives were estimated from the formula: t1/2 = In2 k where t1/2 is the half-life and k is the rate constant. The calculated values are reported in italics. Table 32. Reactions of polycyclic aromatic hydrocarbons with hydroxy radicals Compound Oxidation rate Photooxidation Comments Reference constant half-life (h) Acenaphthene 1 × 10-10 0.879-8.79 Based on estimated reaction rate Atkinson (1987) constant with hydroxy radical in air Acenaphthylene 1.1 × 10-10 0.191-1.27 Based on estimated rate constant for Atkinson (1987) reaction in air Anthracene 1.1 × 10-12cm3 58-580 Rate constant for gas-phase reaction Biermann at al. molec-1s-1 with hydroxy radicals at 298 ± 1 K, based (1985) the relative rate technique for propane 0.501-5.01 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Benz[a]anthracene 0.801-8.01 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Benzo[a]pyrene 0.428-4.28 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Benzo[b]fluoranthene 1.43-14.3 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Benzo[ghi]perylene 0.321-3.21 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Benzo[k]fluoranthene 1.1-11 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Chrysene 0.802-8.02 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Dibenz[a,h]anthracene 0.428-4.28 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Fluoranthene 2.02-20.2 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Fluorene 1.3 × 10-11 6.81-68.1 Based on estimated rate constant for Atkinson (1987) reaction with hydroxy radical in air Table 32. (continued) Compound Oxidation rate Photooxidation Comments Reference constant half-life (h) Naphthalene 2.16 × 10-11 cm3 2.7-27 Rate constant for reaction with hydroxy Atkinson (1989) molec-1s-1 radicals using relative rate technique at 294 K 2 × 10-19 cm3 19-321 Upper limit was obtained for reaction molec-1s-1 with O3 2.35 × 10-11 cm3 2.7-27 Rate constant for gas-phase reaction Biermann et al. molec-1s-1 with hydroxy radicals at 298 K, based (1985) on relative rate technique from propene Phenanthrene 3.4 × 10-11 cm3 2-20 Rate constant for gas-phase reaction Biermann et al. molec-1s-1 with hydroxy radicals at 298 K, based (1985) on relative rate technique for propene 3.1 × 10-11 2.01-20.1 Half-life based on measured rate Atkinson (1987) constants for reaction with hydroxy radical in air Pyrene 0.802-8.02 h Based on estimated rate constant for Atkinson (1987); reactions with hydroxy radical in air and Atkinson & Carter with hydroxy radical and ozone (1984) To allow comparison when only rate constants are reported, half-lives were estimated from the following formula: t1/2 = In 2 [x] × k where t1/2 is the half-life, [x] is the concentration of the radical with which the compounds react (i.e. hydroxyl or ozone), and k is the rate constant. The calculated values are reported in italics. For the concentrations of the radicals, the following ranges of values were used; the lower values are estimates for rural areas and the higher ones for urban areas (Howard et al., 1991): [OH]air = 3-30 × 105 radicals/cm3 [O3]air = 3-50 × 1012 molecules/cm3 [OH]water = 5-200 × 10-17 mol/litre [RO2]water = 1-50 × 10-11 mol/litre [1O2]water = 1-100 × 10-15 mol/litre In a study of the fate of 18 PAH on 15 types of fly ash, carbon black, silica gel, and alumina, the PAH were stabilized, depending on the colour, which is related to the carbon content: the higher the carbon content, the more stable the PAH. The authors suggested that radiation energy is adsorbed by the organic matter of particulates, and PAH therefore do not achieve the excited state in which they can be degraded (Behymer & Hites, 1988). The half-lives for direct photolysis of various PAH adsorbed onto silica gel are in the range of hours (Vu-Duc & Huynh, 1991). A two-layer model has been proposed for the behaviour of naturally occurring PAH on airborne particulate matter, in which photooxidation takes place in the outer layer, and much slower, 'dark' oxidation takes place in the inner layer (Valerio et al., 1987). This model is in line with the results of Kamens et al. (1991), who reported that PAH on highly loaded particles degrade more slowly than those on particles with low loads. As PAH occur mainly on particulate matter with a high carbon content, their degradation in the atmosphere is slower than that of PAH in the vapour phase under laboratory conditions or adsorbed on synthetic materials like alumina and silica gel that have no or a low carbon content. Formation of nitro-PAH was found from the low-molecular-mass two- to four-ring PAH that occur in the atmosphere, predominantly in the vapour phase. The rate constants range from 5.5 × 10-12 cm3/molecule × s for acenaphthylene to 3.6 × 10-28 cm3/molecule × s for naphthalene, with corresponding half-lives ranging from 6 min to 1.5 years. The yields were 1% or less (Atkinson et al., 1991; Atkinson & Arey, 1994). The rate of degradation of absorbed individual PAH seems to be independent of their physicochemical characteristics but dependent on their molecular structure. Thus, activated carbon from graphite particles effectively stabilized pyrene, phenanthene, fluoranthene, anthracene, and benzo [a]pyrene adsorbed onto coal fly ash against photochemical decomposition, but no stabilization was seen for fluorene, benzo [a]fluorene, benzo [b]fluorene, 9,10-dimethyl-anthracene, or 4-azafluorene. The authors suggested that PAH that contain benzylic carbon atoms are less reactive than others (Hughes et al., 1980). PAH with vinylic bridges appear to degrade by direct photolysis more rapidly than those with only aromatic rings, both in air and in the aquatic environment (Hites, 1981). In measurements of the photodegradation of benz [a]anthracene and benzo [a]pyrene, addition of humic acids and purging of the solution with nitrogen reduced the reaction rates significantly (Mill et al., 1981). The authors concluded that light screening and quenching occurred with humic acids. The reduction in rate with exclusion of oxygen was probably due to a decrease in photooxidative processes. The first metabolites were mainly quinones. 4.2.2.2 Hydrolysis PAH are chemically stable, with no functional groups that result in hydrolysis. Under environmental conditions, therefore, hydrolysis does not contribute to the degradation of PAH (Howard et al., 1991). 4.3 Ultimate fate after use The main sinks for PAH are sediment and soil. The available information indicates that high-molecular-mass PAH are especially persistent in groundwater, soil, and sediment under environmental conditions. 5. ENVIRONMENTAL LEVELS AND HUMAN EXPOSURE Appraisal Polycyclic aromatic hydrocarbons (PAH) occur in all environmental compartments. Ambient air, residential heating, and vehicle traffic are the main sources. The levels of individual substances vary over several orders of magnitude but are generally in the range < 0.1-100 ng/m3. Surface waters are contaminated by PAH mainly through atmospheric deposition, urban runoff, and industrial activities such as coal coking and aluminium production. Apart from highly industrial polluted rivers, the concentrations of individual substances are generally < 50 ng/litre. High concentrations of PAH have been measured in rainwater and especially in snow and fog. The concentrations of PAH in sediments are in the low microgram per kilogram range. PAH levels in soils near industrial sources (e.g. coal coking) are especially high, sometimes up to grams per kilogram. In contrast, soils contaminated by atmospheric deposition or runoff have concentrations of 2-5 mg/kg of individual PAH, and the concentrations in unpolluted areas are in the low microgram per kilogram range. PAH have been detected in vegetables but are mainly formed during food processing, roasting, frying, or baking. The highest levels were detected in smoked meat and fish, at up to 200 µg/kg food for individual PAH. Five-fold increases in the concentrations of PAH in soil have been observed over a 150-year period, although there are indications that the concentrations of some PAH are decreasing. Similar findings have been reported for sediments, perhaps because of measures to reduce emissions. Aquatic animals are known to adsorb and accumulate PAH. Especially high concentrations were found in aquatic organisms from highly polluted rivers, at levels up to milligrams per kilogram. Of the terrestrial animals, earthworms are a good indicator of soil pollution with PAH. The benzo[a]pyrene concentrations in the faeces of earthworms living in a highly industrialized region were in the low milligrams per kilogram range. The main sources of exposure for the general population appear to be food and air. The estimated intake of individual PAH in the diet is 0.1-8 µg/d. The main contribution appears to be that of cereals and cereal products, due to the large amounts consumed. In ambient air, the main sources are residential heating and environmental tobacco smoke; exposure to PAH from environmental tobacco smoke in indoor air is estimated to be 6.4 µg/day. Occupational exposure to PAH occurs via the lung and skin. High exposure occurs during the processing and use of coal and mineral oil products, such as in coal coking, petroleum refining, road paving, asphalt roofing, and impregnation of wood with creosotes; high concentrations are also found in the air of aluminium production plants and steel and iron foundries. No measurements were available for the primary production and processing of PAH. 5.1 Environmental levels 5.1.1 Atmosphere Relevant data on the occurrence of PAH in ambient air are compiled in Tables 33-36. The concentrations were determined mainly by gas chromatography and high-performance liquid chromatography, usually with enrichment by filtration through a solid sorbent. The amount of particle-bound PAH is therefore given. In studies in which vapour-phase PAH were also sampled, the results for the vapour and particulate phases were combined (for reviews, see Grimmer, 1979; Ministry of Environment, 1979; Grimmer, 1983b; Lee & Schuetzle, 1983; Daisey et al., 1986; Baek et al., 1991; Menichini, 1992a). 5.1.1.1 Source identification Qualitative indications of different sources can be obtained by comparing the PAH profiles, i.e. the ratio between the total PAH concentration and that of a selected PAH, in air with those of samples representative of the emitting sources or by determining PAH that are emitted mainly from a specific source (Menichini, 1992a). Quantitative assignments are difficult to make, however, owing to the complexity of factors that affect the variability of PAH concentrations and profiles. Measurements were made at selected sources of PAH in the area of Chicago, USA, in 1990-92, in order to identify them: Five samples were taken 100 m directly downwind of a coke plant in an area that was not affected by steel-making facilities, four samples from diesel buses at a parking garage, three samples from petrol vehicles under warm-engine operating conditions at a public parking garage, five samples in heavily travelled tunnels during evening rush hours, and two samples from the roof directly downwind of the chimney of fireplaces burning seasoned oak. The authors give a source distribution pattern in percent related to the total mass of 20 PAH. Naphthalene made by far the largest contribution to petrol engine and coke oven emissions (55 and 89%, respectively). The three-ring compounds acenaphthylene, acenaphthene, fluorene, phenanthrene, anthracene, and retene were detected in large amounts in diesel motor emissions (56%) and in wood combustion exhausts (69%). The four-ring fluoranthene, pyrene, benz [a]anthracene, chrysene, and triphenylene and the five-ring cyclopenta [cd]pyrene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [e]pyrene, and dibenzo [ghi]perylene together contributed 28% to diesel engine emissions, 25% to petrol engine emissions, and 20% to wood combustion emissions (Khalili et al., 1995). The winter levels of PAH are higher than the summer levels (Gordon, 1976; Lahmann et al., 1984; Greenberg et al., 1985; Chakraborti et al., 1988; Catoggio et al., 1989), due to more intensive domestic heating and to meteorological (lower inversions during the winter) and physicochemical factors (temperature-dependent partition between gaseous and particulate phases). The ratios of benzo [a]pyrene:CO, in which CO was used as an 'inert' tracer of automotive emissions, in Los Angeles, USA, were higher at night (0.18-0.34) than in the day (0.12-0.14), and substantially more so during winter (0.14-0.34) than in summer (0.12-0.18), consistent with daytime loss of PAH by chemical degradation (Grosjean, 1983). In studies of sources of PAH at commercial, industrial, and urban sampling sites in Athens, Greece, the effects of wind velocity and thermal inversion were studied. There seemed to be no direct correlation between benzo [a]pyrene and lead levels, which would be expected if exhaust from cars run on leaded petrol were the preponderant source of PAH (linear regression coefficient, 0.32-0.38) (Viras et al., 1987). Differences in the composition of profiles of PAH from different sources can also be standardized by giving the concentrations relative to that of a specific PAH. For particle-bound PAH, benzo [e]pyrene has often been used as a reference compound, since it is photochemically stable and found mainly in the particulate phase (Baek et al., 1991). Cyclopenta [cd]pyrene is emitted particularly from petrol-fuelled automobiles (Grimmer et al., 1981c). Fluoranthene, pyrene, benzo [ghi]perylene, and coronene are also found in higher concentrations in condensates of vehicle exhausts (Baek et al., 1991). The contribution of vehicles and domestic heating has also been estimated as the ratio of indeno[1,2,3- cd]pyrene to benzo [ghi]-perylene concentrations. The ratio should be 0.37 for the PAH profile in traffic exhaust and 0.90 for domestic heating (Lahmann et al., 1984; Jaklin & Krenmayr, 1985). In a comparison of the PAH ratios determined in New Jersey, USA, with those reported in the literature for samples collected under similar conditions in street tunnels, the ratios coronene:benzo [a]pyrene and benzo [ghi]perylene:benzo [a]pyrene indicated that vehicle traffic was the major source of PAH during the summer (Harkov et al., 1984). Measurements in ambient air in North Rhine Westphalia, Germany, in 1990 indicated that coronene is the most characteristic PAH for automobile traffic. At a ratio of benzo [a]pyrene:coronene of < 3.5, vehicle traffic is the dominant PAH source, whereas emissions with ratios > 3.5 are influenced by other sources. The benzo [a]pyrene levels were 0.66-5.0 ng/m3, and those of coronene 0.57-2.5 ng/m3 (Pfeffer, 1994). In a study of the PAH concentrations during weekdays and weekends in South Kensington, London, United Kingdom, no distinct differences were observed in winter, but the average concentrations were 1.5-2.5 times higher during the week than during the weekends in summer. Likewise, the diurnal variations appeared to be less distinct during winter than summer (Baek et al., 1992). Measurements in streets with high traffic density in Stockholm, Sweden, showed that the concentration of PAH decreased by 25-50% during holidays in comparison with weekdays. Benzo [a]pyrene in street air was all particle-bound, while chrysene and lighter PAH occurred both on particles and in the vapour phase (Östman et al., 1991, 1992a,b). In a study of 15 PAH in the air of various areas in an industrial city in Germany with 700 000 inhabitants, the highest levels were detected in air affected by a coke plant, where benzo [a]pyrene was found at 1.4-400 ng/m3 and cyclopenta [cd]pyrene at none detected to 120 ng/m3. The concentrations measured in air affected by vehicle traffic were 11-110 ng/m3 benzo [a]pyrene and 0.1-440 ng/m3 cyclopenta [cd]pyrene. Within 4 km, the average concentration of 88 ng/m3 cyclopenta [cd]pyrene had dropped to 1.6 ng/m3. The levels were lower in areas where hand-stoked residential coal heating predominated (0.37 µg/m3 benzo [a]pyrene and none detected to 39 µg/m3 cyclopenta [cd]pyrene) and where oil heating predominated (0.2-66 ng/m3 and none detected to 15 ng/m3, respectively). The concentration of PAH was three to four times higher between 7:43 and 10:00 than between 10:00 and 15:46. Benzo [c]phenanthrene, cyclopenta [cd]pyrene, benzo [ghi]perylene, and coronene dominated the PAH in areas with heavy traffic, whereas chrysene, benzo [b]fluoranthene, and benzo [a]pyrene occurred at the highest concentrations in an area surrounding a coke plant (Grimmer et al., 1981c). The use of receptor-source apportionment modelling was examined, despite its limited applicability to reactive species, for the PAH profiles of emissions from a variety of sources (Daisey et al., 1986; Pistikopoulos et al., 1990). In one study, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, indeno[1,2,3- cd]pyrene, and coronene were measured in the ambient air of the centre of Paris, France. The concentrations of PAH varied from 42% in winter to 72% in summer for petrol-fuelled vehicles, from 25 to 40% for diesel-fuelled vehicles, and from about 30 to 2% for domestic heating. The winter-summer differences were due mainly to different emission patterns and not to changes in the rate of decay of PAH (Pistikopoulos et al., 1990). In another study, the contributions of PAH from five sources to ambient air were distinguished by use of fuzzy clustering analysis (Thrane & Wikström, 1984). The information on PAH levels in ambient air is discussed below according to possible source: background and rural, industrial emissions, and diffuse sources like automobile traffic and residential heating. Attribution of different studies to these sections was difficult because the sources of PAH emissions are often mixed. For example, Seifert et al. (1986) determined PAH in Dortmund 200 m from a coke plant; this study was deemed to relate to PAH levels resulting from industrial emissions. The concentrations of PAH attributable to mobile sources can be estimated by monitoring near areas with heavy traffic in the summer, but it is difficult to estimate the contribution of home heating, because in winter PAH in ambient air derive from both mobile sources and home heating. Furthermore, emissions from mobile sources may differ in winter from those in the summer because of meteorological and physicochemical factors (Greenberg et al., 1985; see also section 5.1.1.3). 5.1.1.2 Background and rural levels The levels in ambient air of rural areas are summarized in Table 33. Background levels were measured about 25 km from La Paz, Bolivia, at an altitude of 5200 m (Cautreels & van Cauwenberghe, 1977) and on the island of Mallorca, Spain, at an altitude of 1100 m (Simó et al., 1990). The concentrations were generally 0.01-0.1 ng/m3. The average values in rural areas are usually 0.1-1 ng/m3. Average concentrations of 0.34 and 0.27 ng/m3 benzo [a]pyrene were measured in two rural areas in Japan in 1989, with a maximum concentration of 1.1 ng/m3 (Okita et al., 1994). 5.1.1.3 Industrial sources PAH levels in ambient air resulting mainly from industrial emissions are summarized in Table 34. The average concentrations of individual PAH at ground level were 1-10 ng/m3. In general, aluminium smelters and industrial processes for the pyrolysis of coal, such as coking operations and steel mills, result in higher levels of PAH than most other point industrial sources. Furthermore, the levels of PAH are much higher downwind from major sources than upwind. The highest levels of individual PAH were measured near an aluminium smelter in Hoyanger, Norway, with maximum concentrations of 10-100 ng/m3. Phenanthrene was present at very high levels in ambient air contaminated by industrial emissions (Thrane, 1987). In Sundsvall, Sweden, near an aluminium production facility, 310 ng/m3 phenanthrene, 190 ng/m3 naphthalene, 120 ng/m3 pyrene, and 84 ng/m3 fluorene were detected (Thrane & Wikström, 1984). The concentration of benzo [a]pyrene in ambient air near an oil processing plant in Moscow was up to 13 ng/m3 (Khesina, 1994). Benzo [a]pyrene was detected at 15-120 ng/m3 and perylene at 3-37 ng/m3 at 39 measuring stations in the heavily polluted area of Upper Silesia, Poland. The maximum values were 950 ng/m3 for benzo [a]pyrene and 270 ng/m3 for perylene (Chorazy et al., 1994). Table 33. Polycyclic aromatic hydrocarbon concentrations (ng/m3) in ambient air of background and rural areas Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] Acenaphthene 0.32 6.3-23 Anthracene 0.004 0.05 0.03 < 0.05 1.2-3.9 ND-0.05 Anthanathrene 0.004-0.16 0.08 0.07 ND-0.2 ND-0.04 Benz[a]anthracene 0.005 0.12 0.4 0.40 0.07 1.8-3.2 0.16-0.39 Benzo[a]fluorene 0.8-3.3 Benzo[a]pyrene 0.006 0.005 0.002-0.12 0.33/0.47 0.6 ND-0.52 0.45 0.08 0.8-2.5 0.41-0.45 Benzo[b]fluoranthene 0.02 1.2 0.45-0.58 Benzo[b]fluorene 0.24 0.5-2.4 Benzo[c]phenanthrene 0.15-0.20 Benzo[e]pyrene 0.022 0.006 0.007-0.26 0.6 0.59 1.8-5.8 0.44-0.65 Benzo[ghi]fluoranthene ND-0.2 Benzo[ghi]perylene 0.009 0.002 0.005-0.40 0.6 ND-0.58 1.4-3.0 0.89-1.4 Benzo[k]fluoranthene 0.02 0.002-0.088 0.48 0.17-0.25 Chrysene 0.07a 1.0 0.13-0.19 Coronene 0.005-0.23 0.24 ND-0.22 0.4-0.9 0.16-0.26 Cyclopenta[cd]pyrene 0.2 0.16-0.39 Dibenzo[a,h]pyrene 0.14 0.02-0.07 Dibenzo[a,l]pyrene 0.53 Fluoranthene 0.041 0.030 0.18 0.20/0.26 1.2 ND 0.93 1.3 11-47 0.19-0.23 Fluorene 0.45 0.66 14-32 Indeno[1,2,3-cd]pyrene 0.006 0.02 0.7 0.72 0.43-0.65 1-Methylphenanthrene 0.09 0.7-2.8 Naphthalene ND 3.0-98 Perylene 0.001-0.026 0.09 0.08 ND-0.4 Phenanthrene 0.026 2.66 0.4 ND-0.43 4.2 26-70 ND-0.03 Pyrene 0.034 0.024 0.34 0.010-0.15 0.15/0.15 1.3 ND 0.60 0.73 8.8-26 0.16-0.26 Table 33 (continued) ND, not detected; /, single measurements; [1] About 25 km from La Paz, Bolivia, at 5200 m (Cautreels & van Cauwenberghe, 1977); [2] Mallorca, Spain, 1989 (Simo et al.,1991); [3] Lake Superior, USA, 1986; sum of vapour and particulate phases (Baker & Eisenreich,1990); [4] Latrobe Valley, Australia, (Lyall at al.,1988); [5] Belgium, (Van Vaeck et al.,1980); [6] Denmark (Nielsen, 1984); [7] Western Germany, 1981 (Pflock et al.,1983); [8] Oostvoorne, Netherlands, (De Raat et al.,1987b); [9] Canada, 1989-91 (Environment Canada, 1994); [10] Sidsjon, Sweden, 1980-81, sum of vapour and particulate phases (Thrane & Wikstrom, 1984); [11] Folkestone, Ashford, United Kingdom, 1986 (Baek et al., 1992) a With triphenylene Analysed by high-performance liquid chromatography or gas chromatography; only particulates sampled, unless otherwise stated Table 34. Polycyclic aromatic hydrocarbon concentrations (ng/m3) in ambient air near industrial emissions Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] Acenaphthene 23 9.8-372 15-122 3.7 Acenaphthylene 747 0.01 Anthracene 2.9/3.4 158 4.5-6.1 4.1-43 0.12/0.15 0.01-3.4 0.08-0.19 Anthanthrene 0.001/3.0 0.2/1.1 ND-3.0 0.15/0.15 0.13-0.22 Benz[a]anthracene 0.28/1.2 7.6 2.0-158 2.5-58 0.8/3.1 0.02-1.2 1.3-4.7 Benzo[a]fluorene 1.1-179 Benzo[a]pyrene 0.002/1.5 0.5/3.5 25/37 6.3-6.7 5.3 1.1-61 2.1-36 0.14/0.11 0.20-0.11 1.8-3.1 1.1-2.6 Benzo[b]fluoranthene 0.9/1.8 4.8 2.7-6.4 Benzo[b]fluorene 0.7-122 0.61-1.4 Benzo[e]pyrene 0.004/1.4 1.8/3.2 11.6 2.5-86 1.3-3.1 Benzo[ghi]fluoranthene ND-0.5 0.26/0.35 Benzo[ghi]perylene 0.003/1.5 4.2/7.1 O.7 2.2-45 0.35/0.33 0.25 Benzo[j]fluoranthene 0.3/0.8 Benzo[k]fluoranthene 0.001/0.67 0.3/1.3 8.0 1.0-2.2 Chrysene 1.6/3.8 14.7 0.22/0.29 0.01-1.6 2.5-7.5 Coronene 0.003/1.5 3.2/2.8 1.3-1.5 ND 0.6-9.0 0.25/0.26 Cyclopenta[cd]pyrene 2.2 Dibenzo[a,h]pyrene ND 277 Dibenzo[a,l]pyrene 1.0-1.5 Fluoranthene 0.8/3.4 88.3 20-812 22-272 0.12/0.20 0.02-10 2.3-3.3 Fluorene 502 27-419 16-46 0.02-0.86 Indeno[1,2,3-cd]pyrene 0.4/0.3 1.1 3.8-38 0.28/0.27 0.10-7.7 1.4-2.4 1-Methylphenanthrene 2.5-58 Naphthalene 22 400 9.0-193 3.1-26 0.03-0.06 Perylene 0.001/0.2 0.3/1.2 0.1-8.3 0.05/0.05 22 0.23-0.61 Phenanthrene 500 54-1760 58-390 0.11/0.16 0.02-152 Pyrene 1.4/3.8 56.3 16-491 14-207 0.17/0.35 0.006-28 1.6-2.1 Table 34 (continued) ND, not detected; /, single measurements; [1] Three sampling sites near various industries in Latrobe Valley, Australia (Lyall et al., 1988); [2] Near various industries, USA, 1971-72 (Gordon & Bryan, 1973); [3] Near a coke plant, Dortmund, Germany, 1982-83 (Seifert et al., 1986); [4] Near a coke plant, Dortmund, Germany, 1989 (Buck, 1991); [5] 100 m directly downwind of a coke plant, Chicago, USA, 1990-92 (Khalili et al., 1995); [6] Near aluminium smelters, Norway and Sweden, 1980-82 (analytical method not given) (Thrane, 1987); vapour and particulate phase (Thrane & Wikstrom, 1984); [7] Near aluminium smelter, Canada, 1989-91 (Environment Canada, 1994); [8] Near incineration plant, Sweden (Colmsjo et al., 1986a,b); [9] Near refinery, USA, 1981-83 (Karlesky et al., 1987); [10] Brown coal industry area, western Germany, 1983 (Seifert et al., 1986); [11] Near harbours, Netherlands (De Raat et al., 1987b) Analysed by high-performance liquid chromatography or gas chromatography; only particulates sampled, unless otherwise stated In Ontario, Canada, up to 140 ng/m3 benzo [k]fluoranthene, 110 ng/m3 perylene, 110 ng/m3 benzo [a]pyrene, 90 ng/m3 benzo [ghi]perylene, and 43 ng/m3 fluoranthene were found near a steel mill (Potvin et al., 1980). The benzo [a]pyrene concentrations near coke ovens in urban areas of the USA were more than double those in urban areas without coke ovens (Faoro & Manning, 1981). These results are consistent with those of Grimmer et al. (1981c), who detected maximum levels of benzo [a]pyrene, chrysene, benzo [b]fluoranthene, benzo [j]fluoranthene, and benzo [k]fluoranthene in the area surrounding a coke plant. The PAH concentrations in ambient air 900 and 2500 m from a municipal incineration plant were of the same order of magnitude, and no significant contribution from the plant to the ambient PAH concentrations was observed (Colmsjö et al., 1986a). The PAH levels in an industrial area of Ahmedabad City, India, were significantly higher than those in a residential area. The highest levels were found during winter, and the rate of degradation of airborne PAH was predicted to be lowest in the monsoon season. The most striking finding was the high concentration of dibenz [a,h]anthracene in urban air (5.3-23 ng/m3) (Raiyani et al., 1993a). The limited resolution of PAH may have resulted in overestimation: for instance, the concentrations of benzo [ghi]perylene and indeno[1,2,3- cd]pyrene reported are one order of magnitude higher than that of dibenz [a,h]anthracene. 5.1.1.4 Diffuse sources A special situation of local importance was the pollution of ambient air in Kuwait after the war in the Persian Gulf, due to burning of oil fields. The mean concentrations of benzo [a]pyrene at three sampling sites were 0.27-9.2 ng/m3, and the maximum was 26 ng/m3 (Okita et al., 1994). These values are within the range of those detected in urban areas (see below). (a) Motor vehicle traffic The concentrations of PAH in the ambient air of various urban areas are listed in Table 35. The average levels of individual PAH were 1-30 ng/m3. Relatively high concentrations of benzo [a]pyrene, benzo [ghi]perylene, phenanthrene, fluoranthene, and pyrene were measured. Total PAH concentrations of 43-640 ng/m3 were measured in London, United Kingdom, in 1991, nearly 80% of which consisted of phenanthrene, fluorene, and fluoranthene; benzo [a]pyrene and benz [a]anthracene were present at 1% or less (Clayton et al., 1992). Table 35. Polycyclic aromatic hydrocarbon concentrations (ng/m3) in ambient air of urban areas Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] Acenaphthene 0.4-101 2.7-6 Acenaphthlene 0.9-39 4.4-130 Anthracene 34 0.6-36 0.3-2.1 3.5-25 Anthanthrene 2.5 0.1-4.7 30 < 0.1-0.6 0.003-0.76 Benz[a]anthracene 10 0.3-27 1.2-13 0.2-1.4 0.10-25 0.3-7.6 Benzo[a]fluorene 0.1-0.9 0.8-6.9 Benzo[a]pyrene 9.3 0.3-20 29 1.2-11 < 0.1-1.9 0.074-15 0.2-5.7 Benzo[b]fluoranthene 43 1.0-36 Benzo[b]fluorene 0.1-0.8 0.6-7.3 Benzo[c]phenathrene 4.0 0.2-5.0 Benzo[e]pyrene 8.4 0.4-17 16 1.7-15 < 0.1-1.2 0.40-27 0.4-6.5 Benzo[ghi]fluoranthene 12 0.3-5.0 0.1-1.5 0.5-7 Benzo[ghi]perylene 14 0.5-12 1.6/13 27 2.1-11 0.2-3.5 0.45-31 0.9-2.4 0.6-18 Benzo[j]fluoranthene 0.17-13 Benzo[k]fluoranthene 23 0.29-25 Chrysene 0.3-2.5 0.56-29 3.6-5.6 3.3 Coronene 10 0.3-5.5 12 0.88-2.0 0.1-2.4 0.22-3.3 0.4-19 Cyclopenta[cd]pyrene 11 0.1-4.8 71 < 0.1-1.1 0.1-6 Dibenzo[a,h]pyrene 0.22-3.4 0.29-2.8 Fluoranthene 72 6.2-108 0.40/14 1.4-10 0.80-14 1.3-2.0 6.9-38 15 Fluorene 1.3-61 16-86 Indeno[1,2,3-cd]pyrene 8.6 0.4-12 31 < 0.1-2.9 0.39-30 0.4-7.6 1-Methylphenanthrene 0.3-2.5 5-16 Naphthalene 14-63 Perylene 2.3 0.1-4.3 4.8 < 0.1-0.4 0.011-4.4 0.1-1.3 Phenanthrene 153 18-223 3.6-41 32-105 111 Pyrene 74 2.9-67 0.34/12 1.2-5.5 0.34-10 5.5-45 20 Triphenylene 0.15-6.9 Table 35 (continued) ND, not detected; /, single measurements; [1] Vienna, Austria, 1983-84; vapour and particulate phase (Jaklin & Krenmayr, 1985); [2] Linz, Austria, 1985; vapour and particulate phase (Jaklin et al., 1988); [3] Antwerp, Belgium (Van Vaeck et al., 1980); [4] Berlin, western Gemany, 1984-85 (Seifert et al., 1986); [5] Rhein/Ruhr area, western Germany, 1985-88; analytical method not stated (Buck et al., 1989); [6] Kokkola, Finland (Pyysalo et al., 1987); [7] St Denis, France, 1979-80 (Muel & Saguem, 1985); [8] Various cities, Greece, 1984-85 (Viras et al., 1987); [9] Oslo, Norway, 1981-83, vapour and particulate phase (Larssen, 1985); [10] Barcelona, Spain, 1988-89, vapour and particulate phase (Albaiges et al., 1991) Analysed by high-performance liquid chromatography or gas chromatography; only particulates sampled, unless otherwise voted Table 35 (continued) Compound [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] Acenaphthene 0.07-3.58 0.05-31.1 Acenaphthylene 9.1 0.8 0.9 Anthracene 21 1.4 2.8 0.01-8.28 0.20-39.8 0.1-0.9 ND-4.8 6.1/11 Anthanthrene 0.63 Benz[a]anthracene 4.1 0.4 1.4 0.24-10.6 0.12-18.5 0.2-5.8 5-21 0.07-2.1 Benzo[a]fluorene 5.0 0.7 Benzo[a]pyrene 2.9 0.2 0.99/1.4 1.6 0.01-7.02 0.18-13.7 0.3-3.4 1-17 0.04-3.2 0.6/1.6 Benzo[b]fluoranthene 1.8 0.01-3.04 0.13-14.8 0.2-3.7 5-30 0.10-3.7 Benzo[c]phenanthrene 2.8 Benzo[e]pyrene 3.5 0.4 1.1/2.0 2.3 2.1/2.1 Benzo[ghi]fluoranthene 7.3 0.8 Benzo[ghi]perylene 6.6 0.5 2.9/3.3 3.3 0.02-6.90 0.15-85.3 Benzo[k]fluoranthene 0.75 0.23-16.5 0.3-0.8 3-22 0.07-0.85 Chrysene 5.1 0.8 1.6 0.04-4.97 0.13-24.3 0.2-5.5 ND-2.3 Coronene 4.1 0.3 2.4/1.7 1.7 0.02-3.72 0.17-6.92 ND-16 Cyclopenta[cd]pyrene 3.9 0.11 4.1 Dibenz[a,h]pyrene 0.12 Fluoranthene 24 3.9 3.5 2.03-62.4 22-23 14-54 0.24-2.0 8.0/9.7 Fluorene 0.07-27.6 0.07-161 Indeno[1,2,3-cd]pyrene 3.8 0.5 1.6 0.3-4.4 4.24 Naphthalene 15/75 Perylene 1.0 0.1 0.2/0.5 Phenanthrene 76 11 5.1 0.06-111 2.25-492 0.1-2.4 78/81 Pyrene 28 32 18 0.39-17.4 0.33-64.4 0.1-7.5 0.48-3.6 8.0/12 Triphenylene Table 35 (continued) ND, not detected;/, single measurements; [11] Stockholm, Sweden, April 1991; vapour and particulate phases (Ostman et al.,1992a,b); [12] Stockholm, Sweden; 1992 vapour and particulate phases (Ostman et al.,1992a,b); [13] London, United Kingdom, 1985-87(Baek et al.,1992); [14] London, United Kingdom, 1987; vapour and particulate phases (Baek et al.,1992); [15] Manchester, United Kingdom, 1990-91; vapour and particulate phases (Clayton et al.,1992); [16] Various cities, United Kingdom, 1991-92; vapour and particulate phases (Halsall et al.,1994); [17] Lake Baikal shore, Russian Federation, 1993-94 (Grachev et al.,1994); [18] Zagreb, Croatia, 1977-82; determined by thin-layer chromatography and fluorescence detector (Bozicevic et al.,1987); [19] Los Angeles, USA, 1981-82 (Grosjean, 1983); [20] Los Angeles basin, USA, 1986; vapour and particulate phases (Arey et al.,1987) Table 35 (contd) Compound [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] Acenaphthene 3.3-9.0 0.06-5.2 0.6 Acenaphthylene < 11-47 1.9 Anthracene 1.9-4.5 0.45-3.8 0.17-0.57 0.12-0.52 0.2 2.5-5.5 Anthanthrene 0.006-3.3 1-11 Benz[a]anthracene 0.07-1.4 0.19-0.40 0.19-4.4 0.99-7.0 0.37-1.7 1.9 20-66 Benzo[a]fluorene 1.8-6.3 Benzo[a]pyrene 0.11-1.6 ND-0.03 0.09-1.7 0.006-1.8 8-38 1.6-8.4 ND-2.3 3.4 30-120 Benzo[b]fluoranthene 0.17-1.7 3.1-12 3.0 109-200 Benzo[b]fluorene 0.19-0.94 Benzo[e]pyrene 0.03-11 ND-0.04 0.016-2.3 4-19 2.7-9.0 2.3 49-182 Benzo[ghi]fluoranthene 0.12-1.3 Benzo[ghi]perylene 0.24-2.7 0.027-4.7 11-33 3.2-12 3.4 34-141 Benzo[j]fluoranthene 0.08-1.1 22-66 Benzo[k]fluoranthene 0.09-0.97 0.005-0.85 1.8-7.7 2.7 Chrysene 0.22-5.3 0.38-0.57 3-15 0.29-1.4 2.4 Coronene 0.14-1.6 0.020-2.3 5.16 Dibenzo[a,a]pyrene 0.06-2.7 Dibenzo[a,h]pyrene 0.46-1.2 5.3-23 Dibenzo[a,l]pyrene 0.05-0.35 Fluoranthene 5.7-10 1.6-11 14-79 1.5-8.3 1.0 11-26 Fluorene 7.4-14 0.94-5.5 0.08-0.15 0.31-1.2 2.8 Indeno[1,2,3-cd]pyrene 0.20-2.9 6-24 2.6-12 3.1 Naphthalene 280-940 ND 4.5-13 Perylene 0.01-0.15 0.001-0.24 2-9 0.51-1.2 Phenanthrene 21-35 2.2-35 0.79-2.6 0.52-2.4 0.7 12-21 Pyrene 0.12-2.8 4.8-10 1.4-6.9 0.008-0.66 16-69 1.5-9.0 0.46-4.0 3.8 20-44 Triphenylene 22-60 Table 35 (continued) ND, not detected; /, single nwasureme4s; [21] New Jersey, USA, 1981-82 (Greenberg et al, 1985); [22] Portland, Oregon, USA, 1984 (Ligocki et al.,1985); [23] Urban area (not specified), Canada, 1989-91 (Environment Canada,1994); [24] Latrobe Valley, Australia (Lyall et al., 1988); [25] Christchurch, New Zealand, 1979 (Cretney et al., 1985); [26] Osaka, Japan, 1977-78; vapour and particulate phases (Yamasaki et al., 1982); [27] Osaka, Japan, 1981-82 (Matsumoto & Kashimoto, 1985); [28] La Plata, Argentina, 1985 (Catoggio et al., 1989); [29] Ahmedabad City, India, 1984-85 (Raiyani at al.,1993a); [30] Calcutta, India, 1984 (Chakraborti et al.,1988) Table 35 (continued) Compound [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] Acenaphthene 4.5 Anthraceene 14-16 2.5 1.8 ND-34 8.7-23 Anthanthrene 0.15-0.63 0.001-0.21 2-24 Benz[a]anthracene 2.9-4.8 99-139 23 6.5 0.028-4.8 3.1-9.8 Benzo[alpyrene 3.8-5.5 0.005-1.3 67-73 15 5.6 0.023-4.6 Trace-9.3 ND-44 1.9-7.7 19-72 Benzo[blfluoranthene 1.0-3.1 130-133 0.46-16 Benzo[b]fluorene 0.07-0.18 Benzo[c]phenanthrene 33-37 Benzo[e]pyrene 5.5-7.4 0.016-3.3 96 19 9.1 0.18-8.8 0.17-4.2 ND-370 9-41 Benzo[ghi]fluoranthene 3.0-4.9 0.024-0.98 30-33 Benzo[ghi]perylene 7.0-13 0.004-3.2 49-61 12 7.9 0.21-12 ND-74 11-49 Benzo[j]fluoranthene 2.6-5.5 Benzo[k]fluoranthene 3.4-5.0 0.12-7.4 Chrysene 4.3-6.5 0.34-0.49 237-261 43 16 0.22-8.9 0.22-6.4 ND-170 7-71 Coronene 0.002-1.4 14-16 3.1 2.8 0.14-2.1 Trace-2.1 8-96 4-18 Cyclopenta[cd]pyrene ND 3.1 1.6 Dibenzo[a,h]pyrene 0.012-0.98 Fluoranthene 3.4-4.9 0.14-1.2 0.32-8.6 8-520 15-51 Fluorene 15-26 Indeno[1,2,3-cd]pyrene 5.1-9.1 0.022-2.0 57 11 5.5 0.16-9.6 9-43 Naphthalene 44 Perylene 0.01-0.20 7.6-10 0.004-0.88 ND-28 3-21 Phenanthrene 0.002-1.1 4-170 50-271 Pyrene 3.6-6.6 0.002-0.58 0.13-6.7 0.21-8.6 ND-540 12-49 Triphenylene 1.4-1.9 0.07-0.24 0.11-2.9 ND-50 ND, not detected; /, single measurements; [31] Various cities, China (Chen et al.,1981); [32] Various cities, China, 1986-88; determined by thin-layer chromatography and gas chomatography-mass spectroscopy (Chang et al., 1988; Simoneit et al., 1991); [33] Various locations with predominantly coal heating; Germany (analytical method not given) (Grimmer, 1980); [34] Essen, Germany, predominantly coal heating, 1978-79 (Buck, 1983); [35] Essen, Germany, predominantly oil heating, 1978-79 (Buck, 1983); [36] Antony, France, 1979-80 (Muel & Saguem, 1985); [37] Sutton Coldfield, United Kingdom, 1976-78 (Butler & Crossley, 1982); [38] Barrow, USA, fossil fuel combustion area, 1979 (Daisey et al., 1981); [39] Wood-heating area, Canada, 1989-91 (Environment Canada, 1994); [40] Christchurch, New Zealand, 1979 (Cretney et al., 1985) In Delft, the Netherlands, benzo [a]pyrene levels of up to 140 ng/m3 were measured on a foggy day with low wind velocity near a major road. High concentrations of pyrene (220 ng/m3), benzo [ghi]perylene (130 ng/m3), and coronene (21 ng/m3) were also found. At border crossings between the Netherlands and Germany on days with heavy traffic, the maximum levels of individual PAH were 1-54 ng/m3 (Brasser, 1980). PAH concentrations were determined in the centre of Paris, France, at the top of a 55-m tower and thus less likely than ground-level samples to be affected by traffic emissions and street dust; they can therefore be considered to be homogeneous and representative. The maximum levels found were 98 ng/m3 benzo [ghi]perylene, 60 ng/m3 indeno[1,2,3- cd]pyrene, 34 ng/m3 coronene, 28 ng/m3 benzo [b]fluoranthene, 13 ng/m3 benzo [a]pyrene, and 13 ng/m3 benzo [k]fluoranthene (Pistikopoulos et al., 1990). The average concentration of individual PAH in particulate and vapour phases during a nine-day photochemical pollution episode in California, USA, in 1986 was 1 ng/m3. The maximum levels of acenaphthene, acenaphthylene, fluorene, and phenanthrene ranged from 30 to 64 ng/m3 (Arey et al., 1991). In 1989, the average benzo [a]pyrene concentrations in five Japanese cities (Sapporo, Tokyo, Kawasaki, Nagoya, and Osaka) were 1.2-3.1 ng/m3. A maximum level of 15 ng/m3 was detected in Tokyo (Okita et al., 1994). A detailed examination was undertaken of the molecular composition of PAH in street-dust samples collected from the Tokyo metropolitan area. Unsubstituted ring systems (i.e. parent PAH) ranging from phenanthrene with three rings to benzo [ghi]perylene with six rings were the primary components, three- and four-ring PAH (i.e. phenanthrene, fluoranthene, and pyrene) predominating. The concentrations of total PAH were of the order of a few micrograms per gram of dust. On the basis of the PAH profile, it was suggested that PAH in the dust of busy streets arose mainly from automobile exhausts, while residential areas received a greater contribution from stationary sources. In both types of dust, asphalt was thought to contribute to only a minor extent (Takada et al., 1990). Giger & Schaffner (1978) had come to the same conclusion some 20 years earlier. Benzo [a]pyrene was detected in ambient air in Moscow, Russian Federation, at concentrations of 5.4 ng/m3 at a regular traffic site and 20 ng/m3 at a crossroads with heavy traffic (Khesina, 1994). (b) Road tunnels In road tunnels, the concentrations of individual PAH were usually 1-50 ng/m3 (Table 36). Higher levels were reported in tunnels in western Germany, with concentrations of 84 and 96 ng/m3 cyclopenta [cd]pyrene (Buck (1983) and 76 ng/m3 (Brasser, 1980) and 110 ng/m3 pyrene (Benner et al., 1989). Table 36. Polycyclic aromatic hydrocarbon concentrations (ng/m3) in ambient air polluted predominantly by vehicle exhaust Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] Acenaphthene 168 Acenaphthylene 32 445 Anthracene 8.6/9.8 2.3 55 0.6-12 177 Anthanthrene 7.2 1 500 0.1-4.5 2-82 Benz[a]anthracene 37/44 0.6-1.9 16 20 12 000 102 1.9-2.9 90.2 Benzo[a]fluorene 18 2 800 Benzo[a]pyrene 30 2-14 0.2-0.8 16 12 9 600 66 1.3-26 62.6 0.1-14 1-57 Benzo[b]fluoranthene 2.3 8.8 12 000 43.6 Benzo[e]pyrene 28/32 11 9 600 69 1.5-19 55.5 01-12 3-43 Benzo[ghi]fluoranthene 29 18 3.2-26 Benzo[ghi]perylene 40/47 4-16 0.4-2.6 44 30 19 000 85 1.8-18 17.0 0.6-27 20-213 Benzo[k]fluoranthene 8.1 9.7 9 000 41.2 Chrysene 54/58 25 15 9 500 77.9 Coronene 26/27 2-17 0.3-1.1 29 20 7 500 1.0-10 ND 0.3-14 9-156 Cyclopenta[cd]pyrene 84/96 40 31 7.6-65 100 Dibenzo[a,h]pyrene 14.7 Fluoranthene 35 83 93 6.4-69 117 Fluorene 406 Indeno[1,2,3-cd]pyrene 18/22 0.3-1.3 16 13 9 400 0.3-15 20.0 6-70 1-Methylphenanthrene 2.6-43 Naphthalene 8030 Perylene 3.4 3.1 1 500 1-18 Phenanthrene 8.1 243 4.4-56 300 Pyrene 33-114 47 122 16 000 120 9.7-76 193 0.2-29 Table 36 (continued) ND, not detected; /, single measurements; [1] Street tunnel (location not specified), western Germany, 1978-79 (Buck, 1983); [2] Coen Tunnel, Netherlands (Brasser, 1980); [3] Street tunnel in Lincoln, Netherlands, 1981 (Kebbekus et al., 1983), [4] Klara Tunnel, Sweden, 1983 (Colmsjo et al., 1986b); [5] Soderleds Tunnel, Sweden, 1991; vapour and particulate phases (Ostman et al., 1991); [6] Craeybeckx Highway Tunnel, Belgium, 1991 (De Fré et al., 1994); [7] Baltimore Harbor Tunnel, USA, 1975 (Fox & Staley, 1976); [8] Baltimore Harbor Tunnel, USA, 1985-86 (Benner et al., 1989); [9] Heavily travelled tunnel, Chicago area, USA, 1990-92 (Khalili et al., 1995); [10] Diesel bus garage, United Kingdom, 1979 (Waller et al., 1985); [11] Inside car park, New Zealand (Cretney et al., 1985) Analysed by high-performance liquid chromatography or gas chromatography; only particulates sampled, unless otherwise stated PAH were found at levels of up to 4 ng/m3 in an underground bus terminal in Stockholm, Sweden; and 21 ng/m3 fluoranthene, 11 ng/m3 pyrene, and 8.1 ng/m3 phenanthrene were found in a subway station (Colmsjö et al., 1986b). Very high concentrations of PAH were found in the air of the Craeybeckx Highway Tunnel in Belgium, which was used daily by an average of 45 000 vehicles, of which 60% were petrol-fuelled passenger cars, 20% diesel-fuelled cars, and 20% trucks. Of the cars, only 3% had three-way catalysts (De Fré et al., 1994). (c) Residential heating The PAH levels in ambient air resulting mainly from residential heating are included in Table 35, as the source cannot be identified properly (see section 5.1.1.1). The use of wood and coal for heating was the source of high levels of benzo [a]pyrene in Calcutta, India (up to 120 ng/m3; Chakraborti et al., 1988). The concentrations of individual PAH in Calcutta ranged from 1.3 to 200 ng/m3, the highest levels being those of benzo [e]pyrene, benzo [ghi]perylene, and benzo [b]fluoranthene. The average levels of individual PAH resulting from domestic heating in Christchurch, New Zealand were 1-210 ng/m3, benzo [ghi]perylene and coronene showing the highest levels (Cretney et al., 1985), and up to 43 ng/m3 were measured in Essen-Vogelheim, Germany (Buck, 1983). High concentrations of individual PAH were determined in a residential area heated primarily by coal, with levels of up to 260 ng/m3 chrysene, benz [a]anthracene, and benzo [b]fluoranthene (Grimmer, 1980). The following PAH levels were measured on a roof directly downwind of the chimney of a fireplace burning seasoned oak in the Chicago area, USA: 1.8 µg/m3 acenaphthylene, 0.40 µg/m3 naphthalene, 0.35 µg/m3 anthracene, 0.22 µg/m3 phenanthrene, 0.20 µg/m3 benzo [a]pyrene, 0.20 µg/m3 benzo [e]pyrene, 0.13 µg/m3 fluorene, 0.10 µg/m3 pyrene, 0.096 µg/m3 fluoranthene, 0.052 µg/m3 acenaphthene, 0.045 µg/m3 benzo [k]fluoranthene, 0.033 µg/m3 chrysene, 0.030 µg/m3 cyclopenta [cd]pyrene, 0.023 µg/m3 benzo [b]fluoranthene, and 0.019 µg/m3 benz [a]anthracene. The levels of indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, and coronene were below the limit of detection (Khalili et al., 1995). In a comparison of the PAH concentrations in ambient air in eastern and western Germany, the concentrations in rural areas were 3-12 times higher in eastern than in comparable western parts of the country. The PAH profiles were slightly different: the concentrations of the lower-boiling-point PAH fluoranthene and pyrene were 110 and 68 ng/m3 in eastern and 36 and 28 ng/m3 in western Germany. The differences may be due to the different types of brown and hard coal burnt (Jacob et al., 1993a). In 1991, PAH were determined in the air of Berchtesgaden, a national park in Germany, and of the Oberharz (Ministry of Environment, 1993). The concentration of phenanthrene, fluoranthene, and pyrene (about 14 ng/m3) in the Oberharz was two to three times higher than in Berchtesgaden, due to the use of brown coal for heating. The levels of the other PAH were of the same order of magnitude: benz [a]anthracene and benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, about 5 ng/m3; and benzo [ghi]fluoranthene, benzo [c]phenanthrene, benzo [e]pyrene, benzo [a]-pyrene, indeno(1,2,3- cd)pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, anthanthrene, and coronene, < 1 ng/m3. A model calculation for Germany showed that 5000 oil-heated houses contributed to the pollution of ambient air by benzo [a]pyrene to the same extent as one coal-heated house. It was assumed that one German household consumes annually about 5000 litre of heating oil, producing a maximum of 5 mg of benzo [a]pyrene (about 1 µg/litre combusted oil). On the basis of a consumption of a similar amount of hard coal, the same household would have an output of 25 g benzo [a]pyrene (about 5000 µg/kg combusted hard coal) annually (J. Jacob, 1994, personal communication). 5.1.2 Hydrosphere PAH are found in the hydrosphere (Borneff & Kunte, 1983; Müller, 1987), mostly as a result of urban runoff, with smaller particles from atmospheric fallout and larger ones from asphalt abrasion (Hoffman et al., 1984). Long-range atmospheric transport of PAH has been well documented in different countries (Lunde & Bjrseth, 1977; see also section 4.1.2). After PAH are emitted into the atmosphere, for example in motor vehicle exhaust, they are transferred into water by direct surface contact or as a result of rainfall (Grob & Grob, 1974; Van Noort & Wondergem, 1985a,b; Kawamura & Kaplan, 1986). The higher levels of PAH that are found during winter months reflect increased emissions resulting from domestic heating (Quaghebeur et al., 1983; Thomas, 1986; see also section 5.1.1.1); however, the major source of PAH varies for each body of water. Anthropogenic combustion and pyrolysis and urban runoff containing atmospheric fallout, asphalt particles, tyre particles, automobile exhaust condensate and particulates, and lubricating oils and greases were the major sources of PAH in lakes in Switzerland (Wakeham et al., 1980a,b). Comparisons between the levels of individual PAH in precipitation and those in surface water showed that all of the precipitation samples were more highly polluted with PAH, because they had been 'washed out' of the atmosphere. Nearly all of the samples contained > 100 ng/litre of fluoranthene, benzo [b]fluoranthene, pyrene, indeno[1,2,3- cd]pyrene, phenanthrene, and naphthalene. The highest levels of PAH in rainwater were found in Leidschendam, the Netherlands, where pyrene concentrations < 2000 ng/litre, fluoranthene concentrations < 1700 ng/litre, and benzo [a]pyrene and benzo [b]fluoranthene concentrations < 390 ng/litre were detected (van Noort & Wondergem, 1985b). Most surface water samples contained concentrations of < 50 ng/litre of individual PAH. The levels in rainwater were 10-200 ng/litre, whereas those in snow were < 1000 µg/kg, with a maximum of 6800 µg/kg for an individual PAH (Lygren et al., 1984). In one fog sample, benzo [a]pyrene was found at 880 ng/litre and fluoranthene at 3800 ng/litre (Schrimpff, 1983: see section 5.1.2.4). In sediment the levels of individual PAH were usually 1000-10 000 µg/kg dry weight, which are one order of magnitude higher than those in precipitation. Triphenylene was detected in samples of sediment from the Mediterranean Sea (France) at 2-600 µg/kg (Milano et al., 1985) and in samples from Lake Geneva (Switzerland) at 25 µg/kg (Dreier et al., 1985; see section 5.1.3). 5.1.2.1 Surface and coastal waters The levels of individual PAH found in surface and coastal waters at various locations are summarized in Table 37. Rivers in Germany contained some PAH at concentrations of 1-50 ng/litre (Grimmer et al., 1981b; Ernst et al., 1986; Regional Office for Water and Waste Disposal, 1986; Kröber & Häckl, 1989) and fluoranthene, pyrene, chrysene, benzo [a]pyrene, and benzo [e]pyrene at concentrations < 100 ng/litre. The PAH levels in seawater from the German coast varied over one order of magnitude depending on the sampling site. In open seawater, the concentrations of two- to four-ring PAH - naphthalene, fluorene, phenanthrene, fluoranthene, and pyrene - were 0.1-5 ng/litre, and those of five- to six-ring PAH ranged from < 0.01 to 0.2 ng/litre. Near the coast, the concentration of five- to six-ring PAH increased with the content of particles, to which they have greater affinity than two- to four-ring PAH (German Federal Office for Sea Navigation and Hydrography, 1993). The maximum levels of PAH in the Rivers Thames and Trent in the United Kingdom were > 130 ng/litre. The highest levels of individual PAH in the River Thames were 360 ng/litre fluoranthene, 350 ng/litre benzo [a]pyrene, 210 ng/litre indeno[1,2,3- cd]pyrene, 160 ng/litre benzo [ghi]perylene, 140 ng/litre benzo [k]fluoranthene, and 130 ng/litre perylene (Acheson et al., 1976). More recent data were not available. In Norway, the levels of most individual PAH were > 100 ng/litre. For example, surface water from Bislet Creek near Oslo contained fluoranthene, pyrene, phenanthrene, methylphenanthrene, naphthalene, acenaphthene, acenaphthylene, and fluorene at concentrations > 1000 ng/litre (Berglind, 1982). Table 37. Polycyclic aromatic hydrocarbon concentrations (ng/m3) in surface and coastal waters Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] Acenaphthene 14-1232 Acenaphthylene 0.4-0.9 12-1024 Anthracene 1 10 18-932 Anthanthrene 0.2-0.5 15/1.8 Benz[a]anthracene ND 0.16 2.2-6.8 24/66 40/10 71-582 Benzo[a]fluorene 43/330 Benzo[a]pyrene 1-23 0.8 0.39 1.2-7.3 87/25 18 10/60 ND-40 19-311 0.9 Benzo[b]fluoranthene 0.1-0.5 0.07 80/20 ND-42 70-678 0.5-0.9 Benzo[b]fluorene 38 17 Benzo[c]phenanthrene 2.3-4.2 13/34 23-172 Benzo[e]pyrene 2-40 0.06 7.1-11 108/36 40-551 Benzo[ghi]fluoranthene Benzo[ghi]perylene ND ND < 0.05 3.7-7.0 61/16 50/10 ND-61 33-636 ND Benzo[k]fluoranthene 0.7-0.8 0.02 3.6-6.1 59/22 40/10 ND-24 0.2-0.5 Chrysene 11-15 36/87 14 10/10 Coronene ND-2.4 15/4.3 Cyclopenta[cd]pyrene ND ND Dibenzo[a,h]pyrene <0.03 30/10 Fluoranthene 4-616 1.0-3.5 0.35 5.2/9.1 28/102 2.3-13 50/130 2-110 285-3269 3.4-5.1 Fluorene 2 0.63 0.6-1.2 25-1995 Indeno[1,2,3-cd]pyrene Trace < 0.03 2.8-6.1 63/13 50/20 ND-39 17-299 ND 1-Methylphenanthrene 30-1281 5-Methylcholanthrene Naphthalene 4 50-2090 Perylene 0.8-1.4 27 20 9/28 Phenanthrene 3-136 3.5 1.5-9.1 101-5656 Pyrene 5-402 0.28 4.8/8.5 25/90 2.2-13 100/30 485-3099 Triphenylene Table 37 (continued) ND, not detected; /, single measurements; [1] Lake water, Norway, 1981-82 (Gjessing et al., 1984); [2] Lake water, Switzerland (Vu Duc & Huynh, 1981); [3] Lake Superior, USA, 1986 (Baker & Eisenreich, 1990); [4] Elbe River, Germany, 1980 (Grimmer et al., 1981b); [5] Elbe River, main drainage channel, Germany, 1980 (Grimmer et al., 1981b); [6] Water in various rivers, Germany, 1981-83 (Ernst et al., 1986); [7] Water in various rivers, Germany, 1985; analytical method not given (Regional Office for Water and Waste Disposal, 1986); [8] Water in various rivers, Germany, 1985-86; analytical method not given (Krober & Hackl (1989); [9] River water, Norway, 1979 (Berglind, 1982); [10] River water, Switzerland (Vu Duc & Huynh, 1981) Analysed by high-performance liquid chromatography or gas chromatography, unless otherwise stated. The results of studies in which water samples were filtered through solid sorbents may be underestimates of the actual PAH content (see section 2.4.1.4). Table 37 (continued) Compound [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] Acenaphthene ND-3 10 0.08-1.1 50-100 Acenaphthylene ND-5 0.02-1.7 80-1300 Anthracene ND-4 0.2 0.8-9.5 0.01-1.5 < 1-25 ND Anthanthrene NR Benz[a]anthracene ND-5 0.3 ND-9.6 0.04-6.8 ND Benzo[a]fluorene NR Benzo[a]pyrene 0.1-1.8 130-150 0.1/0.2 ND-10 0.2-1.0 0.03-8.8 ND Benzo[b]fluoranthene ND-8 0.04-12 Benzo[b]fluorene 4.0-19 NR Benzo[c]phenanthrene NR Benzo[e]pyrene 0.02-8.8 ND Benzo[ghi]fluoranthene NR Benzo[ghi]perylene 0.2-11 30-160 0.7/0.8 ND-10 0.02-3.8 < 0.3-16 50 Benzo[k]fluoranthene 0.1-1.7 80-140 0.2/0.3 ND-13 0.02-7.7 Chrysene ND-12 NR Coronene 0.01-1.4 NR Dibenzo[a,h]pyrene ND-1 100 Fluoranthene 0.7-508 20-360 1.1/3.7 3-12 0.8 10-25 1.4-2.6 0.40-14 NR Fluorene ND-2 0.7-15 1.9-5.2 0.33-3.2 70-2500 Indeno[1,2,3-cd]pyrene 0.1-8.0 50-210 ND/0.2 ND-8 0.01-3.5 NR 1-Methylphenanthrene NR 5-Methylcholanthrene NR Naphthalene 4-34 3.6 0.4-9.2 NR Perylene 40-130 0.01-5.7 NR Phenanthrene 6-34 21-18 8.0-93 2.4-2.7 0.24-5.8 < 1-3 ND Pyrene 50-260 1-15 0.3-15 8.8-25 0.82-1.7 0.12-15 < 1-53 10-65 Triphenylene NR Table 37 (continued) ND, not detected; /, single measurements; [11] River water, United Kingdom, 1974 (Lewis, 1975); [12] Water in various rivers, United Kingdom, analytical method not given (Acheson et al.,1976); [13] Water in various rivers, United Kingdom; analytical method not given (Sorrell et al., 1980); [14] River water, USA, 1984 (De Leon et al., 1986); [15] Surface water, Canada (Environment Canada, 1994); [16] River water, China, 1981 (Wu et al., 1985); [17] Coastal water, Germany, 1982 (Ernst et al., 1986); [18] Seawater, Germany, 1990 (German Federal Office for Sea Navigation and Hydrography, 1993); [19] Coastal water, Australia, 1983 (Smith et al., 1987); [20] Water (no further specification), Japan, 1974-91 (Environment Agency, Japan, 1993) Analysed by high-performance liquid chromatography or gas chromatography, unless otherwise stated. The results of studies in which water samples were filtered through solid sorbents may be underestimates of the actual PAH content (see section 2.4.1.4). The highest concentrations of PAH in water in Canada were reported for water samples from ditches next to utility and railway lines near Vancouver. The highest mean concentrations were measured near utility poles treated with creosote, with values of 2000 µg/litre for fluoranthene, 1600 µg/litre for phenanthrene, and 490 µg/litre for naphthalene (Environment Canada, 1994). Four individual PAH were detected in seawater from Green Island, Australia. The highest levels of PAH found were 53 ng/litre pyrene, 25 ng/litre anthracene, 16 ng/litre benzo [ghi]perylene, and 3 ng/litre phenanthrene, (Smith et al., 1987). The total content of phenanthrene, anthracene, fluoranthene, pyrene, benzo [b]fluorene, and benz [a]anthracene in the Yellow River, China, was 170 ng/litre (Wu et al., 1985; for individual PAH concentrations, see Table 37). The PAH levels found in the River Rhine in Germany and the Netherlands and in some of its tributaries are summarized in Table 38. Many investigators have detected PAH in the Rhine. The lowest concentrations of benzo [a]pyrene, < 10-20 ng/litre, were found in the Rhine at Lobith and Hagestein in Germany and at Lek in the Netherlands in 1987-90 (Association of Rhine and Meuse Water Supply Companies, 1987-90), when the levels of fluoranthene were 70-140 ng/litre. In 1976-79, the Rhine at Lek and Waal contained < 10-580 ng/litre of benzo [a]pyrene (Association of Rhine and Meuse Water Supply Companies, 1976-79), so that the levels had decreased by one order of magnitude within 14 years. The sum of fluoranthene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, and indeno[1,2,3- cd]pyrene) was 9-40 ng/litre at km 30 and 130-5700 ng/litre at km 853, indicating that the level of pollution increased markedly between the source and the estuary (Borneff & Kunte (1983). The average concentrations of individual PAH were 1-50 ng/litre, although individual PAH were found at concentrations in the range 100-200 ng/litre near Mainz, an industrialized town (Borneff & Kunte, 1964, 1965). In general, the PAH levels in the Rhine decreased by a factor of 3 between 1979 and 1989. The Emscher and Ruhr waterways in Germany have been heavily polluted (see Table 38). In 1985, the Emscher River contained 6400 ng/litre fluoranthene, 6000 ng/litre pyrene, 2000 ng/litre benz [a]anthracene, 1100 ng/litre dibenz [a,h]anthracene, 910 ng/litre benzo [a]pyrene, 880 ng/litre chrysene, 630 ng/litre indeno[1,2,3- cd]pyrene, 510 ng/litre benzo [ghi]perylene, 270 ng/litre anthracene, 220 ng/litre perylene (Regional Office for Water and Waste Disposal, 1986), but by 1989 the levels had decreased by about one order of magnitude (Regional Office for Water and Waste Disposal, 1990 ). The PAH concentrations in the Emscher were three times higher than those in the Rhine near Mainz. Between 1985 and 1989, the PAH levels in the Emscher decreased further by a factor of 15; however, the levels in the Ruhr remained about the same or increased slightly between 1979 and 1985 (Regional Office for Water and Waste Disposal, 1986, 1988, 1990). Table 38. Polycyclic aromatic hydrocarbon concentrations (ng/m3) in the River Rhine and some highly polluted tributaries Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Anthracene 10 270 25-260 10 Anthanthrene 0.9-11 1.3 Benz[a]anthracene 6.1-31 11-50 1970 100-780 13 20 Benzo[a]pyrene 0.8-36 ND-7 6-30 12-40 < 10-20 910 59-280 15 30 Benzo[b]fluoranthene ND-8 7-30 12-40 < 10-30 880 62-310 40 Benzo[c]phenanthrene 1.5-9.1 1.9 Benzo[a]pyrene 18-31 33 Benzo[ghi]fluoranthene 1.0-11 2.2 Benzo[ghi]perylene 15-29 ND-8 6-30 9-30 < 10-20 510 30-210 17 30 Benzo[k]fluoranthene ND-4 2-14 6-20 < 10-40 440 36-150 20 Chrysene 21-62 1080 27 30 Dibenzo[a,h]pyrene 10-40 1100 32-310 30 Fluoranthene 4-18 15-61 25-77 20-140 6420 207-1700 60 Indeno[1,2,3-cd]pyrene 9.5-27 ND-6 2-26 10-40 < 10-20 630 28-220 17 30 Perylene ND-8.1 10 220 13/80 2.1 10 Pyrene 20-50 6010 155-1100 50 ND, not detected; /, single measurements; [1] Rhine, Germany, 1979 (Grimmer et al.,1981b); [2] Rhine, Germany, 1985-88, analytical method not given (Krober & Hackl, 1989); [3] Rhine, Netherlands, 1985-88 (Netherlands' Delegation, 1991); [4] Rhine, Germany, 1987-89, analytical method not given (Regional Office for Water and Waste Disposal, 1988, 1989, 1990); [5] Rhine, Netherlands; 1987-90, analytical method not given (Association of Rhine and Meuse Water Supply Companies, 1987-90); [6] Emscher, Germany, 1985, analytical method not given (Regional Office for Water and Waste Disposal, 1986); [7] Emscher, Germany, 1987-89, analytical method not given (Regional Office for Water and Waste Disposal, 1988, 1989, 1990); [8] Ruhr, Germany, 1979 (Grimmer et al., 1981b); [9] Ruhr, Germany, 1985, analytical method not given (Regional Office for Water and Waste Disposal, 1986) The PAH levels in the main drainage channels of the River Elbe, Germany, were one order of magnitude higher than in the river water (Grimmer et al., 1981b), owing to the high input of rainwater to the channels. 5.1.2.2 Groundwater The PAH concentrations in uncontaminated groundwater in the Netherlands generally did not exceed 0.1 µg/litre, but levels of about 30 µg/litre naphthalene, 10 µg/litre fluoranthene, and 1 µg/litre benzo [a]pyrene were reported in contaminated groundwater (Luitjen & Piet, 1983). Benzo [a]pyrene levels in groundwater in western Germany ranged from 0.1 to 0.6 ng/litre and those of total PAH from 34 to 140 ng/litre (Andelman & Suess, 1970). Benzo [a]pyrene was also detected at levels of 0.1-5.0 ng/litre in groundwater (Woidich et al., 1976). More recent data were not available. Groundwater in the USA contained maximum concentrations of 0.38-1.8 ng/litre naphthalene, 0.02-0.04 ng/litre acenaphthene, and 0.008-0.02 ng/litre fluorene (Stuermer et al., 1982). Near a refinery at Pincher Creek, Alberta, Canada, the pyrene concentrations in groundwater showed a maximum of 300 µg/litre (median, 30 µg/litre); the maximum concentration of fluorene was 230 µg/litre (median, 40 µg/litre). At Newcastle, New Brunswick, Canada, naphthalene was detected at concentrations up to 2.8 µg/litre and benzo [a]pyrene up to 0.32 µg/litre in groundwater near a wood-preserving plant (Environment Canada, 1994). 5.1.2.3 Drinking-water and water supplies PAH levels were determined in drinking-water in samples from Canada, Scandinavia, and the USA up to 1982. The concentration of naphthalene was 1.2-8.8 ng/litre, that of benzo [a]pyrene was 0.2-1.6 ng/litre, and that of the sum of the six 'standard WHO' PAH (fluoranthene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]perylene, and indeno[1,2,3- cd]pyrene) was 0.6-24 ng/litre. The highest levels of naphthalene (1300 ng/litre), benzo [a]pyrene (77 ng/litre), and the six WHO standard PAH (660 ng/litre) were detected in raw water sources in the USA and in the Great Lakes area of Canada (Müller, 1987). More recent measurements are given in Table 39. Most samples contained 0.38-16 ng/litre naphthalene and < 0.04-2.0 ng/litre benzo [a]pyrene. In one set of water samples from the Netherlands, no PAH were detected, with a limit of detection for individual PAH of 4 ng/litre (de Vos et al., 1990). In a study of the changes in PAH concentrations after passage of water through tar-coated major distribution pipes, the level increased from an initial concentration of none detected-13 ng/litre to none detected-62 ng/litre. The finding that water in a few distribution lines had lower concentrations of PAH may be due to sorption of PAH on the surfaces of distribution pipes, chemical interaction with oxidants in water, or a dilution effect (Basu et al., 1987). Of 101 German drinking-water samples analysed in 1994, four exceeded the German drinking-water standard of 0.2 µg/litre for the sum of fluoranthene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]-perylene, and indeno[1,2,3- cd]pyrene. Heavy contamination had occurred after repairs to a pipeline coated with tar, and one drinking-water sample taken in a household contained 2.7 µg/litre of these PAH, in addition to phenanthrene at 2.8 µg/litre and pyrene at 1.2 µg/litre (State Chemical Analysis Institute, Freiburg, 1995). The report stated that abrasion of particles from tar-coated drinking-water pipelines poses a hazard that is often difficult to judge since it is often not known what material was used decades previously. In Canada, the PAH concentrations in drinking-water were usually below or near the detection limits of 1-5 ng/litre, although concentrations of 5.0-21 ng/litre benzo [ghi]perylene, 1.0-12 ng/litre fluoranthene, 1.0-5.0 ng/litre benzo [b]fluoranthene, 1.0-3.0 ng/litre benzo [k]fluoranthene, and 1.0-3.0 ng/litre benzo [a]pyrene were detected in some areas (Environment Canada, 1994). 5.1.2.4 Precipitation (a) Rain The concentrations of PAH found in precipitation in 1979-91 are summarized in Table 40. The levels of benzo [a]pyrene were < 1-390 ng/litre. In an analysis of PAH in rainfall in Hanover, Germany, between July 1989 and March 1990, fluoranthene was the dominant component, followed by pyrene. The average concentration of all PAH increased from 351 ng/litre in summer to 765 ng/litre in the autumn of 1989, while a slight decrease was observed in the winter of 1989-90. These results indicate that the increase in the level of PAH in precipitation in cold weather is due to an increase in residential heating and a slower rate of photochemical degradation (Levsen et al., 1991). Table 39. Polycyclic aromatic hydrocarbon concentrations (ng/litre) in drinking-water Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Acenaphthene 0.6-4.0 7.4-14 Acenaphthylene 0.4-4.4 0.40-1.6 Anthracene 0.5-7 < 1.3-9.7 Anthanthrene 0.2 Benz[a]anthracene ND-1.9 0.4-5.5 0.12-1.5 Benzo[a]fluoranthene 0.1-3.3 0.05-4.2 Benzo[a]pyrene 0.1-0.7 < 0.1-2.0 < 0.04-0.29 Trace-1.9 0.2-0.3 0.2-1.6 < 5.0 Benzo[b]fluoranthene 0.5-1.3 2.4-4.0 0.05-0.34 0.1-14 < 5-40 Benzo[b]fluorene 0.9 0.04-<1.4 Benzo[e]phenanthrene 0.9-1.5 0.28 Benzo[e]pyrene 0.2-4 < 0.1-0.41 Benzo[ghi]fluoranthene 0.36 Benzo[ghi]perylene 0.3-0.9 0.4-1.1 ND 0.4-0.7 0.4-4.0 < 5.0 Benzo[j]fluoranthene 0.03-0.14 0.2-1.2 Benzo[k]fluoranthene 0.2-0.8 0.02-0.10 0.2-4.9 0.1-0.3 0.1-0.7 < 5-40 Chrysene 21-62 1080 27 30 Dibenz[a,h]anthracene 1.2 Fluoranthene 3.5-6.5 1.7-18 < 0.58-24 0.7-3400 3.4-4.2 5-24 2.4-9.0 < 5-623 Fluorene 0.9-4 < 1.1-21 4-16 Indeno[1,2,3-cd]pyrene Trace-0.7 0.4-1.2 ND-1.1 < 0.5 0.7-2.2 < 5.0 1-Methylphenanthrene 0.5-1.0 0.14-13 Naphthalene 1.8-5 < 6.3-8.8 8 6-16 Perylene Trace-0.2 0.2 Phenanthrene 2.5-46 < 2.2-64 24-90 Pyrene 1.6-3.7 1.1-15 < 0.30-12 40/40 Table 39 (continued) ND, not detected; /, single measurements; [1] Austria; analytical method, in-situ fluorescence determination (Woidich et al., 1976); [2] Norway, 1978-80 (Berglind, 1982); [3] Norway, 1980-81 (Kveseth et al., 1982); [4] Switzerland, 1973 (Grob & Grob, 1974); [5] Switzerland (Vu Duc & Huynh, 1981); [6] United Kingdom; water reservoirs after treatment, 1974 (Lewis, 1975); [7] USA, 1976; analytical method, high-performance liquid chromtography and gas chromatography (Thruston, 1978); [8] USA, 1976-77; analytical method, thin-layer chromatography and gas-liquid chromatography with flame ionization detection (Basu & Saxena, 1978a,b); [9] Canada, treated drinking-water, 1987-90 (Environment Canada, 1994) Analysed by high-performance liquid chromatography or gas chromatography, unless otherwise stated. The results of studies in which water samples were filtered through sold sorbernts may be underestimates of the actual PAH content (see section 2.4.1.4). Table 40. Polycyclic aromatic hydrocarbon concentrations (ng/litre) in rainwater Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Acenaphthene 3.2 1.2/16 2.5-8.5 Acenaphthylene 130-200 14 4.7/55 23-59 Anthracene 8-19 0.88/23 2.0-7.9 Benz[a]anthracene 1.2-86 140 6-100 9-33 7-17 20-65 1.6-4.5 Benzo[a]fluoranthene 14-52 Benzo[a]pyrene 5-17 1.1-187 ND-390 10-37 7-26 5-36 ND-0.18 Benzo[b]fluoranthene 2.9-166 15-390 45-70 17-65 Benzo[b]fluorene 15 Benzo[c]phenanthrene 802 Benzo[e]pyrene < 0.5a-149 217-290 7-62 ND-0.51 Benzo[ghi]perylene 7-29 1.7-109 40-70 15-56 22 Benzo[k]fluoranthene 1.0-142 6-190 17-30 9-28 Chrysene 2.9-141 30-120 ND-67 21-29 3.3-12 Dibenz[a,h]anthracene < 0.5a-12 7-20 3-12 Fluoranthene 23-66 23-392 240-270 14-1650 66-180 87-189 115-162 1.7/110 28-70 Fluorene 10-200 6-50 3.2/43 9.1-22 Indeno[1,2,3-cd]pyrene < 0.5a-137 ND-80 50-110 24-72 12 1-Methylphenanthrene 8-26 Naphthalene 8-77 20/72 46-140 Perylene 2 Phenanthrene 130-600 30-133 79-113 158-238 24/140 61-130 Pyrene 9.5-304 25-60 ND-2000 ND-37 36-108 77-175 24-56 Table 40 (continued) ND, not detected; /, single measurements; [1] Bavaria, Germany, 1979-80; analytical method, high-performance thin-layer chromatography (Thomas, 1986); [2] Hanover, Germany, 1989-90 (Levsen et al., 1991); [3] Italy (Morselli & Zappoli, 1988); [4] Leidschendam, Netherlands, 1982 (Van Noort & Wondergem, 1985b); [5] Rotterdam, Netherlands, 1983 (Van Noort & Wondergem, 1985b); [6] Netherlands, 1983 (Den Hollander et al., 1986); [7] Oslo, Norway, 1978 (Berglind, 1982); [8] Oregon, USA, 1982 (Pankow et al., 1984); [9] Portland, USA, 1984 (Ligocki et al., 1985) a Detection limit for benzo[a]pyrene Analysed by high-performance liquid chromatography or gas chromatography, unless otherwise stated. The results of studies in which water samples were filtered through solid sorbents may be underestimates of the actual PAH content (see section 2.4.1.4). The concentrations of phenanthrene and fluoranthene in rainwater were noticeably higher than those at 200 m when sampled simultaneously, but no significant differences in the concentrations of benzo [k]fluoranthene, benzo [b]fluoranthene, benzo [a]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, or indeno[1,2,3- cd]pyrene were found. The authors suggested that scavenging in and below clouds was responsible for the presence of PAH in rainwater (Van Noort & Wondergem, 1985b). The deposition rates of individual PAH in Cardiff, London, Manchester, and Stevenage, United Kingdom, were 0.3-20 µg/m2 per day. Anthracene accounted for about 25% of the deposition in London, followed by pyrene (16%), benzo [b]fluoranthene (16%), and benz [a]anthracene (13%) (Clayton et al., 1992). The rate of precipitation containing PAH after gravitational deposition by rain, snow, and particles was not affected by the type or structure of the receiving surface. Precipitation in a beech and spruce stand contained concentrations of 23-52 ng/litre fluoranthene, 8.9-30 ng/litre benzo [ghi]-perylene, 6.4-27 ng/litre indeno[1,2,3- cd]pyrene, and 2.0-8.4 ng/litre benzo [a]pyrene. The deposition of PAH is in general higher under spruce stands because the rates of interception are higher than those in beech stands. Substantial amounts of PAH are transferred to the soil by litterfall, indicating adsorption of PAH on the surfaces of leaves and needles (Matzner, 1984). (b) Snow The concentrations of PAH in snow samples are summarized in Table 41. A sample collected in Hanover, Germany, contained fluoranthene at 55 ng/litre, pyrene at 31 ng/litre, and other PAH at concentrations up to 9 ng/litre (Levsen et al., 1991). A sample of snow from Bavaria contained 200 ng/litre fluoranthene, 50 ng/litre benzo [ghi]perylene, and 29 ng/litre benzo [a]pyrene (Schrimpff et al., 1979). In Norwegian snow samples, the average concentrations of individual PAH were 10-100 ng/litre, but levels up to 6800 ng/litre were found of phenanthrene, 1-methylphenanthrene, fluoranthene, benzo [b]fluoranthene, and fluorene (Berglind, 1982; Gjessing et al., 1984; Lygren et al., 1984). Snow taken near a steel plant in Canada contained average levels of 50-500 ng/litre of individual PAH but higher amounts of phenanthrene, fluoranthene, and pyrene (Boom & Marsalek, 1988). Table 41. Polycyclic aromatic hydrocarbon concentrations (ng/litre) in snow Compound [1] [2] [3] [4] [5] [6] Acenaphthene 10-13 <50-98 Acenaphthlene 19-47 <50-153 Anthracene 13-28 9-379 165-246 Benz[a]anthracene 2.6 21-47 15-677 228 Benzo[a]fluoranthene 13 179-396 Benzo[a]pyrene 29 3.0 23-77 54-602 250 <100-558 Benzo[b]fluoranthene 9.2 799-1501 <100-647 Benzo[b]fluorene 11 192 Berzo[e]pyrene 5.5 30-64 609 360-630 Benzo[ghi]perylene 50 4.8 29-85 98-551 319-391 <100-466 Benzo[k]fluoranthene 2.8 <100-990 Chrysene 6.2 Dibenz[a,h]anthracene <0.5a Fluoranthene 200 55 108-211 86-2665 1820-3143 <50-7020 Fluorene 13-85 96 485-1237 <50-237 Indeno[1,2,3-cd]pyrene <0.5a 20-82 <100-496 I-Methylphenanthrene 1366-2117 Naphthalene 50-94 36-67 123-195 Perylene 12 Phenanthrene 119-276 45-1385 4055-6787 <50-3560 Pyrene 31 68-143 55-2002 <50-3750 Analysed by high-performance liquid chromatography or gas chromatography, unless otherwise stated. The results of studies in which water samples were filtered through solid sorbents may be underestimates of the actual PAH content (see section 2.4.1.4). a Detection limit for benzo[a]pyrene [1] Bavaria, Germany, 1978; analytical method, high-performence thin-layer chromatography and gas chromatography-mass spectroscopy (Schrimpff et al., 1979); [2] Hanover, Germany, 1990 (Levsen et al., 1991); [3] Norway, 1979-81 (Berglind, 1982); [4] Norway, 1981-82 (Gjessing et al., 1984); [5] Norway (Lygren et al., 1984); [6] Near steel plant, Canada, 1986 (Boom & Marsalek; 1988) (c) Hail The PAH levels in a hail sample collected in Hanover, Germany, were of the same order of magnitude as those in rain samples: fluoranthene, 170 ng/litre; pyrene, 98 ng/litre; benzo [b]fluoranthene, 58 ng/litre; chrysene, 47 ng/litre; benzo [e]pyrene, 40 ng/litre; indeno[1,2,3- cd]pyrene, 29 ng/litre; benzo [ghi]perylene, 27 ng/litre; benzo [k]fluoranthene, 19 ng/litre; benz [a]an-thracene, 16 ng/litre; benzo [a]pyrene, 12 ng/litre; and dibenz [a,h]anthracene, 3.3 ng/litre (Levsen et al., 1991). (d) Fog The concentrations of PAH in fog are higher than those in rain. A fog sample collected in western Germany contained 360-3800 ng/litre fluoranthene and 130-880 ng/litre benzo [a]pyrene (Schrimpff, 1983). In fog samples collected during the autumn of 1986 in Zürich, Switzerland, the average concentrations of PAH found were 4400 ng/litre fluoranthene, 2700 ng/litre benzo [b]fluoranthene, 2500 ng/litre pyrene, 2200 ng/litre phenanthrene, 2100 ng/litre benzo [e]pyrene, 1400 ng/litre benz [a]anthracene, 1400 ng/litre indeno[1,2,3- cd]pyrene, 1200 ng/litre benzo [a]pyrene, 920 ng/litre anthracene, 860 ng/litre 1-methylphenanthrene, 750 ng/litre benzo [b]fluorene, 750 ng/litre perylene, 590 ng/litre benzo [k]fluoranthene, 540 ng/litre benzo [ghi]perylene, 340 ng/litre anthanthrene, 260 ng/litre fluorene, and 160 ng/litre benzo [a]fluorene (Capel et al., 1991). 5.1.3 Sediment PAH levels in sediments from rivers, lakes, seas, estuaries, and harbours are summarized in Tables 42-46. 5.1.3.1 River sediment The concentrations of individual PAH in river sediments in 1987-91 (Table 42) varied over a wide range; the maximum values were in the high nanogram per gram range. The levels of individual PAH in sediments from German rivers were about 4000 µg/kg for benzo [a]pyrene, fluoranthene, and benzo [b]fluoranthene and about 1500 µg/kg for pyrene, indeno[1,2,3- cd]pyrene, and benz [a]anthracene. The levels of other PAH generally did not exceed 500 µg/kg (Kröber & Häckl, 1989; Regional Office for Water and Waste Disposal, 1989). PAH were determined in many German river sediments. Table 42 gives data for three rivers: the Rhine and Neckar rivers are highly polluted, whereas the Gersprenz is relatively uncontaminated. The concentrations of PAH in the sediments of rivers around Aachen, Germany, were determined in different size fractions, which allowed the authors to locate where the sediment became contaminated (Lampe et al., 1991). The PAH concentrations in sediment from the River Elbe in Germany in 1991 were of the same order of magnitude as those in Lake Plöner and Lake Constance, but the river sediment contained more PAH with a low boiling-point than the lake sediments. The ratio of fluoranthene to benzo [e]pyrene, taken as a marker of the emission of PAH from the combustion of brown coal, was 2.8-5.1, similar to those found in the Elbe sediment. It was concluded that the PAH in the sediment were due mainly to brown-coal combustion (German Ministry of Environment, 1993). Table 42. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in river sediments Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] Acenaphthene ND-140 14.5 1 100 ND 0.04-130 Acenaphthylene ND 9.7 1 540 ND 0.7-671 Anthracene ND-1010 80-640 670/NR ND/NR 82.1 8-200 152 4 700 10-1200 Benz[a]anthracene 620-1700 10000/NR 50-90/NR 450 ND-100 541 6 600 3.2-2100 Benzo[a]pyrene Q 400-1250 ND-8000/ 20-90/10-80 1-760 70-11 960 454 ND-80 570 4 400 5-3700 ND-5300 Benzo[b]fluoranthene 460-1290 ND-8700/ 50-190/ 620 ND-50 ND-5600 26-150 Benzo[e]pyrene 596 4 900 0.9-1800 Benzo[ghi]fluoranthene 253 NR Benzo[ghi]perylene Q-578 340-750 ND-2900/ ND 10-70 60-7480 358 ND 353 7 400 3-1310 ND-1900 Benzo[j]fluoranthene 749 Benzo[k]fluoranthene 230-650 ND-4000/ 20-90/10-80 408 ND-60 608 ND-2700 Chrysene ND-1549 6700/NR ND-30/NR 597 904 NR Coronene 20-260 150-2460 284 NR Cyclopenta[cd]pyrene 15 1 100 NR Dibenz[a,h]anthracene 500-1070 2600/NR ND-20/NR 21 ND-200 2 800 8.1-340 Fluoranthene ND-4455 900-2470 ND-19000/ 2-2360 190-29300 904 100-400 1013 13 000 ND-60 NR 100-380/ ND-12 600 52-310 Fluorene ND-260 25.4 ND-2 26 3 000 ND/50 3-130 Indeno[1,2,3-cd]pyrene 360-910 ND-6300/ ND/ND 332 486 16 000 NR ND-4200 1-Methylphenanthrene 145 NR Naphthalene ND-2630 7.0 3 800 ND Perylene 120-320 ND-100 2 400 NR Phenanthrene ND-220 3300/NR ND-40/NR 361 10-400 563 10 000 ND/220 9-2800 Pyrene ND-2526 680-3450 17000/NR ND-130/NR 736 80-300 940 9200 ND-160 20-3900 Triphenylene 10-80 NR Table 42 (continued) NQ not detacted; /, single measurements; NR, not reported; Q, qualitative; [1] Czechoslovakia, 1988; reference weight not given (Holoubek et al., 1990); [2] Rhine, Germany, 1982-83 and 1987-88; analytical method and reference weight not given (Regional Office for Water and Waste Disposal, 1989); [3] Neckar, Germany, 1985-88; fine, unsieved sediment; analytical method not given (Krober & Hackl, 1989); [4] Gersprenz, Germany, 1985-88; fine, unsieved sediment; analytical method not given (Krober & Hackl, 1989); [5] Wildbach, Germany, 1989 (Lampe et al., 1991); [6] Haarbach, Germany, 1989 (Lampe et al., 1991); [7] River, Bremen, Germany, 1994 (Riess & Wefers, 1990; [8] Rhone, France, 1985 (Milano & Vernet, 1988); [9] Sweden, 1985 (Broman et al., 1987); [10] Black River, USA, 1984 (Fabacher et al., 1991); [11] Rainy River, Canada, 1986; reference weight not given (Merriman, 1988); [12] Japan, 1974-91 (Environment Agency, Japan, 1993) Analysed by high-performance liquid chromatography or gas chromatography and concentration in micrograms per kilogram dry weight The maximum levels of individual PAH in sediments in Czechoslovakia were 4500 µg/kg fluoranthene, 2600 µg/kg naphthalene, 2500 µg/kg pyrene, 1500 µg/kg chrysene, 1000 µg/kg anthracene, 580 µg/kg benzo [ghi]perylene, 260 µg/kg fluorene, 220 µg/kg phenanthrene, and 140 µg/kg acenaphthene (Holoubek et al., 1990). The levels of individual PAH in sediments from some of the most polluted areas in continental USA were summarized by Bieri et al. (1986). The levels usually ranged from 1000 to 10 000 µg/kg, but that in sediment from the Elizabeth River, Virginia, contained concentrations up to 42 000 µg/kg. Up to 39 000 µg/kg wet weight were found in the Detroit River (Fallon & Horvath, 1985). The concentrations of individual PAH in sediments from the Trenton Channel of the Detroit River, a waterway in a highly industrialized area, connecting Lake St Clair with Lake Erie. varied from not detected (< 4 µg/kg) to 22 000 µg/kg in different locations. Sediments from the southwest shore of Grosse Ile had low levels of contamination, while those in the vicinity of Monguagon Creek had high levels (Furlong et al., 1988). 5.1.3.2 Lake sediment The concentrations of individual PAH found in lake sediments in 1984-91 (Table 43) ranged from 1 to about 30 000 µg/kg dry weight. The total PAH concentrations in surface sediments from Lake Michigan, USA, were 200-6200 µg/kg dry weight (Helfrich & Armstrong, 1986). 5.1.3.3 Marine sediment The concentrations of individual PAH in marine sediments in 1985-91 (Table 44) varied widely, with maximum values up to about 4000 µg/kg. Sediments near power-boat moorings at the coral reef around Green Island, Australia, were found to contain measurable amounts of several PAH, strongly suggesting that they originated from fuel spillage or exhaust emissions (Smith et al., 1987). The benzo [a]pyrene level was 104-106 times higher in bottom sediments from the Baltic Sea than in water at the same location. The bottom sediments also contained more individual PAH than the corresponding water samples (Veldre & Itra, 1991). Maximum levels of 460 µg/kg benzo [a]pyrene and 400 µg/kg benzo [e]pyrene were determined in northern North Sea sediments in the vicinity of oil fields. The hydrocarbon concentrations were above the background levels only in water and sediments within a 2-km radius of platforms, where diesel-coated drill cuttings were dumped. The contribution of five- and six-ring compounds to the total PAH in sediments was unexpectedly high in samples unlikely to be contaminated by oil. Their source was probably windborne combustion products (Massie et al., 1985). Table 43. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in lake sediments Compound [1] [2] [3] [4] Anthracene 160 41-620 Benz[a]anthracene ND 150-1700 41 Benzo[a]pyrene 180-2000 45 Benzo[b]fluoranthene 200 Benzo[e]pyrene 80 140-1500 75 Benzo[ghi]fluoranthene 75 18-270 Benzo[ghi]perylene 21-1600 107 Benzo[k]fluoranthene 126 Chrysene 250 124 Coronene 1 Dibenz[a,h]anthracene 70 Fluoranthene 66-248 390 330-3900 103 Fluorene 5.9 Indeno[1,2,3-cd]pyrene 100 25-1500 279 Naphthalene ND Perylene 50 47-540 Phenanthrene 70-180 100 300-6600 81 Pyrene 110-122 340 210-3500 60 Triphenylene 25 ND, not detected; [1] Lake Padderudvann, Norway; 1981-82; reference weight not given (Giessing et al., 1984); [2] Lake Geneva, Switzerland (Dreier et al., 1985); [3] Cayuga Lake, USA, 1978; concentrations are given as ng/g deepwater (Heit, 1985); [4] Lake Superior, USA (Hamburg Environment Office, 1993) Analysed by high-performance liquid chromatography or gas chromatography; concentration in micrograms per kilogram dry weight Table 44. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in sea sediments Compound [1] [2] [3] [4] [5] [6] [7] [8] Acenaphthene ND-6 NR Acenaphthylene ND-2000 0.6-4.3 NR Anthracene 3-800 0.3-2.1 6-42 5-313 < 0.06-1.0 Anthanthrene 29-74 NR Benz[a]anthracene 5-39 1-900 0.8-19 9-150 15-250 < 0.01-6.0 Benzo[a]fluoranthene 2-41 NR Benzo[a]pyrene 16-25 6-2200 0.4-13 7-160 1100 14-265 0.2-460 < 0.004-4.3 Benzo[b]fluoranthene 13-26 ND-3800 1300 51-490 Benzo[b]fluorene 2-38 NR Benzo[e]pyrene 5.8-18 0.6-15 9-125 21-345 0.4-396 < 0.1-0.6 Benzo[ghi]perylene ND-400 12-225 700 < 10-623 < 0.01-2.6 Benzo[k]fluoranthene 4.0-9.8 ND-3400 600 10-180 < 0.001-2.5 Chrysene 49 1.0-12 8-165 21-398 < 0.04-0.8 Coronene 11-36 NR Dibenzo[a,e]pyrene 7-79 NR Dibenz[a,h]anthracene 2-7 ND-400 0.5-4.2 4-74 NR Fluoranthene ND/159 4-2000 0.4-31 12-230 2300 36-1913 < 0.1-7.2 Fluorene ND-100 0.5-3.1 1-12 NR Indeno[1,2,3-cd]pyrene 8-200 17-510 Naphthalene ND-100 0.7-8.6 1-2b 18-1074 Perylene 1-2200 5-105 24-178 Phenanthrene 1-1500 0.8-29 23-93 11-971 < 0.06-4.2 Pyrene 8-160 5-1600 1.6-40 10-145 30-1697 < 0.1-15 Triphenylene 2-600 NR ND, not detected /, single measurements; NR, not reported; [1] Baltic Sea, Estonia, reference weight not given (Veldre & Itra, 1991); [2] Mediterranean Sea, France (Milano et al., 1985); [3] Adriatic Sea, Italy, 1983 (Marcomini et al., 1986); [4] Ligurian Sea, Italy (Desideri et al., 1988); [5] Ketelmeer, Netherlands, 1987 (Netherlands' Delegation, 1991); [6] North Sea, Netherlands, within 70 km from the coast; 1987-88 (Compaan & Laane, 1992); [7] North Sea, United Kingdom, 1980 (Massie et al., 1985); [8] Great Barrier Reef, Australia, 1983 (Smith et al., 1987) Analysed by high-performance liquid chromatography or gas chromatography The following background concentrations have been reported in North Sea sediments: < 0.01-20 µg/kg dry weight benzo [a]pyrene, < 30 µg/kg fluoranthene, < 6 µg/kg benzo (b)fluoranthene plus benzo (k)fluoranthene, < 5 µg/kg benzo [ghi]-perylene, and < 3 µg/kg indeno[1,2,3- cd]pyrene (Compaan & Laane, 1992). 5.1.3.4 Estuarine sediments The concentrations of individual PAH in estuarine sediments in 1981-92 (Table 45) varied widely, with maximum values in the high microgram per gram range. Measurements in sediments from the Continental Shelf of the Atlantic Ocean and the Gironde Estuary, France, showed relatively little contamination with PAH when compared with sediments from more polluted European estuaries (Garrigues et al., 1987). The levels of PAH in estuarine sediments in the United Kingdom were 10-500 µg/kg. Higher amounts of fluoranthene (1000-1900 µg/kg) and pyrene (790 µg/kg) were reported in estuaries of the River Mersey and the River Tamar (Readman et al., 1986).The total PAH concentrations in sediments from the Penobscot Bay region of the Gulf of Maine, USA, ranged from 290 to 8800 µg/kg, with a distinct gradient that decreased seawards. The PAH composition was uniform throughout Penobscot Bay. Particulates of combustion products transported in the atmosphere were suggested to be a major source of PAH contamination. The levels in Penobscot Bay sediments were significantly higher than expected for an area previously considered to be uncontaminated and fell within the range found in industrialized regions throughout the world (Johnson et al., 1985). The Saguenay Fjord is the major tributary that empties into the St Lawrence River estuary, and the area is highly industrialized. The PAH concentrations were maximal near the aluminium smelting plants that dominate the industrial sector and which were considered to be the major source of PAH, and the levels decreased with distance from this industrial zone. The concentrations of benzo [a]pyrene, benzo [e]pyrene, fluoranthene, benzo [b]fluoranthene, benzo [j]-fluoranthene, benzo [k]fluoranthene, chrysene and triphenylene, pyrene, indeno[1,2,3- cd]pyrene, benz [a]anthracene, dibenz [a,h]anthracene, perylene, benzo [ghi]perylene, and dibenzo [a,e]pyrene in sediments from the Saguenay Fjord ranged from 2000 to 3800 µg/kg (dry or wet weight basis not given) (Martel et al., 1986). 5.1.3.5 Harbour sediment The levels of individual PAH found in harbour sediments (Table 46) were higher than those in rivers, lakes, or oceans, concentrations < 650 µg/g being reported. Table 45. Polycyclic aromatic hydrocarbon concentrations (µg/kg) found in estuarine sediments Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] Acenaphthene NR NR 210-670 310 Acenaphthylene NR NR <10-160 Anthracene 0.1-18 10-50 30-210 11-93 ND-49 60-860 610 Benz[a]anthracene 10-790 0.2-68 30-160 30-650 23-189 14-540 70-3200 5-140 2000 Benzo[a]fluoranthene NR NR 2-150 Benzo[a]pyrene 10-560 <0.1-52 30-210 30-760 33-313 10-540 160/7200 4-150 2300 60-6800 20-60 Benzo[b]fluoranthene 0.2-79 100-500 53-346 17-1000 Benzo[e]pyrene 10-620 103 40-180 30-550 56-323 120-8200 1-150 2500 Benzo[ghi]perylene 1-72 120-490 70-410 66-403 23-641 <70-4200 3-96 1300 Benzo[k]fluoranthene <0.1-24 20-100 33-189 14-696 Chrysene 20-1210 0.2-46 30-180 37-263 9-578 2900 Cyclopenta[cd]pyrene NR NR 300/830 Dibenz[a,h]anthracene 0.5-12 NR 8-50 2-120 550-4900 470 Fluoranthene 30-1920 1-100 50-180 80-1880 85-506 156-3700 60-7200 14-410 3900 Fluorene 15 40-120 NR 15-1500 390 Indeno[1,2,3-cd]pyrene 20-630 61 60-240 30-420 50-343 9-228 <130-9000 1800 1-Methylphenanthrene NR NR 240 Naphthalene 43 NR NR 80-2200 400 Perylene 2-52 NR NR 270/880 650 60-4200 50-60 Phenanthrene 30-1470 0.5-74 40-130 60-790 119-413 17-252 60-8700 5-300 2400 Pyrene 20-1980 0.5-102 50-220 60-1510 93-425 16-539 50-5400 4-380 4800 ND, not detected; /, single measurements; NR, not reported; [1] Estuarine sediment of the River Elbe, Germany (Japenga at al., 1987); [2] Continental Shelf and Gironde estuary, France (Garrigues et al., 1987); [3] Wadden Sea, Netherlands, 1988 (Compaan & Laane, 1992); [4] Mersey, Dee and Tamar estuaries, United Kingdom, 1984 (Readman at al., 1986); Table 45 (continued) [5] Humber Estuary/the Wash, United Kingdom, 1990 (Compaan & Laane, 1992); [6] Gulf of Maine, Penobscot Bay, USA, 1982 (Johnson et al., 1985); [7] Great Lake tributaries, USA, 1984 (Fabacher at al., 1991); [8] Chesapeake Bay; USA, 1984-86 (Huggett et al., 1988); [9] Puget Sound, USA (Varanasi at al., 1992); [10] Yarra River estuary, Australia, 1976; analytical method: thin-layer chromatography with fluorescence detector (Bagg at al., 1981); [11] Mallacoota Inlet, Australia, 1976; analytical method: thin-layer chromatography with fluorescence detector (Bagg at al., 1981) Analysed by high-performance liquid chromatography or gas chromatography and concentration in micrograms per kilogram dry weight, unless otherwise stated Table 46. Polycyclic aromatic hydrocarbon concentrations (µg/kg) found in harbour sediments Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Acenaphthene <260-2509 50 3800 Acenaphthlene <240-2700 Anthracene <30-27 200 1800/1700 ND-507 110-17 000 120 10 900 Benz[a]anthracene <50-1991 3400/3400 310-20 000 240 8800/414 000 Benzo[a]pyrene 600-1500 400 <30-16 486 1800/2100 <70-94 984 300-19 000 340 8900/109 000 Benzo[b]fluoranthene 450 <35-17 182 ND-4103 410-15 000 Benzo[e]pyrene 930/930 120-11 000 Benzo[ghi]perylene 300 <35-1079 210-12 000 Benzo[k]fluoranthene 200 <35-1430 150-22 000 Chrysene <30-13 900 3900/3800 580-21 000 Coronene 130 Fluoranthene 2000-3600 850 <70-21 566 900/5800 <5-84 514 640 34 200/60 700 Fluorene <60-24 530 370 810-65 000 100 7000 Indeno[1,2,3-cd]pyrene 300 <50-372 180-14 000 160 157 000/715 000 1-Methylphenanthrene 2100/2300 Naphthalene <310-1564 1300/2000 <10-43 628 400 198 000 Perylene 1100/1200 Phenanthrene <50-5001 4200/4000 45-63 683 510 26 000/655 000 Pyrene <70-5179 6300/6400 196-66 831 610-40 000 740 22 800/413 000 ND, not detected; /, single measurements; [1] Rotterdam, Netherlands (Japenga et al., 1987); [2] Rotterdam, Netherlands, 1990 (Netherlands' Delegation, 1991); [3] Hampton Roads, USA, 1982 (Alden & Butt, 1987); [4] New York Bight, USA, 1979; reference weight not given (Boehm & Fiest, 1983); [5] Boston, USA (Shiaris & Jambard-Sweet, 1986); [6] Black Rock, USA (Rogerson, 1988); [7] Various harbours of the Rhine, Germany, 1990 (Hamburg Environment Office, 1993); [8] Vancouver Harbour, Canada (Environment Canada, 1994); [9] Various harbours near steel mills, Canada (Environment Canada, 1994) Analysed by high-performance liquid chromatography or gas chromatography and concentration in micrograms per kilogram dry weight, unless otherwise stated 5.1.3.6 Time trends of PAH in sediment The PAH levels in sediments taken at various depths indicate changes and trends in the sources of PAH, e.g. from coal combustion to oil and gas heating. Measurements in sediments from Plöner Lake, Germany, showed that the concentration of PAH in samples from the northern part of the lake, which is in a populated region situated near a railway, had increased fivefold since 1920, whereas those in the southern part had remained constant. The increase in the northern part was attributed to an increase in the number of PAH emitters. As most of the PAH in the sediment originated from coal combustion, the concentrations decreased when coal-fired railway engines were replaced in this area. The benzo [a]pyrene levels ranged from 240 to 2400 µg/kg dry weight (Grimmer & Böhnke, 1975). These findings are consistent with the results of time-dependent analyses of sediments from Lake Constance (Müller et al., 1977). A general trend in decreasing PAH concentrations from north to south was found in bottom sediments from the main stem of Chesapeake Bay, USA, thought to be due to the higher human population density in the northern region. Most of the compounds appeared to be derived from the combustion or high-temperature pyrolysis of carbonaceous fuels rather than from crude or refined oils. The levels of PAH remained relatively constant over the period 1979-86 at the locations examined. Naturally occurring PAH usually comprised less than 20% of the total; the finding of higher proportions may reflect riverine transport of older sediments to the area or scouring and removal of recently deposited sediments. The benzo [a]pyrene concentrations were 12-150 µg/kg dry weight (Huggett et al., 1988). Similar results were reported for sediments from Buzzard's Bay, USA (Hites et al., 1977). In a study of PAH in sediment samples from the lagoon of Venice, Italy, a historical reconstruction of the PAH depositions in a dated drilling core made it possible to distinguish between natural and anthropogenic combustion and between different PAH sources, including direct petroleum spills and sedimentary diagenesis. The predominance of unsubstituted homologues and the relative abundance of some individual components suggested combustion as the predominant source. The lowest values determined in the deepest strata were assumed to be background concentrations resulting from pre-industrial pyrolytic sources, such as forest fires and wood burning. The benzo [a]pyrene levels were 2.2-17 µg/kg dry weight (Pavoni et al., 1987). 5.1.4 Soil A rough distinction can be made between local sources of pollution (point sources) and diffuse sources. Point sources can obviously give rise to significant local contamination of soil, whereas diffuse sources usually affect more widespread areas, though to a lesser extent. The main sources of PAH in soil are: - atmospheric deposition after local emission, long-range transport, and pollution from combustion gases emitted by industry, power plants, domestic heating, and automotive exhausts (Hembrock-Heger & König, 1990; König et al., 1991) and from natural combustion like forest fires (Hites et al., 1980); - deposition from sewage (sewage sludge and irrigation water) and particulate waste products (compost) (Hembrock-Heger & König, 1990; König et al., 1991); and - carbonization of plant material (Grimmer et al., 1972). The extent of soil pollution by PAH also depends on factors such as the cultivation and use of the soil, its porosity, its lipophilic surface cover, and its content of humic substances (Windsor & Hites, 1978). There is a correlation between the organic content of a soil and the PAH concentration: humus contains more PAH than a soil with little humic content, such as sand (Grimmer et al., 1972; Matzner et al., 1981; Grimmer, 1993). This section addresses PAH in soil resulting mainly from industrial sources, automobile exhaust, and other diffuse sources and gives background values. Attribution of a study to a particular section was difficult, as the sources of PAH emissions are often mixed. 5.1.4.1 Background values Table 47 gives background levels of PAH in soil in rural areas. In non-polluted areas, PAH concentrations were usually in the range 5-50 µg/kg. 5.1.4.2 Industrial sources The PAH levels in soil resulting mainly from industrial sources are given in Table 48. The PAH levels were determined in soil near one American plant where animal by-products and brewer's yeast had been processed since 1957. The operation had subsequently expanded to include the handling of solvents, flue dust, chips, acids, cyanides, and a wide variety of industrial waste. Extremely high PAH concentrations were found in the soil (Aldis et al., 1983). PAH were detected in the soil at the sites of former coking plants in Canada (Environment Canada, 1994). For example in Lasalle, Quebec, the benzo [a]-pyrene levels in 1985 ranged from none detected to 1300 µg/g dry weight. The facility closed in 1976, and by 1991 the benzo [a]pyrene concentration was below 10 000 µg/kg. In Pincher Creek, Alberta, high levels of alkylated PAH were measured after a refinery was dismantled. Maximum concentrations of 260 µg/g dry weight each of fluoranthene and pyrene were measured; benzo [a]pyrene was not detected. Table 47. Polycyclic aromatic hydrocarbon concentrations (µg/kg dry weight) in soil of background and rural areas Compound [1] [2] [3] [4] Acenaphthene 1.7 < 1-21 Acenaphthylene ND/3.0 Anthracene 1.2/4.2 Benzo[a]pyrene 15 6-12 13/22 ND-4.0 Benzo[b]fluoranthene 14/25 Benzo[ghi]perylene 49/28 ND-3.3 Benzo[k]fluoranthene 0.2-3.3 Fluoranthene 22 8-28 35/73 ND-28 Fluorene ND < 1-10 Indeno[1,2,3-cd]pyrene 0.5-4.0 Naphthalene 46 13-60 3.8/11 Phenanthrene 30 17-21 18/39 ND-76 Pyrene 20 9-25 29/42 ND, not detected; /, single measurements; [1] Norway (depth, 0-10 cm), reference weight not given (Vogt at al., 1987); [2] Norway (Aamot et al., 1987); [3] Wales, United Kingdom (depth, 5 cm) (Jones et al., 1987); [4] Green Mountain (depth, 0-5 cm), USA (Sullivan & Mix, 1985) Analysed by high-performance liquid chromatography or gas chromatography PAH profiles were found to depend on the depth of soil from which the samples were taken. A comparison of soil samples from an area of clean air and from an industrialized area showed that the concentrations of PAH with lower boiling-points (fluoranthene, chrysene, and pyrene) decreased with depth, whereas those of PAH with higher boiling-points (indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, and coronene) were relatively greater. The opposite would have been expected on the basis of the solubility of these PAH (Jacob et al., 1993b). Table 48. Polycyclic aromatic hydrocarbon concentrations (µg/kg dry weight) in soil near industrial emissions Compound [1] [2] [3] [4] Acenaphthene 54 5 090 000 Anthracene 144 000 1 600 70 Benz[a]anthracene 79 000 200 000 Benzo[a]pyrene 38 000 321 100 Benzo[b]fluoranthene 200 Benzo[e]pyrene 35 000 Benzo[ghi]perylene 100 Benzo[k]fluoranthene 130 000 100 Chrysene 1 210 000 Fluoranthene 340 000 573 234 000 200 Fluorene 80 8 600 000 Indeno[1,2,3-cd]pyrene 100 Naphthalene 48 5 200 2.4 Perylene 12 000 Phenanthrene 506 000 353 20 000 000 40 Pyrene 208 000 459 16 000 000 100 [1] Near coal gasification plant, Netherlands, concentrations in µg/kg wet weight (de Leeuw et al., 1986); [2] Norway, reference weight not given (Vogt et al., 1987); [3] Near processing plant, USA, 1982; maximum (Aldis et al., 1983); values, analytical method, and reference weight not given; [4] Area of an abandoned coal gasification plant, USA; reference weight not given (Dong & Greenberg, 1988) Analysed by high-performance liquid chromatography or gas chromatography 5.1.4.3 Diffuse sources (a) Motor vehicle and aircraft exhaust The concentrations of individual PAH in soil resulting mainly from motor vehicle exhaust (Table 49) usually range between 1 and 2000 µg/kg. The PAH content of soil often decreased with increasing depth (Matzner et al., 1981; Wang & Meresz, 1982; Butler et al., 1984). Near a motorway in the Midlands, United Kingdom, PAH were determined at depths of 0-4 cm and 4-8 cm. Extremely high concentrations were found in the surface layer, but soil at a depth of 4-8 cm was two times less contaminated (Butler et al., 1984). The pollution may have been a result of airborne transport or of microbial or photochemical degradation (Hembrock-Heger & König, 1990). Comparably high levels of PAH were found at Reykjavik Airport, Iceland (Grimmer et al., 1972; see Table 49). Table 49. Polycyclic aromatic hydrocarbon concentrations (µg/kg dry weight) in soil of areas predominantly polluted by vehicle exhaust Compound [1] [2] [3] [4] [5] Acenaphthylene 71 Anthracene 0.2 13 11 Anthanthrene 0.4 149 Benz[a]anthracene 2.3 430 169-3297 13 Benzo[a]pyrene 3.2 785 165-3196 38 24 Benzo[b]fluoranthene 41 Benzo[e]pyrene 4.5 870 159-2293 29 Benzo[ghi]perylene 7.1 1450 168 46 Benzo[k]fluoranthene 78 Chrysene 4.1 436 251-2703 39 Coronene 1.8 410 40-322 37 Dibenz[a,h]anthracene 1.1 351 2 Fluoranthene 6.5 1290 200-3703 91 37 Fluorene 5 Indeno[1,2,3-cd]pyrene 36 Naphthalene 3 Perylene 0.6 157 6 Phenanthrene 17 1735 92 45 Pyrene 3.5 1610 145-4515 72 61 [1] Iceland (depth, 20 cm; reference weight not given) (Grimmer et al., 1972); [2] Reykjavik Airport, Iceland (surface soil; reference weight not given) (Grimmer et al., 1972); [3] United Kingdom, surface soil near motorway; analytical method, adsorbance measurement, reference weight not given) (Butler et al., 1984); [4] United Kingdom (urban soil; depth, 5 cm) (Jones et al., 1987); [5] Brisbane, Australia (Pathirana et al., 1994) Analysed by high-performance liquid chromatography or gas chromatography (b) Other diffuse sources Table 50 gives the levels of PAH from unpecified sources in soil. Benzo [a]pyrene levels of 800 µg/kg were found in humus, 100-800 µg/kg in garden soil, 35 µg/kg in forest soil, and 0.8-10 µg/kg in sand (Fritz, 1971). Table 50. Polycyclic aromatic hydrocarbon concentrations (µg/kg dry weight) in soil from areas polluted by various diffuse sources Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] Acenaphthylene NR NR 3.8 Anthracene NR NR ND-1.4 22-70 Anthanthrene 27 0.50 ND 10-38 Benz[a]anthracene 80 0.60 ND 47-101 Benzo[a]pyrene 273 10/6.2 24 0.8/357 116 1.50 ND-1.4 157 54-108 Benzo[b]fluoranthene 49-97 Benzo[e]pyrene 23 20/22 50 143 3.10 ND-5.0 47-116 Benzo[ghi]perylene 106 15/33 32 0.9-339 98 3.0 ND 64-147 Benzo[k]fluoranthene 31-62 Chrysene NR NR ND-2.1 50-128 Coronene 49 0.70 ND-1.7 32-66 Dibenz[a,h]anthracene 266 8.4/22 44 44 0.60 ND-1.4 11-29 Fluoranthene 2.5-444 254 2.1 ND-2.1 83 73-170 0.3-75 Fluorene NR NR 14 Indeno[1,2,3-cd]pyrene 30 6.4/7.9 21.4 1.2-545 127 3.3 32-80 Naphthalene NR NR 58 Perylene 3537 4.0/8.5 5.0 NR NR ND 19-71 Phenanthrene NR NIR ND-18 78 31-106 Pyrene 150 0.80 ND-0.5 90 80-183 0.1-64 ND, not detected; /, single measurements; NR, not reported; [1] Germany, birch tree peat (Ellwardt, 1976); [2] Germany, black and white peat (Ellwardt, 1976); [3] Germany, sandy loam (Ellwardt, 1976); [4] Soiling mountain, Germany; depth, 0-15 cm; analytical method, high-performance thin-layer chromatography; reference weight not given (Matzner et al., 1981); [5] Germany, forest, brown soil, surface layer (Bachmann et al., 1994); Table 50 (continued) [6) Germany, forest, brown soil; depth, 0-2 cm (Bachmann et al., 1994); [7] Iceland; depth, 3-30 cm; reference weight not given (Grimmer et al., 1972); [8] Norway, bog soil; depth, 0-10 cm; reference weight not given (Vogt et al., 1987); [9] Toronto, Canada, virgin and cultivated soil; reference weight not given (Wang & Meresz 1982); [10] Nova Scotia, Canada (Windsor & Hites, 1978) Analysed by high-performance liquid chromatography or gas chromatography The PAH concentrations of cultivated soil were slightly higher than those in virgin soil. For example, the benzo [a]pyrene concentrations were 65-87 µg/kg in cultivated soil and 54 µg/kg in virgin soil (Wang & Meresz, 1982). The PAH levels in cultivated soils from German gardens at a maximum depth of 25 cm decreased from industrial areas (fluoranthene, 590-2500 µg/kg; benzo [a]pyrene, 220-1400 µg/kg) to rural areas (fluoranthene, 100-390 µg/kg; benzo [a]pyrene, 30-150 µg/kg) and with soil depth (benzo [a]pyrene concentration, 280-3000 µg/kg at 0-30 cm, 60-4600 µg/kg at 30-60 cm, and 10-7900 µg/kg at 60-100 cm). High PAH concentrations were found at a depth of 100 cm in soil from an old industrial area and from an area filled with contaminated soil. In compost soil, benzo [a]pyrene was present at a concentration of 0.10-2.5 mg/kg in 1986 and 0.02-1.3 mg/kg in 1987 (Crössmann & Wüstemann, 1992). Fluoranthene and pyrene were measured in soil samples, from a wooded area in Maine, a marshy area of South Carolina, a grassy, uncultivated meadow in Nebraska, a mossy area with pine needles in Wyoming, and a sandy desert area in Nevada, USA, and in dark brown, red clay, and light brown loam from Samoa. The highest levels of individual PAH (up to 80 µg/kg) were found in the soil from the wooded area in Maine. The levels in the marshy and grassy soils of South Carolina and Nebraska were 8.4-26 µg/kg. The other soils sampled contained fluoranthene and pyrene at levels < 1 µg/kg (Hites et al., 1980). In Iceland, the concentrations of individual PAH in lava soil at depths of 3 and 25 cm were near the limit of detection. Similar levels were found in vegetable soil at depths of 10 and 30 cm, but the concentrations at 10 cm were twice as high as those at 30 cm (Grimmer et al., 1972). Higher levels of PAH were found in the humus layer of spruce and beech forest ecosystems than in the mineral soil, but the spruce stand contained and stored more PAH than the beech stand (Matzner et al., 1981). Forest soils in Germany contain many PAH in large amounts; Table 48 shows the PAH concentrations in one forest brown soil. The first humic layer of the soil had the highest PAH concentration, and the level decreased with depth to below the limit of detection in layers below 10 cm (Bachmann et al., 1994). The concentrations of PAH were no higher in soil that had been treated with sewage sludge than in untreated soil, indicating that sewage sludge is not a major source of PAH (Hembrock-Heger & König, 1990; König et al., 1991). 5.1.4.4 Time trends of PAH in soil Soil samples collected from Rothamsted Experimental Station in southeast England over a period of about 140 years (1846-1980) were analysed for PAH (Jones et al., 1987). All of the soils were collected from the plough layer (0-3 cm) of an experimental plot for which atmospheric deposition was the only source of PAH. The total PAH burden of the plough layer had increased by approximately fivefold since 1846. The concentrations of most of the individual PAH (anthracene, anthanthrene, fluorene, benzo [a]pyrene, benzo [e]pyrene, fluoranthene, benzo [b]fluoranthene, benzo [k]fluoranthene, chrysene, pyrene, indeno[1,2,3- cd]pyrene, phenanthrene, and benz [a]-anthracene) had increased by about one order of magnitude. For example, the benzo [a]pyrene level was 18 µg/kg in 1846 and 130 µg/kg in 1980, and the anthracene level was 3.6 µg/kg in 1846 and 13 µg/kg in 1980. The levels of coronene, acenaphthylene, acenaphthene, perylene, and benzo [ghi]perylene remained approximately the same, whereas the naphthalene content decreased from 39 µg/kg in 1846 to 23 µg/kg in 1980. 5.1.5 Food In the past, benzo [a]pyrene was the most common PAH determined in foods and was used as an indicator of the presence of PAH (Tilgner, 1968). The earliest measurements of PAH, in particular of benzo [a]pyrene, date to 1954; these were reviewed by Lo & Sandi (1978) and by Howard & Fazio (1980). The levels of individual PAH in foods in more recent studies are summarized in Tables 51-56. 5.1.5.1 Meat and meat products The concentrations of individual PAH found in meat are shown in Table 51. In a comparison of home and commercially smoked meats in Iceland, very little benzo [a]pyrene was detected in smoked sausage and mutton, but considerable amounts of benzo [a]pyrene and other PAH were found in home-smoked mutton and lamb, independently of whether they were covered with cellophane or muslin. About 60-75% of the total benzo [a]pyrene was detected in the superficial (outer) layers of the meat (Thorsteinsson, 1969). These findings are in agreement with those of Rhee & Bratzler (1970) for smoked bologna and bacon and with those of Tilgner (1958) and Gorelova et al. (1960). The amount of PAH formed during roasting, baking, and frying depends markedly on the conditions (Lijinsky & Shubik, 1964). In an investigation of the effect of the method of cooking meat, including broiling (grilling) on electric or gas heat, charcoal broiling, and broiling over charcoal in a no-drip pan, it was shown that the formation of PAH can be minimized by avoiding contact of the food with flames, cooking meat at lower temperatures for a longer time, and using meat with minimal fat (Lijinsky & Ross, 1967). The most likely source of PAH is melted fat that drips onto the heat and is pyrolysed (Lijinsky & Shubik, 1965). The exact chemical mechanism for the formation of PAH is unknown. Table 51. Polycyclic aromatic hydrocarbon concentrations (µg/kg fresh weight) in meat and meat products Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] Anthracene 0.9 20-31a ND-2 0.5-133 Anthanthrene 5-8 ND ND-66.5 Benz[a]anthracene 0.5 0.5 0.02-0.64 0.03 Trace-0.33a 0.02-0.03 O.04-0.38 0.04-0.13 0.05 16-37 ND-1 0.2-144 Benzo[a]fluorene 17-28 1-2 ND-174 Benzo[a]pyrene 0.1 0.6 0-02-0.45 0.02 0.05 O.01-0.14 0.01-0.04 0.04-0.26 0.03-0.26 0.05 26-42 ND-1 ND-212 Benzo[b]fluoranthene 0.3 1.0 0.30 0.04 16-24 ND-92.3 Benzo[b]fluorene 10-12 2-7 ND-71.9 Benzo[c]phenathrene 1.4 0.03-0.36 0.06 Trace-0.18 0.03-0.04 0.05-0.21 0.05-0.10 Benzo[e]pyrene 0.03 6-9 ND-2 ND-80.9 Benzo[ghi]perylene 0.2 0.6 0.03-0.31 0.03 3.75 Trace-0.12 0.03-0.04 0.06-0.27 0.05-0.19 0.05 10-17 ND-2 ND-153 Benzo[j]fluoranthene 5-7 Benzo[k]fluoranthene 0.2 0.2 0.05 0.01 8-14 ND-172b Chrysene 0.6 0.15 0.3-140a Dibenz[a,h]anthracene 0.01 ND-8.8 Fluoranthene 0.9 1.1 7.8 0.48 57-103 6-9 1.1-376 Indeno[1,2,3-cd]pyrene 0.2 0.7 0.04-0.38 0.03 2.5 Trace-0.11 0.01-0.03 0.04-1.40 0.05/0.1 15-22 ND-5 ND-171 1-Methylphenanthrene 4-5 ND-3 0.5-57.6 Perylene ND-3 ND ND-27.9 Phenanthrene 3.0 22-64 10-16 3.5-618 Pyrene 0.55 38-63 5-7 1.2-452 ND, not detected; /, single measurements; [1] Poultry and eggs, Netherlands, reference weight not given (de Vos et al., 1990); [2] Meat and meat products, Netherlands, reference weight not given (de Vos et al., 1990); [3] Smoked beef, Netherlands, reference weight not given (de Vos et al., 1990); [4] Unsmoked beef, Netherlands (de Vos et al., 1990); [5] Bacon, United Kingdom (Crosby et al., 1981); [6] Smoked meat, United Kingdom (McGill et al., 1982); [7] Unsmoked meat, United Kingdom (McGill et al., 1982); Table 51 (continued) [8] Smoked sausages, United Kingdom (McGill et al., 1982); [9] Unsmoked sausages, United Kingdom (McGill et al., 1982); [10] Meat, United Kingdom, reference weight not given (Dennis et al., 1983); [11] Mesquite wood cooked pattie (70-90 % lean), USA, reference weight not given (Maga, 1986); [12] Hardwood charcoal cooked pattie (70-90% lean), USA, reference weight not given (Maga, 1986); [13] Grilled sausages, Sweden, reference weight not given (Larsson et al., 1983) High-performance liquid chromatography or gas chromatography a In sum with triphenylene b In sum with benzo[j]fluoranthene In one study, the highest concentration of benzo [a]pyrene (130 µg/kg) in cooked meat was found in fatty beef, and the concentration appeared to be proportional to the fat content (Doremire et al., 1979). Levels of about 50 µg/kg were detected in a charcoal-grilled T-bone steak (Lijinsky & Ross, 1967), in heavily smoked ham (Toth & Blaas, 1972), and in various other cooked meats (Potthast, 1980). Usually, benzo [a]pyrene levels up to 0.5 µg/kg have been found (Prinsen & Kennedy, 1977). In meat, poultry, and fish in Canada, benzo [k]fluoranthene was detected at concentrations up to 0.30 µg/kg and benzo [a]pyrene up to 1.1 µg/kg (Environment Canada, 1994). Benzo [a]pyrene was found in some German meat products in 1994 at concentrations generally < 1 µg/kg . The highest concentration, 9.2 µg/kg, was found in a ham from the Black Forest (State Chemical Analysis Institute, Freiburg, 1995). 5.1.5.2 Fish and other marine foods Benzo [a]pyrene was found at levels ranging from none detected to 18 µg/kg in smoked fish. The differences were probably due to factors such as the type of smoke generator, the temperature of combustion, and the degree of smoking (Draudt, 1963). The highest concentration of benzo [a]pyrene (130 µg/kg) in seafood was found in mussels from the Bay of Naples (Bourcart & Mallet, 1965), and a level of about 60 µg/kg was detected in smoked eel skin. Most of the fish analysed contained 0.1-1.5 µg/kg (Steinig, 1976). Benzo [a]pyrene was also detected at levels up to 3.3 µg/kg in 21 samples of smoked fish, oysters, and mussels of various origins (Prinsen & Kennedy, 1977). The levels of individual PAH are summarized in Table 52. 5.1.5.3 Dairy products: cheese, butter, cream, milk, and related products PAH were detected in considerable amounts in smoked cheese (Prinsen & Kennedy, 1977; Lintas et al., 1979; McGill et al., 1982; Osborne & Crosby, 1987a). The benzo [a]pyrene content of a smoked Italian Provola cheese was 1.3 µg/kg (Lintas et al., 1979). Concentrations of 0.01-5.6 µg/kg fresh weight fluoranthene, benz [a]anthracene, benzo [c]phenanthrene, benzo [a]pyrene, benzo [ghi]perylene, and indeno[1,2,3- cd]pyrene were found in a smoked cheese sample and 0.01-0.06 µg/kg in unsmoked cheese from the United Kingdom (McGill et al., 1982). In other unsmoked cheese samples from the United Kingdom, the individual PAH levels were between < 0.01 µg/kg for dibenz [a,h]anthracene and 1.5 µg/kg for pyrene. Similar concentrations of PAH were found in British butter and cream samples (Dennis et al., 1991). Table 52. Polycyclic aromatic hydrocarbon concentrations (µg/kg) found in fish and marine foods Compound [1] [2] [3] [4] [5] [6] Acenaphthene Anthracene 0.9 1.3-64.3 1.4-49.6 Benz[a]anthracene 1.3 ND-11.2 ND-6.3 ND-86 Trace-0.09 Benzo[a]pyrene 1.4 ND-5.5 ND-5.4 0.10 ND-18 Trace-0.35 Benzo[b]fluoranthene 2.0 ND-3.9 ND-3.6 0.35 Benzo[c]phenanthrene ND-15 0.01-0.09 Benzo[e]pyrene ND-2.8 ND-3.0 Benzo[ghi]perylene 0.9 ND-2.8 ND-1.8 4.3 ND-25 Trace-0.39 Benzo[k]fluoranthene 0.7 ND-6.7a ND-5.1a 0.10 Chrysene 2.9 ND-13.0b ND-9.4b Dibenz[a,h]anthracene Fluoranthene 1.8 1.4-79.9 1.7-48.4 2.4 Fluorene Naphthalene Indeno[1,2,3-cd]pyrene 1.6 ND-7.1 ND-2.4 2.7 ND-37 ND-0.33 Perylene ND-1.2 ND-1.0 Phenanthrene 3.5 5-330 10.4-277 Pyrene 1.3-67.8 2.1-38.4 ND, not detected; NR, not reported; [1] Fish, Netherlands (de Vos et al., 1990); [2] Herring, whitefish, mackerel, eel, salmon, salmon trout, various fillets; all smoked; Sweden (Larsson, 1982); [3] Fish and fish products: sprats, herring, rainbow trout, caviar, herring paste, salmon paste; all smoked or canned; Sweden (Larsson, 1982); [4] Kippers, United Kingdom (Crosby et al., 1981); [5] Fish (smoked), United Kingdom, concentration in µg/kg wet weight (McGill et al., 1982); [6] Fish, unsmoked, United Kingdom, concentration in µg/kg wet weight (McGill et al., 1982) Table 52 (continued) Compound [8] [9] [10] [11] [12] [13] [14] Acenaphthene < 2-5.13 2.22-22.3 Anthracene < 2-78.4 ND-5.88 ND-0.6 ND-1.9 < 0.05 Benz[a]anthracene 0.14 1.6-7.5 < 2 0.14-5.31 0.8-3.0 0.8-20.9 Benzo[a]pyrene 0.13 t-4.5 < 2-7.63 ND-5.33 0.4-1.0 0.2-12.2 < 0.004 Benzo[b]fluoranthene 0.13 0.13-5.77 4.5-12.2c 1.2-24.3c Benzo[e]pyrene 0.12 2.4-6.3 0.7-7.6 Benzo[ghi]perylene 0.12 0.17-30.9 0.4-0.8 0.3-5.7 Benzo[k]fluoranthene 0.04 NR NR < 0.002 Chrysene 0.65 < 2 ND-15.9 3.2-8.8b 3.9-30.8b < 0.03 Dibenz[a,h]anthracene 0.03 0.21-39.3 0.1-0.2d <0.1-0.5d Fluoranthene 0.1 < 2-123.5 ND-32.7 5.1-17.5 4.5-18.7 Fluorene < 2-18.5 ND-65.7 Napthalene < 2-67.4 2.06-156.1 Indeno[1,2,3-cd]pyrene 0.28-28.6 0.3-0.6 0.2-6.4 Perylene 0.2-2.7 0.1-3.1 < 0.05 Phenanthrene < 2-100.8 5.84-87.2 2.1-4.2 1.9-19.6 Pyrene 0.79 < 2-144.9 ND-68.0 3.1-12.4 2.6-11.2 < 0.03 ND, not detected; NR, not reported; [8] Fish, United Kingdom (Dennis et al., 1983); [9] Fish, Nigeria (Emerole et al., 1982); [10] Fresh fish from the Arabian Gulf: andag, sheim, gato, sheiry, faskar, chaniedah; after an oil spill (Al-Yakoob et al., 1993); [11] Fresh fish and shrimps, Kuwait, after Gulf war (Saeed et al., 1995); [12] Fresh oysters, various origins, concentration in µg/kg wet weight (Speer et al., 1990); [13] Canned or smoked oysters and mussels, various origins, concentration in µg/kg wet weight (Speer at al., 1990); [14] Clam, Australia; analytical method: fluorescence spectrophotometry: concentration in µg/kg wet weight (Smith et al., 1987) Analysed by high-performance liquid chromatography or gas chromatography; reference weight not given, unless otherwise stated a In sum with benzo[j]fluoranthene b In sum with triphenylene c Benzo[b+k]fluoranthenes d Dibenz[a,h+a,c]anthracenes In Finnish butter samples, most of the measured PAH (phenanthrene, 1-methylphenanthrene, fluoranthene, pyrene, benzo [a]fluorene, benzo [ghi]-fluoranthene, cyclopenta [cd]pyrene, perylene, anthanthrene, benzo [ghi]pyrene, and indeno[1,2,3- cd]pyrene) occurred at levels < 0.1 µg/kg. The maximum level was 1.4 µg/kg fluoranthene (Hopia et al., 1986). The concentrations of fluoranthene, pyrene, benz [a]anthracene, chrysene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [e]pyrene, perylene, benzo [ghi]pyrene, indeno[1,2,3- cd]pyrene, and dibenz [a,h]anthracene were measured in milk, milk powder, and other dairy products in Canada (Lawrence & Weber, 1984), the Netherlands (de Vos et al., 1990), and the United Kingdom (Dennis et al., 1983, 1991). The concentrations ranged from < 0.01 µg/kg for benzo [k]fluoranthene and dibenz [a,h]anthracene to 2.7 µg/kg for pyrene. Canadian infant formula was found to contain 8.0 µg/kg fluoranthene, 4.8 µg/kg pyrene, 1.7 µg/kg benz [a]anthracene, 0.7 µg/kg benzo [b]fluoranthene, 1.2 µg/kg benzo [a]pyrene, 0.6 µg/kg perylene, 0.3 µg/kg anthanthrene, and 1.2 µg/kg indeno[1,2,3- cd]pyrene (Lawrence & Weber, 1984). Slightly lower levels were detected in British samples in 1982-83 (Dennis et al., 1991). PAH were detected at levels of 0.003-0.03 µg/kg in human milk (Heeschen, 1985). 5.1.5.4 Vegetables The levels of PAH found in vegetables in recent studies are listed in Table 53. Fluoranthene, but no other PAH, was reported to have been found in unspecified fruits and vegetables in Canada at levels of none detected to 1.8 µg/kg (Environment Canada, 1994). Kale was found to contain high concentrations of fluoranthene (120 µg/kg), pyrene (70 µg/kg), chrysene (62 µg/kg), and benz [a]anthracene (15 µg/kg), and PAH concentrations up to 7 µg/kg were determined in other vegetables (Vaessen et al., 1984). The differences in PAH content have been attributed to variations in the ratio of surface area:weight, in location (rural or industrialized), and in growing season. Washing (at 20°C) vegetables contaminated by vehicle exhausts did not reduce the PAH contamination (Grimmer & Hildebrandt, 1965). In a comparison of the PAH contents of terrestrial plants grown in chambers containing 'clean air' and in the open field, the contamination was shown to be due almost exclusively to airborne PAH, which were not synthesized by the plants (Grimmer & Düvel, 1970) . Table 53. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in vegetables Compound [1] [2] [3] [4] [5] [6] [7] [8] Anthracene 0.09-0.19 <0.1-0.3 Benz[a]anthracene 15 0.7-4.6 0.05-3.17 0.05-3.2 0.4 0.3 Benzo[a]fluoranthene 0.08-2.6 Benzo[a]pyrene 4.2 0.05-1.4 5.6 0.3-6.2 ND-1.42 0.05-3.0 0.2 Benzo[b]fluoranthene 6.1 0.5-7.3 0.9-3.2 0.2 Benzo[b]fluorene 0.11-2.8 Benzo[c]phenanthrene 9.2 0.05-1.5 Benzo[e]pyrene 7.9 0.07-2.2 0.5-6.7 0.2 Benzo[ghi]perylene 7.7 0.13-2.1 10 0.5-10.8 ND-1.39 3.7-10 0.1 Benzo[k]fluoranthene 3.7 ND-17 0.1 Chrysene 62 2.4-4.0 0.8 0.5 Dibenz[a,h]anthracene 1.0 0.04 Dibenzo[a,h]pyrene 0.7 Dibenzo[a,i]pyrene 0.3 Fluoranthene 117 1.1-28 28 2.8-9.1 9.2-17 Indeno[1,2,3-cd]pyrene 7.9 0.14-0.72 2.4 0.3-8.3 ND-1.92 1.8-4.2 1-Methylphenanthrene 0.10-2.1 0.7-1.6 Perylene 0,05-0.75 <0.1-1.7 Phenanthrene 0.47-12 1.8-7.5 Pyrene 70 0.9-18 3.4-10.4 ND, not detected; [1] Kale, Netherlands (Vaessen et al., 1984); [2] Lettuce, Finland, concentration in µg/kg fresh weight (Wickstrom et al., 1986); [3] Lettuce, Germany, from an industrial area (Ministry of Environment, 1994); [4] Lettuce, Sweden, concentration in µg/kg fresh weight (Larsson & Sahlberg, 1982); [5] Lettuce and cabbage, United Kingdom, concentration in µg/kg fresh weight (McGill et al., 1982); [6] Lettuce, India (Lenin, 1994); [7] Potatoes, Netherlands (de Vos et al., 1990); [8] Tomatoes, Netherlands (Vaessen et al., 1984) Analysed by high-performance liquid chromatography or gas chromatography; reference weight not given, unless otherwise stated The benzo [a]pyrene levels in potatoes in eastern Germany were 0.2-400 µg/kg. The highest concentrations were detected in the peel of potatoes grown in soil containing 400 µg/kg benzo [a]pyrene, 750 µg/kg benzo [e]pyrene, 1000 µg/kg benz [a]anthracene, 600 µg/kg chrysene, 160 µg/kg dibenz [a,h]anthracene, 1000 µg/kg benzo [b]fluoranthene, 2300 µg/kg phenanthrene, 1800 µg/kg pyrene, 220 µg/kg benzo [k]fluoranthene, 500 µg/kg indeno[1,2,3- cd]pyrene, 2500 µg/kg fluoranthene, and 120 µg/kg anthracene (Fritz, 1971, 1972, 1983). High concentrations of PAH were detected in lettuce grown close to a highway; the levels of individual PAH decreased with distance from the road. Washing the vegetables reduced their content of high-molecular-mass PAH but not of phenanthrene (Larsson & Sahlberg, 1982). In another study, the profiles of PAH in lettuce were similar to those in ambient air, indicating that deposition of airborne particles was the main source of contamination (Wickström et al., 1986). PAH concentrations were determined in fenugreek, spinach beet, spinach, amaranthus, cabbage, onion, lettuce, radish, tomato, and wheat grown on soil that had been treated with sewage sludge. The levels of individual PAH in lettuce leaves (Table 53) were one to two orders of magnitude lower than those in the sewage sludge and the soil on which the lettuce was grown (Lenin, 1994). The PAH levels in carrots and beans grown near a German coking plant were below 0.5 µg/kg wet weight. The levels of fluoranthene were 1.6- 1.7 µg/kg and those of pyrene 1.0-1.1 µg/kg. Vegetables with large, rough leaf surfaces, such as spinach and lettuce, had PAH levels that were 10 times higher, perhaps due to deposition from ambient air (Crössmann & Wüstemann, 1992). 5.1.5.5 Fruits and confectionery (Table 54) Higher concentrations of PAH were found in fresh fruit than in canned fruit or juice, and especially high concentrations of phenanthrene (17 µg/kg) and chrysene (69 µg/kg) were found in nuts (de Vos et al., 1990). In 1982-83 in the United Kingdom, high PAH levels were found in samples of puddings, biscuits, and cakes, which were probably derived from vegetable oil. Similar concentrations of individual PAH were detected in samples of British chocolate (Dennis et al., 1991). 5.1.5.6 Cereals and dried foods Wheat, corn, oats, and barley grown in areas near industries contained higher levels of PAH than crops from more remote areas. Drying with combustion gases increased the contamination by three- to 10-fold; use of coke as fuel resulted in much less contamination than oil (Bolling, 1964). Cereals contained benzo [a]pyrene at levels of 0.2-4.1 µg/kg (Table 55). The highest concentrations, up to 160 µg/kg, were found in smoked cereals (Tuominen et al., 1988). Table 54. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in fruits and confectionery Compound [1] [2] [3] [4] [5] [6] [7] Anthracene 0.4 0.3 Benz[a]anthracene 0.5 0.11 4.2 0.2 4.2 0.08-2.73 Benzo[a]pyrene 0.1 0.07 0.2 0.3 0.4 0.04-2.20 Benzo[b]fluoranthene 0.1 0.1 0.06 0.4 0.4 3.5 0.03-1.27 Benzo[c]phenanthrene 0.5 12 2.2 Benzo[e]pyrene 0.03 0.08-2.92 Benzo[ghi]fluoranthene 0.9 0.9 Benzo[ghi]perylene 0.1 0.06 0.4 1.1 0.2 0.11-2.55 Benzo[k]fluoranthene 0.1 0.1 0.02 0.1 0.1 0.5 0.04-1.36 Chrysene 0.5 0.23 69 0.5 36 0.09-2.84 Dibenzo[a,h]pyrene 0.01 < 0.01-0.23 Fluoranthene 3.6 1.0 0.93 3.0 1.9 2.3 0.52-3.57 Indeno[1,2,3-cd]pyrene 0.4 0.4 0.2 0.10-3.18 Phenanthrene 7.8 17 2.9 3.2 Pyrene 0.83 0.59-2.37 [1] Fresh fruit, Netherlands (de Vos et al., 19900: [2] Canned fruit and juices, Netherlands (de Vos et al., 1990); [3] Fruit and sugar, United Kingdom (Dennis et al., 1983); [4] Nuts, Netherlands (de Vos et al., 1990); [5] Biscuits, Netherlands (de Vos et al., 1990); [6] Sugar and sweets, Netherlands (de Vos et al., 1990); [7] Puddings, biscuits and cakes, United Kingdom (Dennis et al., 1991) Analysed by high-performance liquid chromatography or gas chromatography; reference weight not given Table 55. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in cereals and dried foods Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Acenaphthene 1.6 NR NR 0.7 2.3 Anthracene 9.4 NR NR 1.3 19/150 Anthanthrene NR NR Benz[a]anthracene 0.1-42 11 0.69 0.11-0.21 2.5/3.7 0.6 0.3 <0.1/0.2 6.3/110 Benzo[a]pyrene ND-0.3 5.4 0.40 0.10-0.12 0.5/0.8 0.2 0.3/0.4 0.6/160 Benza[b]fluoranthene 0.1-0.5 0.28 0.07-0.09 0.9 0.2 0.1 Benzo[e]pyrene 0.42 0.06-0.17 0.1/0.7 Benzo[ghi]perylene 0.54 0.13-120 Benzo[k]fluoranthene 0.50 0.1-0.14 Dibenz[a,h]anthracene ND-1.2 0.06 0.01-0.02 3.6 Fluoranthene 0.8-26 130 0.71 0.58-0.69 18/28 1.9 1.4 1.5/13 70/790 Fluorene 5.9 NR NR 2.3/2.7 6.4/87 Indeno[1,2,3-cd]pyrene ND-0.4 1.08 0.24-0.33 1.4 0.2 Perylene 0.1-0.4 0.7 NR NR 94 NR NR 14/2983/1 Pyrene 1.1-48 47 0.10 0.38-0.62 20/21 2.2 3.4 1.6/5.4 60/630 ND, not detected; /, single measurements; NR, not reported; [1] Barley malt, Canada (Lawrence & Weber, 1984); [2] Bran, Finland (Tuominen et al., 1988); [3] Bran, United Kingdom (Dennis et al., 1991); [4] High bran and granary bread, United Kingdom (Dennis et al., 1991); [5] Bran, Canada (Lawrence & Weber, 1984); [6] Corn bran, Canada (Lawrence & Weber, 1984); [7] Flaked milled corn, Canada (Lawrence & Weber, 1984); [8] Oats, Finland (Tuominen et al., 1988); [9] Smoked oats, barley and beans, Finland (Tuominen et al., 1988) Analysed by high-performance liquid chromatography or gas chromatography; reference weight not given Table 55 (contd) Compound [10] [11] [12] [13] [14] [15] [16] [17] Acenaphthene 0.6/0.7 NR NR 0.6 Anthracene 0.5 NR NR Anthanthrene 0.05-0.08 NR NR Benz[a]anthracene 0.4 ND-0.2 0.14-0.25 <0.1/<O.1 0.3-0.8 0.06-0.15 0.33-1.26 0.1 Benzo[a]pyrene < 0.1 0.17-0.30 0.2/0.4 0.1 0.03-0.05 0.15-0.34 Benzo[b]fluoranthene 0.1/0.2 0.02-0.05 0.1-0.27 Benzo[c]phenanthrene NR NR Benzo[e]pyrene ND-0.1 0.16-0.29 0.2/0.4 0.06-0.16 0.28-0.81 Benzo[ghi]fluoranthene 0.05 NR NR Benzo[ghi]perylene 0.20-0.35 0.06-0.08 0.15-0.28 Benzo[k]fluoranthene ND-0.2a 0.02-0.07 0.15-0.31 Chrysene 0.3-0.7b NR NR Coronene 0.03-0.06 NR NR Cyclopenta[cd]pyrene 0.07-0.13 NR NR Dibenz[a,h]anthracene 0.03-0.05 3.0 < 0.01 0.01-0.02 Fluoranthene 2.9 0.9-1.3 0.32-0.57 1.8/3.0 1.5-7.4 0.22-0.60 0.82-6.17 3.8 Fluorene 1.3/1.7 NR NR 2.0 Indeno[1,2,3-cd]pyrene 0.16-0.29 3.0 0.08-0.15 0.30-0.65 1-Methylphenanthrene 0.3 Perylene 0.1 < 0.1/0.1 0.1-0.3 NR NR Phenanthrene 1.3-1.5 9.9/10 NR NR 14 Pyrene 2.8 1.6-2.3 0.22-0.39 1.6/5.5 2.6-8.5 0.26-1.18 1.41-10.86 2.6 ND, not detected; /, single measurements; NR, not reported; [10] Whole grain oats, Canada (Lawrence & Weber, 1984); [11] Whole-grain rye, Sweden, concentration in µg/kg fresh weight (Larsson, 1982); [12] Wheat grain, United Kingdom (Jones et al., 1989b); [13] Wheat, Finland (Tuominen et al., 1988); [14] Wheat, Canada (Lawrence & Weber, 1984); [15] Breakfast cereal, United Kingdom (Dennis et al., 1991); [16] Bran-enriched cereals, United Kingdom (Dennis et al., 1991); [17] Bolted rye flour, Finland (Tuominen et al., 1988) Analysed by high-performance liquid chromatography or gas chromatography; reference weight not given, unless otherwise specified a Benzofluoranthenes b In sum with triphenylene Table 55 (contd) Compound [18] [19] [20] [21] [22] [23] [24] [25] Acenaphthene NR NR NR Anthracene NR NR NR Anthanthrene NR NR NR Benz[a]anthracene 0.04-0.19 0.64 0.8 0.10-0.14 0.5 0.1 0.4 Benzo[a]pyrene 0.02-0.09 0.43 0.8 0.05-0.15 0.2 0.3 0.8 Benzo[b]fluoranthene 0.02-0.06 0.25 1.2 0.04-0.06 0.5 0.6 1.0 0.05 Benzo[c]phenanthrene NR NR NR 0.7 Benzo[e]pyrene 0.10-0.23 0.35 0.06-0.12 Benzo[ghi]fluoranthene NR NR NR Benzo[ghi]perylene 0.06-0.19 0.39 0.5 0.04-0.21 0.5 0.9 0.6 Benzo[k]fluoranthene 0.03-0.08 0.35 0.6 0.04-0.1 0.1 0.3 0.5 0.08 Chrysene NR NR 1.0 NR 2.0 1.3 0.4 Coronene NR NR NR Cyclopenta[cd]pyrene NR NR NR Dibenz[a,h]anthracene <0.01-011 0.05 <0.01-0.01 Fluoranthene 0.07-0.40 0.66 2.8 0.23-2.03 3.7 0.6 2.5 Fluorene NR NR NR Indeno[1,2,3-cd]pyrene 0.06-0.24 0.84 0.6 0.11-0.25 0.3 0.6 0.5 1-Methylphenanthrene Perylene NR NR NR Phenanthrene NR NR 3 NR 4.2 3.0 2.1 Pyrene 0.04-0.88 0.67 0.23-0.87 NR, not reported; [18] White four, United Kingdom (Dennis et al., 1991); [19] Granary flour, United Kingdom (Dennis et al., 1991); [20] Bread, Netherlands (de Vos et al., 1990); [21] White bread, 1982-83, United Kingdom (Dennis et al., 1991); [22] Noodles, pizza, Netherlands (de Vos et al., 1990); [23] Potato products, Netherlands (de Vos et al., 1990); [24] Rice, macaroni, Netherlands (de Vos et al., 1990); [25] Soups, Netherlands (de Vos et al., 1990) Analysed by high-performance liquid chromatography or gas chromatography; reference weight not given The PAH concentration in rye grown near a highway with high traffic density decreased slightly 7-25 m away from the road (Larsson, 1982). 5.1.5.7 Beverages Benzo [a]pyrene was found at 0.8 µg/kg in coffee powder, 0.01 µg/litre in brewed coffee, 9.51 µg/kg in tea leaves, and 0.02 µg/litre in brewed tea (Lintas et al., 1979). In 40 samples of tea leaves from India, China, and Morocco, the concentration of benzo [a]pyrene was generally 2.2-60 µg/kg, although concentrations up to 110 µg/kg were found in smoked teas (Prinsen & Kennedy, 1978). In samples of whisky and beer, the concentrations of six of 11 PAH (benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [ghi]-perylene, dibenz [a,h]anthracene, and indeno[1,2,3- cd]pyrene) were below or slightly above 0.01 µg/kg. The highest level determined (0.24 µg/kg) was that of pyrene (Dennis et al., 1991). The PAH content of the water used in the preparation of whisky and beer was not described. 5.1.5.8 Vegetable and animal fats and oils The levels of PAH in oil products, butter, and margarine are listed in Table 56. Vegetable oils are reported to be naturally free of PAH, and contamination is due to technological processes like smoke drying of oil seeds or environmental sources such as exhaust gases from traffic. The PAH content of native olive oils was particularly high (Speer et al., 1990). The PAH content of coconut, soya bean, maize, and rapeseed oil was radically reduced during refining, particularly by treatment with activated charcoal (Larsson et al., 1987). This method is now widely used (Dennis et al., 1991). Benzo [a]pyrene was detected in 30 vegetable oils from Italy and France in 1994, including 17 grape-seed oils and one pumpkin-seed oil. The average concentration was 59 µg/kg, and the maximum value was 140 µg/kg. Benzo [b]fluoranthene, benzo [k]fluoranthene, dibenz [a,h]anthracene, and indeno[1,2,3- cd]pyrene were also found in measurable amounts. The source of these high levels was the smoke in drying ovens (State Chemical Analysis Institute, Freiburg, 1995). Lard and dripping were found to contain levels of individual PAH ranging from < 0.01 µg/kg dibenz [a,h]anthracene) to 6.9 µg/kg fluoranthene (Dennis et al., 1991). High PAH levels were found in margarine samples in studies in Finland (Hopia et al., 1986), the Netherlands (Vaessen et al., 1988), New Zealand (Thomson et al., 1996), and the United Kingdom (Dennis et al., 1991) (see Table 56). 5.1.6 Plants PAH with low molecular masses are more readily taken up by vegetation than those with higher molecular masses (Wang & Meresz, 1982). Table 56. Polycyclic aromatic hydrocarbon concentrations (µg/kg) in vegetable oils and related products Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] Acenaphthene NR < 0.02-45 NR NR NR NR 0.29 NR NR < 0.1 -11 Anthracene NR < 0.02-460 <0.1-0.1 ND-4.8 ND-8 NR 0.04-0.92 NR NR < 0.2-5.6 Anthanthrene Trace-0.1 NR NR NR NR NR 0.03-0.53 NR NR < 0.1-2.7 Benz[a]anthracene NR NR 0.7-6.1 ND-6.1 ND 0.30-7.46 NR 0.22-3.98 0.28-0.96 < 0.1-5.2 Benzo[a]fluorene NR < 0.02-130 NR NR ND-2 NR 0.07-3.8 NR NR NR Benzo[a]pyrene Trace-0.3 < 0.02-24 0.5-2.3 ND-4.1 ND 0.29-4.92 0.05-2.2 0.19-6.0 0.17-0.83 < 0.2-5.2 Benzo[b]fluoranthene Trace-0.1 < 0.02-91a NR ND-8.9a ND 0.20-2.39 NR 0.16-3.0 0.09-0.37 < 0.2-9.2 Benzo[b]fluorene NR < 0.02-45 NR NR ND NR 0.03-2.1 NR NR NR Benzo[e]pyrene NR < 0.02-23 0.7-2.4 ND-3.8 ND 0.26-6.06 0.09-2.1 0.42-6.11 0.36-0.87 NR Benzo[ghi]fluoranthene NR < 0.02-1.3 NR NR ND NR 0.14-4.9 NR NR NR Benzo[ghi]perylene NR < 0.02-10 0.5-1.7 ND-4.2 NR 0.06-5.23 0.02-1.4 0.38-5.21 0.17-1.16 < 0.2-10.6 Benzo[k]fluoranthene NR NR NR NR ND 0.24-3.17 NR 0.20-3.40 0.16-0.55 < 0.1-11.4 Chrysene NR NR NR 0.1-8.6b ND 0.39-10.3 NR 0.26-7.36 0.31-0.97 < 0.2-7.5 Coronene NR < 0.02 NR NR NR NR NR NR NR NR Cyclopenta[cd]pyrene NR < 0.02-1.4 NR NR ND NR 0.10-1.1 NR NR NR Dibenz[a,h]anthracene 0.7-1.1 < 0.02-1.1c NR ND-0.2c NR <0.01-0.82 NR 0.05-1.02 0.04-0.11 < 0.1-9.2 Fluoranthene 0.2-7.5 < 0.02-460 1.2-4.8 0.2-18.2 3-15 0.21-12.4 0.52-9.0 0.09-4.50 0.44-1.56 < 0.1-1.6 Fluorene NR < 0.02-200 NR NR ND-7 NR 0.08-1.6 NR NR < 0.2-2.1 Indeno[1,2,3-cd]pyrene Trace-0.5 < 0.02-0.85 0.3-1.7 ND-4.3 NR 0.59-6.78 0.03-1.1 0.49-9.14 0.43-1.17 < 0.2-9.7 Naphthalene NR NR NR NR NR NR NR NR NR < 0.2-52 1-Methylphenanthrene NR < 0.02-190 NR NR NR NR 0.08-1.8 NR NR NR Perylene Trace-0.2 < 0.02-5.9 0.1-0.4 ND-0.9 NR NR 0.02-0.57 NR NR NR Phenanthrene NR 0.09-1400 0.9-1.6 ND-69.4 4-38 NR 0.29-6.0 NR NR < 0.2-4.6 Pyrene 0.2-1.4 < 0.02-330 1.1-4.2 0.1-13.6 2-14 0.58-17.2 0.59-15 0.29-6.03 0.44-1.88 < 0.1-1.7 Table 56 (continued) ND, not detected; /, single measurements; NR, not reported; [1] Corn oil, canola, soya bean oil (Lawrence & Weber, 1984); [2] Corn oil, coconut oil (crude and deodorized), olive oil, soya bean oil, sunflower oil, sesame oil, flaw oil, wheatseed oil (Hopia et al., 1986); [3] Coconut oil (pure) (Sagredos et al., 1988); [4] Various olive oils, safflower oils, sunflower oils, maize germ oils, sesame oil, linseed oil, wheat germ oil (all native) (Speer et al., 1990); [5] Various olive oils (Menichini et al., 1991b); [6] Various unspecified oils (Dennis et al., 1991); [7] Four cooking margarines, seven table margarines (Hopia et al., 1986); [8] Margarine (Dennis et al., 1991); [9] Low-fat spread (Dennis et al., 1991); [10] Margarine (Thomson et al., 1996) Analysed by high-performance liquid chromatography or gas chromatography a Benzo[b+j+k]fluoranthenes b In sum with triphenylene c Dibenz[a,h+a,c]anthracenes In a study of PAH levels in soil (see section 5.1.4), leaf litter, and soil fauna (see section 5.1.7) from a roadside in Brisbane, Australia, vegetation height, soil depth, and distance from the roadside were found to be important in the distribution of PAH. The concentration of PAH in leaf litter declined exponentially with distance from the roadway, few PAH being detectable 30 m away. A decrease in PAH levels with height was found in the roadside vegetation canopy. In leaf litter, fluorene, phenanthrene, fluoranthene, pyrene, chrysene, benzo [k]fluoranthene, and benzo [ghi]perylene were present at concentrations of about 100 µg/kg wet weight. Naphthalene, benz [a]anthracene, benzo [e]pyrene, benzo [a]pyrene, and indeno[1,2,3- cd]pyrene were present at about 50 µg/kg wet weight, whereas anthracene was present at concentrations below 10 µg/kg wet weight. Perylene and dibenz [a,h]anthracene were not detectable. The tree Casuarina littorina contained high levels of pyrene and chrysene (100 µg/kg wet weight each) and benzo [k]fluoranthene (72 µg/kg wet weight); the concentrations of fluoranthene, phenanthrene, and benzo [ghi]-perylene were about 40 µg/kg wet weight. Perylene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, and coronene were not detectable (Pathirana et al., 1994). The benzo [a]pyrene levels in spruce sprouts from a rural area of Germany (Bornhövede, Schleswig-Holstein) decreased from 2.6 µg/kg in 1991 to 1.3 µg/kg in 1993. The concentrations of PAH with low boiling-points significantly decreased between 1991 and 1993; for example, that of fluoranthene decreased from 44 µg/kg in 1991 to 11 µg/kg in 1993, perhaps due to a decrease in coal burning. The levels of phenanthrene, fluoranthene, pyrene, and benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene were about 10 µg/kg; those of benzo [ghi]fluoranthene, benzo [c]phenanthrene, benz [a]anthracene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, benzo [ghi]perylene, and coronene were about 2 µg/kg; and those of anthracene, dibenz [a,h]anthracene, and anthanthrene were < 1 µg/kg. The PAH levels in spruce sprouts from the Saarland, an industrial area in Germany, were about 10 times higher than those in the Bornhöveder area, although these levels also decreased between 1991 and 1993: from 5.9 to 4.1 µg/kg for benzo [a]pyrene and 97 to 58 µg/kg for fluoranthene. the concentrations of pyrene were 40-50 µg/kg, those of benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene were 20 µg/kg, and those of benzo [ghi]perylene, benzo [c]phenanthrene, benz [a]anthracene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, anthanthrene, and coronene were < 10 µg/kg (Jacob & Grimmer, 1994, 1995). In 1994, the PAH levels had decreased further. Overall, a 25% decrease in the PAH levels in spruce sprouts was seen over the previous 10 years (Jacob & Grimmer, 1995). The PAH profiles in spruce sprouts and poplar leaves were reasonably similar in areas with clean air (Bavarian forests) and in industrialized areas (Saarland) of Germany, indicating that one emission source is predominantly responsible for air pollution by PAH. Hard-coal combustion resulted in a characteristic PAH profile (Jacob et al., 1993a). The concentrations of PAH in pine needles from Dübener Heide near Leipzig (Saxony, Germany) were similar to those from the Bornhöveder area (Schleswig-Holstein, Germany), with an average benzo [a]pyrene level of 2.3 µg/kg (Jacob & Grimmer, 1995). Beech leaves from the Harz mountains in Germany contained fluoranthene at a level of 5 µg/kg, whereas the concentrations of phenanthrene, pyrene, benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, anthracene, benz [a]anthracene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, anthanthrene, and coronene were all < 2 µg/kg. Beech sprouts in an industrial area in eastern Germany contained 10-15 times higher levels of PAH, with fluoranthene at about 60 µg/kg, pyrene at about 30 µg/kg, benzo [b]fluoranthene plus benzo [j]-fluoranthene plus benzo [k]fluoranthene at about 10 µg/kg, and anthracene, benz [a]anthracene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, benzo [ghi]perylene, coronene, dibenz [a,h]anthracene, and anthanthrene at < 2 µg/kg (Jacob & Grimmer, 1995). Comparable results were obtained in poplar leaves: those from the Saarland analysed in 1989, 1991, and 1993 had 10 times lower concentrations of PAH than those in Dübener Heide. The concentrations of phenanthrene, fluoranthene, and pyrene were about 20 µg/kg, those of benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene were about 10 µg/kg, and those of anthracene, benz [a]anthracene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, anthanthrene, and coronene were < 5 µg/kg (Jacob & Grimmer, 1995). 5.1.7 Animals 5.1.7.1 Aquatic organisms Aquatic invertebrates are known to adsorb and accumulate PAH from water (see section 4.1.5). The concentrations of PAH in aquatic organisms collected from various sites are listed in Tables 57-64. All of the sampling sites listed in Tables 57-60 were contaminated with industrial effluents, the major components of the PAH profile being benzo [b]fluoranthene, benz [a]anthracene, benzo [a]pyrene, benzo [e]pyrene, fluoranthene, pyrene, and phenanthrene. The average levels of PAH in aquatic organisms from these sites ranged from 1 to 100 µg/kg; the differences in levels generally corresponded to the degree of industrial and urban development and shipping movements. Table 57. Polycyclic aromatic hydrocarbon concentrations (µg/kg dry weight) in bivalves and gastropods; main source, industrial emissions Compound [1] [2] [3] [4] [5] [6] [7] Acenaphthene ND ND 7 2.1/8.8 Anthracens 9 9.0/25 Benz[alanthracene 172 203 3 5-41 25-229 Benzo[alpyrene 12 21 1 8.1 2-8 Trace-28 2.6/2.8 Benzo[b]fluoranthene 23 25 3-30 48-90 Benzo[ejpyrene 17 10 Trace-30 231-356 Benzo[ghilperylene ND 4 5 BenzoUlfluoranthene 1.3 Benzolk]fluoranthene 2.3 Chrysene 209 205 Coronene 4 Dibenzo(a,elpyrene 2 Dibenzo[e,ilpyrane 4 Dbenzo[a,lpyrene Trace Fluoranthene 18 62 7 43-407 300-4992 26/61 Fluorene 2 1.3/6.3 1-Methylphenanthrene 2.9 Naphthalene 15/3 Perylene 8 Phenanthrene 733 462 9 4.4 115-258 55-2542 66/194 Pyrene 85 131 4 32-204 141-3128 23/40 Triphenylene ND ND, not detected; /, single measurement; [1] Whole cooked clam (Mya arenaria); oil-contaminated area (tanker accident), Canada, 1979; concentration in µg/kg wet weight (Sirota & Uthe, 1981); [2] Whole cooked mussel (Mytilus edulls); oil-contaminated area (tanker accident), Canada, 1979; concentration in µg/kg wet weight (Sirota & Uthe, 1981); [3] Whole mussel (Mytilus galloprovincialis); Thermaikos Gulf, Aegean Sea, Greece (agricultural and industrial area); concentration in µg/kg wet weight (Iosifidou et al., 1982); [4] Whole scallop (Amusium pleuronectes); Gulf of Thailand, Thailand; reference weight not given (Hungspreugs et al., 1984); [5] Whole periwinkle (Littorina littorea); moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Borland, 1982); [6] Whole limpet (Patella vulgata); moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [7] Whole snails (Thais haemostoma), Pensacola Bay, USA; creosote contaminated; concentration in µg/kg wet weight (Rostad & Pereira, 1987) High-performance liquid chromatography or gas chromatography Table 58. Polycyclic aromatic hydrocarbon concentrations (µg/kg dry weight) in algae and water plants; main source, industrial emissions Compound [1] [2] [3] [4] [5] [6] Benz[a]anthracene 5 4 31-325 45-431 3-40 Benzo[alpyrene 4 5 Trace-64 Trace-<2 2-20 Benzo[b]fluoranthene 4 5 7-76 6-12 5-31 Benzo[e]pyrene 7 14 Trace-100 Trace-8 8-50 410 Benzo[ghi]perylene 4 79 Fluoranthene 45 32 40-412 15-900 <4-236 Phenanthrene 87 34 31-325 45-431 109-146 Pyrene 36 20 28-286 15-388 <4-224 260 [1] Laminaria saccharins (whole); moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [2] Ceramium rubrum (whole), moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [3] Bladder wrack (Fucus vesiculosus, whole), moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [4] Knotted wrack (Ascophyllum nodosum, whole), moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [5] Toothed wrack (Fucus serratus, whole), moderately polluted parts of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [6] Wakame seaweed, Japan (Obana et al., 1981a) High-performance liquid chromatography or gas chromatography Table 59. Polycyclic aromatic hydrocarbon concentrations (µg/kg wet weight) in lobsters; main source, industrial emissions Compound [1] [2] [3] [4] [5] [6] Acenaphthene ND ND Benz[a]anthracene 684 ND/23 1620-23 400 34-604 762-32 700 17-900 Benzo[a]pyrene 24 0.2/2.6 35-1000 2.0-40 711-1430 27-43 Benzo[b]fluoranthene 24 1 155-2350 6-78 1020-3820 29-835 Benzo[e]pyrene 57 5/8 415-9330 15-165 1550-3600 35-36 Benzo[ghi]perylene ND ND/2 7-493 1.6-31 232-769 10-20 Benzo[k]fluoranthene 7.6 0.3/0.6 43-588 1.6-25 502-955 15-26 Chrysene 445 ND 360-5050 5-79 252-1240 15-24 Fluoranthene 318 ND/0.2 1910-12400 103-545 4220-15 200 68-442 Indeno[1,2,3-cd]pyrene 5 38-855 3-45 486-931 12-40 Phenanthrene 1588 ND Trace-3470 Trace-650 Pyrene 488 ND 730-6710 32-265 2910-13 100 59-333 Triphenylene ND/244 2520-23100 Trace-330 ND, not detected; /, single measurements; [1] Homarus americanus (digestive gland), oil-contaminated area (tanker accident), Canada, 1979 (Sirota & Uthe, 1981); [2] Homarus americanus (tail muscle), oil-contaminated area (tanker accident), Canada, 1979 (Sirota & Uthe, 1981); [3] Homarus americanus (hapatopancreas), Sydney Harbour, near coking plant, Canada (Sirota et al., 1983); [4] Homarus americanus, (tail muscle), Sydney Harbour, near coking plant, Canada (Sirota et al., 1983); [5] Homarus americanus, (digestive gland), Sydney Harbour, near coking plant, Canada, 1982-84 (Uthe & Musial, 1986); [6] Homarus americanus (tail muscle), Sydney Harbour, near coking plant, Canada, 1982-84 (Uthe & Musial, 1986) High-performance liquid chromatography or gas chromatography Table 60. Polycyclic aromatic hydrocarbon levels (µg/kg dry weight) in fish and other aquatic species; main source, industrial emissions Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Acenaphthene 39 Trace-0.9 130 < 25 Acenaphthylene 270 0.1-0.2 Anthracene ND 0.1-0.2 460 < 22 1000 Benz[a]anthracene 22 ND-40 ND-< 0.1 0.1-88 1000 < 21 1-2 800 5 Benzo[a]fluorene 02-0.6 500 Benzo[a]pyrene 7 0.07-8.4 ND-< 0.1 0.1-0.5 570 < 20 ND 8 Benzo[b]fluoranthene <O.1a 28 Benzo[b]fluorene O.1-0.2 Benzo[o]phenanthrene Trace Benzo[e]pyrene 14 ND-< 0.1 0.1-1.6 840 < 25 25 Benzo[ghi]perylene ND-< 0.1 0.2-18 75 < 25 23 Chrysene 61 < 0.1-2.1b 1500 < 22 Dibenz[a,h]anthracene ND-< 0.1c <100 < 25 Fluoranthene 1800 0.1-9.1 1.2-5.6 4800 < 20 13-18 800 48 Fluorene 0.2-2.4 200 < 25 NDc Indeno[1,2,3-cd]pyrene ND-< 0.1 0.3-3.7 150 < 25 1-Methylphonanthrene 85 < 10 Naphthalene 2.5-11 610 < 25 Perylene 6 ND-< 0. 1 Trace-0.2 75 < 20 Phenanthrene 2700 28-15 313 0.1-2.4 0.7-9.1 1400 < 20 32-50 900 71 Pyrene 1500 ND-10.0 0.7-3.7 2300 < 20 10-8 800 39 Triphenylene 800 Table 60 (continued) ND, not detected; [1] Bullhead catfish (Ictalurus nebulosus, whole); Black River, USA, near coking plant; concentration in µg/kg wet weight (Vassilaros et al., 1982); [2] Whole fish (unspecified); Hersey River, USA, creosote polluted; concentration in µg/kg wet weight (Black et al., 1981); [3] Bream (fillet and liver); River Elbe, Germany, industrial region of city of Hamburg (Speer et al., 1990); [4] Dabs (Limanda limanda, muscle, North Sea, United Kingdom, near Beatrice oil platform; concentration in µg/kg wet weight (McGill et al., 1987); [5] English sole (Parophrys vetulus, stomach contents); Mukilteo, USA, near petroleum storage tanks (Malins et al., 1985); [6] English sole (Parophrys vetufus, liver); Mukilteo, USA, near petroleum storage tanks (Malins et al., 1985); [7] Whole starfish (Asterias rubens), moderately polluted areas of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982); [8] Whole holothurians, Toulon, France; urban sewage (Milano et al., 1986); [9] Whole crumb-of-bread sponge (Hafichondria panicea); moderately polluted areas of North Sea coast, Norway, 1978-79 (Knutzen & Sortland, 1982) High-performance liquid chromatography or gas chromatography a Benzo[b+j+k]fluoranthenes b In sum with triphenylene c Dibenz[a,h+a,c]anthracenes Table 61. Polycyclic aromatic hydrocarbon concentratrations (µg/kg dry weight) in bivalves (mussels and clams); background values Compound [1] [2] [3] [4] [5] [6] [7] [8] Acenaphthene NR 24/46 Acenaphthylene NR 34/130 Anthracene 0.7-19 9-15 149-243 36/43 Benz[a]anthracene NR 0.1-0.8 2.9/42 < 1 31/94 Benzo[a]pyrene 4.6-451 3/5 <0.8-2 3.5/8.7 < 1 1.3/26 Benzo[b]fluoranthene 3.0-120 1.5/12 2.5/18 Benzo[c]phenanthrone 5.3-280 3.1/55 < 1 26/94 Benzo[e]pyrene NR 5-25 Benzo[ghi]perylene 3.4-57 5.4/4.2 3 0.4/8.1 Benzo[k]fluoranthene 1.0-43 1-2 2.6/9.6 1.7/17 Chrysene NR 7.6/27 86 Coronene < 10-24 1.3/2.7 0.7/4.6 Dibenz[a,h]anthracene NR 4.7/6.9 2.1/9.6 Fluoranthene 16-288 23/43 8-23 0.7-7.2 11/111 17 47/180 72 Indeno[1,2,3-cd]pyrene ND-9.9 5.9/3.9 0.3/5.7 1-Methylphenanthrene 22-708 Naphthalene NR 5-4 51/120 Perylene 4.2-59 < 5-26 36 Phenanthrene 21-570 7-109 0.1-1.7 12/155 18 108/216 Pyrene 6.6-394 9-77 15-38 0.3-6.6 6.2/62 23 25/109 Triphenylene 7.5-300 7.9/43 27/106 Table 61 (contd) ND, not detected; /, single measurements; NR, not reported; [1] Mussel (Mytilus edulis), Danish, German and Dutch Wadden Sea, 1989 (Compaan & Laane, 1992); [2] Mussel (Mytilus edulis); Finnish archipelago, Finland, 1978-79; concentration in µg/kg wet weight (Rainio et al., 1986); [3] Mussel (Mytilus edulis L.); North Sea coast, Netherlands; concentration in µg/kg wet weight (Boom, 1987); [4] Hard shell clam (Mercenaria mercenaria), Rhode Island (seafood stores), USA; concentration in µg/kg wet weight (Pruell et al., 1984); [5] Softshell clam (Mya arenaria), Coos Bay, Oregon, USA, 1978-79; reference weight not given (Mix & Schaffer, 1983); [6] Clam (Mya mercenaria); Chesapeake Bay, USA, 1984 (Bender & Huggett, 1988); [7] Mussel (Mytilus edulis); Yaquina Bay, USA, 1979-80; concentration in µg/kg wet weight (Mix & Schaffer, 1983); [8] Rangia cuneata; Lake Pontchartrain, USA, 1980; concentration in µg/kg wet weight(McFall et al., 1985) Table 61 (contd) Compound [9] [10] [11] [12] [13] [14] [15] Acenaphthene 16 Acenaphthylene 18 Anthracene < 0.05-3.2 <0.05 Benz[a]anthracene < 1-6 < 10 1.0-1.8 ND-2.3 Benzo[a]pyrene 30-168 < 10 < 0.003-0.02 < 0.004 0.41-1.8 0.40-2.6 1.0 Benzo[b]fluoranthene 1.0-1.8 0.83-1.9 Benzo[c]phenanthrene < 1-9 Benzo[e]pyrene Benzo[ghi]perylene < 1-10 < 0.05-0.3 <0.05 0.53-1.9 0.83-2.3 Benzo[k]fluoranthene < 0.002-0.02 < 0.002 0.29-0.80 0.32-1.2 Chrysene < 0.03-1.4 <0.03 Coronene Dibenz[a,h]anthracene Fluoranthene < 1/52 < 1-370 < 0.04-0.70 Fluorene Indeno[1,2,3-cd]pyrene 1-Methylphenanthrene Naphthalene Perylene < 1-10 < 10-300 < 0.01-0.08 Phenanthrene < 1-15 < 1-60 14 Pyrene 17/165 < 1-450 < 0.03-1.4 <0.03 Triphenylene Table 61 (contd) [9] Rangia cuneaya, Chesapeake Bay, USA, 1984 (Bender & Huggett, 1988); [10] Lampsilus radiata, Elliptio complanatus, Anodonata grandis; Lake George, Heats Bay USA (Heit et al., 1980); [11] Tridacna maxima, Great Barrier Reef, Australia, 1980-82; concentration in µg/kg wet weight (Smith et al., 1984); [12] Clam; Green Island, Great Barrier Reef, Australia, concentration in µg/kg wet weight (Smith et al., 1984); [13] Shortnecked clam; near Miyagi Prefecture, Japan, concentration in µg/kg wet weight (Takatsuki et al., 1985); [14] Mussel; near Miyagi Prefecture; Japan, reference weight not given (Takatsuki et al., 1985); [15] Perna viridis; Gulf of Thailand (mussel farm), Thailand, reference weight not given (Hungspreugs et al., 1984) High-performance liquid chromatography or gas chromatography; Table 62. Polycyclic aromatic hydrocarbon concentrations (µg/kg wet weight) in bivalves (Oysters); background values Compound [1] [2] [3] [4] [5] [6] [7] Acenaphthene 46 < 0.2-2.0 16 Acenaphthylene 36 < 0.4-3.0 Anthracene 44 < 1-40 < 0.08-0.9 < 0.25-4.2 Benz[a]anthracene 9.9 0.3-12 < 1-135 1.1 1.5-10 Benzo[a]pyrene 0.5-1.6 50-285 < 0.01-5 0.6-2.6 0.78 3.5 Benzo[b]fluoranthene 0.3-5.2 < 0.03-6 3.0-20 2.2 Benzo[c]phenanthrene < 1-70 Benzo[e]pyrene < 1-453 2.8-32 Benzo[ghi]perylene O.4-1.2 < 1-73 < 0.05-5 0.87 < 0.20-2.8 Benzo[k]fluoranthene 12 0.1-0.9 < 0.06-5.1 1.2 < 0.01-< 3 Chrysene 58 1.3-14 < 0.1-3 Dibenz[a,h]anthracene < 1-20 < 0.01-< 4 < 0.06 Fluoranthene 80 0.9-94 < 1-450 0.4-22 470 Fluorene 21 0.1-0.8 Indeno[1,2,3-cd]pyrene 1.7 < 0.01-5 1-Methylphenanthrene 3.5 Naphthalene 35 5-48 0.8-7 Perylene < 1-130 Phenanthrene 220 4.9-77 < 1-45 2-38 6.7 Pyrene 200 1.6-50 < 1-645 < 0.4-15 7.0-52 Triphenylene 0.03 Table 62 (continued) [1] Crassostrea virginica, Lake Pontchartrain, USA, 1980 (McFall et al., 1985); [2] Crassostrea virginica; Palmetto Bay (Marina), USA (Marcus & Stokes, 1985); [3] Crassostrea virginica; Chesapeake Bay, USA, 1983-84; concentration in µg/kg dry weight (Bender & Huggett, 1988); [4] Saccostrea cucculata, Mermaid Sound, Australia, 1982 (Kagi et al., 1985); [5] Oyster, Japan (local market); 1977-78 (Obana et al., 1981a); [6] Oyster, near Miyagi Prefecture, Japan; reference weight not given (Takatsuki et al., 1985); [7] Ostrea plicatula; Gulf of Thailand, Thailand; reference weight not given (Hungspreugs et al., 1984) High-performance liquid chromatography or gas chromatography Table 63. Polycyclic: aromatic hydrocarbon concentrations (µg/kg wet weight) in crustacea (lobsters); background values Compound [1] [2] [3] [4] [5] [6] Acenaphthene ND ND Benz[a]anthracene 655 179 9-38 Trace-133 6-79 6-17 Benzo[a]pyrene 18 3.8 0.4-2.1 Trace-2 1.6-8 ND-1.6 Benzo[b]fluoranthene 17 28 3-6.5 Trace-5.3 7-16 ND-0.8 Benzo[e]pyrene ND 170 12-23 ND-22 15-29 ND-3.6 Benzo[ghi]perylene 11 63 1.4-6.8 Trace-2.0 2.4-10 ND-0.8 Benzo[k]fluoranthene 2 4.4 0.8-1.9 Trace-11.6 1.9-8 ND-0.8 Chrysene 140 113 2.5-12 ND-14 2-43 ND Fluoranthene ND 147 46-407 5.5-12 90-162 ND-34 Fluorene ND 194 Indeno[1,2,3-cd]pyrene 22 77 2.1-5.0 ND-3.7 Trace-5 ND-0.8 Phenanthrene ND 1197 20-345 ND-15 Pyrene ND 174 ND-197 ND-5 35-46 ND-22 Triphenylene ND 1373 ND-141 ND-Trace ND, not detected [1] Homarus americanus (digestive gland); Port Hood, Canada, 1979 (Sirota & Uthe, 1981); [2] Homarus americanus (digestive gland); Brown Bank (offshore), Canada, 1979 (Sirota & Uthe, 1981); [3] Homarus americanus (hepatopancreas); Morien Bay and Mira Bay, Canada (Sirota et al.,1983); [4] Homarus americanus (tail muscle); Moran Bayand, Mira Bay, Canada (Sirota et al., 1983); [5] Homarus americanus (digestive gland); Port Morien, Canada, 1982-84 (Uthe & Musial, 1986); [6] Homarus americanus (tail muscle); Port Morien, Canada, 1982-84 (Uthe & Musial, 1986) Analysed by high-performance liquid chromatography or gas chromatography Table 64. Polycyclic aromatic hydrocarbon concentrations (µg/kg wet weight) in fish and other aquatic species (background values) Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] Anenaphthene ND-83 11 7 1-500 Acenaphthylene 43 0.8-24 Anthracene 10 2.0-2.2 ND 20 Benz[a]anthracene 4 4.0-26 1.2 1.6-7.5 20 Benzo[a]fluorene ND Benzo[e]pyrene 0.04-0.84 1 1.9-15 8 Trace-4.5 5 Benzo[b]fluoranthene 3.2-17 Benzo[e]pyrene ND Benzo[ghi]perylene 2.0-14 16 Benzo[k]fluoranthene 2.1-11 Chrysene 6 3 3.4-26 NR Dibenz[a,h]anthracene 1.2-4.13 Fluoranthene 4-95 4 85 9 ND-732 20 Fluorene 8.9 ND-15 1-370 ND Indeno[1,2,3-cd]pyrene ND-15 NR 1-Methylphenanthrene NR Naphthalene 45-215 ND-117 Perylene NR Phenanthrene 8-142 2 157 2.3-35 36 23-43 ND 40 Pyrene 2-62 4 30 31 2.4-74 1.3-9.6 ND Triphenylene 20 ND, not detected; NR, not reported; [1] Various seafish (muscle, liver, gall), Finnish archipelago, Finland, 1979 (Rainio et al., 1986); [2] Edible tissues of various seafish, Arabian Gulf, Iraq (DouAbdul et al., 1987); [3] Whole bullhead catfish (ictalurus nebulosus), Buckeye Lake, USA (Vassilaros et al., 1982); [4] Whole bullhead catfish (Ictalurus, nebulosus;, whole), Black River, USA (West et al, 1985); [5] Whole fish, Hersey River, USA (Black et al., 1981); [6] Whole striped bass (Morone saxatillis); Potomac River, USA (Vassilaros et al., 1982); [7] White suckers (Catastomus commersoni); stomach contents; Lake Erie, USA (Maccubbin et al., 1985); [8] Various fish, Japan, 1970-91 (Environment Agency, Japan, 1993); [9] Fish bought in market, Ibadan, Nigeria; reference weight not given (Emerole et al., 1982); [10] Whole holothurians, France; concentration in µg/kg dry weight (Milano et al., 1986) Analysed by high-performance liquid chromatography or gas chromatography The levels in holothurians from urban sewage were 1-15 mg/kg (Milano et al., 1986). Concentrations of 1-5 mg/kg individual PAH were found in limpets (Patella vulgata) in the North Sea (Knutzen & Sortland, 1982). The PAH concentrations in two species of bivalves in Saudafjorden (Norway) near an iron alloy smelter decreased rapidly with distance from the source, but the compounds could still be detected more than 15 km away. High levels of individual PAH were reported in mussels (Modiolus modiolus), with maximum levels of 57 000 µg/kg benzo [b]fluoranthene, 25 000 µg/kg benz [a]anthracene, 23 000 µg/kg benzo [e]pyrene, 21 000 µg/kg benzo [a]pyrene, 20 000 µg/kg fluoranthene, 8200 µg/kg pyrene, 6000 µg/kg benzo[ghi]perylene, 4000 µg/kg perylene, 2900 µg/kg benzo [a]fluorene, 2300 µg/kg benzo [b]fluorene, 2200 µg/kg dibenz [a,h]anthracene, 2000 µg/kg benzo [c]phenanthrene, 1100 µg/kg phenanthrene, 524 µg/kg anthracene, and 360 µg/kg anthanthrene (Bjrseth, 1979). A very high level of anthracene (243 µg/kg) was found in mussels (Mytilus edulis L.) in the North Sea near the Dutch coast (Boom, 1987). Mussels in the USA frequently contained up to 500 µg/kg of individual PAH (Heit et al., 1980; Mix & Schaffer, 1983). The levels of PAH in pooled mussel samples in 1986, 1988, and 1990 in Germany were about 10 µg/kg for fluoranthene, pyrene, chrysene plus triphenylene, benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, and benzo [e]pyrene and < 4 µg/kg for benzo [ghi]fluoran-thene plus benzo [c]phenanthrene, benz [a]anthracene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, anthanthrene, and coronene. The levels were high in the winter months and low in summer, with minima in June and April. The authors concluded that this seasonal variation was due to more intensive metabolic activity (Jacob & Grimmer, 1994). During 1978-79, the average total PAH concentrations in two subpopulations of softshell clams were 555 µg/kg in the industrialized bayfront area of Coos Bay, Oregon, and 76 µg/kg in a more remote environment. During 1979-80, low-molecular-mass, readily water-soluble PAH were one or two orders of magnitude more concentrated then high-molecular-mass, less water-soluble PAH in mussels (M. edulis) (Mix & Schaffer, 1983). Individual PAH levels of 1-20 mg/kg were found in the hepatopancreas of lobsters (Homarus americanus) in the south arm of Sydney Harbour, Canada, near a coking plant (Sirota et al., 1983), and levels of the same order of magnitude were found in the digestive gland (Uthe & Musial, 1986). The levels in digestive gland, tail muscle, and hepatopancreas from lobsters from other areas of Canada were 100-1000 µg/kg (Sirota & Uthe, 1981; Sirota et al., 1983; Uthe & Musial, 1986). High PAH levels were found in oysters (Crassostrea virginica) in Chesapeake Bay, USA, with maximum levels of 650 µg/kg pyrene, 450 µg/kg benzo [e]pyrene, 450 µg/kg fluoranthene, 290 µg/kg benzo [a]pyrene, 130 µg/kg benz [a]anthracene, 130 µg/kg perylene, 73 µg/kg benzo [ghi]-perylene, 70 µg/kg benzo [c]phenanthrene, 48 µg/kg naphthalene, 45 µg/kg phenanthrene, 40 µg/kg anthracene, and 20 µg/kg dibenz [a,h]anthracene. The levels of PAH in clams (Rangia cuneata) from Chesapeake Bay were 170 µg/kg benzo [a]pyrene, 170 µg/kg pyrene, 52 µg/kg fluoranthene, 15 µg/kg phenanthrene, 10 µg/kg perylene, 10 µg/kg benzo [ghi]perylene, 9 µg/kg benzo [c]phenanthrene, and 6 µg/kg benz [a]anthracene (Bender & Huggett, 1988). Phenanthrene was found at 15 mg/kg in lampreys (Pteromyzon sp.) in the Hersey River, USA, which was polluted with creosote used for wood preservation (Black et al., 1981). The viviparous blenny (Zoarces viviparus) fish contained 0.06 µg/kg benzo [a]pyrene and 0.2-3.9 µg/kg phenanthrene and fluoranthene; the concentrations of other PAH were below the detection limit (0.01 µg/kg). In bream (Abramis brama) the levels were < 0.01-0.15 µg/kg benzo [a]pyrene and 1.3-15 µg/kg phenanthrene. Mussels (Mytilus sp.) were shown to accumulate PAH and were thus a better marker for PAH contamination (Jacob & Grimmer, 1994, 1995). The concentrations of individual PAH in English sole (Paraphrys vetulus) taken from near petroleum storage tanks were 1-5 mg/kg (Malins et al., 1985). 5.1.7.2 Terrestrial organisms The liver of wild deer mice (Peromyscus maniculatus) trapped at a PAH-contaminated site in South Carolina, USA (Whidbey Island Naval Air Station) had levels of PAH ranging from 0.075 for benzo [b]fluoranthene to 4.6 mg/kg for benz [a]anthracene. Acenaphthylene, acenaphthene, fluorene, benz [a]-anthracene, chrysene, benzo [b]fluoranthene, benzo [k]fluoranthene, dibenz [a,h]anthracene, and indeno[1,2,3- cd]pyrene were detected. Liver from mice at an uncontaminated reference site contained measurable amounts of only benz [a]anthracene (0.55 mg/kg) and acenaphthylene (2.2 mg/kg) (Dickerson et al., 1994). In a study of PAH levels in terrestrial organisms from a roadside in Brisbane, Australia, 16 PAH were determined: naphthalene, fluorene, phenanthrene, anthracene, fluoranthene, pyrene, benz [a]anthracene, chrysene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, perylene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, and coronene. In the beetle Laxta granicollis, pyrene and benzo [ghi]perylene were present at the highest levels, at 20 µg/kg wet weight each; phenanthrene and fluoranthene were present at about 10 µg/kg; and the concentrations of other PAH were < 5 µg/kg. Naphthalene, anthracene, dibenz [a,h]anthracene, and coronene were not detected. Fluorene, at a concentration of 11 µg/kg wet weight, was the most abundant PAH in the beetle Platyzosteria nitida; the concentrations of other PAH were < 5 µg/kg; whereas naphthalene, dibenz [a,h]anthracene, and coronene were not detected. In millipedes (myriapods), benzo [k]fluoranthene was the most abundant PAH (19 µg/kg wet weight); the pyrene concentration was 12 µg/kg; those of other PAH were < 5 µg/kg wet weight; and dibenz [a,h]anthracene and coronene were not detected. In centipedes (Myriaod sp.), no PAH were detected. In slugs (Arion ater), benzo [k]fluoranthene showed the highest concentration, at 19 µg/kg wet weight; the pyrene and naphthalene levels were about 10 µg/kg; those of other PAH were < 5 µg/kg wet weight; and anthracene, perylene, dibenz [a,h]anthracene, and coronene were not detected. In earthworms (Lumbricus terrestris), benzo [ghi]perylene was the most abundant PAH (28 µg/kg wet weight); phenanthrene, fluoranthene, pyrene, chrysene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene were present at about 10 µg/kg; and naphthalene and dibenz [a,h]anthracene were not detected (Pathirana et al., 1994). The PAH concentrations in earthworms did not seem to be affected by the location in which the worms lived, but those in the faeces showed a significant dependence on location. In a survey of earthworm faeces from the Bornhöveder Lake district in 1988, the concentrations of phenanthrene, fluoranthene, pyrene, and benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene were in the range of 45 µg/kg; those of benz [a]anthracene, chrysene plus triphenylene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, and benzo [ghi]perylene were about 20 µg/kg; and those of anthracene, benzo [ghi]fluoranthene plus benzo [c]phenanthrene, dibenz [a,h]anthracene, anthanthrene, and coronene were < 5 µg/kg. Earthworm faeces in the Saarland contained 250-770 µg/kg benzo [a]pyrene, and Allolobophora longa earthworm faeces from a highly industrialized region of eastern Germany (Halle, Leipzig) contained even higher concentrations: 37-2100 µg/kg benzo [a]pyrene and 36-1700 µg/kg benzo [e]pyrene. The faeces of the earthworm Lumbricus terrestris contained 4.6-55 µg/kg benzo [a]pyrene and 6.5-50 µg/kg benzo [e]pyrene (Jacob & Grimmer, 1995). In insects near the Hersey River, USA, the maximum concentrations of PAH were 5500 µg/kg phenanthrene, 2900 µg/kg benz [a]anthracene, and 730 µg/kg benzo [a]pyrene (Black et al., 1981). The lipid fraction of liver from herring gulls (Larus argentatus) from Pigeon Island and Kingston, Ontario, Canada, contained 0.15 µg/kg anthracene, 0.082 µg/kg fluoranthene, 0.076 µg/kg pyrene, 0.05 µg/kg naphthalene, 0.044 µg/kg fluorene, 0.038 µg/kg acenaphthene, and 0.038 µg/kg benzo [a]pyrene (Environment Canada, 1994). The concentrations of PAH in pooled samples taken from the eggs of herring gulls (Larus argentatus) on the German North Sea islands Mellum and Trischen during 1992-93 were below the limit of detection, except for that of phenanthrene, which was 1 µg/kg wet weight (Jacob & Grimmer, 1994). 5.2 Exposure of the general population Possible sources of nonoccupational exposure to PAH are: - polluted ambient air (main emission sources: vehicle traffic, industrial plants, and residential heating with wood, coal, mineral oil) (see section 5.1.1); - polluted indoor air (main emission sources: open stoves and environmental tobacco smoke) (see Table 65); - tobacco smoking (see Table 66); - contaminated food and drinking-water (see sections 5.1.5 and 5.1.2.3) - use of products containing PAH (coal-tar skin preparations and coal-tar-containing hair shampoos); - ingestion of house dust; and - dermal absorption from contaminated soil and water. 5.2.1 Indoor air, tobacco smoke, and environmental tobacco smoke PAH are found in indoor air (Table 65) mainly as a result of tobacco smoking and residential heating with wood, coal, or, in some developing countries, rural biomass. The PAH levels in indoor air usually range from 1 to 50 ng/m3. The most abundant PAH were phenanthrene and naphthalene, with levels of up to 2300 ng/m3. Homes with gas heating systems had higher indoor levels than those with electric heating systems (Chuang et al., 1991), and even higher levels were detected in indoor air near open fireplaces (Alfheim & Ramdahl, 1984). Airtight residential wood-burning stoves seemed to have a minor effect on the indoor air concentration of PAH (Alfheim & Ramdahl, 1984; Traynor et al., 1987), but in homes with non-airtight wood stoves, 2-46 times higher PAH concentrations were found during heating periods than during periods without heating (Daisey et al., 1989). Emissions from unvented kerosene heaters can significantly affect indoor air quality in mobile homes, with a maximim value for naphthalene of 2300 ng/m3. Four of eight heaters investigated emitted PAH-containing particles at levels that exceeded the USA ambient air standards for airborne particles, with a 50% cutoff at the aerodynamic diameter of 10 µm. When the kerosene heaters were in operation, the concentrations of carcinogenic PAH (with four rings or more) in the mobile homes were increased by 10-fold (Mumford et al., 1991). Table 65. Polycyclic: aromatic hydrocarbon concentrations (ng/m3) in indoor air; main source, residential heating Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Acenaphthene NR 589-1649 Acenaphthylene NR 60-592 Anthracene 5-30 408 5-15 84 NR 9.9-11 Benz[a]anthracene 3-9 2-6 3-13 145 NR 0.9-5.5 Benzo[a]pyrene 13-370 0.3-12 1-7 3-23 150 < 0.009-1.34 0.34-3.5 2.0-490 8.5-29 Benzo[b]fluoranthene < 0.007-0.68 0.17-3.8 1.4-420 5.6-21 Benzo[e]pyrene < 0.06-1.36 Benzo[ghi]perylene 14-340 0.4-10 1-7 3-30 125 < 0.01-6.20 0.37-3.7 2.8-450 0.4-7.5 Benzo[k]fluoranthene 5-150 0.07-7 0.6-3 2-10 63 0.005-0.48 0.07-1.9 0.67-200 0.7-21 Chrysene 2-12 3-6 4-13 115 NR Coronene NR Cyclopenta[cd]pyrene NR Dibenzo[a,e]pyrene NR Dibenz[a,h]anthracene NR 3.3-25 Fluoranthene 16-56 16-24 16-50 208 0.07-1.18 87-268 Fluorene NR Indeno[1,2,3-cd]pyrene 20-560 1-16 1-8 3-22 130 < 0.02-3.54 1.1-6.1 3.9-740 2.3-11 Phenanthrene 120-400 120-200 140-290 555 NR 31-64 Pyrene 0.02-1.53 1.0-20 ND, not determined; NR, not reported; /, single measurements; [1] Wood-burning open fire-place, Netherlands (Slooff et al., 1989); [2] Wood in multi-burner, Netherlands (Slooff et al., 1989); [3] Coal, Netherlands (Slooff et al., 1989); [4] Briquettes, Netherlands (Slooff et al., 1989); [5] 'Icopower' heating, Netherlands (Slooff et al., 1989); [6] Wood heating in seven homes, USA (Daisey et al., 1989); [7] Wood burning in one home; volume, 236 m3; airtight stove, Truckee, USA, (elevation, 1800 m) (Traynor et al., 1987); [8] Wood burning in one home; volume, 236 m3; non-airtight stove, Truckee, USA (elevation, 1800 m) (Traynor et al., 1987); [9] Wood burning in one home with four different heaters, USA (Knight & Humphreys, 1985) Analysed by high-performance liquid chromatography or gas chromatography Table 65 (contd) Compound [10] [11] [12] [13] [14] [15] [16] [17] Acenaphthene NR 1-258 Acenaphthylene 10-120 21/68 25-36 NR 1-753 Anthracene 1.5-15 4.2-5.9 NR 0.1-80 Benz[a]anthracene 0.24-3.4 0.72/2.8 0.55-1.0 ND-3.81 25 100 1000 4000 5-1021 Benzo[a]pyrene 0.28-3.3 0.24/2.0 0.54-1.0 ND-4.13 14 700 600 3100 8-1645 Benzo[b]fluoranthene NR 2-930 Benzo[b]pyrene 0.33-10 1.4-3.0 NR 5-1106 Benzo[ghi]perylene 0.32-2.5 0.22/3.7 0.72-1.0 ND-5.4 4-802 Benzo[k]fluoranthene ND-7.81a 4-824 Chrysene 0.58-7.2 1.5/3.1 1.4-2.2 0.18-8.61 7-1439 Coronene 0.31-1.4 0.07/2.3 0.55-0.58 ND-4.75 NR Cyclopenta[cd]pyrene 0.18-2.0 0.49/4.2 0.36-0.59 ND-2.38 10 700 400 5600 NR Dibenzo[a,e]pyrene NR 11 700 600 200 NR Dibenz[a,h]anthracene NR 8-958 Fluoranthene 6.2-23 16/11 11 2.4-37.4 5-1095 Fluorene NR 3-275 Indeno[1,2,3-cd]pyrene 0.24-1.8 0.15/1.3 0.48-0.79 ND-3.53 8400 500 2000 4-670 5-Methylcholanthrene NR 7300 200 200 NR Naphthalene 750-2200 2300/950 1200-1600 NR NR Phenanthrene 55-210 48/34 93-110 9.2-210 3-667 Pyrene 3.6-17 9.7/13 6.9-7.6 1.4-18.1 7-850 [10] Gas or electridy, USA (Wilson & Chuang, 1991); [11] Kerosene; unvented heaters in mobile homes, Apex, USA (Mumford et al., 1991); [12] Various heating in eight homes, Columbus, USA (Chuang et al., 1991); [13] Various heating in 33 homes, USA (Wilson et al., 1991); [14] Smoky coal, Xuan Wei, China (Mumford et al., 1987); [15] Smokeless coal, Xuan Wei, China (Mumford et al., 1987); [16] Wood, Xuan Wei, China (Mumford et al., 1987); [17] Various cooking fuels (cattle dung, wood, kerosene, liquid petroleum gas) in 60 homes, India (Raiyani et al., 1993b) a Sum of benzofluranthenes Table 66. Polycyclic aromatic hydrocarbon concentrations (ng/m3 in indoor air; main source, environmental tobacco smoke Compound [1] [2] [3] [4] [5] [6] Acenaphthene 2.5 36 Acenaphthylene 14 177 Anthracene 2.8 25 1.5 < 1 Anthanthrene 0.5 1.5 < 1 2.5 3 Benz[a]anthracene 1.3 12 15 13 Benzo[a]fluorene 5.5 39 Benzo[a]pyrene 1.8 7.3 14 4.5 0.04-0.16 22 Benzo[b]fluoranthene 0.06-0.08 Benzo[b]fluorene 2.5 Benzo[e]pyrene 2.3 7.1 11 4.5 18 Benzo[ghi]fluoranthene 4.3 18 8.5 14 Benzo[ghi]perylene 2.5 5.8 7 2 0.09-0.36 17 Benzo[k]fluoranthene 0.02-0.06 Coronene 2.0 3.1 Fluoranthene 14 41 5 16 99 Indeno[1,2,3-cd]pyrene 2.3 5.8 1 1.5 0.13-0.45 1-Methylphenanthrene 6.6 38 < 1 3.5 Perylene 0.5 0.8 4 2.5 11 Phananthrene 38 168 3 1 Pyrene 13 32 13 21 66 [1] Office room (volume, 88 m3; ventilation, 176 m3/h; background sample after weekend, Finland; vapour and particulate phase (Salomaa et al., 1988); [2] Office room (Volume, 88 m3; ventilation, 176 m3/h; 6 h; 96 cigarettes, American type, 10 different brands, both medium- and low tar, Finland; vapour and particulate phase (Salomaa et al., 1988); [3] House in a forest (room volume, 65 m3; air exchange, 2.0-2.3 turnovers/h); background sample, Norway (Alfheim & Ramdahl, 1984); [4] House in a forest (room volume, 65 m3; air exchange, 2.0-2.3 turnovers/h); with tobacco smoking, Norway (Alfheim & Ramdahl, 1984); [5] House in a residential, wooded area of Truckee, USA (elevation, 1800 m); volume, 236 m3; no stove (Traynor et al., 1987); [6] Model room (volume, 36 m3); one air exchange/h, smoking of five cigarattes/h (Ministry of Environment, 1979)) High-performance liquid chromatography or gas chromatography; concentration of particulate phase, unless otherwise stated Emissions from coal and wood combustion in open fires for cooking purposes in unvented rooms in Xuan Wei County, China, contained extremely high PAH concentrations (see also section 8). The highest concentration (benzo [a]pyrene at 15 000 ng/m3) was measured in fumes from smoky coal combustion. Coal combustion in open fires in Xuan Wei homes emitted 15 µg/m3 of carcinogenic PAH, while wood combustion emitted 3.1 µg/m3 (Mumford et al., 1987). Cooking with rural biomass in open fires also led to high PAH levels in indoor air, as measured in rural Indian households. Benzo [a]pyrene was measured at a concentration of about 4 µg/m3 during the cooking period, which occupied about 10% of the household activities over the year. The cooking fuels included baval, neem, mango, rayan, and crop residues (Smith et al., 1983). The total release of PAH into indoor air from this source is unknown but may be of major importance, especially in developing countries. Very low PAH emissions were found when liquid petroleum gas was used as a fuel for cooking (Raiyani et al., 1993b). In contrast, the PAH content of kitchen air in Berlin, in the industrialized part of Germany, was similar to that encountered in ambient air (Seifert et al., 1983). House dust may be another important source of indoor pollution with PAH. In a study of the homes of four smokers and four nonsmokers in Columbus, Ohio, USA, the sum of the concentrations of naphthalene, acenaphthylene, acenaphthene, fluorene, phenanthrene, anthracene, retene, fluoranthene, pyrene, benz [a]anthracene, chrysene, cyclopenta [cd]pyrene, benzo [b]fluoranthene, benzo [j]fluoranthene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, benzo [ghi]perylene, and coronene in house dust and in soil from the entryway, the pathway, and the foundation of the houses was 16-580 mg/kg. The concentrations in house dust correlated well with those in the entryway soil samples, and a weaker correlation was found with the pathway soil samples, but the relationships were not statistically significant (Chuang et al., 1995). A special source of exposure to PAH is wood-heated saunas. The highest concentrations were found in a smoke sauna, the second highest in a preheated sauna where the flues were closed before use, and the lowest concentrations in a sauna heated by continuous burning of wood. Pyrene, fluoranthene, benz [a]anthracene, and phenanthrene were present at the highest levels (100-330 µg/m3 air); other PAH were present at < 50 µg/m3. The concentrations decreased from benzo [e]pyrene > benzo [a]pyrene > benzo [a]fluorene > anthra-cene > benzo [b]fluorene > fluorene (Häsänen et al., 1983). The protocol of a study of total human environmental exposure included direct monitoring of exposure to benzo [a]pyrene by inhalation and ingestion during three periods of 14 days. The range and magnitude of dietary exposure (2-500 ng/day) was much greater than that by inhalation (10-50 ng/day). The levels of benzo [a]pyrene in indoor air were closely correlated with the ambient levels in most homes (Waldman et al., 1991). Indoor air concentrations of individual PAH due mainly to cigarette smoke are shown in Table 66, and the levels in mainstream and sidestream smoke of cigarettes are listed in Table 67. The average PAH levels ranged from 1 to 50 ng per cigarette, and the major components were phenanthrene, naphthalene, benzo [a]pyrene, benzo [e]pyrene, fluoranthene, and pyrene. Sidestream smoke was found to contain 10 times more PAH than mainstream smoke. The levels in sidestream smoke were 42-2400 ng per cigarette (Grimmer et al., 1987). The PAH concentrations in the mainstream smoke from filter cigarettes increased with increasing puff volume (Funcke et al., 1986). In a pilot study in Columbus, Ohio, USA, naphthalene was the most abundant PAH; environmental tobacco smoke appeared to be the most significant source of indoor pollution (Chuang et al., 1991). Table 67. Concentrations of selected polycyclic aromatic hydrocarbons in cigarette smoke Compound Mainstream smoke Sidestream smoke (µg/100 cigarettes) (µg/100 cigarettes) Anthracene 2.3-23.5 Anthanthrene 0.2-2.2 3.9 Benz[a]anthracene 0.4-7.6 Benzo[b]fluoranthene 0.4-2.2 Benzo[b]fluoranthene 0.6-2.1 Benzo[k]fluoranthene 0.6-1.2 Benzo[ghi]fluoranthene 0.1-0.4 Benzo[a]fluorene 4.1-18.4 75.0 Benzo[b]fluorene 2.0 Benzo[ghi]perylene 0.3-3.9 9.8 Benzo[c]phenanthrene Present Benzo[a]pyrene 0.5-7.8 2.5-19.9 Benzo[e]pyrene 0.2-2.5 13.5 Chrysene 0.6-9.6 Coronene 0.1 Dibenz[a,h]anthracene 0.4 Dibenzo[a,e]pyrene Present Dibenzo[a,h]pyrene Present Dibenzo[a,i]pyrene 0.17-0.32 Dibenzo[a,l]pyrene Present Fluoranthene 1.0-27.2 126.0 Fluorene Present Indeno[1,2,3-cd]pyrene 0.4-2.0 5-Methylcholanthrene 0.06 Perylene 0.3-0.5 3.9 Phenanthrene 8.5-62.4 Pyrene 5.0-27 39.0-101.0 Triphenylene Present 1-Methylphenanthrene 3.2 Adapted from International Agency for Research on Cancer (1985) In studies in eight healthy male smokers, aged 20-40 years, the benzo [a]pyrene intake from the smoking of 20 cigarettes per day was calculated to be 150-750 ng/d, assuming a deposition rate for particulate matter of 75% (Scherer et al., 1990). The total concentration of 14 PAH (fluoranthene, pyrene, benzo [a]fluorene, benz [a]anthracene, chrysene, benzo [b]fluoranthene, benzo [j]fluoranthene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, perylene, dibenz [a,h]-anthracene, benzo [ghi]perylene, and anthanthrene) measured in a 36-m3 room into which sidestream smoke from five German cigarettes was introduced every hour, with one air change per hour, was 429 ng/m3. Assuming that the daily inhalation volume for adults is 18 m3 and that 20 h/d are spent indoors, the volume of indoor air inhaled daily is 18 m3 × 20/24 = 15 m3. Thus, passive smokers are exposed daily to 15 × 429 = 6435 ng PAH, including 15 × 22 = 330 ng benzo [a]pyrene (Ministry of Environment, 1979). An intake of 11 ng benzo [a]pyrene was estimated in another study on the basis of an assumed breath volume of 0.5 m3/h , a deposition rate for particulate matter of 11%, and an exposure time of 8 h, after monitoring in an unventilated, 45-m3, furnished room (Scherer et al., 1990). 5.2.2 Food Smoked and barbecued food in particular can contain PAH (Grimmer & Düvel, 1970; McGill et al., 1982; de Vos et al., 1990; Menichini et al., 1991b; see also section 5.1.5 and Tables 51-56). Preparation of food with contaminated drinking-water (see section 5.1.2.3) may also lead to exposure to PAH. In 1989 and 1990, the levels of naphthalene and alkylated derivatives, acenaphthene, acenaphthylene, fluorene, phenanthrene, anthracene, fluoran-thene, 1-methylphenanthrene, pyrene, benz [a]anthracene, chrysene, benzo [b]fluoranthene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, perylene, indeno[1,2,3- cd]pyrene, dibenz [a,h]anthracene, and benzo [ghi]-perylene were measured in salmon, herring, cod, rockfish, and halibut in the area of the Gulf of Alaska where oil spilled from the tanker Exxon Valdez. As only the sums of the concentrations were considered, there was no apparent difference from those in fish samples taken from unpolluted control sites in 1989. In 1990, slightly elevated PAH concentrations were found at the polluted sampling site. Nevertheless, the fish from the area were considered to be safe for human consumption by these investigators (Saxton et al., 1993). In another special exposure situation, the average daily PAH intake of the inhabitants of Kuwait due to consumption of seafood after the war in the Persian Gulf was calculated to be 0.23 µg/day on the basis of the concentrations monitored in local fish and shrimps (Saed et al., 1995). 5.2.3 Other sources Benzo [a]pyrene was detected in coal-tar-containing hair shampoos at levels of 7000-61 000 µg/kg, and a tar bath lotion contained 150 000 µg/kg benzo [a]pyrene. No PAH were detected in hair shampoos made from wood tar (State Chemical Analysis Institute Freiburg, 1995). PAH are absorbed from coal-tar shampoos through the skin during hair washing. Exposure during one washing with this type of shampoo, which contains benzo [a]pyrene at 56 mg/kg, for anti-dandruff therapy results in absorption of 0.45 µg/kg body weight, assuming 20 g coal-tar, 70 kg body weight, and 3% dermal absorption (van Schooten et al., 1994; see also section 8). 5.2.4 Intake of PAH by inhalation Estimates of PAH intake from air are summarized in Table 68. In an assessment of the risk for cancer due to air pollution in Germany, the average volume of air inhaled during heavy work was assumed to be 140 m3 per person per week. The maximum intake of airborne benzo [a]pyrene per week was thus estimated to be 0.21 µg/week in rural areas, 0.84 µg/week in industrial areas, and 7 µg/week near emission sources (State Committee for Air Pollution Control, 1992). On the basis of an average inhalation of 15 m3 air per day, exposure to benzo [a]pyrene was calculated to be 0.05 µg/d. In industrial areas, the exposure was calculated to be four times higher (0.19 µg/d) (Raiyani et al., 1993a). 5.2.5 Intake of PAH from food and drinking-water Estimates of PAH intake from food are shown in Table 69. The values for benzo [a]pyrene range from 0.14-1.6 µg/d. The total dietary intake of some PAH in the United Kingdom was estimated to be (µg/person per day): 1.1 for pyrene, 0.99 for fluoranthene, 0.50 for chrysene, 0.25 for benzo [a]pyrene, 0.22 for benz [a]anthracene, 0.21 for benzo [ghi]perylene, 0.18 for benzo [b]fluoranthene, 0.17 for benzo [e]pyrene, 0.06 for benzo [k]fluoranthene, and 0.03 for dibenz [a,h]anthracene. The major contributors of PAH to the total dietary intake appeared to be oils and fats, with 28% from butter, 20% from cheese, and 17% from margarine, in respective dietary survey groups; cereals provided 56% from white bread and 12% from flour. The oils and fats had the highest individual PAH levels. Although cereals did not contain high levels of individual PAH, they were the main contributor by weight to the total in the diet. Fruits and vegetables contributed most of the rest of the PAH in the diet, while milk and beverages were of minor importance. Smoked meat and smoked fish made very small contributions to the food groups to which they belonged, which themselves were not major components of the diet (Dennis et al., 1983). Table 68 Estimated intake of polycyclic aromatic hydrocarbons (µg/day per person) from ambient air Compound [1] [2] [3] [4] [5] [6] [7] [8] [9] Anthracene 0.005 0.001 Anthanthrene 0.015 Benz[a]anthracene 0.030 0.013 Benzo[a]pyrene 0.01-0.03a 0.0025-0.025 0.025 0.034a 0.0095-0.0435 0.004a 0.017 0.03-0.05 0.0005-0.20 0.02-0.12b 0.06-1.0c Benzo[b]fluoranthene 0.060 0.029 Benzo[b]fluorene 0.002 0.002 Benzo[e]pyrene 0.035 0.022 Benzo[ghi]perylene 0.030 0.027 Benzo[j]fluoranthene 0.010 Benzo[k]fluoranthene 0.015 0.015 Chrysene 0.035 Coronene 0.025 Dibenz[a,h]anthracene 0.020 0.004 Fluoranthene 0.040 0.016 Fluorene 0.0005 Indeno[1,2,3-cd]pyrene 0.030 0.024 Perylene 0.015 0.003 Phenanethrene 0.200 0.007 Pyrene 0.040 0.017 Triphenylene 0.220 Table 68 (continued) [1] Germany (maximum concentrations) (State Committee for Air Pollution Control, 1992); [2] Italy (Menichini, 1992a); [3] Netherlands (maximum concentrations) (Guicherit & Schulting, 1985); [4] United Kingdom (maximum concentrations) (Butler & Crossley, 1979); [5] USA (Santodonato et al., 1980); [6] USA (WHO, 1987); [7] Japan (maximum concentrations) (Matsumoto & Kashimoto, 1985); [8] China (Chen et al., 1980); [9] India (Chakraborti et al., 1988) a Rural areas b Industrial areas c Near emission source Table 69. Estimated intake of polycyciic aromatic hydrocarbons (µg/day per person, maximum values) from food Compound [1] [2] [3] [4] [5] [6] [7] [8] Anthracene 5.6 Anthanthrene 0.30 Benz[a]anthracene 0.14 Benzo[a]pyrene 0.36 0.14-1a 0.1-0.3b 0.12-0.42 0.5 0.5 0.48 0.16-1.6 0.2c Benzo[b]fluoranthene 1.0 Benzo[ghi]perylene 7.6 0.3 0.9 Benzo[j]fluoranthene 0.90 Benzo[k]fluoranthene 0.30 Chrysene 0.90 5.0 Coronene 0.09 Dibenz[a,h]anthracene 0.10 Fluoranthene 4.3 3 10 Indeno[1,2,3-cd]pyrene 0.31 0.4 <0.3 Perylene 0.20 Phenanethrene 2.0 Pyrene 4.0 5.1 [1] Austria (Pfannhauser, 1991); [2] Germany (State Committee for Pollution Control, 1992); [3] Italy (Menichini, 1992a); [4] Netherlands (de Vos et al., 1990); [5] Market basket study, Netherlands (Vaessen et al., 1984); [6] Duplicate diet study, Netherlands (Vaessen et al., 1984); [7] United Kingdom (Dennis et al., 1983); [8] USA (Santodonato et al., 1980) a Concentration in µg/week b Adult non-smoker (70 kg) c Mean concentration In Sweden, the annual intake per person of the sum of fluoranthene, pyrene, benz [a]anthracene, chrysene, triphenylene, benzo [b]fluoranthene, benzo [j]-fluoranthene, benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, and indeno[1,2,3- cd]pyrene was about 1 mg. Cereals again seemed to be the main contributor (about 34%), followed by vegetables (about 18%) and oils and fats (about 16%). Although smoked fish and meat products had the highest PAH levels, they made a modest contribution since they are minor components of the usual Swedish diet (Larsson, 1986). 5.3 Occupational exposure PAH have been measured in the air at various workplaces. Studies in which measurements were reported only as the benzene-soluble fraction or some other summarizing parameter affected mainly by PAH are not covered because they do not refer to individual substances. The presence of PAH metabolites in biological samples (urine, blood) from workers has been used as a biomarker, and 1-hydroxypyrene seems to be a suitable marker in some workplaces (see section 8.2.3). No data were available on occupational exposure during production and use. Occupational exposure to PAH occurs by both inhalation and dermal absorption. In coke-oven workers, 75% of their exposure to total pyrene and 51% of that to benzo [a]pyrene occurs by cutaneous transfer (Van Rooij et al., 1993a; see also section 6). The exposure of workers due to deposition of airborne pyrene on the skin, detected in wipe samples, can be summarized as follows: in refineries, < 0.0045 µg/cm2 (detection limit), 26 samples below detection limit; in hot-mix asphalt facilities, < 0.0045 µg/cm2, 25 samples below detection limit; during paving, < 0.13-0.31 µg/cm2 found in two of nine samples (assuming a body area of 1.8 m2, equivalent to 5600 µg/person per day); in asphalt roofing manufacture, < 0.0045-0.0091 µg/cm2 found in 1 of 29 samples (assuming a body area of 1.8 m2, equivalent to 170 µg/person per day); in application of asphalt roofing, < 0.0045 µg/cm2, 10 samples below detection limit; in a wood preserving plant, 47-1500 µg pyrene per person per day. These data indicate that skin penetration is an important factor in estimating total body exposure to PAH. 5.3.1 Occupational exposure during processing and use of of coal and petroleum products The following section is based on data obtained up to the early 1980s which were compiled by the IARC (1984b, 1985, 1989b). More recent studies are presented in detail. 5.3.1.1 Coal coking In studies of pollution of the atmosphere near coke-oven batteries, the concentration of benzo [a]pyrene varied from < 0.1 in administrative buildings and a pump house to 100-200 µg/m3 on the machinery and discharge side of a battery roof. At the top of a coke battery, the following concentrations of particulate and gaseous PAH were measured by stationary sampling: naphthalene, 0-4.4 (particulate)/ 280-1200 (gaseous) µg/m3; acenaphthene, 0-17/6.0-100 µg/m3; fluorene, 0-58/23-130 µg/m3; phenanthrene, 27-890/6.7-280 µg/m3; anthracene, 9.6-310/6.0-91 µg/m3; 1-methylphenanthrene, 2.7-21/0-7.0 µg/m3; fluoranthene, 45-430/0-24 µg/m3; pyrene, 35-320/0-14 µg/m3; benzo [a]fluorene, 9.7-90/0-6.8 µg/m3; benzo [b]fluorene, 3.1-61/0-0.3 µg/m3; benzo [c]phenanthrene, 2.6-49 µg/m3 (particulate); benz [a]anthracene, 5.4-160/< 0.4-1.6 µg/m3; benzo [b]fluoranthene, 5.5-67/0-0.7 µg/m3; benzo [j]fluoranthene plus benzo [k]fluoranthene, 0-35/0-0.7 µg/m3; benzo [e]pyrene, 8-73/0-0.2 µg/m3; benzo [a]pyrene, 14-130/0-1.5 µg/m3; perylene, 3.3-19/0-0.1 µg/m3; benzo [ghi]perylene, 8.7-45 µg/m3 (particulate); anthanthrene, 2.6-62 µg/m3 (particulate); and coronene, 1.0-19 µg/m3 (particulate) (IARC, 1984b). At eight sites in a German coke plant in 1981, including the top of the oven and the cabin of a lorry driver, the following PAH concentrations were measured: 2.7 µg/m3 fluoranthene, 1.9-170 µg/m3 pyrene, 0.38-37 µg/m3 benzo [c]phenanthrene, 0.22-21 µg/m3 cyclopenta [cd]pyrene, 1.2-120 µg/m3 benz [a]anthracene, 0.71-79 µg/m3 benzo [c]pyrene, 0.88-89 µg/m3 benzo [a]pyrene, 0.21-14 µg/m3 perylene, 0.37-27 µg/m3 benzo [ghi]perylene, 0.18-17 µg/m3 anthanthrene, and 0.93-6.5 µg/m3 coronene. The authors pointed out that the concentrations may have been much higher previously (Manz et al., 1983). Measurements with personal air samplers in Germany and Sweden showed benzo [a]pyrene concentrations varying from 0.16-33 µg/m3 for coke-oven operators to 4.7-17 µg/m3 for lorry drivers. The ranges of exposure to all PAH at different workplaces in the 1970s were: lorry driver, 170-1000 µg/m3; coke-car operator, 4.8-73 µg/m3; jamb cleaner, 62-240 µg/m3; door cleaner, 9.1-17 µg/m3; push-car operator, 9.4-62 µg/m3; sweeper, 110 µg/m3; quench-car operator, 5.7 µg/m3; and wharf man, 360 µg/m3 (IARC, 1984b). Personal air samples taken from 56 Dutch coke-oven workers in 1986 showed pyrene levels of < 0.6 µg/m3 (detection limit) to 9.8 µg/m3 (Jongeneelen et al., 1990). The results of more recent measurements in personal air samples are shown in Table 70. 5.3.1.2 Coal gasification and coal liquefaction The levels of individual PAH in area air samples in Norwegian and British coal gasification plants between the late 1940s and the mid 1950s were in the low microgram per cubic millilitre range. In modern gasification systems, the concentrations of total PAH are usually < 1 µg/m3, but in one of three plants examined the total aerial PAH load was about 30 µg/m3. Personal samples taken in modern coal gasification plants showed similar PAH concentrations (IARC, 1984b). Table 70. Workplace exposures to polycyclic aromatic hydrocarbons in the atmosphere of coke-oven batteries (µg/m3), determined from personal air samples Compound [1] [2] [3] [4] [5] [6] [7] Acenaphthene 3.8 Acenaphthylene 28 Anthracene 65 16 Anthanthrene 2.4 Benz[a]anthracene 0.11-33.19 96 7.5 Benzo[a]fluorene 70 3.7 Benzo[a]pyrene < 0.01-31.15a 0.03-12.63 0.9-46.02 38 0.1-29 7.3 1300 0.01-22.91b Benzo[b]fluoranthene 42 1500 Benzo[b]fluorene 4.3 Benzo[c]phenanthrene 1.4 Benzo[e]pyrene 4.7 Benzo[ghi]fluoranthene 1.6 Benzo[ghi]perylene 4.4 Benzo[k]fluoranthene 42 Chrysene 0.08-13.17 72 Coronene 3.2 Cyclopenta[cd]pyrene 1.9 Fluoranthene 0.12-17.00a 144 22 4400 Fluorene 109 14 Indeno[1,2,3-cd]pyrene 4.5 1-Methylphenanthrene 3.4 Naphthalene 28-445a 650 Perylene 1.8 Phenanthrene 0.07-8.53a 195 49 Pyrene 2.36-98.63 17 Trace Table 70 (continued) [1] Finland; samples from one plant, 1988-90 (Yrjanheikki et al., 1995); [2] Italy; samples from 69 workers, six workplaces (Assennato et al., 1993a); [3] Italy; samples from three workplaces at battery top (Cenni et al., 1993); [4] Sweden; one typical sample (Andersson et al., 1983); [5] United Kingdom; samples from 12 plants (Davies et al., 1986); [6] USA; samples from topside coke-oven workers (Haugen et al., 1986, [7] India; samples from top of coke oven (Rao et al., 1987) a Area air samples b Personal air samples In a pilot coal liquefaction plant in the United Kingdom, monitoring of five operators for vapour-phase PAH gave following results: 1900-3300 ng/m3 phenanthrene, 340-670 ng/m3 pyrene, 270-380 ng/m3 fluoranthene, 29-130 ng/m3 anthracene, 22-1700 ng/m3 fluorene, < 1-1800 ng/m3 naphthalene, < 1-1000 ng/m3 acenaphthene, and < 1-8 ng/m3 acenaphthylene. The higher-molecular-mass PAH were not detected (limit of detection, 1 ng/m3). Pyrene was detected in the particulate phase at concentrations of 630-2900 ng/m3 (Quinlan et al., 1995a). 5.3.1.3 Petroleum refining Personal samples from operators of catalytic cracker units and reaction and fractionation towers in a petroleum refinery showed total PAH levels of 2.6-470 µg/m3. During performance and turn-round operations on reaction and fractionation towers, naphthalene and its methyl derivatives accounted for more than 99% of the total PAH measured; exposure to anthracene, pyrene, chrysene, and benzo [a]pyrene was < 1 µg/m3. Area monitoring for these PAH during normal activities and during shut-down, leak-testing, and start-up operations after turn-rounds gave total PAH concentrations up to 400 µg/m3, most of the measurements being < 100 µg/m3 (IARC, 1989b). The results of personal air sampling of workers at six jobs in seven American refineries in 1990-91 were as follows (mean and range): 5.5 (< 0.25-10) µg/m3 naphthalene, 3.3 (< 0.44-24) µg/m3 acenaphthene, 3.3 (< 0.19-26) µg/m3 acenaphthylene, 0.98 (< 0.085-7.9) µg/m3 fluoranthene, 0.82 (< 0.055-6.7) µg/m3 phenanthrene, 0.78 (< 0.13-5.3) µg/m3 benzo [e]pyrene, 0.65 (< 0.055-5.2) µg/m3 benzo [b]fluoranthene, 0.47 (< 0.14-2.7) µg/m3 fluorene, 0.29 (< 0.11-1.4) µg/m3 indeno[1,2,3- cd]pyrene, 0.18 (< 0.085-0.69) µg/m3 benz [a]anthracene, 0.16 (< 0.11-< 0.59) µg/m3 benzo [a]pyrene, 0.063 (< 0.028-0.26) µg/m3 anthracene, < 0.11- < 0.2 µg/m3 pyrene, < 0.085-< 0.15 µg/m3 chrysene, < 0.085- < 0.15 µg/m3 benzo [k]fluoran-thene, < 0.11-< 0.2 µg/m3 benzo [ghi]perylene, and < 0.11-< 0.2 µg/m3 dibenz [a,h]anthracene. Dermal wipe samples from the back of the hand or from the forehead of workers showed PAH levels of < 0.0011-0.29 µg/cm2, with the highest level for naphthalene and the lowest for anthracene (Radian Corp., 1991). 5.3.1.4 Road paving In early studies on road paving operations, the total PAH concentrations reported in personal air samples were 4-190 µg/m3, and the mean in area air samples was 0.13 µg/m3. The benzo [a]pyrene concentration in stationary samples was < 0.05-0.19 µg/m3 (IARC, 1985). The concentrations of individual PAH in fume condensates from paving asphalt were generally < 2 mg/kg condensate, varying by about seven times depending on the source of crude oil. The levels of benzo [a]pyrene, for example, were between 0.09 and 2.0 mg/kg (Machado et al., 1993). Fourteen stationary air samples from a road paving site in New Zealand in 1983 contained: 0.14-52 µg/m3 benz [a]anthracene plus chrysene, 0.2-14 µg/m3 benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, 0.15-9.0 µg/m3 benzo [a]pyrene, 0.31-5.4 µg/m3 benzo [e]pyrene, 0.039-2.2 µg/m3 perylene, 0.24-5.4 µg/m3 benzo [ghi]perylene, and 0.03-6.3 µg/m3 indeno[1,2,3- cd]pyrene plus dibenz [a,h]anthracene (Swallow & van Noort, 1985). The concentrations in 17 stationary air samples from a road paving operation in New Zealand in another study (year not given) were: 1.2-18 µg/m3 benz [a]anthracene plus chrysene, 1.1-11 µg/m3 benzo [b]fluoranthene plus benzo [j]fluoranthene plus benzo [k]fluoranthene, 0.9-9.0 µg/m3 benzo [a]pyrene, 0.7-5.4 µg/m3 benzo [e]pyrene, and 0.7-6.3 µg/m3 indeno[1,2,3- cd]pyrene (Darby et al., 1986). Concentrations of up to 1.3 µg/m3 were found for acenaphthene, < 0.13 µg/m3 for anthracene, and < 0.54 µg/m3 pyrene in road-paving operations. The workers, and especially the machine driver, were exposed to a mixture of bitumen fumes and diesel exhaust gases for 4-6 h per day (Monarca et al., 1987). The PAH concentrations in personal air samples obtained from workers at six jobs in six paving operations in the USA in 1990 were (mean and range): 6.5 (1.3-15) µg/m3 naphthalene, 2 (< 0.54-6.9) µg/m3 acenaphthene, 2 (< 0.24-8.1) µg/m3 acenaphthylene, 0.58 (< 0.19-0.98) µg/m3 fluorene, 0.55 (< 0.085-1.3) µg/m3 phenanthrene, 0.26 (< 0.11-0.37) µg/m3 fluoranthene, 0.17 (< 0.13-< 0.31) µg/m3 pyrene, 0.16 (< 0.13-0.27) µg/m3 benzo [e]pyrene, 0.13 (< 0.099-< 0.2) µg/m3 chrysene, 0.052 (< 0.034-0.11) µg/m3 anthracene, < 0.099-< 0.12 µg/m3 benz [a]anthracene, < 0.064-< 0.085 µg/m3 benzo [b]fluoranthene, < 0.099-< 0.12 µg/m3 benzo [k]fluoranthene, < 0.13-< 0.25 µg/m3 benzo [a]pyrene, < 0.13-< 0.16 µg/m3 benzo [ghi]perylene, < 0.13-< 0.16 µg/m3 indeno[1,2,3- cd]pyrene, and < 0.13-< 0.16 µg/m3 dibenz [a,h]anthracene. Dermal wipe samples from the back of the hand and from the forehead of workers contained PAH at < 0.00004-0.43 µg/cm2, with the highest level for naphthalene and the lowest for anthracene and pyrene (Radian Corp., 1991). Measurements in the air in France during road paving with different bitumens and tars showed the highest benzo [a]pyrene concentrations with hard-coal tar (1-6 µg/m3) and the lowest with petroleum-based bitumen (0.004-0.007 µg/m3). In general, the benzo [a]pyrene levels in the workplace atmosphere were two to three orders of magnitude higher during paving operations with tar products than with bitumen products (Barat, 1991). 5.3.1.5 Roofing The concentrations of PAH measured during roofing and roofing manufacture are shown in Table 71. The concentrations of individual PAH in fume condensates from roofing asphalt generated at 232 and 316°C the were usually < 10 mg/kg condensate, with higher levels only for naphthalene. They varied with the source of crude oil: those for benzo [a]pyrene were between 0.6 and 2.8 mg/kg (Machado et al., 1993). Acenaphthene was detected at concentrations of 1.4-2.1 µg/m3 in personal samples from roofing workers at two US roofing sites in 1985 (Zey & Stephenson, 1986); 0.8-22 µg/m3 phenanthrene were measured at one US roofing site in 1981 (Reed, 1983). Pyrene was measured at < 190 µg/m3 at three roofing sites in Canada (year not given) (Malaiyandi et al., 1986). Personal air samples from 12 roofers at one US roofing site contained benzo [a]pyrene at 0.53-2.0 µg/m3 in 1987 (Herbert et al., 1990a). The workplace concentrations during bitumen and coal-tar pitch roofing, waterproofing, and flooring operations were of the same order of magnitude (IARC, 1985). Significant, 10-fold differences were found in the levels of anthracene, fluoranthene, pyrene, benz [a]anthracene, benzo [b]fluoranthene, benzo [k]-fluoranthene, benzo [a]pyrene, and benzo [ghi]perylene on skin wipes from the forehead taken before and after a shift in 10 US roofers in 1987 (Wolff et al., 1989a). Comparable results for benzo [a]pyrene levels were obtained for 12 roofers at another US roofing site (Herbert et al., 1990a,b). Dermal wipe samples from the back of the hand or the forehead of workers at six asphalt roofing manufacturing sites in the USA showed PAH levels of < 0.12-5.5 µg/cm2, with the highest level for acenaphthylene and the lowest for fluoranthene, benz [a]anthracene, benzo [k]fluoranthene, and chrysene. Similar samples from workers at six asphalt roofing sites in the USA in 1990-91 showed PAH levels of < 0.0011-0.0045 µg/cm2, with the highest levels for pyrene, chrysene, and benzo [a]pyrene and the lowest for anthracene (Radian Corp., 1991). 5.3.1.6 Impregnation of wood with creosotes Concentrations of PAH ranging from 0.05 µg/m3 benzo [a]pyrene to 650 µg/m3 naphthalene were detected during the handling of creosote-impregnated wood for railroad ties in Sweden. Naphthalene, fluorene and phenanthrene were by far the most abundant compounds (> 100 µg/m3) (Andersson et al., 1983). Concentrations of 0.04-0.28 µg/m3 anthracene and 0.11-7.7 µg/m3 pyrene were found at workplaces in Finland where railroad ties were manufactured (Korhonen & Mulari, 1983), and concentrations of 1-19 µg/m3 anthracene, 6.5-61 µg/m3 phenanthrene, and 0.6-13 µg/m3 pyrene were measured in one plant where railroad sleepers were impregnated and in another where poles Table 71. Exposure to polycyclic aromatic hydrocarbons (µg/m3) during roofing and roofing manufacture Compound [1] [2] [3] [4] Acenaphthene < 0.52-3.2 (0.87) < 0.6-6.7 (1.5) Acenaphthylene < 0.23-29 (7.1) < 0.26-12 (2.9) Anthracene 0.5/1.5 < 0.033-0.069 (0.043) < 0.037-0.042 Anthanthrene < 0.030 Benz[a]anthracene < 0.03-0.130 1.3/2.5 < 0.099-< 0.13 < 0.11-< 0.13 Benzo[a]fluorene 0.03-0.080 Benzo[a]pyrene < 0.03-0.037 0.9/1.5 < 0.13-< 0.18 < 0.11-< 0.13 Benzo[b]fluoranthene < 0.03-0.093a 0.8/1.2 < 0.065-< 0.38 (0.13) < 0.078-< 0.085 Benzo[b]fluorene 0.051-0.093 Benzo[e]pyrene < 0.03-0.110 < 0.13-3 (0.61) < 0.15-< 0.17 Benzo[ghi]fluoranthene < 0.03 Benzo[ghi]perylene < 0.03-0.069 0.6/0.9 < 0.13-< 0.18 < 0.15-< 0.17 Benzo[k]fluoranthene 0.4/0.7 < 0.099-< 0.13 < 0.099-< 0.12 Chrysene 0.038-0.214 < 0.099-< 0.13 < 0.11-< 0.13 Coronene < 0.03 Dibenz[a,h]anthracene < 0.03 < 0.13-< 0.18 < 0.15-< 0.17 Fluoranthene 0.084-0.234 3.1/7 < 0.099-4 (0.64) < 0.11-0.13 Fluorene < 0.16-14 (2.5) < 0.19-1.1 (0.44) Indeno[1,2,3-cd]pyrene < 0.030 < 0.13-< 0.18 < 0.15-0.94 (0.16) Naphthalene < 0.22-9.2 (5.2) 1.2-25 (7.5) Perylene < 0.030 Phenanthrene < 0.065-1.7 (0.53) < 0.078-1.4 (0.38) Pyrene 0.035-0.183 2.6/5.4 < 0.13-3.4 (0.76) < 0.15-< 0.73 (0.25) /, single determinations; mean values shown in parentheses; [1] Germany; personal and area air samples from one bitumen roofing site (Schmidt, 1992); [2] USA; personal air samples from nine workers; 1987 (Wolff, M.S. et al., 1989); [3] USA; personal air samples from six asphalt roofing sites; 1990 (Radian Corp., 1991); [4] USA; personal air samples from six roofing manufacturing sites; 1990 (Radian Corp., 1991) a Benzo[b+j+k]fluoranthenes were preserved (year not given) (Heikkilä et al., 1987). In measurements of personal air samples from 10 workers in a Dutch plant for impregnation of railroad sleepers in 1991, 0.3-1.3 µg pyrene/m3 was measured in the breathing zone and 47-1500 µg/d in pads placed on various areas of the skin of the workers. Dermal exposure was shown to be reduced by up to 90% by the use of protective clothing (Van Rooij et al., 1993b). 5.3.1.7 Other exposures In area air samples taken near the bitumen processing devices of refineries, the total PAH levels varied from 0.004 to 50 µg/m3 (IARC, 1985, 1989b). The use of lubricating oils may result in exposure to PAH. At two Italian glass manufacturing plants, phenanthrene, anthracene, pyrene, and fluoranthene were found in personal air samples at concentrations < 3 µg/m3 (year not given) (Menichini et al., 1990). The pyrene levels resulting from use of lubricating oils in Italian earthenware factories were 0.02-0.09 µg/m3; the benzo [a]pyrene concentration was below the limit of detection (Cenni et al., 1993). Measurable concentrations of individual PAH were detected in indoor air above asphalt floor tiles in e.g. warehouses, factories, and manufacturing plants. The concentrations at six sampling sites in Germany were between < 0.01 ng/m3 for benzo [ghi]perylene and 3.3 ng/m3 for chrysene. The concentrations of phenanthrene, pyrene, fluoranthene, chrysene, and benzo [b]fluorene in particular were higher than those in outdoor air (Luther et al., 1990). In two Swiss plants for the production of silicon carbide, personal air samples from four and five workers, respectively, contained the following PAH levels: 4-140 ng/m3 acenaphthylene, 8-86 ng/m3 acenaphthene, 11-500 ng/m3 fluorene, 88-1400 ng/m3 phenanthrene, 3-250 ng/m3 anthracene, 20-1100 ng/m3 fluoranthene, 30-2500 ng/m3 pyrene, 7-6400 ng/m3 benz [a]-anthracene, 37-14 000 ng/m3 chrysene, 11-3700 ng/m3 benzo [b]fluoranthene plus benzo [j]fluoranthene, 3-470 ng/m3 benzo [k]fluoranthene, 18-3800 ng/m3 benzo [e]pyrene, 4-630 ng/m3 benzo [a]pyrene, 2-250 ng/m3 indeno[1,2,3- cd]pyrene, 2-520 ng/m3 dibenz [a,h]anthracene, 4-550 ng/m3 benzo [ghi]-perylene, and 4-34 ng/m3 coronene (Petry et al., 1994). 5.3.2 Occupational exposure resulting from incomplete combustion of mineral oil, coal, and their products 5.3.2.1 Aluminium production Early measurements of atmospheric benzo [a]pyrene at workplaces in the aluminium industry showed concentrations of 0.02-970 µg/m3 in personal air samples and 0.03-5.3 µg/m3 in area air samples. In the atmosphere of an aluminium production plant, naphthalene, fluorene, phenanthrene, anthracene, fluoranthene, pyrene, benzo [a]fluorene, benzo [b]fluorene, benzo [c]phenan-threne, benz [a]anthracene, chrysene, triphenylene, benzo [b]fluoranthene plus benzo [k]fluoranthene, benzo [e]pyrene, benzo [a]pyrene, benzo [ghi]perylene, anthanthrene, and coronene were found at concentrations < 400 µg/m3. The most abundant compounds were phenanthrene, naphthalene, fluorene, fluoranthene, and pyrene, at concentrations > 100 µg/m3. The other substances occurred at concentrations < 10 µg/m3 (IARC, 1984b). The following concentrations of PAH were found in four stationary air samples from an aluminium smelter in New Zealand in 1979: 0.37-9.6 µg/m3 benz [a]anthracene plus chrysene, 0.34-7.6 µg/m3 benzo [b+j+k]fluoranthenes, 0.12-2.6 µg/m3 benzo [e]pyrene, 0.19-4.1 µg/m3 benzo [a]pyrene, 0.05-1.5 µg/m3 perylene, 0.13-2.7 µg/m3 indeno[1,2,3- cd]pyrene plus dibenz [a,h]anthracene, and 0.12-3.3 µg/m3 benzo [ghi]perylene (Swallow & van Noort, 1985). Similar levels were found in a typical personal air sample from a Söderberg aluminium plant in Sweden (year not given) with, in addition, 27 µg/m3 phenanthrene, 20 µg/m3 fluoranthene, 2.8 µg/m3 fluorene, 2.8 µg/m3 anthracene, 2.8 µg/m3 benzo [a]fluorene, and < 1.0 µg/m3 naphthalene (Andersson et al., 1983). In personal air samples from 38 workers in the Söderberg potroom of an aluminium smelter in the humid tropics (location not given), mean concentrations of < 1.0-48 µg/m3 benzo [a]pyrene and 3.5-130 µg/m3 pyrene were detected (Ny et al., 1993). The arithmetic mean concentrations of PAH in workplace air samples from the Canadian aluminium industry were 1100 µg/m3 naphthalene, 130 µg/m3 acenaphthene, 45 µg/m3 fluorene, 30 µg/m3 phenanthrene, 4.5 µg/m3 anthracene, 1.1 µg/m3 fluoranthene, and 0.58 µg/m3 pyrene. The concentrations of benz [a]anthracene, chrysene, benzo [a]pyrene, and benzo [e]pyrene were < 0.01 µg/m3 (Lesage et al., 1987). Personal air samples from 18 workers in a US plant producing anodes for use in aluminium reduction (year not given) showed pyrene concentrations of 1.2-7.4 µg/m3 (Tolos et al., 1990). Urine samples from 11 workers in Norwegian Söderberg aluminium plants contained very low levels of unchanged PAH, although the concentrations in the workplace air greatly exceeded the concentrations in urban air. The total concentration of PAH metabolites in the samples was 1.5-6 greater than that in a control group (Becher & Bjrseth, 1983). The PAH concentrations in the air of aluminium plants is reduced dramatically by the use of tempered anodes instead of Söderberg anodes. Measurements of benzo [a]pyrene levels in French factories showed 1-36 µg/m3 in potrooms with Söderberg anodes and 0.004-0.6 µg/m3 in potrooms with tempered anodes (Barat, 1991). 5.3.2.2 Foundries In personal air samples from workers in 10 Canadian foundries, mean concentrations of 0.14-1.8 µg/m3 benz [a]anthracene plus chrysene, 0.09-1.2 µg/m3 benzo [a]pyrene, and 0.09-1.9 µg/m3 dibenz [a,h]anthracene were measured. The benzo [a]pyrene levels in stationary air samples from six Finnish foundries were 0.01-13 µg/m3, depending on whether coal-tar pitch or coal powder was used as the moulding sand additive (IARC, 1984b). In another study, the highest individual PAH levels were found in coke making, moulding, and furnaces (Gibson et al., 1977). Personal air samples from 67 Finnish foundry workers in 1990-91 showed benzo [a]pyrene concentrations of 2-60 ng/m3 with a mean of 8.6 ng/m3 (Perera et al., 1994). Depending on the foundry process and sand binder, the mean benzo [a]pyrene level in 29 French foundries varied from 3 to 2300 ng/m3 (Lafontaine et al., 1990). Concentrations of PAH measured in foundries are shown in Table 72. 5.3.2.3 Other workplaces Personal air samples from German chimney sweeps (year not given; 115 samples) showed an average benzo [a]pyrene level of 0.09 µg/m3, but eight of the samples exceeded 2 µg/m3. With an inhaled air volume of 10 m3 per working day, the daily intake of benzo [a]pyrene was estimated to be 0.24-2.7 µg, with a median value of 1.3 µg (Knecht et al., 1989). In an Italian pyrite mine, pyrene levels of 0.03-0.21 µg/m3 were measured in personal and area air samples. The benzo [a]pyrene concentrations were below the limit of detection (Cenni et al., 1993). Area air samples taken in China showed total PAH levels of 3-40 µg/m3 in two iron mines and 4-530 µg/m3 in four copper mines. Individual compounds were not identified, but the main components were naphthalene and acenaphthene in the iron mines and naphthalene, benz [a]anthracene, benzo [b]fluoranthene, benzo [a]pyrene, benzo [e]pyrene, and dibenz [a,h]anthracene in the copper mines. The PAH concentrations probably resulted from the drilling of holes with hydraulic or pneumatic drills and by the transport of broken ore in diesel-powered scoops (Wu et al., 1992). Area and personal air samples from workers in a railway tunnel in Italy showed pyrene levels of 0.04-0.30 µg/m3. The benzo [a]pyrene concentrations ranged from below the limit of detection to 0.04 µg/m3 (Cenni et al., 1993). Table 72. Exposure to polycyclic aromatic hydrocarbons (µg/m3) in the atmosphere of foundries Compound [1] [2] [3] Acenaphthene 0.03 Acenaphthylene ND Anthracene 2.31 0.05 Anthanthrene 0.64 Benz[a]anthracene 0.008-0.221 0.67 0.01 Benzo[a]fluorene 0.48 Benzo[a]pyrene 0.049-0.152 0.47 0.02 Benzo[b]fluoranthene 0.87a 0.003 Benzo[b]fluorene 0.41 Benzo[e]pyrene 0.48 Benzo[ghi]fluoranthene 0.15 Benzo[ghi]perylene 0.72 0.05 Benzo[k]fluoranthene 0.037-0.458 0.02 Chrysene 0.82b 0.02 Coronene 0.21 Dibenz[a,h]anthracene 0.20 ND Fluoranthene 1.56 0.13 Fluorene 0.08 Indeno[1,2,3-cd]pyrene 0.81 ND Naphthalene 9.68 Perylene 0.21 Phenanthrene 4.46 0.32 Pyrene 1.74 0.01 ND, not detected; /, single measurements; [1] Canada, steel foundry: coke making, moulding, furnaces, finishing, and cranes (Gibson et al., 1977); [2] Western Germany, one foundry, area air samples (Knecht et al., 1986); [3] Denmark, 70 workers, personal air samples; melting, machine moulding, casting, sand preparation (Omland et al., 1994) a In sum with benzo(j+k)fluoranthene b In sum with triphenylene In the air of fish and meat smokehouses in Denmark (year not given), the maximum concentration of naphthalene in stationary air samples was about 2900 µg/m3. The most abundant compounds were naphthalene, phenanthrene, pyrene, fluorene, anthracene, and fluoranthene (> 100 µg/m3) (Nordholm et al., 1986). The minimal values were < 1 µg/m3, benzo [a]pyrene being detected at minimal levels of 0.08 µg/m3 in meat smokehouses and 0.4 µg/m3 in fish smokehouses (Hansen et al., 1991b), with a maximum concentration of 78 µg/m3 (Nordholm et al., 1986). In a further study in nine Danish meat smokehouses, naphthalene was detected at 21 µg/m3, fluorene at 6.9 µg/m3, fluoranthene at 6.6 µg/m3, phenanthrene at 5.6 µg/m3, acenaphthene at 5.2 µg/m3, chrysene at 1.2 µg/m3, anthracene at 1.1 µg/m3, pyrene at 0.2 µg/m3, and benzo [ghi]perylene at 0.2 µg/m3 (Hansen et al., 1992). The concentrations of naphthalene, fluorene, anthracene, phenanthrene, pyrene, benzo [a]fluorene, chrysene, benzo [k]fluoranthene, benzo [a]pyrene, benzo [e]pyrene, benzo [ghi]perylene, and dibenz [a,h]anthracene in cooking fumes in a Finnish food factory, three restaurants, and one bakery (year not given) during the frying of meat and during deep-frying ranged between < 0.02 µg/m3 (the limit of detection) and 26 µg/m3. Naphthalene occurred at by far the highest concentration. Stationary air was sampled as close as possible to the active working area and the workers' breathing zone (Vainiotalo & Matveinen, 1993). 6. KINETICS AND METABOLISM IN LABORATORY MAMMALS AND HUMANS Appraisal Polycyclic aromatic hydrocarbons (PAH) are lipophilic compounds and can be absorbed through the lungs, the gastrointestinal tract, and the skin. In studies of the distribution of PAH in rodents, both the parent compounds and their metabolites were found in almost all tissues and particularly those rich in lipids. As a result of mucociliary clearance and hepatobiliary excretion, they were present, for example, in the gastrointestinal tract even when administered by other routes. The metabolism of PAH to more water-soluble derivatives, which is a prerequisite for their excretion, is complex. Generally, the process involves epoxidation of double bonds, a reaction catalysed by cytochrome P450-dependent mono-oxygenases, rearrangement or hydration of the epoxides to yield phenols or diols, respectively, and conjugation of the hydroxylated derivatives. The reaction rates vary widely: interindividual variations of up to 75-fold have been observed, for example, with human macrophages, mammary epithelial cells, and bronchial explants from different donors. All aspects of the absorption, metabolism, activation, and excretion of benzo[a]pyrene have been covered exhaustively in the published literature, but there is a dearth of information on many of the other PAH considered in this publication, particularly in humans. Thus, this overview sets out general principles and describes pathways relevant to benzo[a]pyrene in greater detail. Most biotransformation leads to detoxification products that are conjugated and excreted in the urine and faeces. The human body burden of PAH has not been extensively studied, but tissue samples taken at autopsy were found in one study to contain benzo[a]pyrene at an average of 0.3 µg/100 g dry tissue; lung contained 0.2 µg/100 g. In contrast, the pathways by which several PAH are metabolized to reactive intermediates that bind covalently to nucleic acids have been examined in great detail. Although the commonest mechanism in animals and humans appears to involve the formation of diol epoxides, radical cations and sulfate esters of hydroxymethyl derivatives may also be important in certain cases. 6.1 Absorption PAH are lipophilic compounds, soluble in organic solvents, that are usually devoid of ionizable or polar groups. Like many other xenobiotic substances, they would be expected to dissolve readily in, and be transported through, the external and internal lipoprotein membranes of mammalian cells. This is confirmed by the uptake of PAH in vitro from media in which cells are maintained in culture and modified metabolically by enzymes of the endoplasmic reticulum. Furthermore, PAH are known to be able to cause biological effects in vivo in cells and tissues that are distant from their site of uptake by the organism. In humans, the major routes of uptake of PAH are thought to be through (i) the lungs and the respiratory tract after inhalation of PAH-containing aerosols or of particulates to which a PAH, in the solid state, has become absorbed; (ii) the gastrointestinal tract after ingestion of contaminated food or water; and (iii) the skin as a result of contact with PAH-bearing materials. 6.1.1 Absorption by inhalation Investigations of the pulmonary absorption of PAH have frequently been clouded by the existence of the mucociliary clearance mechanism, by which hydrocarbons absorbed onto particulates that have been inhaled are swept back up the pulmonary tree and are swallowed, thus entering the organism through the gastrointestinal tract. Use of isolated perfused rat lungs, however, provided a clear demonstration that benzo [a]pyrene is absorbed directly through the pulmonary epithelia. After intratracheal administration, both the hydrocarbon and its metabolites were detected in effluent perfusion fluid (Vainio et al., 1976). Other studies have shown that benzo [a]pyrene administered in vivo as an aerosol is cleared from the lungs of rats by a biphasic process in which an initial rapid phase (tracheal clearance) is followed by a much slower second phase (alveolar clearance) (Mitchell, 1982). PAH absorbed onto particles may take very much longer to be cleared from rodent lungs, however, than the free hydrocarbons, and the factors that affect this clearance rate include the structure of the hydrocarbon and the dimensions and chemical nature of the particles onto which the PAH are absorbed (Henry & Kaufman, 1973; Creasia et al., 1976; Nagel et al., 1976). For example, while 50% of the benzo [a]pyrene coated onto carbon particles of 15-30 µm was cleared from hamster lungs within 60 h, it took only 10 h to clear 50% of the benzo [a]pyrene that had been coated onto 0.5-1.0-µm carbon particles. In a comparable experiment, however, when ferric oxide particles of either 0.5-10 or 15-20 µm were used as carriers for benzo [a]pyrene, 50% of the hydrocarbon was cleared in just over 2 h, and carrier particle size did not affect the clearance rates (Henry & Kaufman, 1973). Benzo [a]pyrene was metabolized by the epithelia lining the nasal cavities of hamsters, dogs, and monkeys when 14C-labelled hydrocarbon was instilled as an aqueous suspension (Dahl et al., 1985; Petridou-Fischer et al., 1988). From their studies with hamsters, the authors concluded that when frequent small doses of 650 ng at 10-min intervals were instilled into the nasal cavity, so as to imitate inhalation, some 50% of the benzo [a]pyrene was metabolized; a large fraction of the metabolites could be recovered from the mucus on the epithelial surfaces; and the nasal epithelia were comparable to those of the trachea and lungs in their ability to metabolize benzo [a]pyrene. Metabolites produced nasally would be expected to be swallowed and then absorbed in the gastrointestinal tract. In humans, the concentrations of benzo [a]pyrene and pyrene present in association with soot particles in the lungs were much lower than would have been expected from the soot content. Thus, only a trace of benzo [a]pyrene was found in one of 11 lung samples examined, in which the expected benzo [a]pyrene content ranged from 9 to 200 µg; in the other 10 samples, no benzo [a]pyrene was detected. Pyrene disappeared more slowly: all 11 lung samples contained the compound, at levels of 0.9-4.9 µg, whereas 3-190 µg might have been expected (Falk et al., 1958). The ability of pulmonary epithelial cells to metabolize PAH such as chrysene and benzo [a]pyrene to a variety of hydroxylated derivatives (Jacob et al., 1992) may facilitate the absorption and clearance of PAH from the lungs. 6.1.2 Absorption in the gastrointestinal tract Indirect evidence for the gastrointestinal absorption of PAH was provided by Shay et al. (1949), who found that repeated intragastric instillation of 3-methylcholanthrene led to the development of mammary cancer. Mammary tumours can also be induced in rats by intracolonic adminstration of 7,12-dimethylbenz [a]anthracene (Huggins et al., 1961). (3-Methylcholanthrene and 7,12-dimethylbenz [a]anthracene are synthetic PAH that are potent carcinogens.) More direct investigations by Rees et al. (1971) showed rapid absorption of intragastrically administered benzo [a]pyrene; the highest levels of hydrocarbon were found in the thoracic lymph some 3-4 h after administration. In a report of studies of intact rats and intestinal sacs to examine the mechanisms involved in benzo [a]pyrene absorption, Rees et al. (1971) proposed that two sequential steps were involved, in which a phase of absorption by the mucosa is followed by diffusion through the intestinal lining. In a study with Sprague-Dawley rats, the presence of bile was found to increase intestinal absorption of PAH such as benzo [a]pyrene and 7,12-dimethylbenz [a]anthracene to a greater degree than that of anthracene and pyrene. The effect may be related to differences in the aqueous solubility of the PAH examined (Rahman et al., 1986). The composition of the diet also affects intestinal absorption of co-administered benzo [a]pyrene. Of the dietary components studied, soya bean oil and triolein gave rise to the highest levels of absorption of 14C-benzo [a]pyrene given orally at a dose of 8.7 µg to Wistar rats, while cellulose, lignin, bread, rice flake, and potato flake suppressed it (Kawamura et al., 1988). 6.1.3 Absorption through skin PAH and PAH-containing materials have been applied dermally in solution in solvents such as acetone and tetrahydrofuran. Dermal transfer without use of a solvent was achieved by use of reconstituted vapour-particulate phases emitted from coal-tar and bitumen (Genevois et al., 1995) and by application in oil (Ingram et al., 1995). Absorption of PAH through the skin was observed indirectly when it was found that repeated topical application of 3-methylcholanthrene led to the appearance of mammary tumours in mice (Maisin & Coolen, 1936; Englebreth-Holm, 1941). The percutaneous mechanism of absorption is not universal, however, since although almost all of a dose of 14C-benzo [a]pyrene applied to mouse skin appeared in the faeces within two weeks, very little dibenz [a,h]anthracene was absorbed in this way and most was lost through epidermal sloughing (Heidelberger & Weiss, 1951). Benzo [a]pyrene has been shown to be absorbed percutaneously in vitro, by absorption from soil into human skin (Wester et al., 1990) and, after application as a solution in acetone, into discs of human, mouse, marmoset, rat, rabbit, and guinea-pig skin (Kao et al., 1985). In the latter experiments, marked interspecies differences were noted: 10% of the applied dose (10 µg/5 cm2) of 14C-benzo [a]pyrene permeated mouse skin, 3% crossed human skin, and < 0.5% crossed guinea-pig skin within 24 h. It was concluded that both diffusional and metabolic processes are involved in the percutaneous absorption of benzo [a]pyrene. In Wistar rats that received 14C-pyrene as a solution in acetone on areas of shaved dorsal skin, the rate of uptake was relatively rapid (half-life, 0.5-0.8 d). The concentrations of pyrene were highest in the liver, kidneys, and fat, but those of pyrene metabolites were highest in the lungs. About 50% of an applied dose of 2, 6, or 15 mg/kg bw was excreted in the urine and faeces during the first six days after treatment (Withey et al., 1993). In studies with 32P-postlabelling for the detection of DNA adducts, when complex mixtures of PAH, such as that present in used lubricating oil from petrol engines, in coal-tar, or in juniper-tar, were applied directly to mouse skin, appreciable, persistent levels of DNA adducts (50-750 amol/µg DNA [1 amol/µg DNA equivalent to 3.3 adducts/1010 nucleotides]) were formed in the lungs (Schoket et al., 1989, 1990). The level of adducts in mouse skin was inversely related to the viscosity of the oil applied (Ingram et al., 1995). Evidence for percutaneous absorption of PAH has also been obtained in humans in vivo. When 2% coal-tar in petroleum jelly was applied topically, phenanthrene, anthracene, pyrene, and fluoranthene were detected in peripheral blood samples (Storer et al., 1984). In addition, volunteers treated topically with creosote (100 µl) or pyrene (500 µg, applied as a solution in toluene) and a psoriasis patient who used a coal-tar shampoo excreted 1-hydroxypyrene in their urine. In each case, maximal excretion occurred 10-15 h after treatment (Viau & Vyskocil, 1995). 6.2 Distribution The whole-body distribution of PAH has been studied in rodents. The levels found in individual tissues depend on a number of factors, including the PAH, the route of administration, the vehicle, the times after treatment at which tissues are assayed, and the presence or absence of inducers or inhibitors of hydrocarbon metabolism within the organism. The investigations have shown that (i) detectable levels of PAH occur in almost all internal organs, (ii) organs rich in adipose tissue can serve as storage depots from which the hydrocarbons are gradually released, and (iii) the gastrointestinal tract contains high levels of hydrocarbon and metabolites, even when PAH are administered by other routes, as a result of mucociliary clearance and swallowing or hepatobiliary excretion (Heidelberger & Jones, 1948; Heidelberger & Weiss, 1951; Kotin et al., 1959; Bock & Dao, 1961; Takahashi & Yasuhira, 1973; Takahashi, 1978; Mitchell, 1982). 14C-Benzo [a]pyrene injected intravenously at 11 µg/rat was cleared rapidly from the bloodstream, with a half-life of < 1 min (Kotin et al., 1959), as confirmed by Schlede et al. (1970a,b), who also noted that the rate of clearance was increased when animals were pretreated with 20 mg/kg bw non-radioactive benzo [a]pyrene or 37 mg/kg bw phenobarbital, both of which can induce metabolism. The distribution of 3-methylcholanthrene in mice and their fetuses was studied by whole-body autoradiography. When 1 mg of 14C-labelled hydrocarbon is injected intravenously, it is not only widely distributed in maternal tissues but also crosses the placenta and can be detected in the fetuses (Takahashi & Yasuhira, 1973; Takahashi, 1978), in which it induces pulmonary tumours (Tomatis, 1973; see also Section 7). The distribution of inhaled and intragastrically or intravenously administered benzo [a]pyrene and 7,12-dimethylbenz [a]anthracene in rats and mice has also been studied, with similar results (Shendrikova & Aleksandrov, 1974; Shendrikova et al., 1973, 1974; Neubert & Tapken, 1988; Withey et al., 1992). Rapid transfer of radioactive benzo [a]pyrene across the placenta was confirmed in experiments in which the appearance of radioactivity in the umbilical vein of pregnant guinea-pigs was measured (Kelman & Springer, 1982). Samples of placenta, maternal blood, umbilical cord blood, and milk from 24 women in south India were examined for the presence of selected PAH. Although umbilical cord blood and milk showed the highest levels (benzo [a]pyrene, 0.005-0.41 ppm; dibenz [a,c]anthracene, 0.013-0.60 ppm; chrysene, 0.002-2.8 ppm), only 50% of the samples examined contained detectable levels. The authors concluded that developing fetuses and newborn infants were exposed to these PAH, probably from the maternal diet (Madhavan & Naidu, 1995). After intratracheal administration to mice and rats, the distribution of PAH was essentially similar to that found after intravenous or subcutaneous injection (Kotin et al., 1959), except for the expected high pulmonary levels. Detailed time-concentration curves for several organs have been obtained after inhalation of 3H-benzo [a]pyrene aerosols at 500 µg/litre of air (Mitchell, 1982). For example, 1 h after the end of administration, the highest levels were present in the stomach and small intestine; as these declined, the amounts of radioactivity in the large intestine and caecum increased. The elimination half-times in the respiratory tract were 2-3 h for the initial rapid phase and 25-50 h for the subsequent slow phase. 6.3 Metabolic transformation The metabolism of PAH follows the general scheme of xenobiotic metabolism originally outlined by Williams (1959). The hydrocarbons are first oxidized to form phase-I metabolites, including primary metabolites, such as epoxides, phenols, and dihydrodiols, and then secondary metabolites, such as diol epoxides, tetrahydrotetrols, and phenol epoxides. The phase-I metabolites are then conjugated with either glutathione, sulfate, or glucuronic acid to form phase-II metabolites, which are much more polar and water-soluble than the parent hydrocarbons. The metabolism of PAH has been studied in vitro, usually in microsomal fractions prepared from rat liver, although many other tissue preparations have also been used. Metabolism in such systems might be expected to be simpler than that in whole animals because the enzymes and co-factors necessary for sulfate, glutathione, or glucuronide conjugate formation may be removed, depleted, or diluted during tissue fractionation. Use of these systems appears to be justified, however, because the same types of phase-I metabolites are formed when animals are treated with simple hydrocarbons such as naphthalene as when the same hydrocarbon is incubated with hepatic microsomes or tissue homogenates (Boyland et al., 1964). The metabolism of PAH has thus been studied extensively in cells and tissues in culture, which metabolize hydrocarbons to both phase-I and phase-II metabolites and which probably better represent the metabolism of PAH that occurs in vivo (for reviews see Conney, 1982; Cooper et al., 1983; Dipple et al., 1984; Hall & Grover, 1990; Shaw & Connell, 1994). Particular attention has been paid to the metabolism of PAH in human tissues that might be exposed to hydrocarbons present in food and in the environment and which are, therefore, potential targets for the carcinogenic action of PAH (Autrup & Harris, 1983). The cells and tissues examined include the bronchus, the colon, mammary cell aggregates, keratinocytes, monocytes, and lymphocytes. The metabolism of PAH by human pulmonary macrophages has also received attention (Autrup et al., 1978a; Harris et al., 1978a; Marshall et al., 1979) because it is conceivable that metabolism by these cells might be responsible, at least in part, for the high incidence of bronchial cancer in smokers (Wynder et al., 1970). Macrophages can engulf particulate matter that reaches the terminal airways of the lung and thus would be expected, especially in smokers, to contain PAH (Hoffmann et al., 1978). The macrophages and engulfed particulate matter can then be transported to the bronchi where proximate and ultimate carcinogens, formed by metabolism in the macrophages, could leave the macrophages and enter the epithelial cells lining the bronchi (Autrup et al., 1978a; Harris et al., 1978a). This is an attractive theoretical mechanism which could account for the high incidence of respiratory tumours at the junctions of the large bronchi and which is supported by experimental evidence. Extracts of organic material from isolated perfused lung tissues of rabbits that had been exposed intratracheally to benzo [a]pyrene with or without ferric oxide were analysed for benzo [a]pyrene metabolites and for mutagenicity. Extracts of lung tissue exposed to benzo [a]pyrene only were mutagenic and contained benzo [a]pyrene metabolites. When ferric oxide was co-administered, only the macrophage extracts were mutagenic, owing to relatively large amounts of unmetabolized benzo [a]pyrene. These experiments demonstrate that ferric oxide particles enhance the uptake of benzo [a]pyrene by lung macrophages and slow its metabolism beyond the 3-h period during which perfused lung systems can be maintained (Schoeny & Warshawsky, 1983). Administration of particles in vitro enhances both the uptake and metabolism of benzo [a]pyrene by hamster alveolar macrophages (Griefe et al., 1988). Metabolites were found in both the cells and the culture medium. Subsequent studies showed that concurrent administration of benzo [a]pyrene and ferric oxide particles resulted in increased benzo [a]pyrene metabolism and release of superoxides (Greife & Warshawsky, 1993). In particular, the dihydrodiol fraction was increased. These studies indicate that particulates may act in lung cancer by changing the time frame for metabolism, shifting the site of metabolism to macrophages and enhancing the production of metabolites that are on the pathway to putative ultimate carcinogenic forms. In this context, it has been demonstrated that particles of various sorts exert different toxic effects on rat and hamster pulmonary macrophages in vitro: ferric oxide and aluminium oxide particulates were toxic, while crystalline silica was not (Warshawsky et al., 1994). The conclusion that the macrophage is the principal metabolizing cell is further supported by the studies of Ladics et al. (1992a,b), who demonstrated that the macrophage population was the only one in murine spleen that could metabolize benzo [a]pyrene, while the other splenic cell types examined, including B cells, T cells, polymorphonuclear cells, and the splenic capsule, did not produce benzo [a]pyrene metabolites above the background level. Although the same types of metabolite are formed from PAH in many of the cell and tissue preparations examined in culture, the relative levels and the rates of formation of these metabolites depend on the type of tissue or cell that is being studied and on the species and strain of animal from which the metabolizing systems are prepared. With heterogeneous populations such as humans, the rate of metabolism depends on the individual from whom the tissues or cells are prepared. For example, a 75-fold variation in the extent of hydrocarbon activation was reported in studies of human bronchus (Harris et al., 1976), and similar variations were observed among human mammary cell aggregates (Grover et al., 1980; MacNicoll et al., 1980) and macrophages (Autrup et al., 1978a). The pattern and role of metabolism can also be varied by adding inhibitors of the enzymes that are responsible for metabolism or by pretreating either cells in culture or the animals from which the metabolizing systems are prepared with enzyme inducers. 6.3.1 Cytochromes P450 and metabolism of PAH The cytochromes P450 (CYP) are a superfamily of haemoproteins that catalyse the oxidation of various endogenous molecules as well as xenobiotics, including PAH. To date, about 250 genes that encode these enzymes have been identified in various organisms. For classification purposes, the CYP have been organized into families and subfamilies according to their structural homology (Nelson et al., 1993). Certain CYP belonging to families 1, 2, and 3 are expressed in mammalian cells and are particularly important in xenobiotic metabolism, and one or more member of each family is capable of metabolizing one or more PAH (Guengerich & Shimada, 1991; Gonzalez & Gelboin, 1994). Most studies to compare the catalytic properties of different CYP have been carried out with model compounds such as benzo [a]pyrene. They show that the catalytic properties (e.g. the Vmax) of different CYP in PAH metabolism can differ essentially (Shou et al., 1994). In considering the contribution of a CYP enzyme to PAH metabolism in vivo, two other parameters in addition to the catalytic properties should be taken into account: the mode of regulation and tissue specificity in its expression. Combinations of the three factors should give an idea of the relative importance of an enzyme in PAH metabolism. 6.3.1.1 Individual cytochrome P450 enzymes that metabolize PAH CYP1A: CYP1A appears to be the only enzyme with metabolic capability towards a wide variety of PAH molecules. It is expressed in various tissues but at a generally low constitutive level (Guengerich & Shimada, 1991). The induction of CYP1A1 is controlled by the Ah (aryl hydrocarbon) receptor, a transcription factor that can be activated by several ligands such as 2,3,7,8-tetradichlorobenzo- para-dioxin (TCDD) and PAH, with variable potency (Negishi et al., 1981). Thus, PAH and material containing PAH can regulate their own metabolism by inducing CYP1A1. After induction, CYP1A1 expression may reach high levels, e.g. in the placenta, lung, and peripheral blood cells; however, in the liver, the principal organ of xenobiotic metabolism, the level of expression is low even after induction, and other CYP appear to be more important, at least in the metabolism of benzo [a]pyrene (Guengerich & Shimada, 1991). CYP1A2: The other member of the CYP1A family, CYP1A2, also metabolizes PAH; however, its capacity to metabolize benzo [a]pyrene to the 3-hydroxy metabolite, for example, is about one-fifth that of CYP1A1 (Shou et al., 1994). Human CYP1A2 is nevertheless very active in forming benzo [a]pyrene 7,8-dihydrodiol (Bauer et al., 1995) and in forming diol epoxides from the 7,8-dihydrodiol (Shou et al., 1994). There is also evidence that CYP1A2 can activate 7,12-dimethylbenz [a]anthracene to mutagenic species, albeit at a low rate (Aoyama et al., 1989). The expression of CYP1A2 is also regulated by the Ah receptor, but in not exactly the same way as CYP1A1 (Negishi et al., 1981). In the liver, for example, the level of CYP1A2 expression is much higher than that of CYP1A1 (Guengerich & Shimada, 1991). While the capacity of CYP1A2 to oxidize various PAH is more limited than that of CYP1A1, its role in reactions like diol epoxide formation from benzo [a]pyrene in the liver could be important because of its high level of expression. CYP1B: The CYP1B subfamily was discovered only recently. Once the enzyme had been isolated, it was found to be capable of metabolizing PAH. Interestingly, its expression is also under the control of the Ah receptor. Only limited information is available on its expression and catalytic properties in different tissues, but it seems to be expressed at least in mouse embryo fibroblasts (Savas et al., 1994), rat adrenal glands (Bhattacharyya et al., 1995), and several human tissues (Sutter et al., 1994). A number of PAH may act as substrates for this enzyme (Shen et al., 1994). CYP2B: When recombinant gene technology was used to express human CYP2B6 cDNA in a human lymphoblastoid cell line, this enzyme was shown to be capable of metabolizing benzo [a]pyrene to 3- and 9-phenols and trans-dihydrodiols (Shou et al., 1994). In addition, CYP2B enzymes may be involved in the metabolism of 7,12-dimethylbenz [a]anthracene (Morrison et al., 1991a). The constitutive levels of CYP2B enzymes are extremely low in human liver, but they are strongly induced by phenobarbital and phenobarbital-type inducers of CYP. Accordingly, immunological studies of inhibition have shown that the CYP2B enzymes may play a significant role in the metabolism of PAH, only when they are induced (Hall et al., 1989; Honkakoski & Lang, 1989). CYP2C: The CYP2C subfamily contains several members, some of which are expressed at high levels in human liver. More than one member of this subfamily may be capable of metabolizing PAH; thus, human CYP2C9 and, to a lesser extent, CYP2C8 metabolize benzo [a]pyrene to 3- and 9-phenols and trans-dihydrodiols (Shou et al., 1994). In addition, CYP2C enzymes may play an essential role in the metabolism of benzo [a]pyrene and 7,12-dimethyl-benz [a]anthracene, particularly in phenobarbital-induced liver (Morrison et al., 1991a,b; Todorovic et al., 1991). In view of the relative abundance of CYP in human liver and their role in the metabolism of PAH, it has been suggested that some CYP2C enzymes play an essential role in hepatic PAH metabolism (Morrison et al., 1991b; Yun et al., 1992). CYP3A: CYP3A is one of the most abundant CYP enzymes in human liver, and it can metabolize benzo [a]pyrene and some of its dihydrodiols to several metabolic products (Shimada et al., 1989; Yun et al., 1992; Shou et al., 1994; Bauer et al., 1995). In one study, human CYP3A4 was the most important single enzyme in the hepatic 3-hydroxylation of benzo [a]pyrene (Yun et al., 1992). 6.3.1.2 Regulation of cytochrome P450 enzymes that metabolize PAH All of the enzymes discussed above are inducible, and their level of expression can be enhanced by external stimuli. CYP1A and CYP1B are under the transcriptional control of the Ah receptor, which can be activated by numerous PAH and other planar hydrocarbons, including dioxins (Negishi et al., 1981; Guengerich & Shimada, 1991) CYP2B enzymes can also be induced by foreign compounds but not through the Ah receptor. The mechanism of induction of these enzymes is not well understood, but their prototype inducer is phenobarbital; several other drugs used clinically have similar effects (Gonzalez & Gelboin, 1994). The regulation of CYP2C enzymes is complicated, and both endogenous factors such as steroid hormones and exogenous factors such as phenobarbital may be involved. Furthermore, different members of this subfamily are regulated differently. The CYP3A are also regulated by endogenous and exogenous factors; typical inducers of this subfamily are rifampicin, dexamethasone, certain macrolide antibiotics, and steroid hormones (Guengerich & Shimada, 1991). Genetic polymorphisms of CYP1A1, CYP1A2, and some CYP2C and CYP3A enzymes have also been described. Some of the genetic defects leading to the polymorphism have been identified and can be used to predict an individual's capacity to metabolize drugs, for example by the polymerase chain reaction. Genetic polymorphism may lead to dramatic changes in the capacity to metabolize PAH (Raunio & Pelkonen, 1994). Studies with a few prototype compounds such as benzo [a]pyrene and its metabolites and 7,12-dimethylbenz [a]anthracene indicate that several CYP are involved in PAH metabolism. As each has its own metabolic capacity, mode of regulation, and tissue-specific expression, the one that plays a key role in PAH metabolism in vivo at any one time may vary and will depend on the compound being metabolized, pre-exposure to inducers of the CYP, the tissue and cell type where the metabolism is taking place, and the genotype of the individual in cases of genetic polymorphism. Many PAH that are metabolized by the CYP-dependent mono-oxygenases also induce the enzyme system. This ability of hydrocarbons to induce their own metabolism usually results in lower tissue levels and more rapid excretion of the hydrocarbon (Schlede et al., 1970b; Aitio, 1974). Although CYP1A1 is mainly responsible for activation of PAH in the lung and CYP1A2 in the liver, most recent investigations have shown that other CYP isoforms may also contribute to the metabolism of PAH in mammals (Jacob et al., 1996). Thus, pretreatment of animals with inducers of mono-oxygenase systems is frequently associated with a decreased tumour incidence (Wattenberg, 1978). Conversely, studies with strains of mice that differ genetically in the capacity of their mono-oxygenase systems to be induced by PAH indicate that inducibility may also be associated with an increased tumorigenic or toxicological response (Nebert, 1980). Induction of the mono-oxygenase system by different types of inducers can result in different profiles of hydrocarbon metabolites, although the extent of the effect appears to be variable (Holder et al., 1974; Jacob et al., 1981a,b; Schmoldt et al., 1981). The metabolism of benzo [a]pyrene has been investigated in more detail than that of other hydrocarbons and is used here as an example. 6.3.2 Metabolism of benzo[a]pyrene In early studies, the PAH metabolites isolated from or excreted by experimental animals were shown to consist of hydroxylated derivatives, commonly in the form of conjugates. Thus, the general scheme of xenobiotic metabolism outlined above applies to PAH. One of the principal interests in hydrocarbon metabolism arose, however, from the realization that hydrocarbons, like many other environmental carcinogens, are chemically unreactive and that their adverse biological effects are probably mediated by electrophilic metabolites capable of covalent interaction with critical macromolecules such as DNA. Identification of the biologically active metabolites of PAH, coupled with advances in both the synthesis of known and potential hydrocarbon metabolites and the analysis of metabolites by high-performance liquid chromatography, has led in the last two decades to a greatly enhanced appreciation of the complexity of hydrocarbon metabolism. Most of these metabolic interrelationships are illustrated for benzo [a]pyrene in Figure 3; the structures of some types of metabolites are given in Figure 4. The metabolism of benzo [a]pyrene and other PAH has been reviewed (for example, Sims & Grover, 1974, 1981; Conney, 1982; Cooper et al., 1983; Dipple et al., 1984; Hall & Grover, 1990). Benzo [a]pyrene is metabolized initially by the microsomal CYP-dependent mono-oxygenase system to several epoxides (Figure 3). Once formed, these epoxides (Sims & Grover, 1974) may spontaneously rearrange to phenols, be hydrated to dihydrodiols in a reaction that is catalysed by epoxide hydrolase (see review by Oesch 1973), or react covalently with glutathione, either chemically or in a reaction catalysed by glutathione S-transferase (Chasseaud, 1979). 6-Hydroxybenzo [a]pyrene is further oxidized either spontaneously or metabolically to the 1,6-, 3,6-, or 6,12-quinone, and this phenol is also a presumed intermediate in the oxidation of benzo [a]pyrene to the three quinones that is catalysed by prostaglandin H synthase. Two additional phenols may undergo further oxidative metabolism: 3-hydroxybenzo [a]pyrene is metabolized to the 3,6-quinone, and 9-hydroxybenzo [a]pyrene is oxidized to the K-region 4,5-oxide, which is hydrated to the corresponding 9-hydroxy 4,5-dihydrodiol (Jernström et al., 1978; for a formula showing a K-region, see Figure 11). Phenols, quinones, and dihydrodiols can all be conjugated to yield glucuronides and sulfate esters, and the quinones may also form glutathione conjugates (Figure 5). In addition to being conjugated, dihydrodiols can undergo further oxidative metabolism. The mono-oxygenase system metabolizes benzo [a]pyrene 4,5-diol to a number of metabolites, while the 9,10-dihydrodiol is metabolized predominantly to its 1- and 3-phenol derivatives, only minor quantities of a 9,10-diol-7,8-epoxide being formed. In contrast to 9,10-dihydrodiol metabolism, the principal route of oxidative metabolism of benzo [a]pyrene 7,8-dihydrodiol is to a 7,8-diol 9,10-epoxide, and triol formation is a minor pathway. The diol epoxides can themselves be further metabolized to triol epoxides and pentols (Dock et al., 1986) and can become conjugated with glutathione either through chemical reaction or via a glutathione S-transferase-catalysed reaction (Cooper et al., 1980; Jernström et al., 1985; Robertson et al., 1986). They may also spontaneously hydrolyse to tetrols, although epoxide hydrolase does not appear to catalyse this hydration. Further oxidative metabolism of benzo [a]pyrene 7,8-diol can also be catalysed by prostaglandin H synthase (Marnett et al., 1978; Eling et al., 1986; Eling & Curtis, 1992), by a myeloperoxidase system (Mallett et al., 1991), or by lipoxygenases (Hughes et al., 1989). These reactions may be of particular importance in situations in which there are relatively low levels of CYP (i.e. in uninduced cells and tissues) or when chronic irritation and/or inflammation occurs, as during cigarette smoking (Kensler et al., 1987; Ji & Marnett, 1992). The products detected have included diol epoxides (Mallet et al., 1991; Ji & Marnett, 1992) and tetrols (Sivarajah et al., 1979). Taken together, these reactions illustrate that benzo [a]pyrene in particular, and PAH in general, can undergo a multitude of simultaneous or sequential metabolic transformations; they also illustrate the difficulty in determining which metabolites are responsible for the various biological effects resulting from treatment with the parent PAH. An additional complexity of hydrocarbon metabolism stems from the fact that the compounds are metabolized to optically active products. Figure 6 illustrates the stereoselective metabolism of benzo [a]pyrene to the 7,8-diol-9,10-epoxides. Four isomers may be generated, since each diastereomer can be resolved into two enantiomers. In rat liver microsomes, the (+) 7,8-epoxide of benzo [a]pyrene is formed in excess relative to the (-) isomer, such that more than 90% of the benzo [a]pyrene 7,8-oxide formed consists of the (+) enantiomer (Levin et al., 1982). The epoxide is then metabolized stereospecifically by epoxide hydrolase to the (-) 7,8-dihydrodiol. This metabolically predominant dihydrodiol is metabolized in turn, primarily to a single diol epoxide isomer, the (+) anti-benzo [a]pyrene 7,8-diol-9,10-epoxide. The biological significance of the stereoselective formation of the 7,8-diol-9,10-epoxide isomers is that the metabolically predominant isomer is also the isomer with the highest tumour-inducing activity and that found predominantly to be covalently bound to DNA in a variety of mammalian cells and organs that have been exposed to benzo [a]pyrene. Benzo [a]pyrene metabolism has been examined extensively in human tissue preparations, including human cells, explant cultures, tissue homogenates, and microsomal preparations. Table 73 lists some studies of the metabolism of benzo [a]pyrene in human tissues that included metabolites soluble in organic solvents and water-soluble conjugates. The results show that the metabolites produced by different human tissues are qualitatively similar and that the metabolites detected are the same as those formed in a variety of animal tissues. The metabolic profiles reported in human tissues are almost all identical to those seen for other eukaryotes, indicating the involvement of similar enzyme systems. The same types of reactive electrophilic intermediates found in other experimental systems also appear to be formed in human tissues (Autrup & Harris, 1983). So far, no differences in the metabolism or activation of benzo [a]pyrene have been reported that might account for differences in the susceptibility of different animal and human tissues to its carcinogenic properties (see Section 7). Studies with cultured cells and other substrates such as benz [a]anthracene, however, indicate that bioactivation of PAH is species-dependent (Jacob, 1996). 6.4 Elimination and excretion Most metabolites of PAH are excreted in faeces and urine. As complete breakdown of the benzene rings of which unsubstituted PAH are composed does not occur to any appreciable extent in higher organisms, very little of an administered dose of an unsubstituted hydrocarbon would be expected to appear as carbon dioxide in expired air. The urinary excretion of PAH metabolites has been studied more extensively than faecal excretion, but the importance of the enterohepatic circulation of metabolites has led to increased research on the latter. Detailed studies of the metabolism and excretion of PAH in whole animals have been restricted mainly to the simpler compounds. Because of the toxicity of the larger hydrocarbons and the complexity of their metabolism, most studies on these compounds have been carried out in hepatic homogenates and microsomal preparations or with cultured cells (see above). Metabolism and excretion in whole animals have been examined with regard to naphthalene (Bourne & Young, 1934; Young, 1947; Booth & Boyland, 1949; Corner & Young, 1954; Corner et al., 1954; Boyland & Sims, 1958; Sims, 1959), anthracene (Boyland & Levi, 1935, 1936a,b; Sims, 1964), phenanthrene (Boyland & Wolf, 1950; Sims, 1962; Boyland & Sims, 1962a,b; Jacob et al., 1990b; Grimmer et al., 1991a), pyrene (Harper, 1957, 1958a; Boyland & Sims, 1964a; Jacob et al., 1989, 1990b), benz [a]-anthracene (Harper 1959a,b; Boyland & Sims, 1964b), and chrysene (Grimmer et al., 1988b, 1990). A limited number of studies have been published on more complex compounds such as benzo [a]pyrene (Berenblum & Schoental, 1943; Weigert & Mottram, 1946; Harper, 1958b,c; Falk et al., 1962; Raha, 1972; Jacob et al., 1990b), dibenz [a,h]anthracene (Dobriner et al., 1939; Boyland et al., 1941; La Budde & Heidelberger, 1958), and 3-methylcholanthrene Table 73. Metabolites of benzo[a]pyrene formed by human tissues and cells Tissue or Type of metabolite detected References cell type Dihydrodols Phenols Quinones Tetrols Conjugates Bronchus + + + + + Pal et al. (1975); Cohen et al. (1976); Harris et al. (1977); Autrup et al. (1978a, 1980) Colon + + + + + Autrup et al. (1978b); Autrup (1979) Endometrium + + + Mass et al. (1981) Fibroblasts + Baird & Diamond (1978) Kidney + + + Prough et al. (1979) Liver + + + + Selkirk et al. (1975); Prough et al. (1979); Pelkonen et al. (1977); Diamond et al. (1980) Lung + + + + + Cohen et al. (1976); Stoner et al. (1978); Mehta et al. (1979); Prough et al. (1979); Sipal. et al. (1979) Lymphocytes + + + Booth et al. (1974); Selkirk et al. (1975); Vaught et al. (1978); Okano et al. (1979); Gurtoo et al. (1980) Macrophages + + + + + Autrup et al. (1978a); Harris et al. (1978a,b); Autrup et al. (1979); Marshall et al. (1979) Table 73 (contd) Tissue or Type of metabolite detected References cell type Dihydrodols Phenols Quinones Tetrols Conjugates Mammary + Grover et al. (1980); epithelium MacNicoll et al. (1980) Monocytes + + + Vaught et al. (1978); Okano et al. (1979) Oesophagus + + + + Harris et al. (1979) Placenta + + + Namkung & Juchau (1980); Pelkonen & Saarni (1980) Skin + + + + Fox et al. (1975); Vermorken et al. (1979); Parkinson & Newbold (1980); Kuroki et al. (1980) (Harper, 1959a; Takahashi & Yasuhira, 1972; Takahashi, 1978). Much of the earlier qualitative work was reviewed by Boyland & Weigart (1947) and by Young (1950). The absorption and excretion of different hydrocarbons in vivo can differ. For example, while almost all of a topically applied dose of benzo [a]pyrene appeared in mouse faeces (Heidelberger & Weiss, 1951), little dibenz [a,h]-anthracene was excreted by this route. In rats given PAH either singly or as mixtures, the faecal elimination of chrysene (25% of the dose) was not affected by co-administration of benz [a]anthracene, but that of benz [a]anthracene was doubled, from 6 to 13% of the dose, when chrysene was given (Bartosek et al., 1984). Such effects are relevant to human pharmacokinetics, since exposure is almost always to mixtures of PAH. In workers in a coke plant exposed to mixtures of PAH, the amounts of phenanthrene, pyrene, and benzo [a]pyrene inhaled and the amounts of their principal metabolites excreted in the urine were correlated (Grimmer et al., 1994). In rats, the amount of benzo [a]pyrene 7,8-diol excreted in the urine is related to the susceptibility of individual animals to the carcinogenic effects of benzo [a]pyrene (Likhachev et al., 1992; Tyndyk et al., 1994). In studies of the disposition of benzo [a]pyrene in rats, hamsters, and guinea-pigs after intratracheal administration, the distribution of the hydrocarbon was qualitatively similar but quantitatively different. In Sprague-Dawley and Gunn rats and in guinea-pigs, the rate of excretion was dependent on the dose administered, but in hamsters the rate of excretion was independent of dose (0.16 or 350 µg 3H-benzo [a]pyrene) (Weyand & Bevan 1986, 1987a). Evidence for enterohepatic circulation of benzo [a]pyrene metabolites was obtained in Sprague-Dawley rats with bile-duct cannulae treated by intratracheal instillation with 1 µg/kg bw 3H-benzo [a]pyrene (Weyand & Bevan, 1986). The results of a study of the pharmacokinetics and bioavailability of pyrene in rats strongly suggested that enterohepatic recycling took place after oral or intravenous administration of 14C-labelled compound at 2-15 mg/kg bw (Withey et al., 1991). Other studies on the enterohepatic circulation of PAH in rats and rabbits have also shown that the significant amounts of metabolites excreted in the bile persist in vivo because of enterohepatic circulation (Chipman et al., 1981; Chipman, 1982; Boroujerdi et al., 1981). For example, while some 60% of an intravenous dose of 3 µmol/kg bw 14C-benzo [a]pyrene was excreted in bile, only 3% appeared in urine within the first 6 h after injection (Chipman et al., 1981). Biliary metabolites of xenobiotic compounds are usually polar and nonreactive, but mutagenic or potentially mutagenic derivatives may be excreted by this route into the intestine (for a review, see Chipman, 1982). Glucuronic acid conjugates of biliary metabolites can be hydrolysed by some intestinal flora to potentially reactive species (Renwick & Drasar, 1976; Chipman et al., 1981; Boroujerdi et al., 1981; Chipman, 1982). Thio-ether conjugates of hydrocarbons may also be involved in enterohepatic circulation (Hirom et al., 1983; Bakke et al., 1983), although there is no evidence that these represent a mutagenic or carcinogenic hazard to the tissues through which they pass. In a controlled study in humans, a 100-250-fold increase in dietary exposure to PAH, as measured by benzo [a]pyrene intake, resulted in a 4-12-fold increase in urinary excretion of 1-hydroxypyrene. The authors concluded that dietary exposure to PAH is as substantial as some occupational exposures (Buckley & Lioy, 1992). 6.5 Retention and turnover Very little is known about the retention and turnover of PAH in mammalian species. It can be deduced from the few data available on hydrocarbon body burdens (see below) that PAH themselves do not persist for long periods and must therefore turn over reasonably rapidly. During metabolism, PAH moieties become covalently bound to tissue constituents such as proteins and nucleic acids. Protein-bound metabolites are likely to persist, therefore, for periods that do not exceed the normal lifetime of the protein itself. Nucleic acid adducts formed from reactions of PAH metabolites can be expected to differ in their persistence in the body according to whether they are RNA or DNA adducts. Although most DNA adducts are removed relatively rapidly by repair, small fractions can persist for long periods. The persistence of these adducts in tissues such as mouse skin is of considerable interest since one of the basic features of the two-stage mechanism of carcinogenesis (Berenblum & Shubik, 1947) is that application of the tumour promoter can be delayed for many months without markedly reducing the eventual tumour yield. The persistence of adducts is also consistent with multistage theories of carcinogenesis, in which multiple steps in neoplastic transformation are dependent on the mutagenic and other actions of carcinogens. 6.5.1 Human body burdens of PAH Since the effects of chemical carcinogens are likely to be related to both the dose and the duration of exposure, it is important to determine the human body load of carcinogens during a lifetime. It has been estimated that the total intake of PAH over a 70-year lifespan may amount to the equivalent of 300 mg of benzo [a]pyrene (Lutz & Schlatter, 1992); however, inhabitants of conurbations are likely to inhale additional amounts of PAH. Of course, much of the intake of PAH is metabolized and excreted. Thus, the pulmonary tissues of elderly town dwellers in Russia contained 1000 times less benzo [a]pyrene (< 0.1 µg per individual) than might have been expected from the estimated intake figures alone (Shabad & Dikun, 1959). Some experiments with cows and domestic fowl fed diets containing added benzo [a]pyrene tend to confirm this finding, since the meat, milk, and eggs produced were, after a suitable delay, reported to be much less heavily contaminated than might have been expected from the amounts of benzo [a]pyrene administered (Gorelova & Cherepanova, 1970). More recent data are not available. The average benzo [a]pyrene levels (measured by ultraviolet spectroscopy) in tissues taken at autopsy from normal people of a wide age range were 0.32 µg/100 g dry tissue weight in liver, spleen, kidney, heart, and skeletal muscle and 0.2 µg/100 g in lung (Gräf, 1970; Gräf et al., 1975). When cancer-free liver and fat from six individuals were assayed for nine hydrocarbons by co-chromatography with authentic standards, pyrene, anthracene, benzo [b]fluoranthene, benzo [ghi]perylene, benzo [k]fluoranthene, and benzo [a]pyrene were detected at average levels of 380 ppt (0.38 µg/kg wet weight) in liver and 1100 ppt (1.1 µg/kg wet weight) in fat. Pyrene was the most abundant PAH present (Obana et al., 1981b). Samples of 24 bronchial carcinomas, taken during surgery or at autopsy from smokers and nonsmokers with a variety of occupations, were analysed for the presence of 12 PAH by thin-layer chromatography and fluorescence spectroscopy. Benzo [a]pyrene, benzo [b]fluoranthene, fluoranthene, and perylene were detected. Benzo [a]pyrene was present, but the other three PAH were found in only some of the samples. The average concentrations of benzo [a]pyrene were 3.5 µg/g in carcinoma tissue and 0.09 µg/g in tumour-free tissue (Tomingas et al., 1976). 6.6 Reactions with tissue components The reactions of metabolites of PAH with tissue constituents (Weinstein et al., 1978) are relevant because they may indicate the mechanisms by which the hydrocarbons exert biological effects that include toxicity and carcinogenesis. 6.6.1 Reactions with proteins Covalent interactions of PAH with protein in whole animals were first noted in 1951 (Miller, 1951). It was proposed that reactions with specific proteins might be involved in the initiation of malignancy in liver (Miller & Miller, 1953), skin (Abell & Heidelberger, 1962), and transformable cells in culture (Kuroki & Heidelberger, 1972). These findings were supported by evidence that hydrocarbon metabolites can react covalently with protein in microsomal incubates (Grover & Sims, 1968), in preparations of nuclei (Vaught & Bresnick, 1976; Pezzuto et al., 1976, 1977; Hemminki & Vainio, 1979), and in cells and tissues maintained in culture, including human tissues (Harris et al., 1978b; MacNicoll et al., 1980). Although hydrocarbon metabolites often react at much greater rates with protein than with nucleic acids in the same biological system, relatively little attention has been paid to the nature of the hydrocarbon metabolites involved or to the specificity of these reactions, in terms of which proteins are most extensively modified and where and the effect that such modification might have on protein function. The evidence suggests, however, that the reactive species involved include diol epoxides. Thus, when protein isolated from the skin of mice that had been treated with benzo [a]pyrene was hydrolysed, tetrols were liberated, and the patterns of specific tetrols indicated that both syn and anti isomers of the benzo [a]pyrene 7,8-diol 9,10-oxides are involved in covalent reactions with protein (Koreeda et al., 1978). Studies of the covalent interactions of diol epoxides with nuclear proteins show that a variety of histones and non-histone proteins are modified (Kootstra & Slaga, 1979; Kootstra et al., 1979; Whitlock, 1979). 6.6.2 Reactions with nucleic acids The covalent interactions of electrophilic metabolites of PAH with nucleic acids have been studied in much greater detail than those with protein, partly because characterization of the products might, in theory, be expected to be simpler, partly because the cellular nucleic acids are, as nucleophiles, more 'homogeneous' than proteins, but mainly because it has long been suspected that nucleic acid modifications could lead to a permanent alteration of cell phenotype. The covalent binding of a PAH (dibenz [a,h]anthracene) to DNA in vivo was first reported by Heidelberger & Davenport in 1961. Subsequent studies with naphthalene, dibenz [a,c]anthracene, dibenz [a,h]anthracene, benzo [a]-pyrene, 3-methylcholanthrene, and 7,12-dimethylbenz [a]anthracene showed that the levels of DNA binding in mouse skin are correlated with carcinogenic potency, as measured by Iball's index (Brookes & Lawley, 1964). 6.7 Analytical methods Of the methods used for the detection of carcinogen-DNA adducts (Phillips, 1990; Strickland et al., 1993; Weston, 1993), one of the most widely used is 32P-postlabelling, in which DNA is hydrolysed to nucleotides, modified nucleotides (i.e. adducts) are labelled with 32P-phosphate, and the post-labelled adducts separated by thin-layer chromatography and/or high-performance liquid chromatography (for reviews of the method, see Phillips, 1991, and Phillips et al., 1993). The main advantages of the 32P-postlabelling assay are its high sensitivity and the fact that radiolabelled carcinogens and/or their metabolites need not be synthesized beforehand. A variety of physical methods have been described for the detection of adducts, including fluorescence line narrowing spectroscopy, synchronous fluorescence spectroscopy, and some specialized gas chromatography-mass spectrometry procedures (Weston, 1993). The physical methods combine high sensitivity with no requirement for prior radiolabelling of the carcinogens or their adducts and may be nondestructive. Sensitive methods involving antisera specific for carcinogen-DNA adducts have also been developed. These include radioimmunoassays, enzyme-linked immunosorbent assays, and immuno-affinity chromatography (Poirier, 1994). Information on the pathways thought to be involved in the metabolic activation of several PAH is given in Table 74. For PAH that have been extensively investigated, reviews are cited. In order to provide an overall view of activation, the Table also includes data on PAH not covered elsewhere in this monograph. Most of the metabolites that have been found to react with nucleic acids are vicinal diol epoxides, and most of these are diol epoxides of the 'bay-region' type, although there are certain exceptions (Table 74). For example, activation of benzo [j]fluoranthene in mouse skin involves a diol epoxide that is not of the bay-region type (Weyand et al., 1993). Additionally, methyl-substituted PAH may become bound to hydroxymethyl derivatives which, when conjugated, yield electrophilic sulfate esters (Surh et al., 1989, 1990a,b). The sites of attack on nucleic acid bases are usually the extranuclear amino groups of guanine and adenine. When the reactions of the syn and anti isomers of benzo [a]pyrene 7,8-diol-9,10-oxide with RNA, DNA, and homopolymers were examined in experiments in which the epoxide was incubated with the nucleic acid in a predominantly aqueous solution, RNA, DNA, poly G, poly A, poly C, poly (dG), poly (dA), and poly (dC) were modified, but there was little reaction with poly U, poly I, or poly (dT) (Weinstein et al., 1976; Jennette et al., 1977). Although many of the hydrocarbon-deoxyribonucleoside adducts formed in human cells and tissues treated with PAH have not been completely characterized, the available evidence, which is mostly chromatographic, suggests that in human bronchial epithelium, colon, mammary cells in culture, and skin the patterns of adducts formed are very similar to those formed in corresponding rodent tissues (Autrup et al., 1978a,b; Harris et al., 1979; Autrup et al., 1980; MacNicoll et al., 1980; Weston et al., 1983). The rates of reaction of diol epoxides with nucleic acids was in the general order: poly G > DNA > poly A > poly C (Jennette et al., 1977). Diol epoxides are also strongly suspected to react frequently with the N7 position of guanine. This type of modification has not been detected more often because N7-alkylated adducts are thought to have a relatively short half-life at pH 7 and would therefore be lost during the isolation and hydrolysis of DNA. In experiments in which care was taken to avoid adduct loss, reactions of benzo [a]pyrene diol epoxide with both the N2 and N7 positions of guanine residues in DNA were detected (Osborne et al., 1978). N7 adducts were not, however, detected in cells treated with anti-benzo [a]pyrene 7,8-diol-9,10-oxide (King et al., 1979). In studies of the role of radical cations in the activation of PAH in vitro, adducts were formed in which the 6 position of benzo [a]pyrene was covalently linked to the C8 and N7 positions of guanine and the N7 position of adenine, and the 7-methyl position of 7,12-dimethylbenz [a]anthracene was covalently linked to the N7 positions of guanine and adenine (see Figure 7; Cavalieri et al., 1993; Rogan et al., 1993). All of these adducts are depurination adducts, which may explain why they were not detected earlier Table 74. Pathways involved in the metabolic activation of polycyclic aromatic hydrocarbons to form ultimate carcinogens Compound Derivatives with highest Putative ultimate carcinogen Reference levels of biological activity Aceanthrylene 1,2-Oxidea Nesnow et al. (1991) Benz[j]aceanthrylene ? 1,2-Oxideb Bartczak et al. (1987); Nesnow et al. (1988) Benz[l]aceanthrylene ? 1,2-Oxideb,c Nesnow et al. (1984); Bartzczak et al. (1987); Nesnow et al. (1988) Benz[a]anthracene 3,4-Diold,e,f,g 3,4-Diol 1,2-oxldea,b,c,f,g Sims & Grover (1981); 8,9-Diold 8,9-Diol 10,1-oxidea,h Conney (1982); Wood et al. (1983a) Benzo[b]fluoranthene 9,10-Dlold,f,i ? 910-Diol-11,12-oxide Geddie et al. (1987); and 5/6-hydroxy-9,10- Pfau et al. (1992) diol-11, 12-oxide Benzo[b]fluoranthene ? 9,10-Diolf,j ? 9,10-Diol 11,12-oxidea Rice et al. (1987); Weyand et al. (1993) ? 4,5-Diola ? 4,5-Diol 6,6a-oxidea Weyand et al. (1987) Benzo[c]phenanthrene 3,4-Diold,e,f,g 3,4-Diol 1,2-oxidea,b,c,f,g Conney (1982); Levin et al. (1986); Agarwal et al. (1987); Dipple et al. (1987); Pruess-Schwartz et al. (1987) Benzo[a]pyrene 7,8-Diold,e,f,h 7,8-Diol 9,10-oxidea,b,c,g Cooper et al. (1983); Osborne & Crosby (1987a) Table 74. (continued) Compound Derivatives with highest Putative ultimate carcinogen Reference levels of biological activity Benzo[e]pyrene 9,10-Diolf ? 9,10-Diol 11,12-oxideg Osborne & Crosby (1987b) Chrysene 1,2-Diold,e,f 1,2-Diol 3,4-oxidea,b,c,h Conney (1982); 9-Hydroxy 1,2-diold,e 9-Hydroxy-1,2-diol Hodgson et al. (1983); 3,4-oxideb,c Glatt et al. (1986) Cyclopenta[cd]pyrene - ? 3,4-oxideb,c,h Gold & Eisenstadt (1980); Gold et al. (1980) 15,16-Dihydro-11-methylcyclo- 3,4-Diold,f 3,4-Diol 1,2-oxidea Coombs & Bhatt (1987) penta[a]phenanthren-17-one 15,16-Diydro-1,11-methano- 3,4-Diold 3,4-Diol 1,2-oxide Coombs & Bhatt (1987) cyclopenta[a]phenanthren-17-one Dibenz[a,c]anthracene 10,11-Diold ? 10, 11-Diol 12,13-oxide Sims & Grover (1981) Dibenz[a,h]anthracene 3,4-Diold,f,g,h ? 3,4-Diol 1,2-oxide and Conney (1982); 3,4:10,1 1-bis-diol-epoxides Lecoq et al. (1991, 1992); Carmichael et al. (1993); Nesnow et al. (1994) Dibenzo[a,e]fluoranthene 12,13-Diold,f 12,13-Diol 10-11-oxidea Perin-Roussel et al. (1983,1984); 3,4-Diold,f 3,4-Diol 1,2-oxidea Saguem et al. (1983a,b); Zajdela et al. (1987) Dibenzo[a,h]pyrene 1,2-Diolf,g ? 1,2-Diol 3,4-oxideg Chang et al. (1982) Dibenzo[a,l]pyrene ? 11,12 Diolf ? 11,12-Diol 13,14-oxide Cavalieri et al. (1991) Table 74. (continued) Compound Derivatives with highest Putative ultimate carcinogen Reference levels of biological activity Dibenzo[a,i]pyrene 3,4-Diolf,g ? 3,4-Diol 1,2-oxideg Chang et al. (1982) 7,12-Dimethylbenz[a]anthracene 3,4-Diold,e,f,h 3,4-Diol 1,2-oxidea Sims & Grover (1981); Conney (1982); Sawicki et al. (1983); Dipple et al.; 1984) 7-Ethylbenz[a]anthracene 3,4-Diold ? 3,4-Diol 1,2-oxidea,b McKay et al. (1988); Glatt et al. (1989) Fluoranthene Z3,Diold 2,3-Diol 1,10b-oxidea La Voie et al. (1982a); Rastetter et al. (1982); Babson et al. (1986a); Hecht et al. (1995) Indeno[1,2,3-cd]pyrene 1,2-oxideb,f ? Rice et al. (1985) 1,2-Diolf Rice et al. (1986) 8-Hydroxyd 9-Hydroxyd 7-Methylbenz[a]anthracene 3,4-Diold,e,f,h 3,4-Diol 1,2-oxidea,b Sims & Grover (1981); McKay et al. (1988); Glatt et al. (1989) 3-Methylcholanthrene 9,10-Diold,f,h ? 9,10-Diol 7,13-oxidea,f Sims & Grover (1981); ? 3-Hydroxymethyl-9,10- Conney (1982); diol 7,8-oxide DiGiovanni et al. (1985); Osborne et al. (1986) 5-Methylchrysene 1,2-Diold,f 1,2-Diol 3,4-oxidea,c,h Hecht et al. (1986); Brookes et al. (1986); Reardon et al. (1987); Hecht et al. (1987) Table 74 (continued) a DNA adducts characterized b Directly acting mutagen in S. typhimurium c Directly acting mutagen in V79 Chinese hamster cells d Mutagenic to S. typhimurium with metabolic activation e Mutagenic to V79 Chinese hamster cells with metabolic activation f Tumour initiator in mouse skin g Induces tumours in newborn mice h Transforms cells in culture i Not detected as a metabolite; activation may therefore occur via a different pathway. j Although the 45-diol is the most active derivative so far tested, there is some evidence that adducts arise from the 9,1-diol. in vivo. The formation of apurinic sites in DNA could lead to strand nicking (Gamper et al., 1977, 1980). When the positions of the nicks produced as a result of modification by benzo [a]pyrene 7,8-diol-9,10-oxide were investigated with DNA of a defined sequence, nicking appeared to be the result of the loss of purines and pyrimidines that had been modified at the N7 position of guanine or at the N3 position of adenine and cytosine (Haseltine et al., 1980). In studies of the distribution of covalently bound benzo [a]pyrene moieties in chromatin, more was bound to the inter-nucleosomal spacer regions of DNA than to DNA in nucleosomes (Jahn & Litman, 1977, 1979; Kootstra & Slaga, 1980). One explanation for this finding may be that nucleosomal DNA is better protected from modification by the presence of nucleoproteins; results consistent with this suggestion have been obtained with mitochondrial DNA. Graffi (1940a,b,c) suggested that lipophilic PAH accumulate in lipid-rich mitochondria. Allen & Coombs (1980) and Backer & Weinstein (1980) showed much higher levels of modification of mitochondrial than nuclear DNA in cultured cells treated with either benzo [a]pyrene or the anti-benzo [a]pyrene 7,8-diol-9,10-oxide. The molecular properties of adducts of benzo [a]pyrene 7,8-dihydrodiol-9,10-epoxides with DNA have been described (Geacintov 1988; Jernström & Gräslund, 1994). Although the biological effectiveness of all types of hydrocarbon-nucleic acid adducts has not been determined, it has been shown that differences in the biological activities of 7-ethyl- and 7-methylbenz [a]-anthracene are not due to differences in the mutagenic potential of the adducts formed (Glatt et al., 1989). Similar conclusions were drawn from work with a series of bay-region and fjord-region diol epoxides (Phillips et al., 1991; see section 7.10 for a description of a fjord region). At present, therefore, all hydrocarbon-deoxyribonucleoside adducts should be regarded as potentially damaging to the organism. The relationships between DNA adduct formation and tumour incidence were examined by Poirier & Beland (1992) on the basis of data from long-term studies in rodents administered carcinogens. The tumour incidence was compared with adduct levels measured in target tissues during the first two months of exposure. In most cases, linear increases in DNA adduct levels with dose were reflected in linear increases in tumour incidence, although there were exceptions. In a comparison of the incidence of lung adenomas in strain A/J mice 240 days after they had received a single intraperitoneal injection of benzo [a]pyrene, dibenz [a,h]anthracene, benzo [b]fluoranthene, 5-methyl-chrysene, or cyclopenta [cd]pyrene with the levels of DNA adducts detected in the lungs by 32P-postlabelling between days 1 and 21 after treatment, time-integrated DNA adduct levels were calculated and plotted against lung adenoma frequency. The slopes obtained were essentially similar for benzo [a]pyrene, benzo [b]fluoranthene, 5-methylchrysene, and cyclopenta [cd]pyrene but were different for dibenz [a,h]anthracene. The authors concluded that 'essentially identical induction of adenomas as a function of [time-integrated DNA adduct levels] for these PAH suggests that the formation and persistence of DNA adducts determines their carcinogenic potency' (Ross et al., 1995). 7. EFFECTS ON LABORATORY MAMMALS AND IN VITRO Appraisal Single doses of polycyclic aromatic hydrocarbons (PAH) have moderate to low toxicity, with LD50 values generally > 100 mg/kg bw after intraperitoneal or intravenous injection and > 500 mg/kg bw after oral administration. Because most of the experimental studies have addressed the carcinogenicity of PAH, the database on their short- and long-term toxicity is quite small. In short-term studies, effects on the haematopoietic system were observed, e.g. benzo [a]pyrene caused myelotoxicity and dibenz [a,h]anthracene caused haemolymphatic alterations in mice. Anaemia is a typical effect of naphthalene. Values for a no-observed-adverse-effect level (NOAEL) and a lowest-observed-adverse-effect level (LOAEL) have been obtained in 90-day studies by oral administration. The NOAEL values based on haematological effects and hepato- and nephrotoxicity were 75-1000 mg/kg bw per day for the noncarcinogenic PAH acenaphthene, anthracene, fluoranthene, fluorene, and pyrene. Few studies have been conducted on dermal or ocular irritation. PAH do, however, have adverse effects after dermal administration, such as hyperkeratosis, which are correlated with their carcinogenic potency. Anthracene and naphthalene were reported to cause mild ocular irritation. The ocular toxicity of naphthalene is characterized by cataract formation. Benzo [a]pyrene caused skin hypersensitization. Anthracene and benzo [a]pyrene have been shown to have phototoxic potential and benzo [a]pyrene, dibenz [a,h]anthracene, and fluoranthene to have immunotoxic potential. PAH can cross the placenta and induce adverse effects on the embryo and fetus. Benz [a]anthracene, benzo [a]pyrene, dibenz [a,h]anthracene, and naphthalene were found to be embryotoxic. Benzo [a]pyrene also reduced female fertility and had effects on oocytes and on postnatal development. Studies on the effects of benzo [a]pyrene in mice with different genotypes demonstrated the importance of the genetic predisposition of animals or embryos for the development of overt toxic effects. A crucial genetic property is the presence or absence of the arylhydrocarbon (Ah) receptor, which induces the monooxygenase system; organisms can thus be divided into Ah responders and Ah non-responders. Mutagenicity has been investigated intensively in a broad range of assays. The only compounds that are clearly not mutagenic are naphthalene, fluorene, and anthracene. The evidence for five PAH is considered to be questionable because of a limited database, while the remaining 25 PAH are mutagenic (see Table 87). Mutagenicity is strictly dependent on metabolic activation of parent compounds. In bacteria and other cell systems that have no metabolizing system, a 9000 × g microsomal preparation of liver (S9 mix) must be added as a metabolic activator. Comprehensive work on the carcinogenicity of these compounds has yielded negative results for fluorene, anthracene, 1-methylphenanthrene, triphenylene, perylene, benzo[ghi]fluoranthene and benzo[ghi]perylene, some of which have been shown to be mutagenic. The evidence for a further nine PAH was classified as questionable, while the other 17 compounds were carcinogenic. Generally, the site of tumour development depends on the route of administration but is not restricted to those sites. Tissues such as the skin can metabolize PAH to their ultimate metabolites, thus becoming target organs themselves, and metabolites formed in the liver can reach various sites of the body via the bloodstream. The carcinogenic potency of PAH differs by three orders of magnitude, and toxic equivalence factors have been used to rank individual PAH (see Appendix I). The various theories for the mechanism of the carcinogenicity of PAH take into account chemical structure and ionization potential. The most prevalent theories are those involving the bay region and radical cations. The bay-region theory is based on the assumption that diol epoxides of the parent compounds are the ultimate carcinogens, which react with electrophilic epoxide groups on N atoms of DNA purines. The radical cation theory postulates the one-electron oxidation of PAH to form strong electrophiles which then react with DNA bases. These theories have been confirmed experimentally by detection of the corresponding DNA adducts in the PAH that have been investigated. Nevertheless, there is general agreement that any one theory cannot cover the mechanisms of action of all PAH. 7.1 Toxicity after a single exposure Few studies are available on the acute toxicity of PAH, except for naphthalene. The LD50 values (Table 75) indicate that the acute toxicity is moderate to low. The results of all of these studies are summarized, even when a study was old and followed a non-systematic protocol, in the absence of alternatives. 7.1.1 Benzo [a]pyrene In young rats, a single intraperitoneal injection of 10 mg benzo [a]pyrene per animal caused an immediate, sustained reduction in the growth rate (Haddow et al., 1937). In mice, a single intraperitoneal injection (dose not specified) resulted in small spleens, marked cellular depletion, prominent haemosiderosis, and follicles with large lymphocytes, leading to death (Shubik & Della Porta, 1957). After a single application of 0.05 ml of a 1% solution in acetone to the interscapular area of hairless mice (hr/hr strain), the mitotic rate of epidermal cells was increased (Elgjo, 1968). 7.1.2 Chrysene In young rats, single intraperitoneal injections of 30 mg chrysene per animal did not reduce growth (Haddow et al., 1937). Table 75. Toxicity of single doses of polycyclic: aromatic hydrocarbons Compound Species Route of LD50 (mg/kg) or Reference administration LC50 (mg/litre) Anthracene Mouse Oral 18 000 Montizaan et al. (1989) Mouse Intraperitoneal > 430 Salamone (1981) Benzo[a]pyrene Mouse Oral > 1 600 Awogi & Sato (1989) Mouse Intraperitoneal approx. 250 Salamone (1981) Mouse Intraperitoneal > 1 600 Awogi & Sato (1989) Rat Subcutaneous 50 Montizaan et al (1989) Chrysene Mouse Intraperitoneal > 320 Simmon et al. (1979) Fluoranthene Rat Oral 2 000 Smyth et al. (1962) Rabbit Dermal 3 180 Smyth et al. (1962) Mouse Intravenous 100 Montizaan et al. (1989) Naphthalene Rat Oral 1 250 Sax & Lewis (1984) Rat (M) Oral 2 200 Gaines (1969) Rat (F) Oral 2 400 Gaines (1969) Rat Oral 9 430 US Environmental Protection Agency (1978a) Rat Oral 1 110 Montizaan et al. (1989) Rat Oral 490 Montizaan et al. (1989) Rat Oral 1 800 Montizaan et al. (1989) Rat (M) Dermal > 2 500 Gaines (1969) Rat (F) Dermal > 2 500 Gaines (1969) Rat Intraperitoneal approx. 1 000 Bolonova (1967) Rat (M) Intraperitoneal approx. 1 600 Plopper et al. (1992) Rat Inhalation > 0.5 mg/litre (8 h) US Environmental Protection Agency (1978a) Mouse (F) Oral 354 Plasterer et al. (1985) Mouse (M) Oral 533 Shopp et al. (1984) Mouse (F) Oral 710 Shopp et al. (1984) Mouse Subcutaneous 5 100 Sandmeyer (1981); Shopp et al. (1984) Mouse Subcutaneous 969 Sax & Lewis (1984) Mouse Intraperitoneal 150 Sax & Lewis (1984) Table 75. (continued) Compound Species Route of LD50 (mg/kg) or Reference administration LC50 (mg/litre) Mouse Intraperitoneal 380 Warren et al. (1982) Mouse (M) Intraperitoneal approx. 400 Plopper et al. (1992) Mouse Intravenous 100 Sax & Lewis (1984) Hamster (M) Intraperitoneal approx. 800 Plopper et al. (1992) Guinea-pig Oral 1 200 Sax & Lewis (1984) Phenanthrene Mouse Oral 700 Montizaan et al. (1989) Mouse Oral 1 000 Montizaan et al. (1989) Mouse Intraperitoneal 700 Simmon et al. (1979) Mouse Intravenous 56 Montizaan et al. (1989) Pyrene Mouse Intraperitoneal 514 (7 d) Salamone (1981) Mouse Intraperitoneal 678 (4 d) Salamone (1981) LC50, median lethal concentration; LD50, median lethal dose; M, male; F, female 7.1.3 Dibenz [a,h]anthracene One or two intraperitoneal injections of 3-90 mg dibenz [a,h]anthracene per animal within two days led to a reduction in the growth rate of young rats that persisted for at least 15 weeks (Haddow et al., 1937). 7.1.4 Fluoranthene In young rats, a single intraperitoneal injection of 30 mg fluoranthene per animal did not inhibit growth (Haddow et al., 1937). 7.1.5 Naphthalene After oral administration of 1-4 g/kg bw naphthalene to dogs or 1-3 g/kg bw to cats, diarrhoea was observed. Rabbits given 1-3 g/kg bw showed corneal clouding (Flury & Zernik, 1935). After intravenous injection of 1-6 mg napthalene to white male rabbits weighing 3-4 kg, no haemolytic effect was seen (Mackell et al., 1951) In mice, Clara cells of the bronchiolar epithelium are the primary targets of low doses of naphthalene. Dose-dependent bronchiolar epithelial cell necrosis was detected after intraperitoneal injection of a single dose of 50, 100, or 200 mg/kg bw per day to mice (O'Brien et al., 1989). Severe bronchiolar epithelial cell necrosis was also seen in mice within 2-4 h after intraperitoneal injection of 200-375 mg/kg bw; hepatic and renal necrosis were not observed (Warren et al., 1982). Alterations in the morphology of Clara cells were observed as early as 6 h after intraperitoneal injection of 64 mg/kg bw; ciliated cells were also affected after 24 and 48 h and at doses up to 256 mg/kg bw. After a 4-h inhalation of 1.0 mg/litre naphthalene, bronchiolar necrosis was detected in mice but not in rats (Buckpitt & Franklin, 1989; see also section 7.2.1). After single injections of 50-400 mg/kg bw to mice, 100-800 mg/kg bw to hamsters, and 200-1600 mg/kg bw to rats, Clara cells in mice showed the effects described above; those of rats showed no significant effects, and minor effects were observed in hamsters. The trachea and lobar bronchi showed swelling and vacuolation of non-ciliated cells in mice, no effects in rats, and cytotoxic changes in hamsters. In the nasal cavity, cytotoxicity to the olfactory epithelium with necrosis was observed in mice and hamsters at 400 mg/kg bw and in rats at 200 mg/kg bw (Plopper et al., 1992). Mice injected intraperitoneally with 200-600 mg/kg bw naphthalene showed dose-dependent abnormalities in the bronchial region (Clara cells) in studies in which the lungs were examined by scanning electron micrography. No pulmonary damage was detected at 100 mg/kg bw. Depletion of pulmonary glutathione, which protects against the toxicity of xenobiotics, was observed within 6 h of naphthalene administration (Honda et al., 1990). The doses and detailed findings of experiments with single doses of naphthalene are summarized in Table 76. 7.1.6 Phenanthrene After acute intraperitoneal injection to rats (dose not specified), liver congestion with a distinct lobular pattern was observed as well as alterations in some serum parameters (Yoshikawa et al., 1987). 7.1.7 Pyrene In young rats, single intraperitoneal injections of 10 mg pyrene per animal did not lead to a reduction in growth rate (Haddow et al., 1937). 7.2 Short-term toxicity 7.2.1 Subacute toxicity 7.2.1.1 Acenaphthene Four of five mice given 500 mg/kg bw per day acenaphthene intraperitoneally for seven days survived (Gerarde, 1960). 7.2.1.2 Acenaphthylene Nine of 10 mice given 500 mg/kg bw per day acenaphthylene for seven days survived (Gerarde, 1960). 7.2.1.3 Anthracene Nine of 10 mice given 500 mg/kg bw per day anthracene for seven days survived (Gerarde, 1960). Oral administration of 100 mg/kg bw per day to rats for four days increased carboxylesterase activity in the intestinal mucosa by 13% (Nousiainen et al., 1984). 7.2.1.4 Benzo [a]pyrene Death due to myelotoxicity was observed after daily oral administration of benzo [a]pyrene at 120 mg/kg bw to poor-affinity Ah receptor DBA/2N mice for one to four weeks, whereas high-affinity C57 Bl/6N mice survived with no myelotoxicity for at least six months under these conditions (Legraverend et al., 1983). Rats given 50 or 150 mg/kg bw per day of benzo [a]pyrene orally for four days showed suppressed carboxylesterase activity in the intestinal mucosa. The NOAEL with respect to gastric, hepatic, and renal effects was 150 mg/kg bw per day (Nousiainen et al., 1984) Table 76. Toxicity of single doses of naphthalene Species Sex Route of Dose (purity) Effects Reference (strain) (no./sex administration per group) Dog Oral 1000-2000,4000 1000-2000: Light diarrhoea; 4000 mg: Flury & Zernick or 5000 mg/dog lethal; 5000 my heavy diarrhoea (1935) Cat Oral 1000-3000 Lethal Flury & Zernick mg/kg bw (1935) Rabbit Oral 1000-3000 and 1000-3000 mg: corneal clouding; Flury & Zernick 3000 mg/kg bw 3000 mg death after 24 h (1935) Dog (1) Oral 400 and 1800 400 mg: weakness, severe anaemia; Zuelzer & Apt mg/kg bw 1800 mg: weakness, vomiting, diarrhoea, (1949) slight anaemia; complete recovery within 1-2 weeks Mouse Inhalation 0.1 mg/litre, 4 h Bronchiolar necrosis Buckpitt & Franklin (1989) Mouse M Intraperitoneal 50,100,200, Dose-dependent bronchiolar epithelial-cell O'Brien et al. (Swiss-Webster 300 mg/kg bw necrosis (1989) Mouse M (4-35) Intraperitoneal 50,100,200, Dose-dependent bronchiolar necrosis; Plopper et al. (Swiss-Webster) 300, and 400 300 mg/kg: swollen cells in trachea (1992) mg/kg bw 400 mg/kg: cytotoxicity in olfactory (> 99.9%) epithelium Rat M (4-11) Intraperitoneal 200,400,800, Bronchiolar necrosis not observed; no Plopper et al. (Sprague-Dawley) and 1600 mg/kg changes in trachea; 200 mg/kg: complete (1992) bw (> 99.9%) necrosis of olfactory epithelium Table 76 (continued) Species Sex Route of Dose (purity) Effects Reference (strain) (no./sex administration per group) Rat M Intraperitoneal 400-1600 mg/kg No damage to lungs, liver, or kidneys O'Brien et al. (Wistar) bw (1985) Hamster M (4-6) Intraperitoneal 100,200,400 800 mg/kg: minor alterations in terminal Plopper et al. (Syrian and 800 mg/kg bronchioles; cytotoxic changes in trachea; (1992) golden) bw (99.9%) 400 mg/kg: necrosis of olfactory epithelium Rabbit M Intraperitoneal 0.3-1.7 mg/kg bw No haemolytic effects Mackell et al. (white) (1951) M, male In Fischer 344/Crl rats exposed by inhalation to 7.7 mg/m3 of benzo [a]pyrene dust for 2 h/day, five days per week for four weeks, no respiratory tract lesions were observed, as measured by lung lavage, clearance of tagged particles, and histopathological findings (Wolff, R.K. et al., 1989). 7.2.1.5 Benz [a]anthracene When benz [a]anthracene was given orally to rats daily for four days, the NOAEL with respect to gastric, hepatic, and renal effects was 150 mg/kg bw per day. Carboxylesterase activity in the intestinal mucosa was suppressed (Nousiainen et al., 1984). 7.2.1.6 Dibenz [a,h]anthracene Adverse haemolymphatic changes, including the appearance of extravascular erythrocytes in the lymph spaces and large pigmented cells, were reported after subcutaneous injection of male rats with 0.28 mg per animal on five days per week for four weeks (Lasnitzki & Woodhouse, 1944). 7.2.1.7 Fluoranthene All of 10 mice that received 500 mg/kg bw per day fluoranthene intraperitoneally for seven days survived (Gerarde, 1960). 7.2.1.8 Naphthalene Anaemia was induced in three dogs by single oral doses of 3 or 9 g or a total dose of 10.5 g per animal given over seven days. All three animals showed neurophysiological symptoms and slight to very severe changes in haematological parameters. Full recovery was observed within 7-14 days (Zuelzer & Apt, 1949). No immunosuppressive effects were observed in a number of test systems. Tolerance to the effects of naphthalene was reported in mice after intraperitoneal injection for seven days. A sharp contrast between single and multiple doses was observed in the effects on the morphology of the bronchiolar epithelium. When naphthalene was given intraperitoneally at a dose of 50, 100, or 200 mg/kg bw per day as a single injection, dose-dependent bronchiolar epithelial cell necrosis was detected; however, when these doses were given daily for seven days, no significant effects were observed. Addition of 300 mg/kg bw on day 8 had no effect, whereas recovered sensitivity was observed with increasing time between the last dose and the challenge dose. A single dose of 300 mg/kg bw without pretreatment resulted in substantial denudation of the bronchiolar epithelium. This pattern was attributed to a reduction in metabolic activation of naphthalene due to a decrease in cytochrome P450 mono-oxygenase activity after multiple dosing. A rough correlation was observed in mouse lung (but not liver microsomes) between induction of tolerance and decreased metabolic formation of the 1 R, 2 S-epoxide enantiomer, which is responsible for tissue-selective toxicity. Such toxicity was demonstrated in mice both in vivo and in isolated perfused lung (Buckpitt & Franklin, 1989). These studies are summarized in Table 77. 7.2.1.9 Phenanthrene Oral administration of 100 mg/kg bw per day phenanthrene to rats for four days induced a 30% increase in carboxylesterase activity in the intestinal mucosa (Nousiainen et al., 1984). 7.2.1.10 Pyrene Four of five mice injected intraperitoneally with 500 mg/kg bw per day pyrene for seven days survived (Gerarde, 1960). 7.2.2 Subchronic toxicity 7.2.2.1 Acenaphthene Administration of 175 mg/kg bw per day acenaphthene to mice by gavage for 90 days resulted in a NOAEL of 175 mg/kg bw per day and a LOAEL of 350 mg/kg bw per day for hepatotoxicity (US Environmental Protection Agency, 1989a). 7.2.2.2 Anthracene Four of five rats given 5 mg per animal anthracene subcutaneously for four months survived (Gerarde, 1960). Anthracene was administered to groups of 20 male and female CD-1 (ICR) BR mice by gavage at a dose of 0, 250, 500, or 1000 mg/kg bw per day for at least 90 days. No treatment-related effects were noted on mortality, clinical signs, body weights, food consumption, ophthalmological findings, the results of haematology and clinical chemistry, organ weights, organ-to-body weight ratios, and gross pathological and histopathological findings. The no-observed-effect level (NOEL) was the highest dose tested, 1000 mg/kg bw per day (US Environmental Protection Agency, 1989b). 7.2.2.3 Benzo [a]pyrene Male Syrian golden hamsters were exposed by inhalation to 9.8 or 44.8 mg/m3 benzo [a]pyrene for 4.5 h/day, five days per week for 16 weeks. No neoplastic response was observed in the respiratory tract (Thyssen et al., 1980). The growth of rats was inhibited by feeding a diet enriched with benzo [a]pyrene at 1.1 g/kg for more than 100 days (White & White, 1939). Table 77. Subacute and subchronic effects of naphthalene Species Sex Route of Dose (purity) Effects Reference (strain) (no./sex administration per group) Mouse M,F Oral 27, 53, and 267 In all groups, slight alterations in haemato Shopp et al. (CD-1) (40-112) mg/kg bw, 7 d/ logical parameters; humoral immune response (1984) week, 14 d not affected. 27 and 53 mg/kg: no significant effects; 267 mg/kg: 5-10% mortality (m/f); significantly decreased terminal body weight (m/f); 30% decrease in thymus weight (m); significant decrease in weight of spleen (f); increase in lung weight (f) Mouse M,F Oral 5.3, 53, and 133 No obvious pulmonary effects or Shopp et al. (CDO) mg/kg bw, 7 d/ immunotoxicity; significantly decreased (1984) week, 90 d relative spleen weights (f); tolerance Mouse M Intraperitoneal 50, 100, and 200 No significant alterations in lung morphology; Buckpitt & Franklin (Swiss-Webster) mg/kg, 7 d tolerance to 300 mg/kg on day 8 (1989); O'Brien et al. (1989) Rat Diet 2 g/kg diet, Inhibition of growth; enlarged, fatty livers White & White 100 d (1939) Dog (1) Oral 122 g/kg bw per Diarrhoea, weakness, lack of appetite, ataxia, Zuelzer & Apt day, 7 d very severe anaemia; complete recovery (1949) within 1-2 weeks M, male; F, female 7.2.2.4 Fluorene Groups of 25 male and 25 female CD-1 mice were given 0, 125, 250, or 500 mg/kg bw per day fluorene suspended in corn oil by gavage for 13 weeks. Increased salivation, hypoactivity, and abdomens wetted with urine were observed in all treated males. The percentage of hypoactive mice was dose-related. In mice exposed at 500 mg/kg bw per day, laboured respiration, ptosis (drooping eyelids), and an unkempt appearance were also observed. A significant decrease in erythrocyte count and packed cell volume were observed in females treated with 250 mg/kg bw per day fluorene and in males and females treated with 500 mg/kg bw per day. The latter also showed a decreased haemoglobin concentration and an increased total serum bilirubin level. A dose-related increase in relative liver weight was observed in treated mice, and a significant increase in absolute liver weight was observed in the mice treated with 250 or 500 mg/kg bw per day. Significant increases in absolute and relative spleen and kidney weights were observed in males and females exposed to 500 mg/kg bw per day and in males at 250 mg/kg bw per day. The increases in absolute and relative liver and spleen weights in animals at the high dose were accompanied by increases in the amounts of haemosiderin in the spleen and in Kupffer cells of the liver. No other histopathological lesions were observed. The LOAEL for haematological effects was 250 mg/kg bw per day, and the NOAEL was 125 mg/kg bw per day (US Environmental Protection Agency, 1989c). In a similar study, fluorene at 35, 50, and 150 mg/kg bw increased the weight of the liver by about 20% in a dose-dependent fashion and the mitotic index of hepatocytes by sixfold after 48 h (Danz et al., 1991). 7.2.2.5 Fluoranthene Groups of 20 male and 20 female CD-1 mice were given 0, 125, 250, or 500 mg/kg bw per day fluoranthene by gavage for 13 weeks. A fifth group of 30 male and 30 female mice was used to establish baseline levels in blood. Body weight, food consumption, and haematological and serum parameters were recorded regularly throughout the experiment. At the end of 13 weeks, the animals were killed and autopsied; organs were weighed and a histological evaluation was made. All treated mice had dose-dependent nephropathy, increased salivation, and increased liver enzyme activities, but these effects were either not significant, not dose-related, or not considered adverse at 125 mg/kg bw per day. Mice exposed to 500 mg/kg bw per day had increased food consumption and increased body weight. Mice exposed to the two higher doses had statistically increased alanine aminotransferase activity and increased absolute and relative liver weights. Treatment-related microscopic liver lesions (indicated by pigmentation) were observed in 65% of mice at 250 mg/kg bw per day and 88% of those at the highest dose. On the basis of the increased alanine aminotransferase activity, pathological effects in the kidney and liver, and clinical and haematological changes, the LOAEL was 250 mg/kg bw per day and the NOAEL 125 mg/kg bw per day (US Environmental Protection Agency, 1988). 7.2.2.6 Naphthalene In a 90-day study in mice, naphthalene at oral doses up to 133 mg/kg bw caused neither mortality nor serious changes in organ weights (Shopp et al., 1984). These authors did not observe haemolytic anaemia in CD-1 mice after oral uptake, although this effect had been seen in human patients (Konar et al., 1939; Zuelzer & Apt, 1949; see Section 8). It was suggested that glucose-6-phosphate dehydrogenase deficiency in erythrocytes, a prerequisite of haemolytic anaemia in adult humans, was not present in the mice (Shopp et al., 1984). In rats that ingested 150 mg/kg bw per day naphthalene for the first three weeks and 200-220 mg/kg bw per day for a further 11 weeks, reduced weight gain and food intake were observed. Later, the liver was found to be enlarged, with cell oedema and congestion of the liver parenchyma, and the kidneys showed signs of inflammation (Kawai, 1979). The presence of 1 g/kg naphthalene in the feed of rats and rabbits for 46-60 days led to cataracts (US Environmental Protection Agency, 1984b; see also section 7.8). Administration to rabbits of 0.1-1 mg/kg bw per day naphthalene by subcutaneous injection for 123 days resulted in severe oedema and a high degree of vacuolar and collicular degeneration in the brain; necrosis of nerve cells also occurred. The author suggested that hypoxaemia resulting from haemolytic anaemia was responsible for this finding (Suja, 1967; cited by Kawai, 1979). Subacute and subchronic studies with naphthalene are summarized in Table 77. 7.2.2.7 Pyrene The growth of rats was inhibited by feeding a diet enriched with benzo [a]pyrene at 2 g/kg for more than 100 days. The livers were enlarged and had a fatty appearance indicating hepatic injury (White & White, 1939). Groups of 20 male and 20 female CD-1 mice were given 0, 75, 125, or 250 mg/kg bw per day pyrene in corn oil by gavage for 13 weeks and then examined for changes in body weight, food consumption, mortality, clinical pathological manifestations in major organs and tissues, and changes in haematology and serum chemistry. Nephropathy, characterized by the presence of multiple foci of renal tubular regeneration, often accompanied by interstitial lymphocytic infiltrates and/or foci of interstitial fibrosis, was present in four male control mice, one at the low dose, one at the medium dose, and nine the high dose. Similar lesions were seen in two, three, seven, and 10 female mice, respectively. The renal lesions in all groups were described as minimal or mild. Relative and absolute kidney weights were reduced in mice at the two higher doses. On the basis of nephropathy and decreased kidney weights, the low dose (75 mg/kg bw per day) was considered to be the NOAEL and 125 mg/kg bw per day the LOAEL (US Environmental Protection Agency, 1989d). 7.3 Long-term toxicity Almost all of the long-term studies reported were designed to assess the carcinogenic potency of PAH and are therefore summarized in section 7.7. Information about the non-carcinogenic effects, such as growth inhibition, liver damage, and irritation, which occurred at concentrations that also caused carcinogenic effects is presented here. General effects, such as on mortality, body weight, and pathological findings at sacrifice, were not considered useful. 7.3.1 Anthracene A group of 28 BD I and BD III rats received anthracene in the diet from the age of about 100 days, at a daily dose of 5-15 mg per rat. The experiment was terminated when a total dose of 4.5 g per rat had been achieved, on day 550. The rats were observed until they died; some lived for more than 1000 days. No treatment-related effects on lifespan or on gross or histological appearance of tissues were observed; haematological parameters were not measured (Schmähl, 1955). After weekly subcutaneous injections of anthracene at 0.25 mg per animal for 40 weeks, mice showed deposition of iron in lymph glands and a reduced number of lymphoid cells (Hoch-Ligeti, 1941). 7.3.2 Benz [a]anthracene Weekly subcutaneous injection of 0.25 mg per mouse for 40 weeks resulted in deposition of iron in lymph glands and a reduced number of lymphoid cells (Hoch-Ligeti, 1941). 7.3.3 Dibenz [a,h]anthracene Mice given weekly subcutaneous injections of 0.25 mg per animal for 40 weeks had pale, soft, enlarged livers with signs of fatty degeneration. There was deposition of iron in lymph glands, and the number of lymphoid cells was reduced (Hoch-Ligeti, 1941). 7.4 Dermal and ocular irritation and dermal sensitization The adverse dermatological effects observed in animals after acute and subchronic dermal exposure to PAH included destruction of sebaceous glands, dermal ulceration, hyperplasia, hyperkeratosis, and alterations in epidermal cell growth. Perylene, benzo [e]pyrene, phenanthrene, pyrene, anthracene, naphthalene, acenaphthalene, fluorene, and fluoranthene did not suppress the sebaceous gland index; benz [a]anthracene, dibenz [a,h]anthracene, and benzo [a]pyrene resulted in indices > 1 (Bock & Mund, 1958). In Swiss mice treated daily for three days with solutions of benzo [a]pyrene in acetone, a concentration of 0.1% destroyed less than half of the sebaceous glands, whereas 0.2% destroyed more than 50% (Suntzeff et al., 1955). 7.4.1 Anthracene Anthracene is a primary irritant, and its fumes can cause mild irritation of the skin, eyes, mucous membranes, and respiratory tract. At a concentration of 4.7 mg/m3, mild skin irritation was found in 50% of exposed mice (Montizaan et al., 1989). The median value for dermal irritant activity (ID50) in the mouse ear was 6.6 × 10-4 mmol or 118 µg/ear; in comparison, the ID50 for benzo [a]pyrene was 5.6 × 10-5 mmol per ear (Brune et al., 1978). Anthracene increases the sensitivity of skin to solar radiation (Gerarde, 1960). No contact sensitivity to anthracene was observed (Old et al., 1963). 7.4.2 Benzo [a]pyrene Four adult female guinea-pigs were injected with a total of 250 µg benzo [a]pyrene in Freund's adjuvant, and two to three weeks later were tested for contact sensitivity with solutions of 0.001, 0.01, 0.1, or 1% benzo [a]pyrene in acetone and olive oil. After 24 h, a slight to severe (0.001-1%) contact hypersensitivity was observed (Old et al., 1963). C3H mice were given an epicutaneous administration of 100 µg benzo [a]pyrene in 0.1% acetone solution into the abdominal skin. Five days later, contact hypersensitivity was elicited by applying 20 µg benzo [a]pyrene to the dorsal aspect of the ear. The response was quantified by ear thickness, which reaced a maximum three to five days after challenge. The LOAEL for allergic contact sensitivity was thus 120 µg (Klemme et al., 1987). The ID50 value for dermal irritant activity in the mouse ear was 5.6 × 10-5 mmol per ear (Brune et al., 1978). 7.4.3 Naphthalene A single dose of 100 mg naphthalene to the rabbit eye was slightly irritating, whereas application of 495 mg to rabbit skin, without occlusion, caused mild irritation (Sax & Lewis, 1984). 7.4.4 Phenanthrene No contact sensitization to phenanthrene was observed (Old et al., 1963). 7.5 Reproductive effects, embryotoxicity, and teratogenicity The mechanistic aspects of reproductive and embryotoxic effects are presented in detail and the results summarized in Tables 78-80. The genotype of mice is decisive for the manifestation of effects. Studies have been reported on anthracene, benz [a]anthracene, benzo [a]-pyrene, chrysene, dibenz [a,h]anthracene, and naphthalene. Embryotoxicity was reported in response to benz [a]anthracene, benzo [a]pyrene, dibenz [a,h]-anthracene, and naphthalene. Benzo [a]pyrene also had adverse effects on female fertility, reproduction, and postnatal development. In a study in young mice, an NOEL of 150 mg/kg bw per day was obtained for benzo [a]pyrene on the basis of effects on fertility (sperm in lumen of testes, size of litters) and embryotoxicity (malformations) (Rigdon & Neal, 1965). 7.5.1 Benzo [a]pyrene 7.5.1.1 Teratogenicity in mice of different genotypes Benzo [a]pyrene is embryotoxic to mice, and the effect is partly dependent on the genetically determined induction of the cytochrome P450 mono-oxygenase receptor, Ah, of the mother and fetus by PAH (see also section 7.10). In the case of an inducible mother (Ahb/ Ahb and Ahb/ Ahd, B groups), the genotype of the fetus is not crucial because the active metabolites formed in the mother appear to cross the placenta, causing fetal death or malformation. In contrast, when the mother is non-inducible (Ahd/ Ahd, D group), the genotype of the fetus is important; one litter may contain both inducible and non-inducible fetuses. Another decisive factor is the route by which benzo [a]pyrene is given to the mother. Three studies of the genetic expression of effects are summarized below. Intraperitoneal injection of benzo [a]pyrene at 50 or 300 mg/kg bw on day 7 or 10 of gestation was more toxic and teratogenic in utero in genetically inducible C57Bl/6 (Ahb/ Ahb) than in non-inducible AKR inbred mice (Ahd/ Ahd). In AKR × (C57Bl/6)(AKR)F1 and (C57Bl/6)(AKR)F1 × AKR back-crosses (father × F1 mother), allelic differences at the Ah locus in the fetus correlated with dysmorphogenesis. The inducible fetal Ahb/ Ahd genotype results in more stillborn and resorbed fetuses,decreased fetal weight, increased frequency of congenital anomalies, and enhanced P1-450-mediated covalent binding of benzo [a]pyrene metabolites to fetal protein and DNA, when compared with fetuses of the non-inducible Ahd/ Ahd genotype (not-inducible) from the same uterus (see Table 78). In the case of an inducible mother (Ahb/ Ahd), however, these parameters do not differ in Ahb/ Ahd and Ahd/ Ahd individuals in the same uterus, presumably because the increased benzo [a]pyrene metabolism in maternal tissues and placenta cancels them out (Shum et al., 1979). An inducible genotype is not the only factor involved in the reproductive toxicity of benzo [a]pyrene. In a study in which C57Bl/6 female mice (Ah inducible) were mated with C57Bl/6, DBA/2, or BDF1 male mice (B groups), and DBA/2 females (non-inducible) were mated with C57Bl/6, DBA/2, or BDF1 males (D groups) and received intraperitoneal injections of benzo [a]-pyrene, fetal mortality increased dose-dependently in all groups except the DBA/2 × DBA/2. Fetal body weight was reduced dose-dependently in all experimental groups, but the effect was more pronounced in D than B groups, as was a dose-dependent increase in the frequency of cervical ribs (for experimental details, see Table 78). These results suggest that Table 78. Embryotoxicity of polycyclic aromatic hydrocarbons in experimental animals Species No. per Route of Duration, dose Effects Reference (strain) group administration Anthracene Rat Gavage Day 19 of gestation, F1: no induction of BaP hydroxylase in liver Welch et al. Sprague- 60 mg/kg compared with control (< 0.2 vs <0.2 units in (1972) bw controls) Benz[a]anthracene Rat 2 Subcutaneous Day 1-11 or 1-15 F0: Day 10 and 12: severe vaginal haemorrhage; Wolfe & of gestation, 5 mg/ Day 14: intraplacental haemorrhage Bryan (1939) animal per day F1: fetal death and resorption up to day 18 Rat Gavage Day 19 of gestation, F1: induction of BaP hydroxylase in liver Welch et al. Sprague-Dawley 60 mg/kg bw (12 vs < 0.2 units in controls) (1972) Benzo[a]pyrene Mouse 9 Diet Day 5 or 10 of F1: no malformations Rigdon & White gestation until Neal (1965) Swiss delivery, 50 mg/ky bw Mouse 6-17 Diet Day 2-10 of F1: increased intrauterine toxicity and Legraverend C57BI/6N, gestation, 120 mg/ malformations in Ahd/Ah7dembryos compared et al. (1984) AKR/J, and kg per day with Ahb/Ahd embryos in pregnant Ahd/Ahd back-crosses mice (effect not seen in pregnant Ahb/Ahd mice) (reciprocal) Mouse 5-30 Intraperitoneal Day 7, 10, or 12 of 200 mg/kg bw: F1: increase in stillbirths, Shum et al. C57BI/6, gestation, 50-300 resorptions, malformations (4-fold higher (1979) AKR and mg/kg bw in pregnant C57BI than in AKR mice) back-crosses (reciprocal) Table 78. (continued) Species No. per Route of Duration, dose Effects Reference (strain) group administration Mouse 20 Intraperitoneal Day 8 of gestation, 150 and 300 mg/kg bw: F0: increased fetal Hoshino et al. C57BI/6, 150 or 300 mg/kg mortality (except DBA/2 × DBA/2 offspring); (1981) DBA/2, and reduced fetal body weight; increased number of back-crosses cervical ribs (reciprocal) 300 mg/kg: F1: increased malformations (C57BI/6 × C57BI/6) Mouse Gavage Day 7-16 of F0: no toxicity MacKenzie & CIT1 gestation, 10, 40, 160 F1: no toxicity Angevine (1981) mg/kg bw per day Rat 17 Subcutaneous Day 1-11 or 16 of F0: Days 10 and 12: profuse vaginal Wolfe & Bryan gestation, 5 mg/ haemorrhage; day 14: intraplacental (1939) animal per day haemorrhage; F1: fetal death and resorption up to day 18 Rat Gavage Day 19 of gestation, F1: induction of BaP-hydroxylase in liver Welch al al. Sprague-Dawley 60 mg/kg bw (20 vs < 0.2 units in controls) (1972) Rat 10-15 Subcutaneous Day 6-8 or 6-11 of F1: significant increase in number of resorptions Bui et al. Sprague-Dawley gestation, 50 mg/kg and fetal wastage (dead fetuses plus resorption); (1986) bw per day fetal weight reduced Chrysene Rat Gavage Day 19 of gestation, F1: induction of BaP hydroxylase in liver Welch et al. Sprague-Dawley 60 mg/kg bw (6 vs < 0.2 units in controls) (1972) Dibenzo[a,h]anthracene Rat Gavage Day 19 of gestation, F1: induction of BaP hydroxylase in liver Welch et al. Sprague-Dawley 60 mg/kg bw (15 vs < 0.2 units in controls) (1972) Table 78. (continued) Species No. per Route of Duration, dose Effects Reference (strain) group administration Rat 38 Subcutaneous Day 1-8 or 1-18 of F0: Days 10 and 12: profuse vaginal haemorrhage; Wolfe & gestation, 5 mg/ day 14: intraplacental haemorrhage Bryan (1939) animal per day F1: fetal death and resorption up to day 18 Naphthalene Mouse 50 Gavage Day 7-14 of F0: significant 15% increase in mortality; Plasterer et CD-1 gestation, 300 mg/ significant reduction in weight gain al. (1985) kg bw per day F1: significant reduction in number of live offspring; no malformations Mouse Gavage Day 6-13 of F0 increased mortality 10/50; control: 0/50); Hardin et al. CD-1 gestation, 300 mg/ significant reduction in weight gain (1987) kg bw per day F1: significant reduction in liveborns per litter Rat 10-15 Intraperitoneal Day 1-15 of F0: no toxicity Hardin et al. Sprague-Dawley gestation, 395 mg/ F1: no toxicity (1981) kg per day For genotypes of the mouse strains used see section 7.5.1.1 Ah-inducible fetuses are more sensitive to lethal events, whereas those of non-inducible dams are more susceptible to a decrease in body weight and an increased incidence of cervical ribs. The incidence of external malformations may, however, differ in mice of different genotypes after treatment with benzo [a]-pyrene, even if both dams and fetuses are inducible (Hoshino et al., 1981). The toxicity of benzo [a]pyrene in utero was investigated in pregnant Ahd/ Ahd × Ahb/ Ahd F1 and Ahb/ Ahd × Ahd/ Ahd F1 back-crossed mice fed benzo [a]pyrene in the diet at 120 mg/kg daily on days 2-10 of gestation. Embryos of D females (Ahd/ Ahd genotype; non-inducible) showed more signs of toxicity and malforma-tions than Ahd/ Ahd embryos. Fetuses of B females (Ahb/ Ahd genotype) did not show these changes. The authors suggested that reduced benzo [a]pyrene metabolism in the intestine had caused high concentrations in the embryos, and more toxic metabolites (benzo [a]pyrene-1,6- and -3,6-quinones) were detected in the Ahd/ Ahd embryos than in Ahb/ Ahd embryos (Legraverend et al., 1984). These results were in contrast to those reported after intraperitoneal injection by Shum et al. (1979) and Hoshino et al. (1981). The route of administration can thus affect the magnitude of the observed effects (see also section 7.8.2.2). 7.5.1.2 Reproductive toxicity A single intraperitoneal injection of benzo [a]pyrene reduced fertility and destroyed primordial oocytes of DBA/2N mice in a dose-dependent manner (Mattison et al., 1980; see also Table 79). In experiments with B6 (Ah-inducible) and D2 (non-inducible) mice, primordial oocytes of B6 mice underwent more rapid destruction after treatment with benzo [a]pyrene than those of D2 mice. This effect corresponded to a two- to threefold increase in ovarian arylhydrocarbon hydroxylase (AHH) activity in B6 mice after treatment. This correlation was not found in analogous experiments with D2B6F1 mice, in which AHH activity was increased by two- to threefold, but the oocyte destruction was similar to that observed in D2 mice. This demonstrates an inconsistent consequence of strain differences in genotype (Mattison & Nightingale, 1980; see also Table 79). The sum of activation, detoxification, and repair seems to be decisive for the process of oocyte destruction (Figure 8). Benzo [a]pyrene and its three metabolites, benzo [a]pyrene 7,8-oxide, benzo [a]pyrene 7,8-diol, and benzo [a]pyrene diol epoxide, were administered by injection at a single dose of 10 µg into the right ovary of B6, D2, and D2B6F1 mice. Ovarian volume, weight, and follicle numbers were measured after two weeks; various reductions were observed in all strains. There was also compesatory hypertrophy of the left ovary (Mattison et al., 1989; see also Table 79). Table 79. Effects of benzo[a]pyrene on fertility in experimental animals Species Sex/No. Route of Duration, dose Effects Reference (strain) per administration group Mouse M Diet Up to 30 days before mating, NOEL: 150 mg/kg bw per day Rigdon & Neal (1965) White 5 37.5, 75, or 150 mg/kg bw Parameters: sperm in lumen per day of testes; number of offspring Mouse F Diet 20 days before mating NOEL: 150 mg/kg bw per day Rigdon & Neal (1965) White 5-65 37.5, 75, or 150 mg/kg bw Parameter: number of offspring Swiss per day Mouse F Intraperitoneal Day 14 before mating, 10, 100 mg/kg bw: dose-dependent Mattison et al. (1980) DBA/2N 15 10, 100, 200, or 500 mg/kg decrease in number of pups bw once 200, 500 mg/kg bw: completely infertile; threshold: 3.4 mg/kg bw; 50% effect dose: 25.5 mg/kg bw Mouse F Intraperitoneal Day 21 before sacrifice, Dose-dependent increase in Mattison et al. (1980) DBX2N 5, 10, 50, 100, or 500 mg/kg primordial oocyte destruction; bw once 500 mg/kg: 100% destruction; threshold: 2.7 mg/kg bw; 50% effect dose: 24.5 mg/kg bw Mouse F Intraperitoneal Day 13 before sacrifice, 100 mg/kg bw: significant increase Mattison & Nightingale B6 and D2 5 100 mg/kg bw once in primordial oocyte destruction in (1980) both genotypes; effects in B6 mice greater than in D2 mice Mouse F Intra-ovarian Day 14 before sacrifice, 10 µg: decreased ovarian weight Mattison et al. (1989) C57BI/6N (136), injection 10 µg/right ovary once (D2); decreased ovarian volume (D2 DBA/2N (D2), and F1); decreased antral follicles D2B6F1(F1) (F1) decreased number of small follicles (D2 and F1) Table 79. (continued) Species Sex/No. Route of Duration, dose Effects Reference (strain) per administration group Mouse F Intraperitoneal 1, 2, 3, and 4 weeks 500 mg/kg: 35% mortality Swartz & Mattison, C57BI/6N 5 before sacrifice; 1, 5, 1-500 mg/kg bw: dose- and time- 1985); 10, 50, 100, or 500 dependent decrease in ovarian Miller et al. (1992) mg/kg bw volume, total volume and number of corpora lutea/ovary (for last parameter, after 1 week threshold was about 1 mg/kq bw and ED50 1.6 mg/kg bw);effect transitory in low-dose groups, butnot reversible in two highest by four weeks For genotypes of the mouse strains used see section 7.5.1.1 7.5.1.3 Effects on postnatal development Three studies of the postnatal effects of benzo [a]pyrene on mouse offspring, with administration dermally, intraperitoneally, or orally, showed adverse effects, including an increased incidence of tumours, immunological suppression, and reduced fertility (see also Table 80). 7.5.1.4 Immunological effects on pregnant rats and mice Benzo [a]pyrene given to pregnant rats on day 15 or 19 of gestation caused alterations at the thymic glucocorticoid receptors in the offspring, suggesting binding to the pre-encoded hormone receptors and interference with receptor maturation (Csaba et al., 1991; Csaba & Inczefi-Gonda, 1992; see also section 7.8.2.6). Strong suppression of immunological parameters was found in the progeny of mice that had been treated intraperitoneally with benzo [a]pyrene at mid-gestation (Urso & Johnson, 1987; see also section 7.8.2.6). 7.5.2 Naphthalene 7.5.2.1 Embryotoxicity Naphthalene was administered by gavage at 50, 150, or 450 mg/kg bw per day to pregnant Sprague-Dawley rats on days 6-15 of gestation, i.e. during the main period of organogenesis. The dams showed signs of toxicity including lethargy, slow breathing, prone body posture, and rooting, and these effects persisted after the end of dosing with the high dose. The body-weight gain of treated animals was reduced by 31 and 53% in the groups at the two higher doses. Naphthalene did not induce fetotoxic or teratogenic effects, and the numbers of corpora lutea per dam, implantation sites per litter, and live fetuses per litter were within the range in controls. The maternal NOAEL was < 50 mg/kg bw per day (National Toxicology Program, 1991). In a second study, doses of 0, 20, 80, or 120 mg/kg bw per day were given to rabbits by gavage during days 6-19 of gestation. There were no signs of maternal toxicity, fetotoxicity, or developmental toxicity (National Toxicology Program, 1992a). 7.5.2.2 Toxicity in cultured embryos Mice injected intraperitoneally on day 2 of gestation with 14 or 56 mg/kg bw naphthalene were sacrificed 36 h later, and embryos were cultured in vitro. Maternal doses below the LD50 value inhibited the viability and implantation capacity of the embryos, and attachment and embryonic growth in vitro were markedly decreased (Iyer et al., 1990). Table 80. Effects of benzo[a]pyrene on postnatal development in experimental animals Species Sex/No. Route of Duration, dose Effects Reference (strain) per administration group Mouse F Dermal Entire gestation period F1- F4: sensitization of offspring: increased Andrianova non-inbred 1 drop of 0.5% solution, incidence of papillomas and carcinomas (1971) twice per week; F1-F4 in all generations compared with animals treated with BaP, m not treated in utero 1x/week, f 2x/week Mouse F Intraperitoneal Day 11-13 or 16-18 F1: no difference in birth rate, litter size of Urso & C3H/Anf 25 of gestation, 100 or progeny compared to controls; severe suppression Gengozian 150 mg/kg of anti-SRBC PFC response up to 78 weeks of life (1980) (see also section 7.8.2.6); 11-1 fold increase in tumour incidence (liver, lung, ovaries) after 56-78 weeks Mouse F Gavage Days 7-16 of F1: 10 mg/kg markedly impaired fertility (by 20%) MacKenzie & CD-1 gestation, 10, 40, and reduced testis weight (by 40%), 34% sterility Angevine 160 mg/kg bw per day of females; 40 and 160 mg/kg: fertility impaired (1981) by > 900%/100%; testis weight reduced by > 800%; 100%/100% sterility of females anti-SRBC PFC, anti-sheep red blood cell antibody (plaque)-forming cells In a subsequent study, three-day-old whole mouse embryos were collected at the blastocyst stage, cultured in NCTC 109 medium, and exposed to naphthalene at 0.16, 0.2, 0.39, or 0.78 mmol/litre for 1 h with and without S9. They were then transferred to toxicant-free medium, cultured for 72 h, and evaluated microscopically. Naphthalene was not directly embryotoxic, but growth and viability were decreased in the presence of S9, with 100% embryolethality at doses > 0.2 mmol/litre; furthermore, hatching and attachment rates were significantly decreased. The approximate LC50 in S9-supplemented media was 0.18 mmol/litre (Iyer et al., 1991). 7.6 Mutagenicity and related end-points Benzo [a]pyrene has been used extensively as a positive control in a variety of short-term tests. It is active in assays for the following end-points: bacterial DNA repair, bacteriophage induction, and bacterial mutation; mutation in Drosophila melanogaster; DNA binding, DNA repair, sister chromatid exchange, chromosomal aberration, point mutation, and transformation in mammalian cells in culture; and tests in mammals in vivo, including DNA binding, sister chromatid exchange, chromosomal aberration, sperm abnormalities, and somatic mutation at specific loci (Hollstein et al., 1979; De Serres & Ashby, 1981). Positive effects were seen in most assays for the mutagenicity of benzo [a]pyrene. A selection of these studies is summarized in Tables 81-88. All of the data available on the other PAH considered in this monograph were taken into account. Because of the amount of data, the purities of the chemicals tested and details of the assay conditions are omitted from the tables, but they do show the results obtained when S9 was used. Variations in the S9 metabolic activation component of the assay system, e.g. the age, sex, and strain of the rats used as a source of liver and any pretreatment with enzyme inducers such as Aroclor, 3-methylcholanthrene, or phenobarbital, may markedly affect the results and may account for apparent discrepancies. DNA binding of benzo [a]pyrene was observed in various species. For example, adducts were found in cells from hamsters, mice (Arce et al., 1987), rats (Moore et al., 1982), and chickens (Liotti et al., 1988), in calf thymus DNA (Cavalieri et al., 1988a), and in human cell systems (Moore et al., 1982; Harris et al., 1984). Formation of DNA adducts was inhibited in the presence of scavengers of active oxygen species like superoxide dismutase, catalase, and citrate-chelated ferric iron, indicating that reactive oxygen species such as superoxide, OH radicals, and singlet oxygen may be involved in DNA binding (Bryla & Weyand, 1991). Benzo [a]pyrene at a total dose of 10 mg/kg bw induced gene mutations in mice, as seen in the coat-colour spot test (Davidson & Dawson, 1976). The results of tests for reverse mutation in Salmonella typhimurimum (Ames test) and for forward mutation in S. typhimurimum strain TM677 are presented in Table 81. Bacterial tests for DNA damage in vitro are shown in Table 82. The results of tests for mutagenicity in yeasts and Drosophila melanogaster, including host-mediated assays, are shown in Table 83. The results of various assays carried out on mammalian cells in vitro are summarized in Tables 82-86. The results of tests in vivo are shown in Tables 87 and 88. The activity of PAH in short-term tests is summarized in Table 89, which gives the evaluations of IARC (1983; see also Section 12) and the results of studies reported after 1983. Only three of the 33 PAH considered, i.e. anthracene, fluorene, and naphthalene were inactive in all short-term tests; 16 had mutagenic effects. Eight PAH showed a tendency for mutagenic activity, but the data are still too sparse to permit a final judgement. The available information on acenaphthene, acenaphthylene, benzo [a]fluorene, and coronene is still inadequate. As phenanthrene and pyrene showed inconsistent results in various experiments, they could not be clearly classified as mutagenic. 7.7 Carcinogenicity Most of the studies that have been conducted on PAH were designed to assess their carcinogenicity. Studies on various environmentally relevant matrices such as coal combustion effluents, vehicle exhaust, used motor lubricating oil, and sidestream tobacco smoke showed that PAH are the agents predominantly responsible for their carcinogenic potential (Grimmer et al., 1991b). Because of the abundance of literature, only studies involving the administration of single PAH have been taken into account in this monograph. Benzo [a]pyrene has been tested in a range of species, including frogs, toads, newts, trout, pigeons, rats, guinea-pigs, rabbits, ferrets, ground squirrels, tree shrews, marmots, marmosets, and rhesus monkeys. Tumours have been observed in all experiments with small animals, and the failure to induce neoplastic responses in large animals has been attributed to lack of information on the appropriate route or dose and the inability to observe the animals for a sufficient time (Osborne & Crosby, 1987a). In studies with other PAH, benzo [a]pyrene was often used as a positive control and therefore administered at only one concentration. Benzo [a]pyrene has been shown to be carcinogenic when given by a variety of routes, including diet, gavage, inhalation, intratracheal instillation, intraperitoneal, intravenous, subcutaneous, and intrapulmonary injection, dermal application, and transplacental administration. Assessment of the carcinogenic potency of the selected PAH is restricted for various reasons: Many of the studies performed before about 1970 were carried out without controls, without clearly defined, purified test substances, or using experimental designs and facilities considered today to be inadequate. Despite these shortcomings, all of the available studies were taken into account, except for those on dibenz [a,h]anthracene and benzo [a]pyrene. An overview of the results, as reported by the authors, is given in Table 90. To facilitate appraisal of the studies, the penultimate column gives a classification of the substances as positive, negative, or questionably carcinogenic; indicates whether the tumour incidence was evaluated statistically; and judges that a study is valid or provides reasons suggesting that it is unreliable. The criteria used to classify a study as valid were (i) an appropriate study protocol, i.e. use of concurrent controls (sham or vehicle), 20 or more animals per group, and study duration at least six months; and (ii) sufficient documentation, including detailed description of administration, results, and the survival of animals. As the use of concurrent controls is important for making judgements, data for these are given with the results for treated groups. If control data are not mentioned, it is because they were not given in the original paper. In experiments by topical application, the lower, more volatile PAH partially evaporate, and therefore their doses may have varied. The substances may also decompose. Both features could lead to underestimations of carcinogenic potency if they are not taken into account. Table 91 shows the classification of the compounds as carcinogenic, noncarcinogenic, or questionably carcinogenic. In order to make these classifications, all of the studies were summarized according to species and route of administration. In cases of doubt, the judgement was based on valid studies only. For example, despite one positive but invalid result and two questionable (one valid, one invalid) results from 17 studies, anthracene was classified as negative; however, pyrene, for which one positive, valid result and three questionable, valid results were found in 15 studies, could not be classified as negative and the compromise 'questionable' was chosen. The PAH found not to be carcinogenic were anthracene, benzo [ghi]perylene, fluorene, benzo [ghi]fluoranthene, 1-methylphenanthrene, perylene, and triphenylene. Questionable results were obtained for acenaphthene, benzo [a]-fluorene, benzo [b]fluorene, coronene, naphthalene, phenanthrene, and pyrene. The remaining compounds were found to be carcinogenic. The dermal route was the commonest mode of administration, followed by subcutaneous and intramuscular injection. In most studies, the site of tumour development is related closely to the route of administration, i.e. dermal application induces skin tumours, inhalation and intratracheal instillation result in lung tumours, subcutaneous injection results in sarcomas, and oral administration induces gastric tumours. Tumour induction is, however, not restricted to the obvious sites. For example, lung tumours have been observed after oral administration or subcutaneous injection of benzo [a]pyrene to mice and liver tumours following intraperitoneal injection. In two studies, lung tumours were found in mice after intravenous injection of benzo [a]pyrene and dibenz [a,h]anthracene. Thus, tissues such as the skin must be able to metabolize PAH to their ultimate metabolites and itself become a target organ; however, all PAH that reach the liver via the bloodstream can be metabolized there. The liver in turn is a depot from which the metabolites are distributed all over the body (Wall et al., 1991). The carcinogenic potency of the PAH differs by three orders of magnitude, and several authors have presented tables of toxic equivalence factors based on experimental results in order to quantify these differences. Carcinogenic potency cannot be based only on chemical structure but requires theoretical considerations and calculations (see section 7.10). Although this monograph primarily addresses single PAH, it was considered necessary for risk assessment to present some information on mixtures of PAH, to which humans are almost always exposed, predominantly adsorbed onto inhalable particles. Although the essential results of the studies of carcinogenicity are summarized in Table 90, special aspects and comparisons of individual PAH are presented in more detail below. 7.7.1 Single substances 7.7.1.1 Benzo [a]pyrene Oral administration of benzo [a]pyrene to male and female CFW mice induced gastric papillomas and squamous-cell carcinomas and increased the incidence of pulmonary adenomas (Rigdon & Neal, 1966). In other studies in which mice of the same strain were fed benzo [a]pyrene, pulmonary adenomas, thymomas, lymphomas, and leukaemia occurred, indicating that it can cause carcinomas distal to the point of application (Rigdon & Neal, 1969). The incidence of gastric tumours was 70% or more in mice fed 50-250 ppm benzo [a]pyrene for four to six months. No tumours were observed in controls (Rigdon & Neal, 1966; Neal & Rigdon, 1867; see also Table 90). In a study of the effects of benzo [a]pyrene given in the diet or by gavage in conjunction with caffeine, groups of 32 Sprague-Dawley rats of each sex were fed diets containing 0.15 mg/kg bw benzo [a]pyrene either five times per week or only on every ninth day. Tumours were observed in the forestomach, oesophagus, and larynx, at combined tumour incidences of 3/64, 3/64, and 10/64 in the controls and those at the low and high doses, respectively. In the study by gavage, groups of 32 rats of each sex were treated with benzo [a]pyrene at 0.15 mg/kg bw in a 1.5% caffeine solution every ninth day, every third day, or five times per week. The combined incidences of tumours of the forestomach, oesophagus, and larynx were 3/64 in controls, 6/64 in rats at the low dose, 13/64 in those at the medium dose, and 14/64 among those at the high dose (Brune et al., 1981). In hamsters exposed to 9.5 or 46.5 mg/m3 benzo [a]pyrene by inhalation for 109 weeks, a dose-response relationship was seen with tumorigenesis in the nasal cavity, pharynx, larynx, and trachea. The fact that lung tumours were not detected could not be explained (Thyssen et al., 1981). Hamster lung tissue can activate benzo [a]pyrene to carcinogenic derivatives (Dahl et al., 1985). Table 81. Mutagenicity of polycyclic aromatic hydrocarbons to Salmonella typhimurium Compound Result with Reference Strain metabolic activation Acenaphthene TA98,TA100 - Florin et al. (1980) TM677 + Kaden et al. (1979) TA98,TA100 + Epler et al. (1979) TA100 - Pahlman & Pelkonen (1987) Acenaphthylene TA98,TA100 - Florin et al. (1980) TM677 + Kaden et al. (1979) TA98,TA100 - Bos et al. (1988) Anthanthrene TA98 + Hermann (1981) TA100 + LaVoie et al. (1979); Andrews et al. (1978) TA98 - Tokiwa et al. (1977) TM677 + Kaden et al. (1979) Anthracene TA98,TA100 - Purchase et al. (1976) TA98,TA100 - Epler et al. (1979) TA100 - LaVoie et al. (1979); Gelboin & Ts'o (1978) TA98, TA100, - McCann et al. (1975a); TA1535,TA1537, Salamone et al. (1979); TA1538 Ho et al. (1981); Purchase et al.(1976) TA98,TA100 - Bridges et al, (1981) TA98,TA100, - Simmon (1979) TA1535, TA1536, TA1537,TA1538 TM677 - Kaden et al. (1979) TA97 + Sakai et al. (1985) TA98,TA100 - Probst et al. (1981) TA100 + Carver et al. (1986) TA98,TA100 - LaVoie et al.(1983a(1985) TA1535,TA1538 - Rosenkranz & Poirier (1979) TA100 - Pahlman & Pelkonen (1987) TA98,TA100 - Bos et al. (1988) TA98,TA100 - Florin et al. (1980) Table 81. (continued) Compound Result with Reference Strain metabolic activation Benz[a]anthracene TA100 + Epler et al. (1979); Bartsch et al. (1980) TA98,TA100 + McCann et al. (1975a); Coombs et al. (1976); Simmon (1979); Salamone et al. (1979) TA1535,TA1538 Rosenkranz & Poirier (1979) TA100 + Pahlman & Pelkonen (1987) TA98,TA100 + Hermann (1981); Carver et al.(1986) TA100 + Bartsch et al. (1980) TM677 + Kaden et al. (1979) TA100 + Baker et al. (1980) TA98,TA100 + Bos et al. (1988) TA98,TA100, + Probst et al. (1981) TA1535,TA1537 TA98, TA100, TA1537, TA1538 ± Dunkel et al. (1984) TA1535 - Dunkel et al. (1984) TA98,TA100 + Florin et al. (1980) TA1537,TA1538 - Teranishi et al. (1975) TA98 + Tokiwa et al. (1977) Benzo[b]fluoranthene TA98 + Hermann (1981) TA100 + LaVoie et al. (1979); Hecht et al. (1980) TA100 + Amin et al, (1985a) TA98,TA100 - Mossanda et al. (1979) Benzo[j]fluoranthene TA100 + LaVoie et al. (1980); Hecht et al. (1980) TM677 + Kaden et al. (1979) Benzo[k]fluoranthene TA100 + LaVoie et al. (1980); Hecht et al. (1980) TA98 + Hermann et al. (1980) Table 81. (continued) Compound Result with Reference Strain metabolic activation Benzo[ghi]fluoranthene TA98 ± Karcher et al. (1984) TA100 + Karcher et al. (1984) TA98,TA100 + LaVoie et al. (1979) Benzo[a]fluorene TA98, TA100, Salamone et al. (1979) TA1535, TA1537, TA1538 TA100 + Epler et al. (1979) TA100 - LaVoie et al. (1980) TA98,TA100 - Bos et al. (1988) TA98 + Tokiwa et al. (1977) Benzo[b]fluorene TA98, TA100 - LaVoie et al. (1980) TA98, TA100, - Salamone et al. (1979) TA1535, TA1537, TA1538 TM677 + Kaden et al. (1979) TA98,TA100 + Bos et al. (1988) Benzo[ghi]perylene TA98, TA1538 + Mossanda et al. (1979); Tokiwa et al. (1977); Katz et al. (1981) TA100 + Andrews et al. (1978); Katz et al. (1981); LaVoie et al. (1979); Salamone et al. (1979) TA1537,TA1538 + Poncelet et al. (1978) TM677 + Kaden et al. (1979) TA97 + Sakai et al. (1985) TA100 + Carver et al. (1986) Benzo[c]phenanthrene TA98 + Salamone et al. (1979); Wood et al. (1980) TA100 + Carver et al. (1986) TA100 + Wood et al. (1980) TA98,TA100 + Bos et al. (1988) Table 81. (continued) Compound Result with Reference Strain metabolic activation Benzo[a]pyrene TA98 + Epler et al. (1979) TA100 + Andrews et al. (1978) TA98,TA100 + LaVoie et al. (1979) TA98,TA100, + McCann et al. 1975a,b) TA1537,TA1538 TM677 + Kaden et al. (1979) TM677 + Rastetter et al. (1982) TM677 + Babson et al. (1 986b) TA97,TA98, + Sakai et al. (1985) TA100 TA98,TA100 + Prasanna et al. (1987)); Simmon (1979)); Glatt et al. (1987) TA1535,TA1538 + Rosenkranz & Poirier (1979) TA100 + Norpoth et al. (1984)); Alzieu et al. (1987)); Carver et al. (1986)); Bos et al. (1988); Hermann (1981); Bruce & Heddle (1979); Marino (1987); Alfheim & Ramdahl (1984) TA98 + Lee & Lin (1988) TA100 + Pahlman & Pelkonen (1987) TA97,TA98,TA100 + Marino (1987) TA97,TA98,TA100 + Sakai et al. (1985) TA98,TA100 + Devanesan et al. (1990) TM677 + Skopek & Thilly (1983) TA98,TA100, + Dunkel et al. (1984) TA1535, TA1537, TA1538 TA98, TA100 + Lofroth et al. (1984) TA98,TA100 + Florin et al. (1980) TA98 + Tokiwa et al. (1977) Table 81. (continued) Compound Result with Reference Strain metabolic activation Benzo[e]pyrene TA98 + LaVoie et al. (1979); Hermann (1981) TA100 ± Salamone et al. (1979) TA100 + Andrews et al. (1978); LaVoie et al., 1979) TA100 ± McCann et al. (1975a) TA1535,TA1538 - Rosenkranz & Poirier (1979) TM677 + Kaden et al. (1979) TA100 - Epler et al. (1979) TA98,TA100, + Simmon (1979) TA1538 TA97, TA100 + Sakai et al. (1985) TA98, TA100, ± Dunkel et al. (1984) TA1535,TA1537, TA1538 TA100 + Carver et al. (1986) TA100 - Pahlman & Pelkonen (1987) TA1537,TA1538 - Teranishi et al. (1975) TA98 + Tokiwa et al. (1977) Chrysene TA100 + McCann et al. (1975a); LaVoie et al. (1979) TA98 + McCann et al. (1975a) TA100 + Wood et al. (1977) TA100 + Epler et al. (1979); LaVoie et al. (1979) TA100 + Salamone et al. (1979) TA1535,TA1536, - Simmon (1979) TA1537,TA1538 TA98,TA100 + Bhatia et al. (1987) TM677 + Kaden et al. (1979) TA1535,TA1538 - Rosenkranz & Poirier (1979) TA97,TA100 + Sakai et al (1985) TA98,TA100 + Bos et al. (1988) TA98 + Hermann (1981) TA100 + Carver et al. (1986) TA100 + Pahlman & Pelkonen (1987) TA100 + Florin et al. (1980) TA98 + Tokiwa et al. (1977) Table 81. (continued) Compound Result with Reference Strain metabolic activation Coronene TA98 + Mossanda et al. (1979) TA98 + Hermann (1981) TA98 ± Salamone et al. (1979) TA98 + Florin et al. (1980) TA98, TA1537, + Poncelet et al. (1978) TA1538 TA97 ± Sakai et al. (1985) TM677 - Kaden et al. (1979) Cyclopenta[cd]pyrene TA98 + Wood et al. (1980) TA98,TA100, + Eisenstadt & Gold (1978) TA1537,TA1538 TM677 + Kaden et al. (1979); Cavalieri et al. (1981a) TA98 + Reed et al. (1988) Dibenz[a,h]anthracene TA100 + Andrews et al. (1978); Epler et al. (1979); McCann et al. (1975a,b) TA100 + Salamone et al. (1979) TA98 + Baker et al. (1980) TA98 + Hermann (1981) TM677 + Kaden et al. (1979) TA100 + Wood et al. (1978) TA100 + Pahlman & Pelkonen (1987); Carver et al., 1986) TA98, TA100, + Probst et al. (1981) TA1537,TA1538 TA100 + Platt et al. (1990) TA100 + Lecoq et al. (1989) TA1537,TA1538 - Teranishi et al. (1975) Dibenzo[a,e]pyrene TA100 + LaVoie et al. (1979) TA1537,TA1538 + Teranishi et al. (1975) TA98,TA100 +.± Devanesan et al. (1990) Dibenzo[a,h]pyrene TA100 ± LaVoie et al. (1979) TA98,TA100 + Wood et al. (1981) Table 81. (continued) Compound Result with Reference Strain metabolic activation Dibenzo[a,i]pyrene TA100 + LaVoie et al. (1979); McCann et al. (1975a) TA100 + Baker et al. (1980) TA98 + Hermann (1981) TA98 + Wood et al. (1981) TA1537,TA1538 + Teranishi et al. (1975) Not specified + Sardella et al. (1981) Dibenzo[a,l]pyrene TA98,TA100 + Karcher et al. (1984) TA98 + Hermann (1981) TA98,TA100 +,± Devanesan et al. (1990) Fluoranthene TA98 + Hermann et al. (1980) TA98 + Epler et al. (1979) TA100 - LaVoie et al. (1979) TA100 + LaVoie et al. (1982a) TA98, TA100, - Salamone et al. (1979) TA1535,TA1537, TA1538 TA98,TA100 + Poncelet et al. (1978) TA98,TA100 + Mossanda et al. (1979) TM677 + Kaden et al. (1979) TM677 + Rastetter et al. (1982) TM677 + Babson et al. (1986b) TA97,TA98,TA100 + Sakai et al. (1985) TA98,TA100 + Bos et al. (1988) TA100 + Carver et al. (1986); Hermann (1981); LaVoie et al., 1979) TA98,TA100 + Bos et al. (1987) TA97,TA102, ± Bos et al. (1987) TA1537 TA1535 - Bos et al. (1987) TA98,TA100 + Bhatia et al. (1987) TA98,TA100 - Florin et al. (1980) TA98 - Tokiwa et al. (1977) Table 81. (continued) Compound Result with Reference Strain metabolic activation Fluorene TA98, TA100, - McCann et al. (1975a); TA1535,TA1537 LaVoie et al. (1979, 1980, 1981a) TM677 - Kaden et al. (1979) TA97 - Sakai et al. (1985) TA98,TA100 - Bos et al. (1988) TA100 - Pahlman & Pelkonen (1987) Indeno[1,2,3-cd]pyrene TA98 + Hermann et al. (1980) TA100 + LaVoie et al. (1979) TA100 + Rice et al. (1985) 5-Methylcholanthrene TA100 + Coombs et al. (1976); Gelboin & Ts'o (1978); LaVoie et al. (1979); McCann et al. (1975a); Hecht et al. (1978) TA100 + Amin et al. (1979) TA100 + El-Bayoumy et al. (1986) 1-Methylphenanthrene TA100 + LaVoie et al. (1981b) TM677 + Kaden et al. (1979) TA97,TA98,TA100 + Sakai et al. (1985) TA98,TA100 + LaVoie et al. (1983b) Naphthalene TA98,TA100, - Florin et al. (1980) TA1535,TA1537 TA98, TA100, - McCann et al. (1975a) TA1535,TA1537, TA1538 TA98, TA100, - Purchase et al. (1976) TA1535,TA1538 TA98 - Ho et al. (1981) TM677 - Kaden et al. (1979) G46, E. coli K12 - Kramer et al. (1974) TA98,TA100 - Epler et al. (1979) TA98,TA100 - Mamber et al. (1984) Table 81. (continued) Compound Result with Reference Strain metabolic activation TA97,TA98,TA100 - Sakai et al. (1985) TA100 - Pahlman & Pelkonen (1987) TA98,TA100 - Bos et al. (1988) Perylene TA98 + Ho et al. (1981) TA100 + LaVoie et al. (1979) TA98,TA100, - Salamone et al. (1979) TA1535, TA1537, TA1538 TA98 + Hermann (1981) TA98 + Florin et al. (1980) TM677 + Kaden et al. (1979); Penman et al. (1980) TA100 + Carver et al. (1986) TA97,TA100 + Sakai et al. (1985) TA98,TA100 + Lofroth et al. (1984) TA100 - Pahlman & Pelkonen (1987) Phenanthrene TA100 + Oesch et al. (1981) TA100 - Wood et al. (1979) TA98 + Epler et al. (1979) TA98 - LaVoie et al. (1979, 1980) TA100 - LaVoie et al. (1981b) TA98,TA100 - Probst et al. (1981) TA100 - LaVoie et al. (1979); LaVoie et al. (1980); Gelboin & Ts'o (1978); McCann et al. (1975a) TA98, TA100, - McCann et al. (1975a) TA1535,TA1537 TA100 + Carver et al. (1986) TM677 - Kaden et al. (1979) TA97 + Sakai et al. (1985) TA98,TA100 ± Bos et al. (1988) TA1535,TA1536, - Simmon (1979) TA1537,TA1538 TA1535,TA1538 - Rosenkranz & Poirier (1979) TA100 - Pahlman & Pelkonen (1987) Table 81. (continued) Compound Result with Reference Strain metabolic activation TA98, TA100, - Dunkel et al. (1984) TA1535,TA1537, TA1538 TA98,TA100 - Florin et al. (1980) Pyrene TA98 - Ho et al. (1981); Rice et al. (1988a) TA98,TA100, - McCann et al. (1975a); LaVoie et al. (1979); TA1535,TA1537 Ho et al. (1981) TA1537 + Bridges et al. (1981) TA98,TA100 - Salamone et al. (1979) TA98,TA100 - Probst et al. (1981) TA1537 + Epler et al. (1979) TM677 + Kaden et al. (1979) TA97 + Sakai et al. (1985) TA98,TA100 ± Bos et al. (1988) TA100 - Carver et al. (1986); Hermann (1981) TA98,TA100 + Bhatia et al. (1987) TA98, TA100, - Dunkel et al. (1984) TA1536,TA1537, TA1538 TA100 - Pahlman & Pelkonen (1987) TA98,TA100 - Florin et al. (1980) Triphenylene TA98 + Epler et al. (1979) TA98 + Tokiwa et al. (1977) TA98,TA100 + Mossanda et al. (1979); Wood et al. (1980) TA98 + Hermann (1981) TA98,TA100 + Poncelet et al. (1978) TM677 + Kaden et al. (1979) TA98,TA100 + Bos et al. (1988) TA100 + Pahlman & Pelkonen (1987) TA, used to test reverse mutation to histidine non-auxotrophic mutants); TM, used to test forward mutation to 8-azaguanine-resistant mutants +, positive); -, negative); ±, inconclusive Table 82. DNA damage induced by polycyclic aromatic hydrocarbons in vitro Test system End-point Metabolica Resultb Reference activation Prokaryotes Anthracene E. coli pol A- R + - Rosenkranz & Poirier (1979) E. coli WP2, E. coli WP100 R + - Member et al. (1983) E. coli WP2, E. coli WP67, R +/- - Tweats (1981) E. coli CM871 E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) E. coli WP2s(lambda) R +/- + Rossman et al. prophage induction) (1991) B. subtilis R +/- - Ashby & Kilby (1981) B. subtilis R +/- - McCarroll et al. (1981) E. coli GY5027 (prophage R + - Mamber et al. induction) (1984) Anthranene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Benz[a]anthracene E. coli pol A- R + - Rosenkranz & Poirier (1979) E. coli WP2 uvrA R + - Dunkel et al. (1984) E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Benzo[b]fluoranthene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Table 82. (continued) Test system End-point Metabolica Resultb Reference activation Benzo[ghi]fluoranthene E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) Benzo[j]fluoranthene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Benzo[a]fluoranthene E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) Benzo[b]fluoranthene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Bunzo[ghi]perylene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Benzo[a]pyrene E. coli WP2, E. coli WP100 R + + Mamber et al. (1983) E. coli GY5027 R + + Mamber et al. (1983) E. coli pol A- R + + Rosenkranz & Poirier (1979) E. coli WP2, E. coli WP67, R +/- + Tweats (1981) E. coli CM871 E. coli WP2 uvrA R + - Dunkel et al. (1984) E. coli PQ37 R +/- +/+ Mersch- Sundermann et al. (1992) B. subtilis R +/- + McCarroll et al. (1981) E. coli WP2s(lambda) R +/- + Rossman et al. prophago induction) (1991) Table 82. (continued) Test system End-point Metabolica Resultb Reference activation Benzo[e]pyrene E. coli pol A- R + - Rosenkranz & Poirier (1979) E. coli WP2 uvrA R + - Dunkel et al. (1984) E. coli WP2s(lambda) R +/- + Rossman et al. prophage induction) (1991) E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Chrysene E. coli pol A- R + - Rosenkranz & Poirier (1979) E. coli PQ37 R +/- + Mersch- Sundermann at al. (1992) Coronene E. coli PQ37 R +/- - Mersch- Sundermann at al. (1992) Dibenz[a,h]anthracene E. coli R +/- + Ichinotsubo et al. (1977) E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) B. subtilis R +/- + McCarroll et al. (1981) E. coli WP2s (lambda R +/- + Rossman et al. prophage induction) (1991) Dibenzo[a,i]pyrene E. coli R +/- + Ichinotsubo et al. (1977) E. coli PQ37 R +/- + Mersch- Sundermann et al.(1992) B. subtilis R +/- + McCarroll et al. (1981) Table 82. (continued) Test system End-point Metabolica Resultb Reference activation Dibenzo[a,h]pyrene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Dibenzo[a,i]pyrene E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Fluoranthene E. coli WP2s (lambda R +/- - Rossman et al. prophage induction) (1991) E. coli PQ37 R +/- + Mersch- Sundermann et al. (1992) Fluoranthene E. coli WP2, E. coli WP100 R + - Mamber et al. (1983) E. coli GY5027 R + - Mamber et al. (1984) E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) Indeno[1,2,3-cd]pyrene E. coli PQ37 R + - Mersch- Sundermann et al.(1992) Naphthalene E. coli WP2, E. coli WP 100 R + - Mamber et al. (1983) E. coli GY5027 R + - Mamber et al. (1984) E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) Parylene E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) Table 82. (continued) Test system End-point Metabolica Resultb Reference activation Phenanthrene E. coli pol A- R + - Rosenkranz & Poirier (1979) E. coli WP2 uvrA R + - Dunkel et al. (1984) E. coli PQ37 R +/- + Mersch- Sundermann at al. (1992) E. coli WP2s (lambda R +/- + Rossman et al. prophage induction) (1991) B. subtilis R +/- - McCarroll et al. (1981) Pyrene E. coli R +/- - Ashby & Kilbey (1981; De Serres & Ashby, 1981) E. coli WP2, E, coli WP100 R + - Mamber et al. (1983) E. coli GY5027 R + - Mamber et al. (1984) E. coli WP2 uvrA R + - Dunkel et al. (1984) E. coli WP2, E. coli WP67, R +/- - Tweats (1981) E. coli CM871 E. coli PQ37 R +/- - Mersch- Sundermann et al. (1992) B. subtilis R +/- - Ashby & Kilbey (1981) B. subtilis R +/- - McCarroll et al.