IPCS INCHEM Home


    UNITED NATIONS ENVIRONMENT PROGRAMME
    INTERNATIONAL LABOUR ORGANISATION
    WORLD HEALTH ORGANIZATION


    INTERNATIONAL PROGRAMME ON CHEMICAL SAFETY



    ENVIRONMENTAL HEALTH CRITERIA 188





    Nitrogen Oxides

    (Second Edition)



    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 draft prepared by Drs J.A. Graham, L.D. Grant, L.J. Folinsbee,
    D.J. Kotchmar and J.H.B. Garner, US Environmental Protection Agency



    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, 1997

         The International Programme on Chemical Safety (IPCS) is a joint
    venture of the United Nations Environment Programme, the International
    Labour Organisation, and the World Health Organization. The main
    objective of the IPCS is to carry out and disseminate evaluations of
    the effects of chemicals on human health and the quality of the
    environment. Supporting activities include the development of
    epidemiological, experimental laboratory, and risk-assessment methods
    that could produce internationally comparable results, and the
    development of manpower in the field of toxicology. Other activities
    carried out by the IPCS include the development of know-how for coping
    with chemical accidents, coordination of laboratory testing and
    epidemiological studies, and promotion of research on the mechanisms
    of the biological action 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 UN 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

    Nitrogen oxides - 2nd ed.

    (Environmental health criteria ; 188)

    1.Nitrogen dioxide                 2.Nitrogen oxides
    I.Series

    ISBN 92 4 157188 8                 (NLM Classification: WA 754)
    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 1997

         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.

         The Federal Ministry for the Environment, Nature Conservation and
            Nuclear Safety, Germany, provided financial support for, and
                      undertook the printing of, this publication

    CONTENTS

    ENVIRONMENTAL HEALTH CRITERIA FOR NITROGEN OXIDES

    Preamble

    1. SUMMARY

         1.1. Nitrogen oxides and related compounds
               1.1.1. Atmospheric transport
               1.1.2. Measurement
               1.1.3. Exposure
         1.2. Effects of atmospheric nitrogen species, particularly
               nitrogen oxides, on vegetation
         1.3. Health effects of exposures to nitrogen dioxide
               1.3.1. Studies of the effects of nitrogen compounds on
                       experimental animals
                       1.3.1.1    Biochemical and cellular mechanisms of
                                  action of nitrogen oxides
                       1.3.1.2    Effects on host defence
                       1.3.1.3    Effects of chronic exposure on the
                                  development of chronic lung disease
                       1.3.1.4    Potential carcinogenic or co-carcinogenic
                                  effects
                       1.3.1.5    Age susceptibility
                       1.3.1.6    Influence of exposure patterns
               1.3.2. Controlled human exposure studies on nitrogen
                       oxides
               1.3.3. Epidemiology studies on nitrogen dioxide
               1.3.4. Health-based guidance values for nitrogen dioxide

    2. PHYSICAL AND CHEMICAL PROPERTIES, AIR SAMPLING AND ANALYSIS,
         TRANSFORMATIONS AND TRANSPORT IN THE ATMOSPHERE

         2.1. Introduction
               2.1.1. The nomenclature and measurement of atmospheric
                       nitrogen species
         2.2. Nitrogen species and their physical and chemical properties
               2.2.1. Nitrogen oxides
                       2.2.1.1    Nitric oxide
                       2.2.1.2    Nitrogen dioxide
                       2.2.1.3    Nitrous oxide
                       2.2.1.4    Other nitrogen oxides
               2.2.2. Nitrogen acids
                       2.2.2.1    Nitric acid
                       2.2.2.2    Nitrous acid
               2.2.3. Ammonia
               2.2.4. Ammonium nitrate
               2.2.5. Peroxyacetyl nitrate
               2.2.6. Organic nitrites and nitrates

         2.3. Sampling and analysis methods
               2.3.1. Nitric oxide
                       2.3.1.1    Nitric oxide continuous methods
                       2.3.1.2    Passive samplers for NO
                       2.3.1.3    Calibration of NO analysis methods
                       2.3.1.4    Sampling considerations for NO
               2.3.2. Nitrogen dioxide
                       2.3.2.1    Chemiluminescence (NO + O3)
                       2.3.2.2    Chemiluminescence (luminol)
                       2.3.2.3    Laser-induced fluorescence and tuneable
                                  diode laser absorption spectrometry
                       2.3.2.4    Wet chemical methods
                       2.3.2.5    Other methods
                       2.3.2.6    Passive samplers
                       2.3.2.7    Calibration
               2.3.3. Total reactive odd nitrogen
               2.3.4. Peroxyacetyl nitrate
               2.3.5. Other organic nitrates
               2.3.6. Nitric acid
               2.3.7. Nitrous acid
               2.3.8. Dinitrogen pentoxide and nitrate radicals
               2.3.9. Particulate nitrate
               2.3.10. Nitrous oxide
               2.3.11. Summary
         2.4. Transport and transformation of nitrogen oxides in the air
               2.4.1. Introduction
               2.4.2. Chemical transformations of oxides of nitrogen
                       2.4.2.1    Nitric oxide, nitrogen dioxide and ozone
                       2.4.2.2    Transformations in indoor air
                       2.4.2.3    Formation of other oxidized nitrogen
                                  species
               2.4.3. Advection and dispersion of atmospheric nitrogen
                       species
                       2.4.3.1    Transport of reactive nitrogen species
                                  in urban plumes
                       2.4.3.2    Air quality models
                       2.4.3.3    Regional transport
         2.5. Conversion factor for nitrogen dioxide
         2.6. Summary

    3. SOURCES, EMISSIONS AND AIR CONCENTRATIONS

         3.1. Introduction
         3.2. Sources of nitrogen oxides
               3.2.1. Sources of NOx emission
                       3.2.1.1    Fuel combustion
                       3.2.1.2    Biomass burning
                       3.2.1.3    Lightning
                       3.2.1.4    Soils
                       3.2.1.5    Oceans

               3.2.2. Removal from the ambient environment
               3.2.3. Summary of global budgets for nitrogen oxides
         3.3. Ambient concentrations of nitrogen oxides
               3.3.1. International comparison studies of NOx
                       concentrations
               3.3.2. Example case studies of NOx and NO2
                       concentrations
         3.4. Occurrence of nitrogen oxides indoors
               3.4.1. Indoor sources
                       3.4.1.1    Gas-fuelled cooking stoves
                       3.4.1.2    Unvented gas space heaters and water
                                  heaters
                       3.4.1.3    Kerosene space heaters
                       3.4.1.4    Wood stoves
                       3.4.1.5    Tobacco products
               3.4.2. Removal of nitrogen oxides from indoor environments
         3.5. Indoor concentrations of nitrogen oxides
               3.5.1. Homes without indoor combustion sources
               3.5.2. Homes with combustion appliances
               3.5.3. Homes with combustion space heaters
               3.5.4. Indoor nitrous acid concentrations
               3.5.5. Predictive models for indoor NO2 concentration
         3.6. Human exposure
         3.7. Exposure of plants and ecosystems

    4. EFFECTS OF ATMOSPHERIC NITROGEN COMPOUNDS (PARTICULARLY NITROGEN
         OXIDES) ON PLANTS

         4.1. Properties of NOx and NHy
               4.1.1. Adsorption and uptake
               4.1.2. Toxicity, detoxification and assimilation
               4.1.3. Physiology and growth aspects
               4.1.4. Interactions with climatic conditions
               4.1.5. Interactions with the habitat
               4.1.6. Increasing pest incidence
               4.1.7. Conclusions for various atmospheric nitrogen
                       species and mixtures
                       4.1.7.1    NO2
                       4.1.7.2    NO
                       4.1.7.3    NH3
                       4.1.7.4    NH4+ and NO3- in wet and occult
                                  deposition
                       4.1.7.5    Mixtures
               4.1.8. Appraisal
                       4.1.8.1    Representativity of the data
               4.1.9. General conclusions
         4.2. Effects on natural and semi-natural ecosystems
               4.2.1. Effects on freshwater and intertidal ecosystems
                       4.2.1.1    Effects of nitrogen deposition on
                                  shallow softwater lakes

                       4.2.1.2    Effects of nitrogen deposition on lakes
                                  and streams
               4.2.2. Effects on ombrotrophic bogs and wetlands
                       4.2.2.1    Effects on ombrotrophic (raised) bogs
                       4.2.2.2    Effects on mesotrophic fens
                       4.2.2.3    Effects on fresh- and saltwater marshes
               4.2.3. Effects on species-rich grasslands
                       4.2.3.1    Effects of nitrogen on calcareous
                                  grasslands
                       4.2.3.2    Critical loads for nitrogen in
                                  calcareous grasslands
                       4.2.3.3    Comparison with other semi-natural
                                  grasslands
               4.2.4. Effects on heathlands
                       4.2.4.1    Effects on inland dry heathlands
                       4.2.4.2    Effects of nitrogen on inland wet
                                  heathlands
                       4.2.4.3    Effects of nitrogen on arctic and alpine
                                  healthlands
                       4.2.4.4    Effects on herbs of matgrass swards
               4.2.5. Effects of nitrogen deposition on forests
                       4.2.5.1    Effects on forest tree species
                       4.2.5.2    Effects on tree epiphytes, ground
                                  vegetation and ground fauna of forests
               4.2.6. Effects on estuarine and marine ecosystems
               4.2.7. Appraisal and conclusions

    5. STUDIES OF THE EFFECTS OF NITROGEN OXIDES ON EXPERIMENTAL ANIMALS

         5.1. Introduction
         5.2. Nitrogen dioxide
               5.2.1. Dosimetry
                       5.2.1.1    Respiratory tract dosimetry
                       5.2.1.2    Systemic dosimetry
               5.2.2. Respiratory tract effects
                       5.2.2.1    Host defence mechanisms
                       5.2.2.2    Lung biochemistry
                       5.2.2.3    Pulmonary function
                       5.2.2.4    Morphological studies
               5.2.3. Genotoxicity, potential carcinogenic or
                       co-carcinogenic effects
               5.2.4. Extrapulmonary effects
         5.3. Effects of mixtures containing nitrogen dioxide
         5.4. Effects of other nitrogen oxide compounds
               5.4.1. Nitric oxide
                       5.4.1.1    Endogenous formation of NO
                       5.4.1.2    Absorption of NO
                       5.4.1.3    Effects of NO on pulmonary function,
                                  morphology and host lung defence
                                  function

                       5.4.1.4    Metabolic effects
                       5.4.1.5    Haematological changes
                       5.4.1.6    Biochemical mechanisms for nitric oxide
                                  effects: reaction with iron and effects
                                  on enzymes and nucleic acids
               5.4.2. Nitric acid
               5.4.3. Nitrates
         5.5. Summary of studies of the effects of nitrogen compounds on
               experimental animals

    6. CONTROLLED HUMAN EXPOSURE STUDIES OF NITROGEN OXIDES

         6.1. Introduction
         6.2. Effects of nitrogen dioxide
               6.2.1. Nitrogen dioxide effects on pulmonary function and
                       airway responsiveness to bronchoconstrictive agents
                       6.2.1.1    Nitrogen dioxide effects in healthy
                                  subjects
                       6.2.1.2    Nitrogen dioxide effects on asthmatics
                       6.2.1.3    Nitrogen dioxide effects on patients
                                  with chronic obstructive pulmonary
                                  disease
                       6.2.1.4    Age-related differential susceptibility
               6.2.2. Nitrogen dioxide effects on pulmonary host defences
                       and bronchoalveolar lavage fluid biomarkers
               6.2.3. Other classes of nitrogen dioxide effects
         6.3. Effects of other nitrogen oxide compounds
         6.4. Effects of nitrogen dioxide/gas or gas/aerosol mixtures on
               lung function
         6.5. Summary of controlled human exposure studies of oxides of
               nitrogen

    7. EPIDEMIOLOGICAL STUDIES OF NITROGEN OXIDES

         7.1. Introduction
         7.2. Methodological considerations
               7.2.1. Measurement error
               7.2.2. Misclassification of the health outcome
               7.2.3. Adjustment for covariates
               7.2.4. Selection bias
               7.2.5. Internal consistency
               7.2.6. Plausibility of the effect
         7.3. Studies of respiratory illness
               7.3.1. Indoor air studies
                       7.3.1.1    St Thomas' Hospital Medical School
                                  Studies (United Kingdom)
                       7.3.1.2    Harvard University - Six Cities Studies
                                  (USA)
                       7.3.1.3    University of Iowa Study (USA)

                       7.3.1.4    Agricultural University of Wageningen
                                  (The Netherlands)
                       7.3.1.5    Ohio State University Study (USA)
                       7.3.1.6    University of Dundee (United Kingdom)
                       7.3.1.7    Harvard University - Chestnut Ridge
                                  Study (USA)
                       7.3.1.8    University of New Mexico Study (USA)
                       7.3.1.9    University of Basel Study (Switzerland)
                       7.3.1.10   Yale University Study (USA)
                       7.3.1.11   Freiburg University Study (Germany)
                       7.3.1.12   McGill University Study (Canada)
                       7.3.1.13   Health and Welfare Canada Study (Canada)
                       7.3.1.14   University of North Carolina Study (USA)
                       7.3.1.15   University of Tucson Study (USA)
                       7.3.1.16   Hong Kong Anti-Cancer Society Study
                                  (Hong Kong)
                       7.3.1.17   Recent studies
               7.3.2. Outdoor studies
                       7.3.2.1    Harvard University - Six City Studies
                                  (USA)
                       7.3.2.2    University of Basel Study (Switzerland)
                       7.3.2.3    University of Wuppertal Studies
                                  (Germany)
                       7.3.2.4    University of Tubigen (Germany)
                       7.3.2.5    Harvard University - Chestnut Ridge
                                  Study (USA)
                       7.3.2.6    University of Helsinki Studies (Finland)
                       7.3.2.7    Helsinki City Health Department Study
                                  (Finland)
                       7.3.2.8    Oulu University Study (Finland)
                       7.3.2.9    Seth GS Medical College Study (India)
         7.4. Pulmonary function studies
               7.4.1. Harvard University - Six City Studies (USA)
               7.4.2. National Health and Nutrition Examination Survey
                       Study (USA)
               7.4.3. Harvard University - Chestnut Ridge Study (USA)
               7.4.4. Other pulmonary function studies
         7.5. Other exposure settings
               7.5.1. Skating rink exposures
         7.6. Occupational exposures
         7.7. Synthesis of the evidence for school-age children
               7.7.1. Health outcome measures
               7.7.2. Biologically plausible hypothesis
               7.7.3. Publication bias
               7.7.4. Selection of studies
                       7.7.4.1    Brief description of selected studies
                       7.7.4.2    Studies not selected for quantitative
                                  analysis
               7.7.5. Quantitative analysis

         7.8. Synthesis of the evidence for young children
         7.9. Summary

    8. EVALUATION OF HEALTH AND ENVIRONMENT RISKS ASSOCIATED WITH
         NITROGEN OXIDES

         8.1. Sources and exposure
         8.2. Evaluation of the effects of atmospheric nitrogen species
               on the environment
               8.2.1. Guidance values - critical levels for air
                       concentrations of nitrogen oxides
               8.2.2. Environment-based guidance values - critical loads
                       for total nitrogen deposition
         8.3. Evaluation of health risks associated with nitrogen oxides
               8.3.1. Concentration-response relationships
               8.3.2. Subpopulations potentially at risk
               8.3.3. Derivation of health-based guidance values

    9. CONCLUSIONS AND RECOMMENDATIONS FOR PROTECTION OF HUMAN HEALTH
         AND THE ENVIRONMENT

    10. FURTHER RESEARCH

    REFERENCES

    RESUME

    RESUMEN
    

    NOTE TO READERS OF THE CRITERIA MONOGRAPHS

         Every effort has been made to present information in the criteria
    monographs as accurately as possible without unduly delaying their
    publication.  In the interest of all users of the Environmental Health
    Criteria monographs, readers are requested to communicate any errors
    that may have occurred to the Director of the International Programme
    on Chemical Safety, World Health Organization, Geneva, Switzerland, in
    order that they may be included in corrigenda.



                                *     *     *



         A detailed data profile and a legal file can be obtained from the
    International Register of Potentially Toxic Chemicals, Case postale
    356, 1219 Châtelaine, Geneva, Switzerland (Telephone No. 9799111).

    Environmental Health Criteria

    PREAMBLE

    Objectives

         In 1973 the WHO Environmental Health Criteria Programme was
    initiated 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 and since that time an ever-increasing
    number of assessments of chemicals and of physical effects have been
    produced.  In addition, many EHC monographs have been devoted to
    evaluating toxicological methodology, e.g., for genetic, neurotoxic,
    teratogenic and nephrotoxic effects.  Other publications have been
    concerned with epidemiological guidelines, evaluation of short-term
    tests for carcinogens, biomarkers, effects on the elderly and so
    forth.

         Since its inauguration the EHC Programme has widened its scope,
    and the importance of environmental effects, in addition to health
    effects, has been increasingly emphasized in the total evaluation of
    chemicals.

         The original impetus for the Programme came from World Health
    Assembly resolutions and the recommendations of the 1972 UN 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 UN Conference on Environment and
    Development and the subsequent establishment of the Intergovernmental
    Forum on Chemical Safety with the priorities for action in the six
    programme areas of Chapter 19, Agenda 21, all lend further weight to
    the need for EHC assessments of the risks of chemicals.

    Scope

         The criteria monographs are intended to provide critical reviews
    on the effect on human health and the environment of chemicals and of
    combinations of chemicals and physical and biological agents.  As
    such, they include and review studies that are of direct relevance for
    the evaluation.  However, they do not describe  every study carried
    out.  Worldwide data are used and are quoted from original studies,
    not from abstracts or reviews.  Both published and unpublished reports
    are considered and it is incumbent on the authors to assess all the
    articles cited in the references.  Preference is always given to
    published data.  Unpublished data are only used when relevant
    published data are absent or when they are pivotal to the risk
    assessment.  A detailed policy statement is available that describes
    the procedures used for unpublished proprietary data so that this
    information can be used in the evaluation without compromising its
    confidential nature (WHO (1990) Revised Guidelines for the Preparation
    of Environmental Health Criteria Monographs. PCS/90.69, Geneva, World
    Health Organization).

         In the evaluation of human health risks, sound human data,
    whenever available, are preferred to animal data.  Animal and
     in vitro studies provide support and are used mainly to supply
    evidence missing from human studies.  It is mandatory that research on
    human subjects is 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
    standard setting.  These 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 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., IARC, JECFA,
         JMPR

    Selection of chemicals

         Since the inception of the EHC Programme, the IPCS has organized
    meetings of scientists to establish lists of priority chemicals for
    subsequent evaluation.  Such meetings have been held in: Ispra, Italy,
    1980; Oxford, United Kingdom, 1984; Berlin, Germany, 1987; and North
    Carolina, USA, 1995. The selection of chemicals has been based on the
    following criteria: the existence of scientific evidence that the
    substance presents a hazard to human health and/or the environment;
    the possible use, persistence, accumulation or degradation of the
    substance shows that there may be significant human or environmental
    exposure; the size and nature of populations at risk (both human and
    other species) and risks for environment; international concern, i.e.
    the substance is of major interest to several countries; adequate data
    on the hazards are available.

         If an EHC monograph is proposed for a chemical not on the
    priority list, the IPCS Secretariat consults with the Cooperating
    Organizations and all the Participating Institutions before embarking
    on the preparation of the monograph.

    Procedures

         The order of procedures that result in the publication of an EHC
    monograph is shown in the 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
    layout and language.  The first draft, prepared by consultants or,
    more usually, staff from 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 document acceptable as
    a first draft, it is distributed, in its unedited form, to well over
    150 EHC contact points throughout the world who are asked to comment
    on its completeness and accuracy and, where necessary, provide
    additional material.  The contact points, usually designated by
    governments, may be Participating Institutions, IPCS Focal Points, or

    FIGURE 1

    individual scientists known for their particular expertise.  Generally
    some four months are allowed before the comments are considered by the
    RO and author(s).  A second draft incorporating comments received and
    approved by the Director, IPCS, is then distributed to Task Group
    members, who carry out the 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 health and
    environmental risks from exposure to the chemical.  A summary and
    recommendations for further research and improved safety aspects are
    also required.  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.
    Representatives from relevant national and international associations
    may be invited to join the Task Group as observers.  While observers
    may provide a valuable contribution to the process, they can only
    speak at the invitation of the Chairperson.  Observers do not
    participate in the final evaluation of the chemical; this is the sole
    responsibility of the Task Group members.  When the Task Group
    considers it to be appropriate, it may meet  in camera.

         All individuals who as authors, consultants or advisers
    participate in the preparation of the EHC monograph 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 conflict of
    interest statement. Such a 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 then goes for language editing, reference checking, and
    preparation of camera-ready copy.  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 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 for
    health or environmental effects of the agent because of greater
    exposure; an appreciable time period 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 NITROGEN OXIDES

     Members 

    Dr K. Bentley*, Health and Environment Policy Section, Department
         of Community Services and Health, Canberra ACT, Australia

    Dr S. Dobson, Institute of Terrestrial Ecology, Monks Wood
         Experimental Station, Abbots Ripton, Huntingdon, Cambridgeshire,
         United Kingdom

    Dr L. van der Eerden, Centre "De Bom"  Wageningen, The Netherlands

    Dr L. Folinsbee, Health Effects Research Laboratory, US Environmental
         Protection Agency, Research Triangle Park, North Carolina, USA
          (Rapporteur)

    Dr L. Grant*, National Center for Environmental Assessment, US
         Environmental Protection Agency, Research Triangle Park, North
         Carolina, USA

    Mr L. Heiskanen, Health and Environment Policy Section, Department of
         Community Services and Health, Canberra ACT, Australia

    Mr G.M. Johnson, CSIRO, Division of Coal and Energy Technology, Centre
         for Pollution Assessment and Control, North Ryde, NSW, Australia

    Dr J. Kagawa, Professor of Hygiene and Public Health, Tokyo Women's
         Medical College, Shinjuku-ku, Tokyo, Japan

    Dr R.R. Khan, Ministry of Environment and Forests, Paryavaran Bhawan,
         New Delhi, India

    Dr D.B. Menzel, University of California, Department of Community &
         Environment and Medicine, California, USA

    Dr L. Neas, Department of Environmental Health, Environmental
         Epidemiology Program, Harvard School of Public Health, Boston,
         Massachusetts, USA

    Dr S.E. Paulson, Department of Atmospheric Sciences, University of
         California, Los Angeles, California, USA

    Dr P.J.A. Rombout, Department for Inhalation Toxicology, National
         Institute of Public Health and Environmental Hygiene, Bilthoven,
         The Netherlands  (Chairman)

               

    *  Invited, but unable to attend

    Dr W. Tyler, Veterinary Anatomy and Cell Biology, University of
         California, California, USA

    Dr K. Victorin, Karolinska Institute, Institute of Environmental
         Medicine, Stockholm, Sweden

    Dr A. Woodward, Department of Community Medicine, University of
         Adelaide, Adelaide, Australia

    Dr R. Ye, Deputy Director, National Environmental Protection Agency,
         Xizhimennei Nanziaojie, Beijing, People's Republic of China

     Observers

    Professor M. Moore, National Research Centre for Environmental
         Toxicology, Nathan, Australia

    Dr M. Pain, Department of Thoracic Medicine, Royal Melbourne Hospital,
         Melbourne VIC, Australia

    Dr P. Psaila-Savona, WA Department of Health, Perth WA, Australia

    Mr B. Taylor, Policy and Planning Group, Public and Planning Group,
         Public Health Commission, Wellington, New Zealand

    Mr B. Saxby, AGL Gas Companies, North Sydney NSW, New  Zealand

     Secretariat

    Dr B.H. Chen, International Programme on Chemical Safety, World Health
         Organization, Geneva, Switzerland  (Secretary)

    Dr M. Younes, WHO European Centre for Environment & Health, Bilthoven,
         The Netherlands

    ENVIRONMENTAL HEALTH CRITERIA FOR NITROGEN OXIDES

         A WHO Task Group on Environmental Health Criteria for Nitrogen
    Oxides met in Melbourne, Australia from 14 to 18 November 1994.  The
    meeting was hosted by the Clean Air Society of Australia and New
    Zealand and the Victorian Departments of Health and Environment,
    Australia.  Dr B.H. Chen, IPCS, opened the meeting and welcomed the
    participants on behalf of the Director, IPCS, and the three IPCS
    cooperating organizations (UNEP/ILO/WHO).  The Task Group reviewed and
    revised the draft criteria monograph and made an evaluation of the
    risks for human health and the environment from exposure to nitrogen
    oxides. 

         The first draft of this monograph was prepared by Drs J.A.
    Graham, L.D. Grant, L.J. Folinsbee, D.J. Kotchmar and J.H.B. Garner,
    US EPA.  Drs W.G. Ewald, T.B. McMullen and B.E. Tilton, US EPA,
    contributed to the preparation of the first draft.  The second draft
    was prepared by Dr L.D. Grant incorporating comments received
    following the circulation of the first draft to the IPCS Contact
    Points for Environmental Health Criteria.  Drs R. Bobbink, L. Van der
    Eerden and S. Dobson prepared the final text of the environmental
    section.  Mr G.M. Johnson contributed to the final text of the
    chemistry section.

         Dr B.H. Chen and Dr P.G. Jenkins, both members of the IPCS
    Central Unit, were responsible for the overall scientific content and
    technical editing, respectively.

         The efforts of all who helped in the preparation and finalization
    of the document are gratefully acknowledged.

         Financial support for this Task Group meeting was provided by the
    Department of Community Services and Health, Australia, Victorian
    Departments of Health and Environment, Australia, and the Clean Air
    Society of Australia and New Zealand.

    ABBREVIATIONS

    ADP       adenosine diphosphate
    AM        alveolar macrophages
    AQG       Air Quality Guidelines
    BAL       bronchoalveolar lavage
    BHPN       N-bis (2-hydroxypropyl) nitrosamine
    CI        confidence interval
    CLM       chemiluminescence method
    COPD      chronic obstructive pulmonary disease
    ECD       electron capture detection
    FEF       forced expiratory flow
    FEV       forced expiratory volume
    FTIR      Fourier transformed infrared
    FVC       forced vital capacity
    GC        gas chromatography
    GDH       glutamate dehydrogenase
    (c)GMP    (cyclic) guanosine monophosphate
    GS        glutamine synthetase
    HNO2      nitrous acid
    HNO3      nitric acid
    LIF       laser-induced fluorescence
    MS        mass spectrometry
    N2        nitrogen (elemental)
    NH3       ammonia
    NH4+      ammonium ion
    NHy       the sum of NH3 and NH4+
    NiR       nitrate reductase
    NK        natural killer
    NO        nitric oxide
    NO2       nitrogen dioxide
    NO2-      nitrite ion
    NO3-      nitrate ion
    N2O       nitrous oxide
    N2O5      nitrogen pentoxide
    NOx       nitric oxide plus nitrogen dioxide
    NOy       gas-phase oxidized nitrogen species (except nitrous oxide)
    NPSH      non-protein sulfhydryl
    NR        nitrate reductase
    O3        ozone
    PAN       peroxyacetyl nitrate
    PBzN      peroxybenzoyl nitrate
    PEF       peak expiratory flow
    PFC       plaque-forming cell
    PMN       polymorphonuclear leukocyte
    ppb       parts per billion (10-9)
    ppm       parts per million (10-6)
    ppt       parts per trillion (10-12)
    pptv      parts per trillion (by volume)
    PSD       passive sampling device

    Raw       airway resistance
    ROC       reactive organic carbon
    RUBISCO   ribulose 1,5-biphosphate carboxylase
    SD        standard deviation
    SES       socioeconomic status
    SGaw      specific airway conductance
    SO2       sulfur dioxide
    SOy       sulfur oxides
    SPM       suspended particulate matter
    SRaw      specific airway resistance
    TDLAS     tuneable diode laser absorption spectrometry
    TSP       total suspended particulate
    VOC       volatile organic carbon

    1.  SUMMARY

    1.1  Nitrogen oxides and related compounds

         Nitrogen oxides can be present at significant concentrations in
    ambient air and in indoor air.  The types and concentrations of
    nitrogenous compounds present can vary greatly from location to
    location, with time of day, and with season.  The main sources of
    nitrogen oxide emissions are combustion processes.  Fossil fuel power
    stations, motor vehicles and domestic combustion appliances emit
    nitrogen oxides, mostly in the form of nitric oxide (NO) and some
    (usually less than about 10%) in the form of nitrogen dioxide (NO2). 
    In the air, chemical reactions occur that oxidize NO to NO2 and other
    products. There are also biological processes that liberate nitrogen
    species from soils, including nitrous oxide (N2O).  Emissions of N2O
    can cause perturbation of the stratospheric ozone layer.

         Human health may be affected when significant concentrations of
    NO2 or other nitrogenous species, such as peroxyacetyl nitrate (PAN),
    nitric acid (HNO3), nitrous acid (HNO2), and nitrated organic
    compounds, are present.  In addition, nitrates and HNO3 may cause
    health effects and significant effects on ecosystems when deposited on
    the ground.

         The sum of NO and NO2 is generally referred to as NOx.  Once
    released into the air, NO is oxidized to NO2 by available oxidants
    (particularly ozone, O3).  This happens rapidly under some conditions
    in outdoor air; in indoor air, it is generally a much slower process.
    Nitrogen oxides are a controlling precursor of photochemical oxidant
    air pollution resulting in ozone and smog formation; interactions of
    nitrogen oxides (except N2O) with reactive organic compounds and
    sunlight form ozone in the troposphere and smog in urban areas.

         NO and NO2 may also undergo reactions to form a range of other
    oxides of nitrogen, both in indoor and outdoor air, including HNO2,
    HNO3, nitrogen trioxide (NO3), dinitrogen pentoxide (N2O5), PAN
    and other organic nitrates.  The complex range of gas-phase nitrogen
    oxides is referred to as NOy.  The partitioning of oxides of nitrogen
    among these compounds is strongly dependent on the concentrations of
    other oxidants and on the meteorological history of the air.

         HNO3 is formed from the reaction of OH- and NO2.  It is a
    major sink for active nitrogen and also a contributor to acidic
    deposition.  Potential physical and chemical sinks for HNO3 include
    wet and dry deposition, photolysis, reaction with OH radicals, and
    reaction with gaseous ammonia to form ammonium nitrate aerosol.

         PANs are formed from the combination of organic peroxy radicals
    with NO2.  PAN is the most abundant organic nitrate in the
    troposphere and can serve as a temporary reservoir for reactive
    nitrogen, which may be regionally transported.

         The NO3 radical, a short-lived NOy species that is formed in
    the troposphere primarily by the reaction of NO2 with O3, undergoes
    rapid photolysis in daylight or reaction with NO.  Appreciable
    concentrations are observed during the night.

         N2O5 is primarily a night-time constituent of ambient air as it
    is formed from the reaction of NO3 and NO2.  In ambient air, N2O5
    reacts heterogeneously with water to form HNO3, which in turn is
    deposited.

         N2O is ubiquitous because it is a product of natural biological
    processes in soil.  It is not known, however, to be involved in any
    reactions in the troposphere.  N2O participates in upper atmospheric
    reactions contributing to stratospheric ozone (O3) depletion and is
    also a relatively potent greenhouse gas that contributes to global
    warming.

    1.1.1  Atmospheric transport

         The transport and dispersion of the various nitrogenous
    species in the lower troposphere is dependent on both meteorological
    and chemical parameters.  Advection, diffusion and chemical
    transformations combine to dictate the atmospheric residence times. 
    In turn, atmospheric residence times help determine the geographic
    extent of transport of given species.  Surface emissions are dispersed
    vertically and horizontally through the atmosphere by turbulent mixing
    processes that are dependent to a large extent on the vertical
    temperature structure and wind speed.

         As the result of meteorological processes, NOx emitted in the
    early morning hours in an urban area typically disperses vertically
    and moves downwind as the day progresses.  On sunny summer days, most
    of the NOx will have been converted to HNO3 and PAN by sunset, with
    concomitant formation of ozone.  Much of the HNO3 is removed by
    deposition as the air mass is transported, but HNO3 and PAN carried
    in layers aloft (above the nighttime inversion layer but below a
    higher subsidence inversion) can potentially be transported long
    distances in oxidant-laden air masses.

    1.1.2  Measurement

         There are a number of methods available to measure airborne
    nitrogen-containing species.  This document briefly covers
    methodologies currently available or in general use for  in situ

    monitoring of airborne concentrations in both ambient and indoor
    environments.  The species considered are NO, NO2, NOx, total
    reactive odd nitrogen (NOy), PAN and other organic nitrates, HNO3,
    HNO2, N2O5, the nitrate radical, NO3-, and N2O.

         Measuring concentrations of nitrogen oxides is not trivial. 
    While a straightforward, widely available method exists for measuring
    NO (the chemiluminescent reaction with ozone), this is an exception
    for nitrogen oxides.  Chemiluminescence is also the most common
    technique used for NO2; NO2 is first reduced to NO.  Unfortunately,
    the catalyst typically used for the reduction is not specific, and has
    various conversion efficiencies for other oxidized nitrogen compounds. 
    For this reason, great care must be taken in interpreting the results
    of the common chemiluminescence analyser in terms of NO2, as the
    signal may include many other compounds.  Additional difficulties
    arise from nitrogen oxides that may partition between the gaseous and
    particulate phases both in the atmosphere and in the sampling
    procedure.

    1.1.3  Exposure

         Human and environmental exposure to nitrogen oxides varies
    greatly from indoors to outdoors, from cities to the countryside, and
    with time of day and season.  The concentrations of NO and NO2
    typically present outdoors in a range of urban situations are
    relatively well established. The concentrations encountered indoors
    depend on the specific details of the nature of combustion appliances,
    chimneys and ventilation.  When unvented combustion appliances are
    used for cooking or heating, indoor concentrations of nitrogen oxides
    typically greatly exceed those existing outside.  Recent research has
    shown in these circumstances that HNO2 can reach significant
    concentrations. One report showed that HNO2 can represent over 10% of
    the concentrations usually reported as NO2.

    1.2  Effects of atmospheric nitrogen species, particularly nitrogen
         oxides, on vegetation

         Most of earth's biodiversity is found in (semi-)natural
    ecosystems, both in aquatic and terrestrial habitats.  Nitrogen is the
    limiting nutrient for plant growth in many (semi-)natural ecosystems. 
    Most of the plant species from these habitats are adapted to nutrient-
    poor conditions, and can only compete successfully on soils with low
    nitrogen levels.

         Human activities, both industrial and agricultural, have greatly
    increased the amount of biologically available nitrogen compounds,
    thereby disturbing the natural nitrogen cycle.  Various forms of
    nitrogen pollute the air: mainly NO, NO2 and ammonia (NH3) as dry
    deposition; and nitrate (NO3-) and ammonium (NH4+) as wet
    deposition.  NHy refers to the sum of NH3 and NH4+.  Another

    contribution is from occult deposition (fog and clouds).  There are
    many more nitrogen-containing air pollutants (e.g., N2O5, PAN, N2O,
    amines), but these are neglected here, either because their
    contribution to the total nitrogen deposition is believed to be small,
    or because their concentrations are probably far below effect
    thresholds.

         Nitrogen-containing air pollutants can affect vegetation
    indirectly, via photochemical reaction products, or directly after
    being deposited on vegetation, soil or water surface.  The  indirect
    pathway is largely neglected here although it includes very relevant
    processes, and should be taken into account when evaluating the entire
    impact of nitrogen-containing air pollutants: NO2 is a precursor for
    tropospheric O3, which acts both as a phytotoxin and a greenhouse
    gas.

         The impacts of increased nitrogen deposition upon biological
    systems can be the result of direct uptake by foliage or uptake via
    the soil.  At the level of individual plants, the most relevant
    effects are injury to the tissue, changes in biomass production and
    increased susceptibility to secondary stress factors.  At the
    vegetation level, deposited nitrogen acts as a nutrient; this results
    in changes in competitive relationships between species and loss of
    biodiversity.  The critical loads for nitrogen depend on (i) the type
    of ecosystem; (ii) the land use and management in the past and
    present; and (iii) the abiotic conditions (especially those that
    influence the nitrification potential and immobilization rate in the
    soil).

         Adsorption on the outer surface of the leaves takes place and may
    damage wax layers of the cuticle, but the quantitative relevance for
    the field situation has not yet been proved.  Uptake of NOx and NH3
    is driven by the concentration gradient between atmosphere and
    mesophyll.  It generally, but not always, is directly determined by
    stomatal conductance and thus depends on factors influencing stomatal
    aperture.  There is increasing evidence that foliar uptake of nitrogen
    reduces the uptake of nitrogen by the roots.  Uptake and exchange of
    ions through the leaf surface is a relatively slow process, and thus
    is only relevant if the surface remains wet for longer periods.

         NO is only slightly soluble in water, but the presence of other
    substances can alter the solubility.  NO2 has a higher solubility,
    while that of NH3 is much higher.  NO2- (the primary reaction
    product of NOx), NH3 and NH4+ are all highly phytotoxic, and could
    well be the cause of adverse effects of nitrogen-containing air
    pollutants.  The free radical *N=O may play a role in the phytotoxicity
    of NO.

         More-than-additive effects (synergism) have been found in nearly
    all studies concerning SO2 plus NO2.  With other NO2 mixtures (NO,
    O3 and CO2), interactive effects are the exception rather than the
    rule.

         When climatic conditions and supply of other nutrients allow
    biomass production, both NOx and NHy result in growth stimulation at
    low concentrations and growth reduction at higher concentrations. 
    However, the exposure level at which growth stimulation turns into
    growth inhibition is much lower for NOx than for NHy.

         Evidence exists that plants are more sensitive at low light
    intensity (e.g., at night and in winter) and at low temperatures (just
    above 0°C).  NOx and NHy can increase the sensitivity of plants to
    frost, drought, wind and insect damage.

         An interaction exists between soil chemistry and sensitivity of
    vegetation to nitrogen deposition; this is related to pH and nitrogen
    availability.

         The relative contribution of NO and NO2 to the NOx effect on
    plants is unclear.  The vast majority of information is on effects of
    NO2 but available information on NO suggests that NO and NO2 have
    comparable phytotoxic effects.

         Air quality guidelines refer to thresholds for adverse effects. 
    Two different types of effect thresholds exist: critical levels (CLEs)
    and critical loads (CLOs).  The critical level is defined as the
    concentration in the atmosphere above which direct adverse effects on
    receptors, such as plants, ecosystems or materials, may occur
    according to present knowledge.  The critical load is defined as a
    quantitative estimate of an exposure (deposition) to one or more
    pollutants below which significant harmful effects on specified
    sensitive elements of the environment do not occur according to
    present knowledge.

         According to current practice, critical levels have been derived
    from assessment of the lowest exposure concentrations causing adverse
    effects on physiology or growth of plants (biochemical effects were
    excluded), using a graphical method.

         To include the impact of NO, a critical level for NOx is
    proposed instead of one for NO2; for this purpose it has been assumed
    that NO and NO2 act in an additive manner.  A strong case can be made
    for the provision of critical levels for short-term exposure. However,
    currently there are insufficient data to provide these with sufficient
    confidence.  Current evidence suggests a critical level of about
    75 µg/m3 for NOx as a 24-h mean.

         The critical level for NOx (NO and NO2 added in ppb and
    expressed as NO2 in µg/m3) is considered to be 30 µg/m3 as an
    annual mean.

         Information on organisms in the environment is almost exclusively
    restricted to plants, with minimum data on soil fauna.  This
    evaluation and guidance values are, therefore, expressed in terms of
    nitrogen species effects on vegetation.  However, it is expected that
    plants will form the most sensitive component of natural systems and
    that the effect on biodiversity of plant communities is a sensitive
    indicator of effects on the whole ecosystem.

         Critical loads are derived from empirical data and steady-state
    soil models.  Estimated critical loads for total nitrogen deposition
    in a variety of natural aquatic and terrestrial ecosystems are given. 
    Possible differential effects of deposited nitrogen species (NOx and
    NHy) are insufficiently known to differentiate between nitrogen
    species for critical load estimation.

         The great majority of ecosystems for which there is sufficient
    information to estimate critical loads are from temperate climates. 
    The few arctic and montane ecosystems included, which might be
    expected to be representative of higher latitudes, have the least
    reliable basis.  There is no information on tropical ecosystems and
    little on estuarine or marine ecosystems in any climatic zone. 
    Nutrient-poor tropical ecosystems such as rain forests and mangrove
    swamps are likely to be adversely affected by nitrogen deposition. 
    The lack of both deposition data and effect thresholds make it
    impossible to make risk assessments for these climatic regions.

         The most sensitive ecosystems (ombrotrophic bogs, shallow soft-
    water lakes and arctic and alpine heaths) for which effects thresholds
    can be estimated show critical loads of 5-10 kg N.ha-1.year-1 based
    on decreased biological diversity in plant communities.  A more
    average value for the limited range of ecosystems studied is 15-20 kg
    N.ha-1.year-1, which applies to forest trees.

         The atmospheric chemistry of nitrogen oxides includes the
    capacity for ozone generation in the troposphere, ozone depletion in
    the stratosphere, and contribution to global warming as greenhouse
    gases.  Nitrogen oxides and ammonia contribute to soil acidification
    (along with sulfur oxides) and thereby to increased bioavailability of
    aluminium.

         The phytotoxic effects of nitrogen oxides on plants have little
    direct relevance to crop plants when concentrations marginally exceed
    the critical level. However, the role of NOx in the generation of
    ozone and other phytotoxic substances, e.g., organic nitrates leads to

    crop loss. Nitrogen deposited on growing crops will represent a very
    small increase in total available nitrogen compared to that added as
    fertilizer.

    1.3  Health effects of exposures to nitrogen dioxide

         A large number of studies designed to evaluate the health effects
    of NOx have been conducted.  Of the NOx compounds, NO2 has been
    most studied.  The discussion in this section focuses on NO2, NO,
    HNO2 and HNO3, while nitrates are mentioned briefly.

    1.3.1  Studies of the effects of nitrogen compounds on experimental
           animals

         Extrapolating animal data to humans has both qualitative and
    quantitative components.  As summarized below, NO2 causes a
    constellation of effects in several animal species; most notably,
    effects on host defence against infectious pulmonary disease, lung
    metabolism/biochemistry, lung function and lung structure.  Because of
    basic physiological, metabolic and structural similarities in all
    mammals (laboratory animals and humans), the commonality of the
    observations in several animal species leads to a reasonable
    conclusion that NO2 could cause similar types of effects in humans. 
    However, because of the differences between mammalian species, exactly
    what exposures would actually cause these effects in humans is not yet
    known.  That is the topic of quantitative extrapolation.  Limited
    modelling research on the dosimetric aspect (i.e., the dose to the
    target tissue/cell that actually causes toxicity) of quantitative
    extrapolation suggests that the distribution of the deposition of NO2
    within the respiratory tract of animals and humans is similar,
    without yet providing adequate values to use for animal-to-human
    extrapolation.  Unfortunately, very little information is available on
    the other key aspect of extrapolation, species sensitivity (i.e., the
    response of the tissues of different species to a given dose).  Thus,
    from currently available animal studies, we know which human health
    effects NO2 may cause. We are unable to assert with great confidence
    the effects that are  actually caused by a given inhaled dose of
    NO2.

         With the above issues in mind, the animal toxicology database
    for NO2 is summarized below according to major classes of effects
    and topics of special interest.  Although it is clear that the
    effects of NO2 exposure extend beyond the confines of the lung, the
    interpretation of these systemic effects relative to potential human
    risk is not clear.  Therefore they are not summarized further here,
    but are discussed in later chapters.  Although interactions of NO2
    and other co-occurring pollutants, such as O3 and sulfuric acid
    (H2SO4), can be quite important, especially if synergism occurs, the
    database does not yet allow conclusions that enable assessment of
    real-world potential interactions.

    1.3.1.1  Biochemical and cellular mechanisms of action of nitrogen
             oxides

         NO2 acts as a strong oxidant.  Unsaturated lipids are readily
    oxidized with peroxides as the dominant product.  Both ascorbic acid
    (vitamin C) and alpha-tocopherol (vitamin E) inhibit the peroxidation
    of unsaturated lipids.  When ascorbic acid is sealed within bilayer
    liposomes, NO2 rapidly oxidizes the sealed ascorbic acid.  The
    protective effects of alpha-tocopherol and ascorbic acid in animals
    and humans are due to the inhibition of NO2 oxidation.  NO2 also
    oxidizes membrane proteins.  The oxidation of either membrane lipids
    or proteins results in the loss of cell permeability control.  The
    lungs of NO2-exposed humans and experimental animals have larger
    amounts of protein within the lumen.  The recruitment of inflammatory
    cells and the changes in the lung are due to these events.

         The oxidant properties of NO2 also induce the peroxide
    detoxification pathway of glutathione peroxidase, glutathione
    reductase and glucose-6-phosphate dehydrogenase. Following NO2
    exposure the increase in the peroxide detoxification pathway in
    animals follows an exposure-response relationship.

         The mechanism of action of NO is less clear.  NO is readily
    oxidized to NO2 and peroxidation then occurs.  Because of the
    concurrent exposure to some NO2 in NO exposures, it is difficult to
    discriminate NO effects from NO2.  NO functions as an intracellular
    second messenger modulating a wide variety of essential enzymes, and
    it inhibits its own production (e.g., negative feedback).  NO
    activates guanylate cyclase which in turn increases intracellular cGMP
    levels.  A possible mechanism of action of nitrates may be through the
    release of histamine from mast cell granules.  Acidic nitrogenous air
    pollutants, particularly HNO3, may act by alteration of intracellular
    pH.

         PAN decomposes in water, generating hydrogen peroxide.  Little is
    known of the mechanism of action, but oxidative stress is likely for
    PAN and its congeners.

         Inorganic nitrates may act through alterations in intracellular
    pH.  Nitrate ion is transported into alveolar type 2 cells acidifying
    the cell.  Nitrate also mobilizes histamine from mast cells.  HNO2
    could also act to alter intracellular pH, but this mechanism is
    unclear.

         The mechanisms of action of the other nitrogen oxides are
    unknown.

         Acute exposure to NO2 at a concentration of 750 µg/m3 (0.4 ppm)
    can result in lipid peroxidation.  NO2 can oxidize polyunsaturated

    fatty acids in cell membranes as well as functional groups of proteins
    (either soluble proteins in the cell, such as enzymes, or structural
    proteins, such as components of cell membranes).  Such oxidation
    reactions (mediated by free radicals) are a mechanism by which NO2
    exerts direct toxicity on lung cells.  This mechanism of action is
    supported by animal studies showing the importance of lung antioxidant
    defences, both endogenous (e.g., maintenance of lung glutathione
    levels) and exogenous (e.g., dietary vitamins C and E), in protecting
    against the effects of NO2.  Many studies have suggested that various
    enzymes in the lung, including glutathione peroxidase, superoxide
    dismutase and catalase, may also serve to defend the lung against
    oxidant attack.

    1.3.1.2  Effects on host defence

         Although the primary function of the respiratory tract is to
    ensure an efficient exchange of gases, this organ system also provides
    the body with a first line of defence against inhaled viable and non-
    viable airborne agents.  An extensive database clearly shows that
    exposure to NO2 can result in the dysfunction of these host defences,
    increasing susceptibility to infectious respiratory disease.  The
    host-defence parameters affected by NO2 include the functional and
    biochemical activity of cells in lungs, alveolar macrophages (AMs),
    immunological competence, susceptibility to experimentally induced
    respiratory infections, and the rate of mucociliary clearance.

         Alveolar macrophages are affected by NO2.  These cells
    are responsible for maintaining the sterility of the pulmonary
    region, clearing particles from this region, and participating in
    immunological functions.  Functional changes that have been reported
    include the following: the suppression of phagocytic ability and
    stimulation of lung clearance at 560 µg/m3 (0.3 ppm) 2 h/day for
    13 days; a decrease in bactericidal activity at 4320 µg/m3 (2.3 ppm)
    for 17 h; and a decreased response to migration inhibition factor at
    3760 µg/m3 (2.0 ppm) 8 h/day, 5 days/week for 6 months.  The
    morphological appearance of these defence cells changes after chronic
    exposure to NO2.

         The importance of host defences becomes evident when animals have
    to cope with laboratory-induced pulmonary infections.  Animals exposed
    to NO2 succumb to bacterial or viral infection in a concentration-
    dependent manner.  Mortality also increases with increased NO2
    concentration or duration of exposure.  After acute exposure,
    effects are observed at concentrations as low as 3760 µg/m3 (2 ppm). 
    Exposure to concentrations as low as 940 µg/m3 (0.5 ppm) will cause
    effects in the infectivity model after 6 months.

         Both humoral and cell-mediated defence systems are changed by
    NO2 exposure.  In the cases in which the immune system has been
    investigated, effects have been observed after short-term exposure to

    concentrations > 9400 µg/m3 (5 ppm).  The effects are complex
    since the direction of the change (i.e., increase or decrease) is
    dependent upon NO2 concentration and the length of exposure.

    1.3.1.3  Effects of chronic exposure on the development of chronic
             lung disease

         Humans are chronically exposed to NO2.  Therefore, such
    exposures in animals have been studied rather extensively, typically
    using morphological and/or morphometric methods.  This research has
    generally shown that a variety of pulmonary structural and correlated
    functional alterations occur.  Some of these changes may be reversible
    when exposure ceases.

         Pulmonary function may be altered following chronic NO2 exposure
    of experimental animals. Impaired gas exchange occurred following
    exposure to 7520 µg/m3 (4.0 ppm) NO2 for four months and this was
    reflected in decreased arterial O2 tension, impaired physical
    performance and increased anaerobic metabolism.

         Although NO2 produces morphological changes in the respiratory
    tract, the database is sometimes confusing due to quantitative and
    qualitative variability in responsiveness between, and even within,
    species.  The rat, the most commonly used experimental animal in
    morphological assessments of exposure, appears to be relatively
    resistant to NO2.  Short-term exposures to concentrations of
    9400 µg/m3 (5.0 ppm) or less generally have little effect in the
    rat, where similar exposures in the guinea-pig may result in some
    centriacinar epithelial damage.

         Longer-term exposures result in lesions in some species with
    concentrations as low as 560 to 940 µg/m3 (0.3 to 0.5 ppm).  These
    are characterized by epithelial remodelling similar to that described
    above, but with the involvement of more proximal airways and
    thickening of the interstitium. Many of these changes, however, will
    resolve even with continued exposure, and long-term exposures to
    levels above about 3760 µg/m3 (2.0 ppm) are required for more
    extensive and permanent changes in the lungs.  Some effects are
    relatively persistent (e.g., bronchiolitis), whereas others tend to be
    reversible and limited even with continued exposure.  In any case, it
    seems that for either short- or long-term exposure, the response is
    more dependent upon concentration than duration of exposure.

         There is substantial evidence that long-term exposure of several
    species of laboratory animals to high concentrations of NO2 results
    in morphological lung lesions.  Destruction of alveolar walls, an
    essential additional criterion for human emphysema, has been reliably
    reported in lungs from animals in a limited number of studies.  The

    lowest NO2 concentration for the shortest exposure duration that will
    result in emphysematous lung lesions cannot be determined from these
    published studies.

    1.3.1.4  Potential carcinogenic or co-carcinogenic effects

         NO2 has been shown to be mutagenic in  Salmonella bacteria, but
    was not mutagenic in one study with a mammalian cell culture.  Other
    studies using cell cultures have demonstrated sister chromatid
    exchanges (SCE) and DNA single strand breaks.  No genotoxic effects
    have been demonstrated  in vivo concerning lymphocytes, spermatocytes
    or bone marrow cells, but two inhalation studies with high
    concentrations (50 760 and 56 400 µg/m3, 27 and 30 ppm) for 3 h and
    16 h, respectively, have demonstrated such effects in lung cells.

         Literature searches revealed no published reports of NO2 studies
    using classical whole-animal chronic bioassays for carcinogenesis. 
    Research with mice having spontaneously high tumour rates was
    equivocal.  In one study, NO2 at 18 800 µg/m3 (10 ppm) slightly
    enhanced the incidence of lung adenomas in a sensitive strain of mice
    (A/J).  Although several co-carcinogenesis investigations have been
    undertaken, conclusions are precluded because of problems with
    methodology and interpretation. Reports on whether NO2 facilitates
    the metastasis of tumours to the lung are also inadequate to form
    conclusions.  Other investigations have centred on whether NO2 could
    produce nitrates and nitrites that, by reacting with amines in the
    body, could produce nitrosamines.  A few studies suggest that
    nitrosamines are formed in animals treated with high doses of amines
    and exposed to NO2, but other studies have indicated that nitrosamine
    formation is unlikely.

    1.3.1.5  Age susceptibility

         Investigations into age dependency are inadequate and results so
    far are equivocal.

    1.3.1.6  Influence of exposure patterns

         Several animal toxicological studies have elucidated the
    relationships between concentration (C) and duration (T) of exposure,
    indicating that the relationship is complex.  Most of this research
    has used the infectivity model.  Early C × T studies demonstrated that
    concentration had more impact on mortality than did duration of
    exposure.  An evaluation of the toxicity of NO2 exposures cannot be
    delineated by C × T relationships.

    1.3.2  Controlled human exposure studies on nitrogen oxides

         Human responses to a variety of oxidized nitrogen compounds have
    been evaluated.  By far, the largest database and the one most

    suitable for risk assessment is that available for controlled
    exposures to NO2.  The database on human responses to NO, HNO3
    vapour, HNO2 vapour and inorganic nitrate aerosols is not as
    extensive. A number of sensitive or potentially sensitive subgroups
    have been examined, including adolescent and adult asthmatics, older
    adults, and patients with chronic obstructive pulmonary disease (COPD)
    and pulmonary hypertension.  Exercise during exposure increases the
    total uptake and alters the distribution of the deposited inhaled
    material within the lung. The relative proportion of NO2 deposited in
    the lower respiratory tract is also increased by exercise.  This may
    increase the effects of the above compounds in people who exercise
    during exposure.

         As is typical with human biological response to inhaled particles
    and gases, there is variability in the biological response to NO2. 
    Healthy individuals tend to be less responsive to the effects of NO2
    than  individuals with lung disease.  Asthmatics are clearly the most
    responsive group to NO2 that has been studied to date.  Individuals
    with COPD may be more responsive than healthy individuals, but they
    have limited capacity to respond to NO2 and thus quantitative
    differences between COPD patients and others are difficult to assess. 
    Sufficient information is not available at present to evaluate whether
    age and sex play a role in the response to NO2.

         Healthy subjects can detect the odour of NO2, in some cases at
    concentrations below 188 µg/m3 (0.1 ppm).  Generally, NO2 exposure
    did not increase respiratory symptoms in any of the subject groups
    tested.

         NO2 causes decrements in lung function, particularly increased
    airway resistance in resting healthy subjects at 2-h concentrations as
    low as 4700 µg/m3 (approx.2.5 ppm).  Available data are insufficient
    to determine the nature of the concentration-response relationship.

         Exposure to NO2 results in increased airway responsiveness to
    bronchoconstrictive agents in exercising healthy, non-smoking subjects
    exposed to concentrations as low as 2800 µg/m3 (approx.1.5 ppm) for
    1 h or longer.

         Exposure of asthmatics to NO2 causes, in some subjects,
    increased airway responsiveness to a variety of provocative mediators,
    including cholinergic and histaminergic chemicals, SO2 and cold air. 
    The presence of these responses appears to be influenced by the
    exposure protocol, particularly whether or not the exposure includes
    exercise.  These responses may begin at concentrations as low as
    380 µg/m3 (0.2 ppm).  A meta-analysis suggests that effects may occur
    at even lower concentrations.  However, an unambiguous concentration-
    response relationship is observed between 350 to 1150 µg/m3
    (approx.0.2 to 0.6 ppm).

         The implications of this overall trend are unclear, but increased
    airway responsiveness could potentially lead to increased response to
    aeroallergens or temporary exacerbation of asthma, possibly leading to
    increased medication usage or even increased hospital admissions.

         Modest increases in airway resistance may occur in COPD patients
    from brief exposure (15-60 min) to concentrations of NO2 as low as
    2800 µg/m3 (approx.1.5 ppm), and decrements in spirometric measures
    of lung function (3 to 8% change in FEV1 (forced expiratory volume
    in 1 second)) may also be observed with longer exposures (3 h) to
    concentrations as low as 600 µg/m3 (approx.0.3 ppm).

         Exposure to NO2 at levels above 2800 µg/m3 (approx.1.5 ppm) may
    alter the numbers and types of inflammatory cells in the distal
    airways or alveoli.  NO2 may alter the functioning of cells within
    the lungs and production of mediators that may be important in lung
    host defences.  The constellation of changes in host defences,
    alterations in lung cells and their activities, and changes in
    biochemical mediators is consistent with the epidemiological findings
    of increased host susceptibility associated with NO2 exposure.

         In studies on mixtures of NO2 with other pollutants, NO2 has
    not been observed to increase responses to other co-occurring
    pollutant(s) beyond that which would be observed for the other
    pollutant(s) alone.  A notable exception is the observation that
    pre-exposure to NO2 enhanced the ozone-induced change in airway
    responsiveness in healthy exercising subjects during a subsequent
    ozone exposure.  This observation suggests the possibility of delayed
    or persistent responses to NO2.

         Within an NO2 concentration range that may be of interest with
    regard to risk evaluation (i.e., 100-600 µg/m3), the characteristics
    of the concentration-response relationship for acute changes in lung
    function, airway responsiveness to bronchoconstricting agents or
    symptoms cannot be determined from the available data.

         On the basis of an effect at 400 µg/m3 and the possibility of
    effects at lower levels, based on a meta analysis, a one-hour average
    daily maximum NO2 concentration of 200 µg/m3 (approx.0.11 ppm) is
    recommended as a short-term guideline.

         NO is acknowledged as an important endogenous second messenger
    within several organ systems.  Inhaled NO concentrations above
    6000 µg/m3 (approx.5 ppm) can cause vasodilation in the pulmonary
    circulation without affecting the systemic circulation.  The lowest
    effective concentration has not been established.  Information on
    pulmonary function and lung host defences consequent to NO exposure
    are too limited for any conclusions to be drawn at this time.

    Relatively high concentrations (> 40 000 µg/m3) have been used in
    clinical applications for brief periods (< 1 h) without reported
    adverse reactions.

         Nitric acid levels in the range of 250-500 µg/m3 (97-194 ppb)
    may cause some pulmonary function responses in adolescent asthmatics,
    but not in healthy adults.

         Limited information on HNO2 suggests that it may cause eye
    inflammation at 760 µg/m3 (0.40 ppm).  There are currently no
    published data on human pulmonary responses to HNO2.

         Limited data on inorganic nitrates suggest that there are no lung
    function effects of nitrate aerosols at concentrations of 7000 µg/m3
    or less.

    1.3.3  Epidemiology studies on nitrogen dioxide

         Epidemiological studies on the health effects of nitrogen oxides
    have mainly focused on NO2.  Many indoor and outdoor epidemiological
    studies designed to evaluate the health effects of NO2 have been
    conducted.  Two health outcome measurements of NO2 exposure are
    generally considered: lung function measurements and respiratory
    symptoms and diseases.

         The evidence from individual studies of the effect of NO2 on
    lower respiratory symptoms and disease in school-aged children is
    somewhat mixed.  The consistency of these studies was examined and
    the evidence synthesized in a combined quantitative analysis
    (meta-analysis) of the subject studies.  Most of the indoor studies
    showed increased lower respiratory morbidity in children associated
    with long-term exposure to NO2.  Mean weekly NO2 concentrations
    in bedrooms in studies reporting NO2 levels were predominantly
    between 15 and 122 µg/m3 (0.008 and 0.065 ppm).  Combining the
    indoor studies as if the end-points were similar gives an estimated
    odds ratio of 1.2 (95% confidence limits of 1.1 and 1.3) for the effect
    per 28.3 µg/m3 (0.015 ppm) increase of NO2 on lower respiratory
    morbidity.  This suggests that, subject to assumptions made for the
    combined analysis, an increase of about 20% in the odds of lower
    respiratory symptoms and disease corresponds to each increase of
    28.3 µg/m3 (0.015 ppm) in estimated 2-week average NO2
    exposure.  Thus, the combined evidence is supportive for the effects
    of estimated exposure to NO2 on lower respiratory symptoms and
    disease in children aged 5 to 12 years.
    
         In individual indoor studies of infants 2 years of age or younger,
    no consistent relationship was found between estimates of NO2
    exposure and the prevalence of respiratory symptoms and disease.  Based

    on a meta-analysis of these indoor infant studies, subject to the
    assumptions made for the meta-analysis, the combined odds ratio for the
    increase in respiratory disease per increase of 28.2 µg/m3 (0.015 ppm)
    NO2 was 1.09 with a 95% confidence interval of 0.95 to 1.26, where
    mean weekly NO2 concentrations in bedrooms were predominantly between
    9.4 and 94 µg/m3 (0.005 and 0.050 ppm) in studies reporting levels.
    The increase in risk was very small and was not reported consistently
    by all studies.  We cannot conclude that the evidence suggests an effect
    in infants comparable to that seen in older children.  The reasons for
    these age-related differences are not clear.

         The measured NO2 studies gave a higher estimated odds ratio than
    the surrogate estimates, which is consistent with a measurement error
    effect.  The effect of having adjusted for covariates such as
    socioeconomic status, smoking and sex was that those studies that
    adjusted for a particular covariate found larger odds ratios than
    those that did not.

         Although many of the epidemiological studies that involved
    measured NO2 levels used measurements over only 1 or 2 weeks, these
    levels were used to characterize children's exposures over a much
    longer period.  The standard respiratory symptom questionnaire used by
    most of these studies summarizes information on health status over an
    entire year.  The 28.2 µg/m3 (0.015 ppm) difference in NO2 levels
    used in the meta-analyses relates to a difference in the household
    annual average exposure between gas and electric cooking stoves.
    Some studies measured NO2 levels only in the winter and may have
    overestimated annual average exposures.  This would tend to have
    underestimated the health effect of a 28.2 µg/m3 (0.015 ppm)
    difference in the annual NO2 exposure.  A study based on a household
    annual average exposure measured in both the winter and summer found a
    stronger health effect than many of the other studies.  The true
    biologically relevant exposure period is unknown, but these exposures
    extended over a lengthy period up to the entire lifetime of the child.

         The association between outdoor NO2 and respiratory health is
    not clear from current research.  There is some evidence that the
    duration of respiratory illness may be increased at higher ambient
    NO2 levels.  A major difficulty in the analysis of outdoor studies is
    distinguishing possible effects of NO2 from those of other associated
    pollutants.

         Several uncertainties need to be considered in interpreting the
    above studies and meta-analysis.  Error in measuring exposure is
    potentially one of the most important methodological problems in
    epidemiological studies of NO2.  Although there is evidence that
    symptoms are associated with indicators of NO2 exposure, the quality
    of these exposure estimates may be inadequate to determine a
    quantitative relationship between exposure and symptoms.  Most of the
    studies that measured NO2 exposure did so only for periods of 1 to

    2 weeks and reported the values as averages.  Few of the studies
    attempted to relate the observed effects to the pattern of exposure
    (e.g., transient NO2 peaks). Furthermore, measured NO2 concentration
    may not be the biologically relevant dose; estimating actual exposure
    requires knowledge of pollutant species, levels and related human
    activity patterns.  However, only very limited activity and aerometric
    data are available that examine such factors.  The extrapolation to
    possible patterns of ambient exposure is difficult.  In addition,
    although the level of similarity and common elements between the
    outcome measures in the NO2 studies provide some confidence in their
    use in the quantitative analysis, the symptoms and illnesses combined
    are to some extent different and could indeed reflect different
    underlying processes.  Thus, caution is necessary in interpreting the
    meta-analysis results.

         Other epidemiological studies have attempted to relate some
    measure of indoor and/or outdoor NO2 exposure to changes in pulmonary
    function.  These changes were marginally significant.  Most studies
    did not find any effects, which is consistent with controlled human
    exposure study data.  However, there is insufficient epidemiological
    evidence to draw any conclusions about the long- or short-term effects
    of NO2 on pulmonary function.

         On the basis of a background level of 15 µg/m3 (0.008 ppm) and
    the fact that significant adverse health effects occur with an
    additional level of 28.2 µg/m3 (0.015 ppm) or more, an annual
    guideline value of 40 µg/m3 (0.023 ppm) is proposed.  This value will
    avoid the most severe exposures.  The fact that a no-effect level for
    subchronic or chronic NO2 exposure concentrations has not yet been
    determined should be emphasized.

    1.3.4  Health-based guidance values for nitrogen dioxide

         On the basis of human controlled exposure studies, the
    recommended short-term guidance value is for a one-hour average NO2
    daily maximum concentration of 200 µg/m3 (0.11 ppm).  The recommended
    long-term guidance value, based on epidemiological studies of
    increased risk of respiratory illness in children, is 40 µg/m3
    (0.023 ppm) annual average.

    2.  PHYSICAL AND CHEMICAL PROPERTIES, AIR SAMPLING AND ANALYSIS,
        TRANSFORMATIONS AND TRANSPORT IN THE ATMOSPHERE

    2.1  Introduction

         Nitrogen oxides are produced by combustion processes and are
    emitted to the air mainly as NO together with some NO2.  Natural
    biological processes and lightning also emit NO and N2O.  In the
    atmosphere nitrogen oxides undergo complex chemical and photochemical
    reactions; NO is oxidized to NO2 and other products and eventually to
    HNO3 and nitrates.  Nitrogenous species are removed from the air to
    the ground by wet and dry deposition processes.  Oxidized nitrogen
    compounds can have impacts on human health and the environment, and
    are important to the formation of photochemical smog and tropospheric
    ozone.

         In this chapter the properties of nitrogen compounds are briefly
    described and techniques for their sampling and analysis outlined. 
    Atmospheric chemical reactions that cause the oxidation of NO to NO2
    and the production of ozone, organic nitrates and HNO3 are described.
    The differences between night-time and day-time chemistry and the
    composition of the atmosphere are discussed.  The nature of the
    nitrogen species and their chemical reactions in urban regions, in
    chimney plumes such as those from power stations, in air advected away
    from urban regions and in rural and remote areas are described.  The
    role of nitrogen oxides in photochemical smog production and the
    effects of nitrous oxide on stratospheric ozone are briefly discussed.

    2.1.1  The nomenclature and measurement of atmospheric nitrogen
           species

         There are several methods available for determining nitrogen
    species, but many of these techniques are nonspecific.

         To denote various mixtures of nitrogen species, the terms NOx,
    NOy and NOz are often employed.  It is customary to refer to the sum
    of NO and NO2 emitted from a source as NOx, the unit of measure for
    NOx being the NO2 mass equivalent of the NO plus NO2.

         The term NOy is frequently used to denote the sum of the gas
    phase oxidized nitrogen species (except N2O) and NOz to denote the
    sum of NOy plus the oxidized nitrogen present as particulate matter. 
    Measurement of NOz requires a combination of particulate and gas
    phase sampling and analysis.

         A confusion arises because one of the most commonly used methods
    for determining NO2 in ambient air (thermal conversion of NO2 to NO
    and measurement of the resultant NO by chemiluminescent reaction with
    O3) is nonspecific and responds to several gaseous species in
    addition to NO2.  These include organic nitrogen compounds and,

    depending on the converter, HNO3, although HNO3 can be readily lost
    to the sampling system.  Therefore, depending on the composition of
    the air being sampled, the results from this type of instrument can be
    representative of NOy rather than NOx (or NO2) concentrations. 
    This technique is used in most routine determinations of ambient NOx
    and NO2 concentrations but the discrepancy between these values and
    true NOx and NO2 can be considerable for air in which the pollutant
    emissions have undergone substantial exposure to sunlight.

         Nitrous oxide is ubiquitous in the atmosphere because it is a
    product of biological processes in soil as well as anthropogenic
    activities.  It is not involved to any appreciable extent in chemical
    reactions in the lower atmosphere, but it is an active "greenhouse"
    gas.  In the stratosphere N2O forms NO by reaction with excited
    oxygen atoms, and this NO then acts to deplete the stratospheric O3
    concentration.

         Although NO3, dinitrogen trioxide (N2O3), dinitrogen tetroxide
    (N2O4), and N2O5 may play a role in atmospheric chemical reactions
    leading to the transformation, transport, and ultimate removal of
    nitrogen compounds from ambient air, they are present in very low
    concentrations, even in polluted environments.

         NH3 is generated during decomposition of nitrogenous matter in
    natural ecosystems and may be locally produced in high concentrations
    by human activities such as intensive animal husbandry and feedlots. 
    Under suitable conditions NH3 can react with oxidized nitrogen
    species to form ammonium nitrate aerosol.

    2.2  Nitrogen species and their physical and chemical properties

         There are seven oxides of nitrogen that may be present in ambient
    air, namely: NO, NO2, N2O, NO3, N2O3, N2O4 and N2O5.  In
    addition these can be present as HNO2, HNO3 and various organic
    nitrogen species, such as PAN, other organic nitrates and particles
    containing oxidized nitrogen compounds (particularly adsorbed nitric
    acid).  Of these species, NO and NO2 are the ones most often measured
    and are present in the greatest concentrations in urban and industrial
    air.

         The chemical and physical properties of individual nitrogen
    species are given below and are summarized in Table 1.

        Table 1.  Some physical and thermodynamic properties of oxides of nitrogen and other nitrogen compoundsa
                                                                                                                                              

    Oxide               Relative         Melting point    Boiling point     Solubility in water            Thermodynamic functions
                        molecular        (°C)b,c,d        (°C)b,c           at 0°C (cm3 per 100 g)b        (Ideal gas, 1 atm, 25°C)
                        mass (g/mol)                                                                                                          
                                                                                                           Enthalpy of      Entropy
                                                                                                           formation        (cal/mol-deg)
                                                                                                           (kcal/mol)
                                                                                                                                              

    NO                  30.01            -163.6           -151.8            7.34                              21.58            50.35

    NO2                 46.01            -11.2            21.2              Reacts with H2O forming            7.91            57.34
                                                                            HNO2 and HNO3

    N2O                 44.01            -90.8            -88.5             130.52                            19.61            52.55

    N2O3                76.01            -102             47                Reacts with H2O forming           19.80            73.91
                                                          (decomposes)      HNO2

    N2O4                92.02            -11.3            21.2              Reacts with H2O forming            2.17            72.72
                                                                            HNO2 and HNO3

    N2O5                108.01           30               3.24              Reacts with H2O forming            2.7             82.8
                                                          (decomposes)      HNO2

    HNO2                47.01            -                -                 -                                  -                -

    HNO3                63.01            -42              83                                                 -32.1             63.7
                                                                                                                                              

    Table 1.  (Con't)
                                                                                                                                              

    Oxide               Relative         Melting point    Boiling point     Solubility in water            Thermodynamic functions
                        molecular        (°C)b,c,d        (°C)b,c           at 0°C (cm3 per 100 g)b        (Ideal gas, 1 atm, 25°C)
                        mass (g/mol)                                                                                                          
                                                                                                           Enthalpy of      Entropy
                                                                                                           formation        (cal/mol-deg)
                                                                                                           (kcal/mol)
                                                                                                                                              

    PAN                 121.06           -                -                 -                                  -                -
    (CH3COOONO2)

    NH4NO3              80.04            169.6            210 at            118.3 g/100 cm3                  -87.37            36.11
                                                          11 torr           H2O at 0°C
                                                                                                                                              

    a  Adopted from: US EPA (1993)
    b  Matheson Gas Data Book (Matheson Company, 1966)
    c  Handbook of Chemistry and Physics (Weast et al., 1986)
    d  At 0°C and 1 atm pressure
        2.2.1  Nitrogen oxides

    2.2.1.1  Nitric oxide

         NO is a colourless, odourless gas that is only slightly soluble
    in water.  It is a by-product of combustion processes, arising from
    (i) high temperature oxidation of molecular nitrogen from the
    combustion air, and (ii) from oxidation of nitrogen present in certain
    fuels such as coal and heavy oil.

    2.2.1.2  Nitrogen dioxide

         NO2 is a reddish-orange-brown gas with a characteristic pungent
    odour.  The boiling point is 21.1°C, but the low partial pressure of
    NO2 in the atmosphere prevents condensation.  NO2 is corrosive and
    highly oxidizing.  About 5 to 10% by volume of the total emissions of
    NOx from combustion sources is usually in the form of NO2, although
    substantial variations from one source type to another have been
    observed.

         In the atmosphere, photochemical reactions involving ozone
    and organic compounds convert NO to NO2.  NO2 is an efficient
    absorber of light over a broad range of ultraviolet (UV) and visible
    wavelengths.  Because of its brown colour, NO2 can contribute to
    discoloration and reduced visibility of polluted air.  Photolysis of
    NO2 by sunlight produces NO and an oxygen atom, which usually adds to
    an oxygen molecule to produce ozone.

    2.2.1.3  Nitrous oxide

         N2O is a colourless gas with a slight odour at high
    concentrations.  It is emitted to the atmosphere as a trace component
    from some combustion sources and from the consumption of nitrate by
    an ubiquitous group of denitrification bacteria that use nitrate as
    their terminal electron acceptor in the absence of oxygen (Delwiche,
    1970; Brezonik, 1972; Keeney, 1973; Focht & Verstraete, 1977).  At
    atmospheric concentrations N2O has no significant physiological
    effects in humans, although at higher concentrations it is employed as
    an anaesthetic.

         N2O does not play a significant role in atmospheric reactions in
    the lower troposphere.  In the stratosphere it reacts with singlet
    oxygen to produce NO, which participates in O3 decomposition in
    the stratosphere.  These reactions are of concern because of the
    possibility that increasing N2O concentrations resulting from fossil
    fuel use, and also from denitrification of excess fertilizer, may
    contribute to a decrease in stratospheric O3 (Council for
    Agricultural Science and Technology, 1976; Crutzen, 1976) with
    consequent potential for adverse impacts on ecosystems and human

    health.  Also of concern is the fact that N2O absorbs long-wave
    radiation, and therefore serves as a radiatively important greenhouse
    gas that may contribute to global warming.

    2.2.1.4  Other nitrogen oxides

         Other nitrogen oxides can be present in trace quantities in the
    air.  NO3 has been identified in laboratory systems containing
    NO2/O3, NO2/O and N2O5 as an important reactive transient
    (Johnston, 1966).  It is likely to be present in photochemical smog. 
    In the presence of sunlight, NO3 is rapidly converted to either NO or
    NO2 (Wayne et al., 1991).  Nitrogen trioxide is highly reactive
    towards both NO and NO2.  Its expected concentration in polluted air
    is very low (about 10-6 µg/m3).  However, traces of NO3 may play an
    important role in atmospheric chemistry, especially at night when it
    may serve as a reservoir for NOx (Wayne et al., 1991).  In the
    atmosphere N2O3 is in equilibrium with NO and NO2.  It reacts with
    water to form HNO2.  N2O4 is the dimer of NO2, formed in
    equilibrium with NO2 molecules, and it readily dissociates to NO2. 
    N2O5 can be a trace night-time component of the air because it is
    formed by a reaction between NO2 and NO3.  Since NO3 can exist in
    appreciable quantities only in the absence of sunlight, N2O5 is only
    important at night, when its reaction with water can be a significant
    source of nitric acid.

    2.2.2  Nitrogen acids

    2.2.2.1  Nitric acid

         HNO3 is the most oxidized form of nitrogen.  In the gaseous
    state it is colourless.  It is photochemically stable in the
    troposphere.  HNO3 is volatile, so that at typical concentrations and
    temperatures in the atmosphere the vapour does not coalesce into
    aerosol and is not retained on particles unless the aerosol contains
    reactants such as sodium chloride or ammonium salts to react with the
    acid, when it produces particulate nitrates (Wolff, 1984).

         In the aqueous phase (e.g., rain drops), HNO3 dissociates to
    form the nitrate ion (NO3-).  Because nitrate is chemically
    unreactive in dilute aqueous solution, nearly all of the
    transformations involving nitrate in natural waters result from
    biochemical pathways.  The nitrate salts of all common metals are
    quite soluble.

    2.2.2.2  Nitrous acid

         HNO2 is formed when NO and NO2 are present in the atmosphere,
    as a result of their reaction with water.  In sunlight, the dominant

    pathway for HNO2 formation is the reaction of NO with hydroxyl
    radicals.  During the daytime, atmospheric concentrations of HNO2 are
    limited by the photolysis of HNO2 to produce NO and hydroxyl radical.

         Nitrous acid is a weak reducing agent and is oxidized to nitrate
    only by strong chemical oxidants and by nitrifying bacteria.

    2.2.3  Ammonia

         NH3 is the completely reduced form of nitrogen.  It is a
    colourless gas with a pungent odour.  It is extremely soluble in
    water, forming ammonium (NHy+) and hydroxyl (OH-) ions.  In the
    atmosphere, NH3 has been reported to be converted into NOx by
    reaction with hydroxyl radicals (Soederlund & Svensson, 1976).  In the
    stratosphere, NH3 can be dissociated by irradiation with sunlight at
    wavelengths below 230 nm (McConnell, 1973).

    2.2.4  Ammonium nitrate

         Gas-phase ammonia reacts with nitric acid to form ammonium
    nitrate (NH4NO3).  Ammonium nitrate is a solid at room temperature. 
    Like ammonia, it is very soluble in water and hence will be absorbed
    by any water droplets present.  Thus it readily forms an aerosol in
    the atmosphere.  Pathways to aerosol formation include nucleation and
    condensation on existing particles.  The presence of NH4NO3
    particles can result in a visible haze.

    2.2.5  Peroxyacetyl nitrate

         Of the various peroxy nitrates found in ambient air, peroxyacetyl
    nitrate (CH3COOONO2), or PAN, is found at the highest concentrations.
    PAN undergoes a temperature-dependent decomposition to its precursors,
    NO2 and acetyl peroxy radicals.  At low ambient temperatures PAN
    can have a substantial lifetime in the atmosphere (Cox & Roffey, 1977).
    In polluted air PAN concentrations can reach several parts per billion.

    2.2.6  Organic nitrites and nitrates

         A wide variety of organic nitrites (RNO2) and nitrates (RNO3),
    where R denotes CH3, CH2CH3, benzyl, etc., may be found in ambient
    air.  Some of these are emitted directly while others are formed by
    photochemical reactions in the atmosphere.

    2.3  Sampling and analysis methods

         This section outlines methods for measuring nitrogen-containing
    species in the atmosphere.  The main focus is on methodologies
    currently available and in general use for monitoring concentrations
    in both ambient and indoor air.

         Table 2 summarizes sampling and analytical methods for selected
    species and addresses relevant characteristics, including the type of
    method (i.e.,  in situ, remote, active, passive, continuous or
    integrative), the stage of development of the method, sampling
    duration, precision, accuracy and detection limits.

    2.3.1  Nitric oxide

    2.3.1.1  Nitric oxide continuous methods

         Nitric oxide reacts rapidly with O3 to give NO2 in an excited
    electronic stage.  The transition of excited NO to the grand state can
    be accompanied by the emission of light in the red-infrared spectral
    range.  When this chemiluminescent reaction occurs under controlled
    conditions, the intensity of the emitted light is proportional to the
    concentration of the NO reactant.  This provides the basis of the
    chemiluminescence method (CLM) for analysis of NO.  This method is a
    continuous technique and is the most commonly used method for
    measuring NO in ambient air.  Commercial instruments for measuring NO
    and NO2 are available with detection limits of approximately 5 ppb
    and response times of the order of minutes.  CLM measurement of NO2
    can also be accomplished by firstly converting the NO2 of the sample
    to NO.  This is discussed in section 2.3.2.1.

         Other NO analytical methods include laser-induced fluorescence
    (LIF) (Bradshaw et al., 1985), absorption spectroscopy (e.g., tuneable
    diode laser absorption spectroscopy, TDLAS) and passive samplers.

    2.3.1.2  Passive samplers for NO

         Passive samplers are used for air with higher-than-typical
    ambient concentrations, which may be found indoors or in the
    workplace.  They are often used to obtain data at a large number of
    sites.  Sampling typically lasts a few hours.

         The Palmes tube is a passive sampler that relies on diffusion of
    an analyte molecule through a quiescent diffusion path of known length
    and cross-sectional area to a reactive surface where the molecule is
    captured by chemical reaction (Palmes et al., 1976).  The Palmes tube
    does not measure NO directly.  Two tubes are required; the first one
    has reactive grids coated with triethanolamine (TEA) to collect NO2,
    the second tube is similar but has an additional reactive surface
    coated with chromic acid to convert NO to NO2, which is in turn
    collected by the TEA-coated grids.  The NO concentration of the air is
    determined from the difference in the results from the two tubes.  The
    data is corrected for the effects of the different diffusivities of NO
    and NO2 molecules. To ensure reliable results, contact between the
    chromic-acid-coated surface and the TEA-coated grids for longer than
    24 h must be avoided.  Analysis of the material contained in the TEA

        Table 2.  Selected instruments and methods for determining oxides of nitrogen in ambient air (from: Sickles, 1992)
                                                                                                                                              

    Species        Methodsa   Typeb      Development  Sample                Performance           Comments                 References
                                         stagec       duration                                
                                                                 Precision  Accuracy   MDLd
                                                                                                                                              

    NO             CLM        I, A, C      C          5 min      < 10%      < 20%      < 9 ppb    -                    Finlayson-Pitts &
                   (NO + O3)                                                                                           Pitts (1986)

                   TP-LIF     I, A, C      R          30 sec     -          16%        10 ppt     -                    Bradshaw et al. (1985);
                                                                                                                       Davis et al. (1987)

                   TDLAS      I, A, C      R, C       60 sec     -          -          0.5 ppb    40-m path length     NASA (1983)

                   PSD        I, P, IN     C          24 h       -          -          70 ppb-he

    NO2            CLM        I, A, C      C          5 min      10%        20%        9 ppb      Commonly used        Finlayson-Pitts &
                   (NO + O3)                                                                      method; many         Pitts (1986)
                                                                                                  interferences

                   CLM        I, A, C      R          < 100 sec  20 ppt     30%        10-25 ppt  Uses thermal or      Helas et al. (1987);
                   (NO + O3)                                                                      photolytic           Fehsenfeld et al.
                                                                                                  converters           (1987)

                   CLM        I, A, C      C          100 sec    0.6 ppb    -          10 ppt     Interferences:
                   (Luminol)                                                                      PAN, HNO2, O3

                   TP-LIF     I, A, C      R          2 min      20 ppt     16%        12 ppt     -                    Davis (1988)

                   TDLAS      I, A, C      R, C       60 sec     -          15%        100 ppt    150-m path length    NASA (1983)

                   DOAS       R, A, C      R, C       12 min     -          10%        4 ppb      800-m path length    Platt & Perner (1983)

                   Bubbler    I, A, IN     RM         24 h       6 ppb      10%        8 ppbe                          Purdue & Hauser (1980)
                                                                                                                                              

    Table 2.  (Con't)
                                                                                                                                              

    Species        Methodsa   Typeb      Development  Sample                Performance           Comments                 References
                                         stagec       duration                                 
                                                                 Precision  Accuracy   MDLd
                                                                                                                                              

                   TEA        I, A, IN     L          24 h       15%        10%        0.2 ppbe   Interferences:       Sickles et al. (1990)
                   filter                                                                          PAN and HNO2f

                   Guaiacol   I, A, IN     L          1 h        4%         -          0.1 ppbe   Stability of         Buttini et al. (1987)
                   Denuder                                                                        extract uncertain

                   DPA        I, A, IN     L          8 h        8%         -          0.1 ppbe   DPA may volatilize;  Lipari (1984)
                   Cartridge                                                                      interferences:
                                                                                                  HNO2 and PAN

                   TEA PSD    I, P, IN     L          24 h       30%        -          30 ppb-he  Similar to Palmes
                                                                                                  Tube; interferences
                                                                                                  as abovef

    NOy           CLM        I, A, C      R          10 sec     -          15%        10 ppt     CO with Au           Fahey et al. (1986)
                   (NO + O3)                                                                      reducing catalyst

    PAN            GC-ECD     I, A, IN     R, RM      15 min     -          30%        10 ppte    Sensitivity can be   Vierkorn-Rudolph
                                                                                                  enhanced by using    et al. (1985)
                                                                                                  cryogenic sampling
                                                                                                  and capillary
                                                                                                  columns

                   GC-CLM     I, A, IN     L          -          -          -          -          CLM (NO + O3) and
                                                                                                  (Luminol) reported

    Other organic  GC-ECD/MS  I, A, C      R          24 h       -          -          1 ppte     Sample collected     Atlas (1988)
    Nitrates                                                                                      on charcoal
                                                                                                                                              

    Table 2.  (Con't)
                                                                                                                                              

    Species        Methodsa   Typeb      Development  Sample                Performance           Comments                 References
                                         stagec       duration                                
                                                                 Precision  Accuracy   MDLd
                                                                                                                                              

    NHO3           Filter     I, A, IN     R, RM      24 h       10%        20%        8 ppte     May be nylon or      Finlayson-Pitts &
                                                                                                  calcium chloride     Pitts (1986)
                                                                                                  impregnated filter;
                                                                                                  subject to
                                                                                                  artifactsf

                   Denuder    I, A, IN     R, RM      24 h       8%         -          8 ppte     Not subject to       Sickles (1987);
                                                                                                  above artifactsf     Sickles et al. (1989)

                   TDLAS      I, A, C      R, C       5 min      -          20%        100 ppt    150-m path length    NASA (1983)

    HNO2           Denuder    I, A, IN     R, RM      24 h       15%        -          10 ppte    Annular denuder      Sickles et al. (1989);
    
                                                                                                  preferredf           Vossler et al. (1988)

                   LIF        I, A, C      R          15 min     -          -          20 ppt     OH detected
                                                                                                  following photo-
                                                                                                  fragmentation

                   DOAS       R, A, C      R, C       12 min     -          30%        600 ppt    800-m path length    Biermann et al. (1988)
                                                                                                                                              

    Table 2.  (Con't)
                                                                                                                                              

    Species        Methodsa   Typeb      Development  Sample                Performance           Comments                 References
                                         stagec       duration                                
                                                                 Precision  Accuracy   MDLd
                                                                                                                                              

    NO3            DOAS       R, A, C      R, C       12 min     -          15%        20 ppt     800-m path length    Platt & Perner (1983)

    Particulate    Denuder/   I, A, IN     R, RM      24 h       10%        -          40 ng/m3e  Use of denuders      Vossler et al. (1988)
    NO3            Filter(s)                                                                      avoids artifacts;
                                                                                                  denuders collect
                                                                                                  HNO3 and NH3;
                                                                                                  teflon and nylon
                                                                                                  filters used

    N2O            GC-ECD     I, A, IN     R, RM      15 min     3%         -          20 ppbe    -
                                                                                                                                              

    a  CLM (NO + O3) = Chemiluminescent using NO + O3 reaction       b  I = In situ
       TP-LIF = Two-photon laser-induced                                A = Active
       TDLAS = Tuneable diode laser absorption spectroscopy             C = Continuous
       TTFMS = Two-tone frequency modulated spectroscopy                P = Passive
       PSD = Passive sampling device                                    IN = Integrative
       CLM (Luminol) = Chemiluminescent using reaction with Luminol     R = Remote
       DOAS = Differential optical absorption spectroscopy
       DIAL = Differential absorption lidar                          c  C = Commercially available
       TEA = Triethanolamine                                            R = Research tool
       DPA = Diphenylamine                                              L = Laboratory prototype
       GC-ECD = Gas chromatography with electron capture detector       RM = Routine method
       CG-CLM = Gas chromatography with CLM detector
       LIF = Laser-induced fluorescence                              d  MDL  =  Minimum detection limit
       GC-MS = gas chromatography with mass spectrometer             e  Depends on the sampled air volume (i.e., flow rate and sampling
                                                                        duration)
                                                                     f  Uses ion chromatographic or colorimetric analytical finish
        is accomplished by extracting the grids into solution and analysing
    the extract for NO2- by the use of the spectrophotometric or ion
    chromatographic method (Miller, 1984).  The colorimetric analysis is
    calibrated by dilution of gravimetrically prepared nitrite solutions. 
    The Palmes Tube method was proposed for sampling occupational
    exposures where the dosage does not exceed 25 ppm for 8 h (i.e.,
    200 ppm-h).  The reliability of this method for measuring NO in the
    field at the parts-per-billion or parts-per-million level remains to
    be demonstrated.

         A badge-type sampler similar to the Palmes tube has been devised
    by Yanagisawa & Nishimura (1982).  This device uses a series of
    12 layers of chromium-trioxide-impregnated glass fibre to oxidize NO
    to NO2.  This technique is claimed to be more sensitive by
    approximately a factor of 10 than the Palmes tube and to have a lower
    limit dosage of 0.07 ppm-h.

    2.3.1.3  Calibration of NO analysis methods

         Calibration of CLM, TP-LIF and TDLAS measurement systems for NO
    all rely on compressed gas mixtures of known concentration being
    available.  Typically compressed gas mixtures are supplied in
    passivated aluminium/stainless steel gas bottles certified by the
    manufacturer and with NO diluted with N2 concentration in the rage of
    1 to 50 ppm (Schiff et al., 1983; Carroll et al., 1985; Bradshaw et
    al., 1985).  Calibrations are performed by dynamic dilution of the
    reference NO/N2 mixture with air to give NO concentrations within the
    range of 0.1 to 5 ppm.

         For passive NO samplers, only the analysis portion of the
    procedure is routinely calibrated (using gravimetrically prepared
    nitrite solution).

    2.3.1.4  Sampling considerations for NO

         Oxides of nitrogen are reactive species and exhibit various
    solubilities (Table 1).  The most inert materials (i.e. glass and
    TeflonTM) are recommended for use in sampling trains.  Since ambient
    air contains water vapour that may be sorbed on sampling lines,
    surface effects may influence the integrity of air samples containing
    the more reactive and more soluble NOy species.  In hot, humid
    conditions condensation in the sample lines of liquid water from the
    air can cause difficulties when analysis equipment is installed in an
    air-conditioned environment.  To minimize contamination of the system
    by dust and foreign matter, it is common practice to sample through an
    inert (teflon) sample inlet filter.  Of the NOy species, NO is
    probably the least susceptible to surface effects, whereas surface
    effects are very important in the sampling of HNO3.

         Nitric oxide reacts rapidly with O3 to form NO2.  In the
    presence of sunlight NO2 in air photolyses to yield NO and O3.  Thus
    in daylight NO, O3 and NO2 can exist simultaneously in ambient air
    in a condition known as a "photostationary state".  The relative
    amounts of the three species at any time are influenced by the
    intensity of the sunlight present at that moment.  Photolysis ceases
    when a sample is drawn into a dark sampling line, but NO and O3 can
    continue to react to form NO2.  Therefore residence times in sampling
    lines must be minimized to maintain the intensity of the NO/NO2 ratio
    of the sample.

    2.3.2  Nitrogen dioxide

         Airborne concentrations of NO2 can be determined by several
    methods including CLM, LIF, absorption spectroscopy, including
    differential optical absorption spectroscopy (DOAS) and TDLAS, bubbler
    and passive collection with subsequent wet chemical analysis.  The
    most common techniques are chemiluminescence and passive sampling.

    2.3.2.1  Chemiluminescence (NO + O3)

         Instruments discussed in this section do not detect NO2
    directly.  They sample continuously and rely on the conversion of some
    or all of the NO2 in the air sample to NO, followed by the CLM
    reaction of NO and O3. The NO2 concentration is calculated from the
    difference in the signal given by the sample after passing through the
    converter compared to that when the converter is by-passed.

         Several methods have been employed to reduce NO2 to NO (Kelly,
    1986).  They include catalytic reduction using heated molybdenum or
    stainless steel, reaction with carbon monoxide over a gold catalyst
    surface, reaction with iron sulfate at room temperature, reaction with
    carbon at 200°C, and photolysis of NO2 to NO by light in the
    wavelength range of 320 to 400 nm.

         CLM instruments for the determination of NO2 are readily
    available commercially.  Field evaluation of nine instruments showed
    that the minimum detection limits (MDLs) ranged from 5 to 13 ppb
    (Michie et al., 1983; Holland & McElroy, 1986).

         Converters may be non-specific for NO2 and may convert
    several other nitrogen-containing compounds to NO, giving rise to
    overestimates for NO2 concentrations.  Using commercial instruments,
    Winer et al. (1974) found over 90% conversion of PAN, ethyl nitrate
    and ethyl nitrite to NO with a molybdenum converter, and similar
    responses to PAN and  n-propyl nitrate with a carbon converter.  With
    a stainless steel converter at 650°C, Matthews et al. (1977) reported
    100% conversion for NO2, 86% for NH3, 82% for CH3NH2, 68% for HCN,
    1% for N2O and 0% for N2.  Using a commercial instrument, Joseph &
    Spicer (1978) found quantitative conversion of HNO3 to NO with a

    molybdenum converter at 350°C.  Similar responses to PAN, methyl
    nitrate,  n-propyl nitrate,  n-butyl nitrate and HNO3, substantial
    response to nitrocresol, and no response to peroxybenzoyl nitrate
    (PBzN) were reported with a commercial instrument using a molybdenum
    converter at 450°C (Grosjean & Harrison, 1985).  These results were
    confirmed for PAN and HNO3 by Rickman & Wright (1986) using
    commercial instruments with a molybdenum converter at 375°C and a
    carbon converter at 285°C.

         Interference from species that do not contain nitrogen have also
    been reported.  Joshi & Bufalini (1978), using a commercial instrument
    with a carbon converter, found significant apparent NO2 responses
    to phosgene, trichloroacetyl chloride, chloroform, chlorine (Cl2),
    hydrogen chloride, and photochemical reaction products of a
    perchloroethylene-NOx mixture.  Grosjean & Harrison (1985) reported
    substantial responses to photochemical reaction products of Cl2-NOx
    and Cl2-methanethiol mixtures and small negative responses to
    methanethiol, methyl sulfide, and ethyl sulfide.  Sickles & Wright
    (1979), using a commercial instrument with a molybdenum converter at
    450°C, found small negative responses to 3-methylthiophene,
    methanethiol, ethanethiol, ethyl sulfide, ethyl disulfide, methyl
    disulfide, hydrogen sulfide, 2,5-dimethylthiophene, methyl sulfide
    and methyl ethyl sulfide, and negligible responses to thiophene,
    2-methylthiophene, carbonyl sulfide and carbon disulfide.

         Methods of sample trapping followed by batch measurement of NO
    and NO2 in the desorbed sample using a chemiluminescence instrument
    have been reported.  Gallagher et al. (1985) used cryosampling of
    stratospheric whole-air samples, and Braman et al. (1986) used
    copper(I) iodide coated denuder tubes to sample NO2 in ambient air.

    2.3.2.2  Chemiluminescence (luminol)

         A method for the direct chemiluminescence determination of NO2
    was reported by Maeda et al. (1980) and is based on the CLM reaction
    of gaseous NO2 with a surface wetted with an alkaline solution of
    luminol (5-amino-2,3-dihydro-1,4-phthalazinedione).  The light
    emission is strong at wavelengths between 380 and 520 nm.  The
    intensity of the light can be proportional to the NO2 concentration
    in the sampled air, and the NO2 concentration can be determined by
    calibration of the instrument with air of known NO2 concentration.

         Since the introduction of the luminol method by Maeda et al.
    (1980), improvements have been made to develop an instrument
    suitable for use in the field (Wendel et al., 1983), and additional
    modifications have been made recently to produce a continuous
    commercial instrument (Schiff et al., 1986).  Detection limits of 5 to
    30 ppt and a response time of seconds have been claimed, based on
    laboratory tests (Wendel et al., 1983; Schiff et al., 1986).  Recent
    laboratory evaluation of two instruments has revealed a detection

    limit (i.e., twice the standard deviation of the clean air response)
    of 5 ppt, and 95% rise and fall times of 110 and 15 seconds (Rickman
    et al., 1988).  Field tests of the same instruments have shown an
    operating precision of ± 0.6 ppb.

    2.3.2.3  Laser-induced fluorescence and tuneable diode laser
             absorption spectrometry

         Two newer techniques that show considerable promise for measuring
    NO2 specifically are photofragmentation/2-photon LIF and TDLAS.  The
    LIF and TDLAS techniques provide specific spectroscopic methods to
    measure NO2 directly and compare favourably to the sample photolysis-
    chemiluminescence technique (Fehsenfeld et al., 1990; Gregory et al,
    1990b).  For NO2 concentrations above 0.2 ppb, no interferences were
    found for TDLAS (Fehsenfeld et al., 1990).

    2.3.2.4  Wet chemical methods

         Most wet chemical methods for measuring NO2 involve the
    collection of NO2 in solution, followed by a colorimetric finish
    using an azo dye.  Many variations of this method exist, including
    both manual and automated versions.  These include the Griess-Saltzman
    method, the continuous Saltzman method, the alkaline guiacol
    method, the sodium arsenite method (manual or continuous), the
    triethanolamine-guaiacol-sulfite (TGS) method and the TEA method. 
    These methods have been reviewed by Purdue & Hauser (1980).

    2.3.2.5  Other methods

         Several other methods for the determination of NO2 have been
    reported.  Atmospheric pressure ionization mass spectrometry has been
    investigated for the continuous measurement of NO2 and SO2 in
    ambient air (Benoit, 1983).  Methods employing photothermal detection
    of NO2 have been reported (Poizat & Atkinson, 1982; Higashi et al.,
    1983; Adams et al., 1986).

         A portable, battery-powered analyser specific to NO2, which uses
    an electrochemical cell as the detector, is commercially available. 
    By careful selection and design of the cell, levels down to
    approximately 0.1 ppm (v/v) can be detected, although with
    uncertainties of approximately 20-50%.  The detection cell has a
    finite life, dependent on the time integral of the NO2 concentrations
    measured.  When the cell deteriorates, the instrument typically
    develops a gradual drift.

    2.3.2.6  Passive samplers

         Passive samplers are frequently used in industrial hygiene,
    indoor air and personal exposure studies and are less frequently used
    for ambient air analysis.  Namiesnik et al. (1984) have provided an
    overview of passive samplers.

         One type of passive NO2 sampler for ambient application is the
    nitration plate.  It is essentially an open petri dish containing
    TEA-impregnated filter paper.  Mulik & Williams (1986) have adapted
    the nitration plate concept by adding diffusion barriers in their
    design of a passive sampling device (PSD) for NO2 in ambient and
    personal exposure applications.  The device employs a TEA-coated
    cellulose filter paper, two 200-mesh stainless steel diffusion screens
    and two stainless steel perforated plates on each side of the coated
    filter to act as diffusion barriers and permit NO2 collection on both
    faces of the filter paper.  After sampling, the paper is removed
    from the PSD, extracted in water, and analysed for NO2- by
    ion chromatography.  A sensitivity of 0.03 ppm-h and a rate of
    2.6 cm3/second were claimed. Comparison of PSD results with
    chemiluminescence determinations of NO2 in laboratory tests at
    concentrations between 10 and 250 ppb showed a linear relation and
    high correlation (i.e.,  r = 0.996) (Mulik & Williams, 1987). 
    Interference from PAN and HNO2 would be expected (Sickles, 1987). 
    Results of TDLAS and triplicate daily PSD NO2 measurements in a
    13-day field study showed good agreement between the study average
    values but a correlation coefficient for daily results of only 0.47
    (Mulik & Williams, 1987; Sickles et al., 1990).  The Palmes tube
    described in section 2.3.1.2 has been used to sample air in the
    workplace and indoor environments to assess personal exposure to NO2
    (Palmes et al., 1976; Wallace & Ott, 1982).

    2.3.2.7  Calibration

         Calibration methods for NO2 use permeation tubes or gas-phase
    titration (GPT) to generate known concentrations of NO2. 
    Calibrations are performed dynamically using dilution with purified
    air.

         GPT employs the rapid, quantitative gas-phase reaction between
    NO, usually supplied as a known concentration from a gas cylinder, and
    O3 supplied from a stable O3 generator, to produce one NO2 molecule
    for each NO molecule consumed by reaction.  When O3 is added to
    excess NO in a titration system, the decrease in NO concentration
    (and O3) is equivalent to the increase in NO2 produced (US EPA,
    1987b).

         Use of cylinders of compressed gas containing NO2 for
    calibration purposes (Fehsenfeld et al., 1987; Davis, 1988) is unwise
    because of the uncertain stability of the NO2 concentrations
    delivered; this is a consequence of its relatively high boiling point.

    2.3.3  Total reactive odd nitrogen

         In this monograph, gas-phase total reactive odd nitrogen is
    represented by NOy.  Individual components comprising NOy are gas

    phase NO, NO2, NO3, N2O5, HNO2, HNO3, peroxynitric acid
    (HO2NO2), PAN, and other organic nitrates.  NH3 and N2O are not
    components of NOy.

         Researchers have successfully combined highly sensitive research-
    grade CLM NO detectors with catalytic converters that are sufficiently
    active to reduce most of the important gas phase NOy species to NO
    for subsequent detection (Helas et al., 1981; Dickerson, 1984; Fahey
    et al., 1986; Fehsenfeld et al., 1987).

    2.3.4  Peroxyacetyl nitrate

         Several methods have been used to measure the concentration of
    PAN in ambient air.  Roberts (1990) has provided an overview of many
    of these methods.  A well-developed method is gas chromatography using
    electron capture detection (GC-ECD) (Darley et al., 1963; Smith et
    al., 1972; Stephens & Price, 1973; Singh & Salas, 1983).

    2.3.5  Other organic nitrates

         Other organic nitrates (e.g., alkyl nitrates, peroxypropionyl
    nitrate and PBzN) can also be present in the atmosphere, but usually
    at lower concentrations than PAN (Fahey et al., 1986).  In general,
    similar methods for sampling, analysis and calibration may be used for
    other organic nitrates as are used for PAN (Stephens, 1969).  FTIR,
    GC-ECD and GC-MS may be used to measure these compounds.

    2.3.6  Nitric acid

         Several methods are available for the determination of HNO3
    concentrations in the atmosphere.  These include filtration (Okita et
    al., 1976; Spicer et al., 1978a), denuder tubes (Forrest et al., 1982;
    De Santis et al., 1985; Ferm, 1986), CLM (Joseph and Spicer, 1978) and
    absorption spectroscopy (Tuazon et al., 1978; Schiff et al., 1983;
    Biermann et al., 1988).  Many of these techniques carry significant
    uncertainties, which have been compared by Hering et al. (1988).

    2.3.7  Nitrous acid

         Available techniques for the measurement of HNO2 in ambient
    atmospheres employ denuders (Ferm & Sjodin, 1985), annular denuders
    (De Santis et al., 1985), CLM (Braman et al., 1986), PF/LIF (Rodgers &
    Davis, 1989), absorption spectroscopy (Tuazon et al., 1978; Biermann
    et al., 1988) and FTIR (Finlayson-Pitts & Pitts, 1986).

    2.3.8  Dinitrogen pentoxide and nitrate radicals

         N2O5 is readily reduced to NO at temperatures above 200°C and
    may be measured nonspecifically as NO2 with CLM NO2 analysers
    (Bollinger et al., 1983; Fahey et al., 1986).

         Ambient concentrations of the NO3 radical have been measured
    using DOAS; concentrations between 1 and 430 ppt have been observed
    (Atkinson et al., 1986).

    2.3.9  Particulate nitrate

         Many methods are available for sampling ambient aerosols,
    including impactors, filtration, and filtration coupled with devices
    to remove particles larger than a specified size (e.g., elutriators,
    impactors and cyclones).

         Particulate nitrate samples are generally collected by
    filtration, extracted, and analysed directly or indirectly for nitrate
    by ion chromatography or colorimetry.

    2.3.10  Nitrous oxide

         The most commonly used analytical method for N2O employs GC-ECD. 
    It has a detection limit of 20 ppb (Thijsse, 1978) and a precision of
    ± 3% at the background level of 330 ppb (Cicerone et al., 1978).

    2.3.11  Summary

         Gas-phase CLM instruments have replaced manual (wet) methods
    to a large extent in air quality monitoring network applications. 
    Gas-phase CLM measurement technology permits the determination of NO,
    NO2 and NOy in the low ppt range.  Although CLM NO detectors coupled
    with catalytic NO2 to NO converters are still not specific for NO2,
    they have proved to be useful for measuring NOy.  CLM NO detectors
    coupled with photolytic NO2 to NO converters have shown improved
    specificity for NO2.  Most ambient NO2 monitoring data reported are
    from the nonspecific thermal conversing technique.

         Passive samplers for NO2 have been used primarily for workplace
    and indoor applications, but hold promise for averaged ambient
    measurements as well.  GC-ECD is useful in the determination of PAN,
    other organic nitrates and N2O.

    2.4  Transport and transformation of nitrogen oxides in the air

    2.4.1  Introduction

         Oxides of nitrogen are transformed by and removed from the
    atmosphere by a complex web of reactions that are fundamental to the
    formation and destruction of ozone and other oxidants.  The
    predominant form of oxidized nitrogen (NO, NO2, HNO3, etc.)
    in the lower atmosphere varies, depending upon sunlight intensity,
    temperature, pollutant emissions, period of time since these emissions
    occurred and the meteorological history of an airmass.

    2.4.2  Chemical transformations of oxides of nitrogen

    2.4.2.1  Nitric oxide, nitrogen dioxide and ozone

         The dominant source of nitrogen oxides in the air is combustion
    processes (see chapter 3); 90-95% of these nitrogen oxides are usually
    emitted as NO and 5-10% as NO2.  NO may be oxidized to NO2 by
    atmospheric oxygen according to reaction 2-1:

              NO + NO + O2 -> 2 NO2                                   (2-1)

    However at low NO concentrations this reaction is slow and is
    important only when NO > 1 ppm (Boström C, 1993).  NO concentrations
    greater than 1 ppm are not frequently found in ambient air, but they
    may possibly occur in indoor air and in plumes from industrial sources
    (see Chapter 3).  When concentrations are below 1 ppm, NO is oxidized
    to NO2 by two types of reaction.  The first type of reaction is given
    in equations 2-2 to 2-4.  NO can react with O3:

              NO + O3 -> NO2 + O2                                     (2-2)

    Also O3 is formed when NO2 is photolysed, forming NO plus an O atom

              NO2 + hnu -> O + NO                                     (2-3)

    and O atoms react rapidly with O2 to form ozone:

                     M
              O + O2 -> O3                                            (2-4)

    Thus reactions 2-2, 2-3 and 2-4 recycle O3 rather than producing a
    net increase in O3 concentrations, where the "M" represents a third
    molecule such as N2, O2, etc., that absorbs excess vibrational
    energy from the newly formed O3 molecules.  However, a second
    oxidation path involving the reaction of organic species can lead to
    increases in O3 concentrations and in the conversion rate of NO to
    NO2 (2-9 and 2-10).  Organic compounds in the air are commonly
    referred to as VOC (volatile organic carbon), ROC (reactive organic
    carbon) and non-methane hydrocarbons (NmHC).  Urban areas are usually
    characterized by significant sources of both nitrogen oxides and ROC
    emissions.  With suitable atmospheric conditions this can lead to the
    formation of photochemical smog.  The smog-forming reactions are
    initiated by photolytic reactions which produce free radicals, for
    example:

    (i) the photolysis of O3

              O3 + hnu -> O2 + O*                                     (2-5)

    O* is an excited form of atomic oxygen, which can react with water to
    produce the hydroxyl radical (OH):

              O* + H2O -> 2OH                                       (2-6)


    (ii) the photolysis of aldehydes, which also results in the production
    of OH.  Aldehydes are emitted in motor vehicle exhaust and are
    produced in the air by reaction of ROC species with OH.  OH is the
    most important oxidizing agent in the lower atmosphere; it can react
    with all organic compounds, usually forming water and producing an
    organic radical.

         For a generalized organic compound, R-H (R = CH3, CHO, CH2CH3,
    etc.), the principal elements of the reaction sequence are:

              R-H + OH -> H2O + R                                   (2-7)

                     M
              R + O2 -> RO2 (fast)                                  (2-8)

    RO2 provides a pathway to oxidize NO to NO2 without destroying O3
    (unlike reaction 2-2):

              RO2 + NO -> NO2 + RO                                  (2-9)

    RO can undergo reactions that form additional HO2 or RO2.  HO2
    reacts with NO to form NO2 and regenerate OH:

              HO2 + NO -> NO2 + OH                                  (2-10)

         In the case of photochemical smog episodes, the quantity of NOx
    emitted into the air determines the ultimate quantity of O3 that may
    be produced.  The ROC concentration and sunlight intensity are the
    major determinates of the rates at which NO will be oxidized to
    produce net increases in NO2 and O3 concentrations.  Ozone
    production is terminated when NO and NO2 are consumed by reaction to
    form products such as HNO3 (see below), resulting in insufficient NO
    concentration for reactions 2-9 and 2-10 to proceed at significant
    rates.

         In large cities with sunny climates and poor dispersion of
    emissions (e.g., Los Angeles and Mexico City), O3 concentrations in
    excess of 200 ppb are not uncommon.

    2.4.2.2  Transformations in indoor air

         Oxides of nitrogen in indoor air arise from two sources: a)
    outdoor air; and b) indoor sources, such as combustion appliances and
    heaters.  Photochemical reactions do not take place under artificial
    lighting, so chemical transformations are limited by the amounts of
    oxidizing species (HO2, O3, etc.) that arrive in outdoor air, or are
    generated by combustion sources.

    2.4.2.3  Formation of other oxidized nitrogen species

         Oxidation products of NOx arising from tropospheric
    photochemical reactions include HNO3, HO2NO2, HNO2,
    peroxyacylnitrates (RC(O)O2NO2), N2O5, nitrate radical (NO3) and
    organic nitrates (RNO3).

         Fig. 1 shows a summary for the interconversion pathways for
    oxides of nitrogen.  These pathways govern urban and indoor air, as
    well as "clean" air, but the partitioning between the nitrogen oxide
    species varies according to the specific conditions characteristic of
    each type of airmass.

    a)  Nitric acid

         Nitric acid is a strong mineral acid that contributes to acidic
    deposition from the air.  In terms of atmospheric chemistry, HNO3 is
    a major sink for active nitrogen.  In daylight, HNO3 is formed by the
    reaction of NO2 with the OH radical:

                       M
              NO2 + OH -> HNO3                                       (2-11)

         This reaction is a chain-terminating step in the free radical
    chemistry that produces urban photochemical smog and it removes
    reactive nitrogen as well as the hydroxyl radical.  Reaction 2-11 is a
    relatively fast reaction that can produce significant amounts of HNO3
    over a period of a few hours.  At night, in polluted air containing
    significant ozone concentrations, the heterogeneous reaction between
    gaseous N2O5 and liquid water is thought to be a source of HNO3
    (N2O5 is produced from NO3 (see section 2.4.3.5) and NO2).  This
    pathway to HNO3 production is negligible during daytime, because the
    NO3 radical photolyses rapidly and is not present in sufficient
    quantities to react with NO2.  The NO3 radical can also abstract a
    hydrogen atom from certain organic compounds (such as aldehydes,
    dicarbonyls and cresols) to provide another night-time source of
    HNO3.

         Logan (1983) has estimated a lifetime of 1 to 10 days for
    HNO3 in the lower troposphere.  The primary removal mechanism is
    deposition.  The loss of HNO3 by rain-out is subject to precipitation
    frequency while the loss rate by dry deposition varies with the nature

    FIGURE 2

    of the ground and vegetation and atmospheric mixing characteristics of
    the boundary layer.  Chemical destruction mechanisms for HNO3 also
    exist, but their importance is not well understood and is suspected to
    be minor for the lower troposphere.

         In the presence of NH3, HNO3 may form the salt, ammonium
    nitrate:

              HNO3(g) + NH3(g) -> NH4NO3                             (2-12)

         Ammonium nitrate gas readily condenses to the particulate phase.
    Ammonium nitrate aerosol can be responsible for significant visibility
    reduction and particulate pollution, e.g., where nitric acid is
    produced in air from urban areas and this interacts with NH3 emitted
    from agricultural operations.

    b)  Nitrous acid

         HNO2 is produced from the reaction of NO and OH:

                      M
              NO + OH -> HNO2                                        (2-13)

         In indoor air other reactions (particularly surface reactions)
    may be important sources of nitrous acid.

         There have been a few measurements of nitrous acid in urban
    environments (Harris et al., 1982; Winer et al., l987).  Daytime
    levels of nitrous acid are expected to be low because it photolyses
    rapidly, yielding NO and ·OH.  This reaction probably serves as a
    source of OH radicals during the morning in urban regions, where
    nitrous acid may form (from NO, NO2 and H2O) and accumulate during
    the night-time hours.  Reaction 2-13 may lead to a build up of nitrous
    acid in urban air only during the late afternoon and evening hours
    when sunlight intensities are low but some OH radicals are still
    present.

    c)  Peroxynitric acid

         While peroxynitric acid (HO2NO2) has never been measured in the
    atmosphere, it is expected to be present in the upper troposphere. 
    Models suggest concentrations in the 10 to 100 ppt range at altitudes
    above 6 kilometres (Logan, 1983; Singh, 1987).  HO2NO2 is thermally
    unstable, so that boundary layer concentrations are expected to be
    extremely low (< 1 ppt).  Peroxynitric acid is formed through the
    combination of a hydroperoxy (HO2) radical with NO2.  In the upper
    troposphere, HO2NO2 is destroyed by photolysis or by reaction with
    OH radicals.

    d)  Peroxyacyl nitrates

         Peroxyacetyl nitrate (PAN) is the most abundant of this family of
    organic nitrates.  The second most abundant homologue, peroxypropionyl
    nitrate (PPN), is generally less than 10% of the PAN concentration,
    and species with higher relative molecular mass, such as PBzN, are
    expected to have even lower concentrations.  PAN is a strong oxidant
    and is known to be phytotoxic; it is formed from the reaction of
    acetylperoxy radical with NO:

              CH3C(O)OO + NO2 +M -> CH3C(O)O2NO2 +M                 (2-14)

         PAN is thermally unstable and so its lifetime is very dependent
    on ambient temperature.  For example, PAN lifetimes of about 5 and
    20 h have been calculated for 20°C and 10°C, respectively.

         In cold conditions PAN can serve as a reservoir for reactive
    nitrogen, which is liberated when the temperature of the air is
    increased.  PAN can be lost from the atmosphere by dry deposition over
    land, but it is very likely that a significant fraction of PAN
    produced in urban plumes can be transported into the regional
    environment.

    e)  Nitrate radical

         The nitrate (NO3) radical is a short-lived species formed mainly
    by the reaction of NO2 with O3, although other sources of NO3
    radicals exist (Wayne et al., 1991).

              NO2 + O3 -> NO3 + O2                                  (2-15)

         NO3 also reacts with NO2 to form N2O5

                        M
              NO2 + NO3 -> N2O5                                     (2-16)

         Nitrate radicals rapidly photolyse, resulting in a lifetime of
    about 5 seconds at midday.  They also react rapidly with NO, which
    limits their lifetime both during the day- and night-time hours.  At
    night if atmospheric NO concentrations are approximately 320 pptv,
    then the lifetime of NO3 radicals is similar to that at midday (about
    5 seconds).

         At night, NO3 concentrations range from about 0.3 ppt in clean
    tropospheric air to 70 ppt in urban areas (Biermann et al., 1988).  In
    clean background environments, it has been reported that measured NO3
    radical levels are significantly less than those predicted by the
    above reactions.  Several loss mechanisms have been suggested (Noxon
    et al., 1980; Platt et al., 1981): (i) NO3 radical reaction with

    organic compounds; (ii) heterogeneous losses of NO3 radicals and/or
    N2O5 on particle surfaces; (iii) reactions of NO3 radicals with
    H2O vapour; and (iv) reaction of NO3 radicals with NO.

    f)  Dinitrogen pentoxide

         N2O5 is formed from NO3 and NO2 (reaction 2-15).  Since NO3
    is present only at night, N2O5 is also primarily a night-time
    species.  N2O5 is thermally unstable, decomposing to NO3 and NO2
    (reaction 2-15).  At high altitudes in the troposphere, where
    temperatures are low, N2O5 can act as a temporary reservoir for
    NO3.  Dinitrogen pentoxide photolyses at wavelengths less than 330 nm
    to give NO3 and NO2.

         Dinitrogen pentoxide reacts heterogeneously with water to form
    HNO3.  This serves as the main night-time production mechanism for
    HNO3 and it provides an important route for removal of oxidized
    nitrogen from the atmosphere, since HNO3 is readily removed by dry
    and wet deposition.  Other atmospheric reactions of N2O5 include its
    reaction with gas-phase water to form HNO3 and possible reactions
    with aromatic VOCs such as naphthalene and pyrene (Pitts et al., 1985;
    Atkinson et al., 1986).  Nitroarenes appear to be the product of
    N2O5-aromatic reactions.

    2.4.3  Advection and dispersion of atmospheric nitrogen species

         The transport and dispersion of the various nitrogen species is
    dependent on both meteorological and chemical parameters.  Advection,
    diffusion and chemical transformations dictate the atmospheric
    residence time of a particular trace gas.  Nitrogenous species that
    undergo slow chemical changes in the troposphere and are not readily
    removed by depositional processes can have atmospheric lifetimes of
    several months.  Gases with lifetimes of the order of months can be
    dispersed over continental scales and possibly even over an entire
    hemisphere.  At the other extreme are gases that undergo rapid
    chemical transformation and/or depositional losses limiting their
    atmospheric residence times to a few hours or less.  Dispersion of
    these short-lived species may be limited to only a few kilometres from
    their point of emission.

         Surface emissions are dispersed vertically and horizontally
    through the atmosphere by turbulent mixing processes.  These processes
    are dependent to a large extent on the vertical temperature structure
    of the boundary layers and on wind speed.  In the vertical dimension,
    transport occurs as follows (see also Fig. 2.):

    a)   the daytime and/or night-time mixed layer; this layer can extend
         from the surface up to a few hundred metres at night or to
         several thousand metres during the daytime;

    FIGURE 3

    b)   a layer that can exist during the night-time above a low level
         surface inversion and below the daytime mixing height; this layer
         generally is situated between 200 and 2000 m altitude;

    c)   the free troposphere; this transport zone is above the boundary
         layer mixing region.

         During the warm, summertime period, vertical mixing follows a
    fairly predictable diurnal cycle.  A surface inversion normally
    develops during the evening hours and persists throughout the night-
    time and morning period until broken by sunlight heating the surface
    of the earth.  While the inversion is in place, surface NOx emissions
    can lead to relatively high local concentrations because of restricted
    vertical dispersion.  Following the break-up of the night-time surface
    inversion, vertical mixing will increase and surface-based emissions
    will disperse to higher altitudes.  The depth of the vertical mixing
    during the daytime is often controlled by synoptic weather features. 
    Temperature inversions aloft, associated with high pressure systems,
    are common in many parts of the world.

         The dispersion processes described above, coupled with the
    chemical transformations of reactive nitrogen compounds, determine the
    distances oxidized nitrogen will be transported in the troposphere.  A
    reasonable understanding exists concerning the short-term (daylight
    hours) fate of NOx emitted in urban areas during the morning hours. 
    As described above, NOx emitted in the early morning hours in an
    urban area will disperse vertically and move downwind as the day
    progresses.  On sunny summer days, most of the NOx will have been
    converted to HNO3 and PAN by sunset.  Much of the HNO3 will be
    removed by depositional processes as the air mass moves along.  After
    dusk, an upper portion of the daytime mixed layer will be decoupled
    from the surface because of formation of a low-level radiation
    inversion.  Transport will continue in this upper level during the
    night-time hours and, although photochemical processes will cease,
    dark-phase chemical reactions can proceed.  Peroxyacetyl nitrate and
    HNO3, if carried along in this layer, can be transported long
    distances.

    2.4.3.1  Transport of reactive nitrogen species in urban plumes

         Overall removal rates for reactive nitrogen species during
    daytime at mid-latitudes have been measured or calculated for a few
    areas.  For example, in the plume from Boston, USA, after correction
    for dilution, removal rates ranged from 0.14 to 0.24 h-1 on 4 days
    (Spicer, 1982, Altshuller, 1986).  In Los Angeles and Detroit, the
    removal rate has been estimated to be 0.04-0.1 h-1 (Calvert, 1976;
    Chang et al., 1979; Kelly, 1987).  Formation and removal of HNO3 is
    thought to be the rate-controlling step for removal of reactive
    nitrogen.

    2.4.3.2  Air quality models

         Air quality models are mathematical descriptions of pollutant
    emissions, atmospheric transport, diffusion and chemical reactions
    of pollutants.  However, air quality models are very complex and
    difficult to test for validity.  Inputs include emissions, topography
    and meteorology of a region.  Air quality models represent an
    integration of knowledge for the chemistry and physics of the
    atmospheric system; they offer some predictive capability for the
    effectiveness of pollution control strategies.  Models have also been
    developed for indoor air.

    2.4.3.3  Regional transport

         Transport of reactive nitrogen species in regional air masses can
    involve several mechanisms.  Mesoscale phenomena, such as land-sea
    breeze circulations or mountain-valley wind flows, will transport
    pollutants over distances of ten to hundreds of kilometres.  On a
    larger scale, synoptic weather systems such as the migratory highs
    that cross the eastern USA and other areas of the world in the
    summertime influence air quality over many hundreds of kilometres. 
    The accumulation and fate of nitrogen compounds will differ somewhat
    between the mesoscale and synoptic systems.  Mountain-valley and land-
    water transport mechanisms have dual temporal scales because of their
    dependence on solar heating.  However, in the larger-scale synoptic
    systems, reactive nitrogen species can build up over multiday periods. 
    The residence time of air parcels within a slow-moving high pressure
    system can be as long as 6 days (Vukovich et al., 1977).

         In many cases, the transport mechanisms mentioned above are
    interrelated.  Mountain-valley or land-water breezes can dictate
    pollutant transport in the immediate vicinity of sources, but the
    eventual fate of reactive nitrogen species will be distribution into
    the synoptic system.

    2.5  Conversion factor for nitrogen dioxide

         1 ppm     = 1.88 mg/m3
         1 mg/m3   = 0.53 ppm

    2.6  Summary

         Combustion provides the major source of oxides of nitrogen in
    both indoor and outdoor air, producing mostly NO with some NO2.  The
    sum of NO and NO2 is generally referred to as NOx.  Once released
    into the air, NO is oxidized to NO2 by available oxidants,
    particularly O3, and by photochemical reactions involving reactive
    organic compounds.  This happens rapidly under some conditions in
    outdoor air; for indoor air, it is generally a much slower process. 
    Nitrogen oxides are a controlling precursor of ozone and smog

    formation; interactions of nitrogen oxides (except N2O) with reactive
    organic compounds and sunlight form ozone in the troposphere and smog
    in urban areas.

         In both indoor and outdoor air, NO and NO2 may undergo reactions
    to form a suite of other nitrogenous species including HNO2, HNO3,
    NO3, N2O5, PAN and other organic nitrates.  The complete suite of
    gas-phase nitrogen oxides is referred to as NOy.  The partitioning
    of nitrogen among these compounds is strongly dependent on the
    concentrations of other oxidants, sunlight exposure, the presence of
    reactive organic compounds and the meteorological history of the air.

         A sensitive, specific and reliable analytical method exists for
    measuring NO (by the chemiluminescent reaction with ozone), but this
    is an exception for NOy species.  Chemiluminescence is also the
    most common technique used for NO2, which is first reduced to NO. 
    Unfortunately, the method of reduction usually used is not specific
    for NO2, and it has various conversion efficiencies for other
    oxidized nitrogen compounds that may also be present in the air
    sample.  For this reason, care must be taken in interpreting the NO2
    values given by the common chemiluminescence analyser, as the signal
    may include responses from interfering compounds.  Additional
    difficulties arise from nitrogen species such as HNO3 that may
    partition between the gas and particulate phases both in the
    atmosphere and in the sampling procedure.

    3.  SOURCES, EMISSIONS AND AIR CONCENTRATIONS

    3.1  Introduction

         Oxides of nitrogen can have significant concentrations in ambient
    air and in indoor air.  The types and concentrations of nitrogenous
    compounds present can vary greatly from location to location, with
    time of day, and with the season.  The main sources of nitrogen oxides
    emissions are combustion processes.  Fossil fuel power stations, motor
    vehicles and domestic combustion appliances emit nitrogen oxides,
    mostly in the form of NO but with some (usually less than about 10%)
    in the form of NO2.  In the air chemical reactions occur which
    oxidize NO to NO2 and other products (chapter 2).  Also, there are
    biological processes in soils which liberate nitrogen species,
    including N2O.  Emissions of N2O can cause perturbation of the
    stratospheric ozone layer.

         Human health may be affected when significant concentrations of
    NO2 or other nitrogenous species, such as PAN, HNO3, HNO2 and
    nitrated organic compounds, are present.  In addition, nitrates and
    nitric acid can cause significant effects on ecosystems when deposited
    on the ground.

         Indoors, the use of combustion appliances for cooking and heating
    can give rise to greater NO and NO2 concentrations than are present
    outdoors, especially when the appliance is not vented to the outside. 
    Recent research has shown that in these circumstances nitrous acid can
    reach significant concentrations (Brauer et al., 1993).

         This chapter discusses both ambient and indoor sources
    of nitrogenous compounds, their emissions, and the resulting
    concentrations that may directly affect human health or participate
    in atmospheric chemical pathways leading to effects on human health
    and welfare.  Nitrogen-containing compounds are also of particular
    interest because of their secondary impacts.  For example, production
    of photochemical smog and ozone pollution depends on emissions to the
    air of nitrogen oxides together with volatile organic compounds. 
    Nitric acid, which is produced in the air by the reaction of hydroxyl
    radicals (OH*) with NO2, is one of the major components of acidic
    precipitation.  As well as being present in the gas phase, oxidized
    nitrogen can, by reaction and adsorption, become incorporated into
    aerosol particles.  Graedel et al. (1986) identified 20 inorganic
    nitrogen-containing species detectable in the atmosphere.  Near
    cities and urban regions the species usually present in greatest
    concentrations are NO and NO2, and these are the most reliably
    measured and frequently monitored nitrogen oxide species.

         Knowledge of emission patterns and concentrations of nitrogenous
    compounds is critically important for air quality planning and human
    health and environment risk assessments.  Because nitrogen oxides and

    their reaction products have lifetimes of several days in the
    atmosphere, they can be transported long distances by the wind and
    give rise to environmental impacts far from their source of emission.

    3.2  Sources of nitrogen oxides

         Combustion systems emit NO and NO2 and together these species
    are usually denoted as NOx.

         When NOx emissions are expressed in mass units, the mass is
    expressed as if all the NO had been converted to NO2.  Another
    convention adopted in some of the following sections is to report the
    emissions on a mass basis in terms of the nitrogen content.

    3.2.1  Sources of NOx emission

    3.2.1.1  Fuel combustion

         Annual production of NOx from combustion of fossil fuels is
    typically estimated from emission factors for various combustion
    processes, combined with worldwide consumption data for coal, oil and
    natural gas.  Logan (1983) provided a tabular summary of emission
    factors, which has been updated by the US National Acid Precipitation
    Assessment Program (Placet et al., 1991).  Owing to variations in
    process operating conditions, the emission factors must be considered
    to be uncertain by about ± 30%.  Table 3 provides a summary of global
    emission estimates for NOx according to fuel type.  The estimates of
    Logan (1983) are slightly higher than those of Ehhalt & Drummond
    (1982), the largest discrepancies being in emission estimates for the
    transportation sector.  The differences arise because Logan (1983)
    based estimates of emissions on fuel usage, while Ehhalt & Drummond
    (1982) scaled the totals somewhat indirectly by using world automobile
    population numbers.

         Dignon (1992) has assembled a database for mapping (with a
    resolution of one degree in latitude and longitude) and estimated
    global NOx and sulfur oxides emissions from their common principal
    anthropogenic source, i.e. fossil fuel combustion.  For 1980, the
    global total was estimated to be 22 million tonnes, as nitrogen. 
    Countries heading the list (in millions of tonnes of nitrogen per
    year) were: USA, 6.4; USSR, 4.4; China, 1.7; Japan, 0.80; and Federal
    Republic of Germany, 0.66.  An estimated 95% of NOx emissions from
    fossil fuel combustion originates in the northern hemisphere.

         For oceanic regions, shipping is a source of NOx emissions. 
    Aircraft also emit nitrogen oxides and this may be significant for the
    upper troposphere and stratosphere.

        Table 3.  Estimates of global emissions of nitrogen oxides (NOx) from combustion of fossil fuels and biomass (from: US EPA, 1993)a
                                                                                                                                              

    Source type            Annual consumption                             Emission factorsb             Global source strength
                           (106 tonnes, unless indicated otherwise)                                     (106 tonnes nitrogen/year)
                                                                                                                                              
                           (E & D)          (L)            (C et al.)     (E & D)        (L)            (E & D)         (L)        (C et al.)
                                                                                                                                              

    Fossil fuelsc

      Hard coal            2150             2696           -              1.0-2.8        2.7            3.9 (1.9-5.8)   6.4        -
      Lignite              810                             -              0.9-2.7                       1.6 (0.8-2.3)              -
      Light fuel oil       300              1.39           -              1.5-3.0        2.2d           0.7 (0.5-0.9)   3.1        -
      Heavy fuel oil       470                                            1.5-3.1                       1.1 (0.7-1.5)              -
      Natural gas          1.04             1.2 × 109 m3   -              0.6-3.0        1.9d           1.9 (0.6-3.1)   2.3        -
      Industrial sources   -                               -                                            -               1.2        -
      Automobiles          (4.1-5.4)        1.0 × 109 m3   -              0.9-1.2e       8.0d           4.3 (3.7-6.4)   8.0
                           × 1012 km

      Total                                                                                             13.5 (8.2-18.5) 19.9

    Biomass burningf

      Savanna              (6-14) × 103     2000           1200           1.0            1.7            3.1 (1.8-4.3)   3.4        2.1
      Forest clearings     (2.7-6.7) × 103  4100           2700           1.0-1.6        2.0            2.1 (0.8-3.4)   8.2        4.7
      Fuel wood            -                850            1100           -              0.5            2.0 (1-3)       0.4        0.5
      Agricultural waste   -                15             1900           -              1.6            4.0 (2-6)       0.02       3.3

      Total                                                                                             11.2 (5.6-16.4) 12.0       10.6
                                                                                                                                              

    a  Estimates according to Ehhalt & Drummond (1982) (E & D) and Logan (1983) (L). Ranges are given in parentheses.
    b  Emission factors refer to grams of nitrogen per kg of fuel consumed, unless indicated otherwise
    c  Petroleum refining and manufacture of nitric acid and cement; global emissions were obtained by scaling USA emissions for each
       industrial process
    d  Grams of nitrogen per m3 of fuel consumed
    e  Grams of nitrogen per km
    f  For biomass-burning, Crutzen et al. (1979) (C et al.) have given annual consumption rates differing somewhat from those of the other
       authors.  The data of Crutzen et al. (1979) and the resulting nitrogen oxides production rates are included for comparison
        3.2.1.2  Biomass burning

         Table 3 includes a breakdown of estimates for release of NOx
    from burning of biomass.  In natural fires and the burning of wood,
    temperatures are rarely high enough to cause oxidation of nitrogen
    molecules of the air.  The emissions are thereby more closely related
    to the fixed nitrogen content of the fuel.  Logan (1983) reviewed a
    number of experimental determinations of nitrogen emission factors
    that indicate yields are highest for grass and agricultural refuse
    fires (1.3 g nitrogen/kg fuel), less for prescribed forest fires
    (0.6 g nitrogen/kg fuel), and still lower for burning of fuel wood in
    stoves and fireplaces (0.4 g nitrogen/kg fuel).  The values roughly
    reflect differences in nitrogen content of the materials burned. 
    Biomass burning is mainly associated with agricultural practices in
    the tropics, which include plant, slash, and shift practices as well
    as natural or intentional burning of savanna vegetation at the end of
    the dry season.  Forest wildfires and use of wood as fuel make a
    lesser contribution.

    3.2.1.3  Lightning

         Thunderstorm activity has been viewed as a major NOx source
    since 1827, when Von Liebig proposed it as a natural mechanism for
    fixation of atmospheric nitrogen.  Electrical discharges in air
    generate NOx by thermal dissociation of nitrogen molecules due to
    ohmic heating inside the discharge channel and shockwave heating of
    the surroundings.  Laboratory studies by Chameides et al. (1977) and
    Levine et al. (1981) indicate an NOx yield of 6 × 1016 molecules per
    joule of spent energy.  Great uncertainties exist, however, about the
    total energy generated by lightning in the atmosphere.  Noxon (1976,
    1978) first studied the increase of NOx in the air during a
    thunderstorm.  His results provide the basis for many of the estimates
    shown in Table 4.  Reviews by Kowalczyk & Bauer (1981) Borucki &
    Chameides (1984) and Albritton et al. (1984) provide a best estimate
    of annual generation by lightning: 1 million tonnes of NOx in North
    America and 13 million tonnes globally (Placet et al., 1991).

    3.2.1.4  Soils

         The biochemical release of NOx from soils is poorly understood,
    and the flux estimates must be viewed with caution.  Both rely on the
    observations by Galbally & Roy (1978), who used the flux box method in
    conjunction with chemiluminescence detection of NOx.  They found
    average fluxes of 5.7 and 12.6 µg nitrogen/m2*h on ungrazed and
    grazed pastures, respectively, where NO was the main product.  More
    recent measurements of Slemr & Seiler (1984) indicate that the release
    of NOx from soils depends critically on the temperature and moisture

    content of the soil, which in turn complicates the estimate of the
    global emissions.  Slemr & Seiler (1984) also found an average release
    rate of 20 µg nitrogen/m2 per h for uncovered natural soils, evenly
    divided between NO and NO2. Grass coverage reduced the escape flux,
    whereas fertilization enhanced it.  Ammonium fertilizers were about
    five times more effective than nitrate fertilizers.  This suggests
    that nitrification as a source of NOx is more important than
    denitrification.  According to Slemr & Seiler (1984), an annual global
    flux of 10 million tonnes of nitrogen represents an upper limit to the
    release of NOx from soils.  Galbally et al. (1985) presented more
    detailed estimates for arid lands, and Table 4 provides a compilation
    of current literature used to develop the global budgets.  Soil is
    also a source of N2O and NH3 emissions.

         In the presence of low concentrations, plants can emit NH3,
    rather than absorb it.  This is especially true with scenescing and
    with highly fertilized plants (Grünhage et al., 1992; Holtan-Hartwig &
    Bockman 1994; Fangmeijer et al., 1994).  Release to the atmosphere of
    N2 and NO by plants has also been reported.  In some cases this was
    part of the response following exposure to nitrogen-containing
    pollutants, but other mechanisms are involved (Wellburn, 1990). NO and
    N2O are emitted in significant quantities by the soil.  The reason
    why the deposition velocity of NO is relatively low see (see Table 5)
    is partly due to the fact that the downward flux (and uptake by the
    canopy) is "mathematically" compensated by soil emissions.  In other
    words: a low deposition velocity does not always mean that the uptake
    by the vegetation is low.  In the case of N2O, soil emissions are
    mostly larger than deposition; this emission is the result of
    denitrification and is positively related to the nitrogen and water
    content and the temperature of the soil.  This is why the release of
    nitrogen from the ecosystem in the form of N2O is dependent on the
    ecosystem type, climate and land use (fertilization and water table
    height).  Skiba et al. (1992) estimated for the United Kingdom the NO
    and N2O emissions from agricultural land to be 2-6% of the nationwide
    NOx emissions and 16-64% of the N2O emissions, respectively.

        Table 4.  Global and North America natural emissions (average and range) of nitrogen oxides (NOx)
              from lightning, soils and oceans
                                                                                                          

                  Global               North America        Reference
                  (106 tonnes/year)    (106 tonnes/year)
                                                                                                          

    Lightning     8.6 (2.6-26)                              Borucki & Chameides (1984)
                  18                   1.7                  Albritton et al. (1984)
                  13 (7-26)            1 (0.3-2)            Kowalczyk & Bauer (1981); Placet et al. (1991)

    Soils         50 (as NO2)                               Lipschultz et al. (1981)
                  30 (as NO)                                Levine et al. (1984); Galbally & Roy (1978)
                  36                                        Slemr & Seiler (1984)
                                       2                    Placet et al. (1991)

    Oceans        0.35                                      Zafiriou & McFarland (1981); Logan (1983)
                                                                                                          
    
    Table 5.  Deposition velocity of nitrogen-containing gases and
              aerosols
                                                                        

                 Deposition velocity   Reference
                 (mm/second)
                                                                        

    NO2          0.1-10                Grennfelt et al. (1983);
                                       Anonymous (1991)

    NO           0.2-1                 Prinz (1982)

    NH3          12 (-5 - +40)         Grünhage et al. (1992);
                                       Sutton et al. (1993);
                                       Fangmeijer et al. (1994);
                                       Holtan-Hartwig & Bockman (1994)

    NH4+         1.4 (0.03-15)         Fangmeijer et al. (1994)
                                                                        

         Estimates of global emissions of N2O and ammonia are summarized
    in Table 6.


    Table 6.  Annual global estimates (average and range) of N2O and
              NH3 emissions to the troposphere (106 tonnes
              of nitrogen)
                                                                        

    Source             N2O          NH3      Reference
                                                                        

    Soils              10 (2-20)    15       Dawson (1977);
                                             Boettger et al. (1978)

    Ocean              26 (12-38)            Hahn (1981)

    Biomass burning    2            2-8      Crutzen et al. (1979);
                                             Crutzen (1983)

    Fossil fuels       1.6          0.2      Weiss & Craig (1976);
                                             Boettger et al. (1978)

    Fertilizer         0.1          3        Boettger et al. (1978);
                                             Crutzen et al. (1979);
                                             Crutzen (1983);
                                             Stedman & Shetter (1983)

    Domestic animals                22       Soederlund & Svensson (1976)
                                             Boettger et al. (1978);
                                             Crutzen et al. (1979);
                                             Crutzen (1983);
                                             Stedman & Shetter (1983)
                                                                        

    3.2.1.5  Oceans

         There have been few measurements of NOx, N2O or NH3 fluxes
    over the ocean, and current literature suggests that the sea is a
    negligible source of NO.  Zafiriou & McFarland (1981) observed a
    supersaturation of seawater with regard to NO in regions of relatively
    high concentrations of nitrite, owing to upwelling conditions. The

    excess NO must, in this case, arise from photochemical decomposition
    of nitrite by sunlight.  Logan (1983) estimated a local source
    strength of 1.3 × 1012 molecules/m2 per second under these
    conditions.  Linear extrapolation results in an annual global flux
    estimate of 350 000 tonnes of nitrogen.

    3.2.2  Removal from the ambient environment

         Wet precipitation and dry deposition provide two of the major
    mechanisms for removal of NOx from the atmosphere.  The addition to
    the plant soil ecosystem of nitrate (and ammonium) by rainwater
    constitutes an important source of fixed nitrogen to the terrestrial
    biosphere, and until 1930 practically all studies of nitrate in
    rainwater were concerned with the input of fixed nitrogen into
    agricultural soils.  Eriksson (1952) and Boettger et al. (1978) have
    compiled many of the available data.  Despite the wealth of
    information, it remains difficult to derive a global average for the
    deposition of nitrate, because of an uneven global coverage of the
    data, unfavourably short measurement periods at many locations, and
    inadequate collection and handling techniques for rainwater samples. 
    In addition, the concentration of nitrate in rainwater has increased
    in those parts of the world where the utilization of fossil fuels has
    led to a rise in the emissions of NOx, i.e. primarily western Europe
    and the USA.

         Dry deposition is important as a sink for those gases that are
    readily absorbed by materials covering the earth surface.  In the
    budget of NOx, the gases affected most by dry deposition are NO2 and
    HNO3.  The deposition velocity of NO is too small and the
    concentration of peroxyacetyl nitrates is not high enough for a
    significant contribution.

         According to Grennfelt et al. (1983) and Wellburn (1990), NO3-
    and HNO3 have a higher deposition velocity then NH3, but this was
    not quantified. HNO2 is assumed to have a deposition velocity equal
    to SO2: 1-30 mm/second (Table 5).

         There are several other nitrogen-containing air pollutants with
    relatively high deposition velocities.  These generally add only small
    amounts to the total nitrogen deposition, because most of the time
    their ambient concentrations are relatively low.

         Atmospheric nitrogen deposition can significantly change the
    chemical composition of the soil. In the rooting zone these changes
    have an impact on vegetation. The changes in deeper soil layers
    are particularly relevant if groundwater is used as a source of
    drinking-water. Groundwater under fertilized agricultural land can be
    heavily polluted with nitrate (and aluminium), but this is beyond the
    scope of this chapter. Due to atmospheric nitrogen deposition, the

    groundwater under forests and other non-fertilized vegetation can
    become polluted with nitrate.  For instance, in 20% of the forested
    area of the Netherlands, the nitrate concentration in phreatic
    groundwater is higher than 50 mg/litre (the EC drinking-water
    standard); in 37% it is higher than 25 mg/litre (Boumans & Beltman,
    1991).  The average annual nitrogen deposition in the Netherlands is
    45 kg/ha; approximately 10 kg/ha is from dry deposition of NOx.  The
    nitrate concentration in groundwater is strongly related to the soil
    type.  With the same atmospheric deposition, the nitrate concentration
    increases as follows: peaty soils < moderately drained sandy soils
    < well-drained rich sandy soils (Boumans, 1994).  A distinct relation
    also exists concerning the age of the trees: tree stands in Wales
    showed nitrate leaching (measured in the stream water draining the
    catchments), but only with stands older than 30 years. Younger trees
    used the nitrogen as nutrient, but the nitrogen demand of the older
    trees was lower.  The annual nitrogen deposition in that region was
    estimated to be 20 kg/ha (Emmett et al., 1993).

    3.2.3  Summary of global budgets for nitrogen oxides

         The principal routes to the production of NOx are combustion
    processes, nitrification and denitrification in soils, and lightning
    discharges.  The major removal mechanism is oxidation to HNO3,
    followed by wet and dry deposition.  In developing Table 7, the dry
    deposition velocities for NO2 over bare soil, grass and agricultural
    crops were assumed to fall in the range of 3 to 8 mm/second.  However,
    over water the velocities are significantly smaller, so that losses of
    NO2 by deposition onto the ocean surface can be ignored.  The
    absorption of nitric acid by soil, grass and water is rapid, and dry
    deposition correspondingly important, but the global flux is difficult
    to estimate because information on HNO3 mixing ratios is still
    sparse.  Logan (1983) adopted NO mixing ratios of 50 pptv over the
    oceans and 100 pptv over the continents.  The mixing ratios assumed
    for NO2 were 100 and 400 pptv, respectively.  Allowance was made for
    higher mixing ratios in industrialized areas affected by pollution. 
    Logan (1983) included the deposition of particulate nitrate over the
    oceans, using a settling velocity of 3 mm/second.  This process
    contributes 2 million tonnes nitrogen/year to a total dry deposition
    rate of 12 to 22 million tonnes nitrogen/year.

         Efforts by Boettger et al. (1978), Ehhalt & Drummond (1982),
    Galbally et al. (1985) and Warneck (1988) to quantify the sources and
    sinks have led to an improved understanding of the global budget of
    NOx, in which the flux of NOx into the troposphere and the rate of
    nitrate deposition are approximately balanced.  Ehhalt & Drummond
    (1982) relied on the detailed evaluation of data by Boettger et al.
    (1978).  Their analysis emphasized measurements from the period 1950
    to 1977, and they prepared a world map for nitrate deposition rates, 

    which were then integrated along 5° latitude belts.  Logan (1983)
    considered recent network data from North America and Europe; Galloway
    et al. (1982) reported measurements of nitrate in precipitation at
    remote locations in Alaska, South America, Australia and the Indian
    Ocean.  Both estimates gave wet nitrate deposition rates in the range
    of 2 to 14 million tonnes nitrogen/year for the marine environment and
    8 to 30 million tonnes nitrogen/year on the continents.  An earlier
    appraisal by Soederlund & Svensson (1976) led to rather similar
    values, i.e. 5 to 16 and 13 to 30 million tonnes nitrogen/year,
    respectively, although it was primarily based on Eriksson's (1952)
    compilation of data from the period 1880 to 1930.

    Table 7.  Global budget (average and range) of nitrogen oxides
              in the troposphere (from US EPA, 1993)a
                                                                        

    Type of source or sink                       Global flux
                                         (106 tonnes nitrogen/year)
                                                                        

                                            Ehhalt &       Logan (1983)
                                         Drummond (1982)
                                                                        

    Production

     Fossil-fuel combustion              13.5 (8.2-18.5)  21 (14-28)
     Biomass burning                     11.2 (5.6-16.4)  12 (4-24)
     Release from soils                  5.5 (1-10)       8 (4-16)
     Lightning discharges                5.0 (2-8)        8 (2-20)
     NH3 oxidation                       3.1 (1.2-4.9)    uncertain (1-10)
     Ocean surface (biologic)            -                < 1
     High-flying aircraft                0.3 (0.2-0.4)    -
     Stratosphere                        0.6 (0.3-0.9)    approx. 0.5

    Total production                     39 (19-59)       50 (25-99)

    Losses

     Wet deposition of NO3-, land        17 (10-24)       19 (8-30)
     Wet deposition of NO3-, oceans      8 (2-14)         8 (4-12)
     Wet deposition, combined            24 (15-33)       27 (12-42)
     Dry deposition of NOx               -                16 (12-22)

    Total loss                           24 (15-40)       43 (24-64)
                                                                        

    a  Derived from estimates according to Ehhalt & Drummond (1982)
       and Logan (1983)

         On continents, one should also consider the interception of
    aerosol particulates by high growing vegetation.  The interception of
    nitrate is expected to be particularly effective.  Hoefken &
    Gravenhorst (1982) studied the enrichment of nitrate in rainwater
    collected underneath forest canopies compared to that collected in
    open areas outside forests.  The effect is caused by the wash-off of
    dry-deposited material from foliage.  Hoefken & Gravenhorst (1982)
    found that, in a beech forest, nitrate was enhanced by a factor of
    1.4, whereas in a spruce forest enhancement by a factor of 4.1
    occurred.  Unfortunately, they were unable to differentiate between
    contributions of particulate nitrate versus gaseous nitrate to the
    total dry deposition.

         If losses of NO2 and HNO3 by dry deposition are included in the
    total budget of NOx, one obtains a reasonable balance between the
    sources and sinks, as Table 7 shows.  Ehhalt & Drummond (1982) noted
    that an appreciable part of their dry deposition is already included
    in their wet deposition rates, because rain gauges frequently are left
    open continuously, so that the collection of nitrate occurs during
    both wet and dry periods.  For NO2, they estimated a dry deposition
    rate of 7 million tonnes nitrogen/year.  Because of the uncertainty,
    they chose to include it in the error bounds and not in the mean value
    of total NOx-derived nitrogen deposition.  Clearly, the total budget
    of NOx is far from being well defined.  Moreover, in view of the
    relatively short residence times of chemical species involved in the
    NOx cycle, it is questionable whether a global budget gives an
    adequate description of the tropospheric behaviour of NOx and its
    reaction products.  Supplemental regional budgets could be more
    appropriate.

    3.3  Ambient concentrations of nitrogen oxides

         Because cities usually have an aggregation of emissions sources
    ambient concentrations of NO and NO2 tend to be greatest in cities. 
    High concentrations of NO are common in street canyons, owing to motor
    vehicle emissions.  In rural areas the emissions may have spent
    considerable time in the atmosphere and have undergone reactions to
    produce significant concentrations of other species, such as HNO3 and
    PAN.

    3.3.1  International comparison studies of NOx concentrations

         Data for monthly average concentrations of NOx collected by the
    World Meteorological Organization at five background locations in
    Europe for the period 1983 to 1985 are summarized in Fig. 3 (WMO,
    1988, 1989).  Fig. 4 presents published monthly averages of NO2 in
    1987 for 12 stations in a cooperative network under the Organisation
    for Economic Co-operation and Development (OECD) (Grennfelt et al.,
    1989).  These two figures show that concentrations of both NOx and
    NO2 tend to be higher during winter months. 

    FIGURE 4

    FIGURE 5

         Measurements of NO2 in several countries during the late 1970s
    and early 1980s are summarized in "Assessment of Urban Air Quality"
    (WHO, 1988).  The trends in composite annual averages for urban NO2
    monitoring stations in five countries are portrayed in Fig. 5 for the
    period 1975 to 1985.  The trend in the Canadian data appears to have
    been downward, but essentially stable trends were evident for data
    from the other countries.  Annual averages in the 1980-1984 period for
    42 cities around the world are summarized in the same report (WHO,
    1988).  During that period, only one city, Sao Paulo, reported an
    annual average greater than 0.053 ppm (100 µg/m3).

         Short-term peak values (1-h or 30-min maxima, or 98th or 95th
    percentile values) have been reported for 18 cities during the
    1980-1984 period (WHO, 1988).  Ten of these cities (Amsterdam,
    Brussels, Hamilton, Hong Kong, Jerusalem, Montreal, Munich, Rotterdam,
    Tel Aviv and Toronto) reported values above the WHO 1-h guideline
    level of 400 µg/m3 (0.21 ppm) for at least one year during that
    5-year period.  For eleven cities in the WHO report, both the annual
    average and a "1-hour" peak statistic were reported for the 1980-1984
    period.  Fig. 6 compares these two statistics.  It shows that three
    cities, Amsterdam, Jerusalem and Tel Aviv, reported an average peak
    value above the WHO 1-hour guideline value of 400 µg/m3 (0.21 ppm). 
    It should be kept in mind that the peak-value statistic is more
    susceptible to undetected spurious measurements than is the annual
    average.  Data from the remaining eight cities place them in the
    quadrant below the target levels for both the annual average and the
    1-hour peak.  A similar situation is seen in the majority of cities in
    the USA and is discussed in the next section.

         More recent data on NO2 trends in the world's largest cities
    have been reported by WHO/UNEP (1992) in the monograph "Urban Air
    Pollution in Megacities of the World".  Such trends for six selected
    cities from various regions of the world are illustrated in Fig. 7, a
    composite of figures extracted directly from the WHO/UNEP (1992)
    report.  In general, the overall trends appeared to be relatively
    stable for most of the cities (and/or specific neighbourhoods). 
    However, there were a few exceptions, e.g., an apparent decrease in
    the late 1980s for Bombay and an apparent increase during the same
    period for some areas of Moscow.  There are substantial differences in
    the concentrations reported for different cities.

         Table 8 summarizes emissions of nitrogen oxides and ambient
    monitoring data from the WHO/UNEP (1992) report for the years
    indicated.  Included are estimates for total emissions and percentages
    attributed to mobile sources, primarily private motor vehicles and
    public land transport systems.  However, the quality and type of
    information contained in the report is mixed, reflecting a variety of
    monitoring methods and reporting policies in different countries. 
    Ambient data in some cities was reported as NOx, and in others as
    NO2; reporting periods varied from one hour to one year.

    FIGURE 6

    FIGURE 7

    FIGURE 7a

    FIGURE 7b

    FIGURE 7c

        Table 8.  Estimated mobile and stationary source emissions of nitrogen oxides in 
              megacities (from: WHO/UNEP, 1992)a
                                                                                                    

    City                Total emissions of       Mobile source       Ambient concentration
                        nitrogen oxides          contribution        (µg/m3)
                        (tonnes/year)            (%)
                                                                                                    

    Bangkok             60 000 (1990)            30                  max 1 h NOx (as NO2)
                                                                     270 at one site; < 320 at
                                                                     three stations (1987)

    Beijing             na

    Bombay              56 000 (1990)            52                  NO2 70-85 (annual 98th
                                                                     percentile, 1990)

    Buenos Aires        27 000 (1989)            48                  na

    Cairo               24 700 (1989)            23                  NOx 380-1400 (1979,
                                                                     monthly means; single
                                                                     study)

    Calcutta            36 550 (1990)            29

    Delhi               73 000 (1990)            20                  NO2 500 (1990, 8 h)
                                                 (mostly diesel)

    Jakarta             20 500 (1989)            75                  NOx 28 (1990, annual mean)

    Karachi             50 000 (1989)            38                  38-544 (12-13 June 1988;
                                                                     single study)
                                                                                                    

    Table 8.  (Con't)
                                                                                                    

    City                Total emissions of       Mobile source       Ambient concentration
                        nitrogen oxides          contribution        (µg/m3)
                        (tonnes/year)            (%)
                                                                                                    

    London              79 000 (1983)            75 (1984)           NO2 max 1 h 867; > 600
                                                                     for 8 h; > 205 for 72 h
                                                                     (episode 12-15 Dec. 1991);
                                                                     98th percentile > 135;
                                                                     50th percentile > 50 (1989);
                                                                     NO recorded but not
                                                                     reported

    Los Angeles         440 000 (1987)           76                  NO2 max 1 h 526; > 400
                                                                     at 8 out of 24 stations (1990)

    Manila              119 000 (1990 -          90                  na
                        dubious accuracy)

    Mexico City         177 300 (1991)           75                  NO2 hourly maxima
                                                                     301-714 (1986-91)

    Moscow              210 000 (1990)           19                  NO2 max daily means
                                                                     100-150

    New York            120 000 New York         na                  NO2 1 h max 402; daily
                        City; 513 000 New                            max 160; annual mean 87
                        York metropolitan                            (1990)
                        area (1985) 

    Rio de Janeiro      63 000 (1978)            92                  na

    Sao Paulo           245 000 (1988)           82                  NO2 max 1 h
                                                                     600-1500 (1988)
                                                                                                    

    Table 8.  (Con't)
                                                                                                    

    City                Total emissions of       Mobile source       Ambient concentration
                        nitrogen oxides          contribution        (µg/m3)
                        (tonnes/year)            (%)
                                                                                                    

    Seoul               270 000 (1990)           78                  NO2 annual means only

    Shanghai            127 000 (1983);          na                  NOx annual mean 50;
                        1991 emissions                               indoor level 90
                        assumed 50%
                        higher, i.e.
                        approx. 190 000

    Tokyo               52 700 (1985)            67% from motor      daily mean 98th percentile
                                                 vehicles; 5% from   > 115 tolerable level at
                                                 ship and aircraft   25% of stations
                                                                                                    

    a  na = not available
    
         As shown in Table 8, of importance for air quality management is
    the large contribution of NOx from motor vehicles reported for some
    cities and the continuing growth in this contribution.  For example,
    emissions from vehicles in Bombay (about 29 000 tonnes per year in
    1990) are expected to increase by an additional 14 600 tonnes/year by
    the year 2000 (WHO/UNEP, 1992).

         Estimates for Jakarta attribute some three-quarters of NOx
    emissions to motor vehicles, which is comparable with London, Los
    Angeles and Mexico City.  Data from Manila indicate that some 90% of
    NOx originates from motor vehicles.

    3.3.2  Example case studies of NOx and NO2 concentrations

         Data from a range of countries and locations are given in Table 9
    (Agra, India) and Tables 10 and 11 (various cities in China).


    Table 9.  Concentrations of NO2 measured in the vicinity of the
              Taj Mahal, Agra Indiaa
                                                                        

              Year           Mean monthly concentration range (µg/m3)
                                                                        

              1987                        5.5 to 41.9
              1988                        6.3 to 33.1
              1989                        4.2 to 15.2
                                                                        

    a  Highest concentrations tend to occur in winter
       Personal communication from R.R. Khan, Ministry of Environment and
       Forests, New Delhi, India (1994)

         In urban areas in the USA, hourly patterns at fixed-site ambient
    air monitors often follow a bimodal pattern of morning and evening
    peaks, related to motor vehicular traffic patterns. Sites affected by
    large stationary sources of NO2 (or NO that reacts to produce NO2)
    are often characterized by short episodes at relatively high
    concentrations, as the plume moves to downwind areas.

         Since 1980, the annual average level among NO2-reporting
    stations in the USA has been below 0.03 ppm, with no significant
    trend evident.  This is exemplified in Fig. 8 (US EPA, 1991) by
    annual averages for the period 1980 to 1989 for 60 metropolitan
    areas subdivided into three population categories: 16 areas with a
    population of 250 000 to 500 000, 14 with 500 000 to one million, and

        Table 10.  Annual average NOX concentration (µg/m3) in China from 1981 to 1990a
                                                                                                                      

        Year      Cities all over China               Southern cities                    Northern cities
                                                                                                                      

                  Concentration    Annual             Concentration       Annual         Concentration       Annual
                  range            average            range               average        range               average
                                                                                                                      

        1981      10-90            50                 10-80               40             20-90               60
        1982      10-110           45                 10-90               40             30-110              50
        1983      9-94             46                 9-79                36             29-94               55
        1984      10-95            42                 13-75               37             10-95               46
        1985      13-49            50                 13-84               41             22-49               59
        1986      14-108           48                 14-98               41             18-108              55
        1987      17-199           56                 17-60               43             30-199              69
        1988      9-110            45                 9-110               42             8-120               48
        1989      10-140           47                 10-133              43             12-140              51
        1990      7-130            43                 12-71               38             7-130               47
                                                                                                                      

    a  General Environmental Monitoring Station of China (1991)

    Table 11.  Statistical data for the percentiles of ambient annual average NOx concentrations (µg/m3) for Chinese cities (1986-1990)a
                                                                                                                               

    Year      Number      Minimum   Percentile                            Maximum    Arithmetic              Geometric
              of cities   value                                           value                                                
                                     5    10   25   50   75   90   95                Average   Standard      Average  Standard
                                                                                               deviation              deviation
                                                                                                                               

    1986      71           14       17    20   30   43   60   81   88       108        48         22          43         488

    1987      71           13       16    21   33   46   60   74   80       105        48         20          44         478

    1988      73           8        11    18   30   43   58   67   84       120        45         22          40         547

    1989      63           10       14    19   30   44   58   64   87       140        47         26          41         546

    1990      59           7        13    17   27   38   51   71   86       130        43         23          37         554
                                                                                                                               

    a  General Environmental Monitoring Station of China (1991)
        FIGURE 8

    30 with over one million.  No group exhibited a time trend, but the
    areas with more than one million people clearly reported levels higher
    than the smaller metropolitan areas.  For 103 Metropolitan Statistical
    Areas (MSA) reporting a valid year's data for at least one station in
    1988 and/or 1989, peak annual averages ranged from 0.007 to 0.061 ppm
    (Fig. 9). The only recently measured concentrations exceeding the USA
    annual average standard (0.053 ppm) have occurred at stations in
    southern California.

         The seasonal patterns at stations in California are usually quite
    marked and reach their highest levels through the autumn and winter
    months.  Stations elsewhere in the USA usually have less prominent
    seasonal patterns and may peak in the winter or summer, or may contain
    little discernable variation (Fig. 10) (US EPA, 1991).

         One-hour NO2 values at stations in the USA can exceed 0.2 ppm,
    but in 1988 only 16 stations (12 of which are in California) reported
    an apparently credible second high 1-h value above 0.2 ppm (Fig. 11). 
    Because at least 98% of 1-h values at most stations are below 0.1 ppm,
    these values above 0.2 ppm are quite rare excursions whose validity
    should be verified (US EPA, 1991).

    3.4  Occurrence of nitrogen oxides indoors

         This section summarizes emissions of NOx from sources that
    affect indoor air quality and are commonly found in residential
    environments. There are several reasons for considering these
    emissions.  Firstly, examining emissions from several types of sources
    and source categories can help identify the relative impact of each
    source on indoor air quality and thus its influence on human exposure. 
    Secondly, such information is needed to understand the fundamental
    physical and chemical processes influencing emissions.  This
    understanding can be used to help develop strategies for reducing
    emissions.  Finally, studying emissions from indoor sources can
    provide source strength input data needed for indoor air quality
    modelling.  Knowledge of indoor concentrations is an important
    component in estimating the total exposure of individuals to nitrogen
    oxides.

         An important factor for indoor air quality is how (or if) the
    combustion products from appliances are vented outside the building. 
    It should be noted that several common types of vented appliances
    usually emit NOx to the outdoors; examples include gas-fired
    furnaces, water heaters and clothes dryers, as well as stoves and
    furnaces using wood, coal and other fuels.  Under some circumstances
    even these vented emissions may filter back inside and contribute to
    elevated NOx levels indoors.  For example, Hollowell et al. (1977)
    reported high NO and NO2 concentrations in a house where a vented
    forced-air gas-fired heating system was used.  Elevated concentrations
    may also be a problem with malfunctioning vented appliances.  Other

    FIGURE 9

    FIGURE 10a

    FIGURE 10b

    FIGURE 11

    data (e.g., Fortmann et al., 1984), however, suggest that fugitive
    emissions of NOx from vented appliances are small.  The importance of
    unvented appliances to indoor NOx levels is well documented; this
    section focuses on emissions from such appliances.

    3.4.1  Indoor sources

    3.4.1.1  Gas-fuelled cooking stoves

         Several research programmes have investigated NOx emissions
    from stoves fuelled with natural and liquid petroleum gas (Himmel &
    DeWerth, 1974; Cote et al., 1974; Massachusetts Institute of
    Technology, 1976; Yamanaka et al., 1979; Traynor et al., 1982b; Cole
    et al., 1983; Caceres et al., 1983; Fortmann et al., 1984;
    Moschandreas et al., 1985; Cole & Zawacki, 1985; Tikalsky et al.,
    1987; Borrazzo et al., 1987a).  Most of these studies have included
    investigations of several other pollutants, including CO, aldehydes
    and unburned hydrocarbons.  Table 12 lists average emission factors
    for range-top burners and for oven and broiler burners operated at
    maximum heat input rate.  Data are shown for both well-adjusted blue
    flames and for poorly adjusted yellow flames.  Each of the averages is
    based on the total number of stoves tested for that category, using
    data from the above studies.  For top burners with blue flames, a
    total of 27 values are represented; for yellow flames, there are 23
    values (24 for NOx).  Averages for the oven and broiler burners
    represent 20 blue flame and 16 yellow flame values.  Values are
    generally very similar for emissions from these two types of burners
    on the same stove.  Overall, the results show that well-adjusted blue
    flames emit more NO but less NO2 than poorly adjusted yellow flames. 
    Emission factors from range-top burners are comparable to those from
    oven and broiler burners.

    Table 12.  Average emission factors for nitric oxide (NO),
               nitrogen dioxide (NO2) and nitrogen oxides (NOx)
               from burners on gas stoves
                                                                           

                        Flame     Factor for    Factor for     Factor for
                        type      NO (µg/kJ)    NO2 (µg/kJ)    NOx (µg/kJ)
                                                                           

    Top burners         blue      20.0 ± 4.5    10.2 ± 3.1     41.0 ± 8.2
    Top burners         yellow    16.9 ± 4.5    15.0 ± 4.8     42.0 ± 9.1
    Ovens and broilers  blue      21.9 ± 6.3    7.23 ± 3.01    40.9 ± 8.6
    Ovens and broilers  yellow    19.8 ± 9.6    11.4 ± 5.7     39.0 ± 10.8
                                                                           

    3.4.1.2  Unvented gas space heaters and water heaters

         The findings of several investigators (Thrasher & DeWerth, 1979;
    Traynor et al., 1983a, 1984b; Zawacki et al., 1986) are summarized in
    Table 13.  The most significant result is the markedly lower emissions
    from heaters equipped with catalytic burners, radiant ceramic tile
    burners and improved-design steel burners (radiant and Bunsen),
    compared to emissions from simpler convection designs using
    conventional cast-iron Bunsen burners.  Equipping convective heaters
    with radiant tiles does not make much difference to emission levels,
    nor does the choice of natural gas or liquid petroleum gas fuel. 
    Other studies by Billick et al. (1984), Zawacki et al. (1984) and
    Moschandreas et al. (1985) produced similar results.

    3.4.1.3  Kerosene space heaters

         The data presented in Table 14 show that emission factors of NO
    and NO2 for radiant kerosene heaters are generally much smaller than
    those for convective kerosene heaters.  Emissions of NO from two-stage
    heaters are only slightly greater than those from radiant heaters,
    whereas emissions of NO2 are the lowest of the three heater types. 
    Most of the emissions from radiant heaters are in the form of NO2;
    for convective heaters that are two-stage heaters, the emissions of NO
    and NO2 are of comparable magnitude.  There are insufficient data
    to evaluate changes in emissions as kerosene heaters age.  Other
    products, including particles, present in these emissions may also be
    of concern for their possible health effects.

    3.4.1.4  Wood stoves

         A number of studies have examined pollutant emissions from wood
    stoves.  Some of these studies have developed emission factors based
    on concentrations in the flue gases; such information would be useful
    for assessing the contribution of wood stove emissions to ambient air
    quality.  Very little information is available, however, on fugitive
    emissions from wood stoves into the indoor living space.

         In a detailed literature survey, Smith (1987) reported that
    emissions of pollutants from wood stoves are highly variable,
    depending on the type of wood used, stove design, the way the stove is
    used and other factors.  He reported emission factors for NOx and
    other pollutants for wood stoves used in developing countries.  Many
    of these stoves are unvented, which results in excessive indoor
    concentrations as the combustion products are exhausted into the room. 
    The major health concerns for wood fires without chimneys arise from
    pollutants other than NO2, such as particulate matter.

        Table 13.  Summary of studies with unvented convective (C) and infrared (I) space heaters
                                                                                                                         

    Type of         Number    Heat input     NO emission     NO2 emission     NOx emission    Reference
    heater                    (kJ/min)       (µg/kJ)         (µg/kJ)          (µg/kJ)
                                                                                                                         

    Convective      5         86-661         24-47           2.2-7.3          39-77           Thrasher & DeWerth (1979)

    Convective      8         188-830        9.5-22          9.5-20           34-47           Traynor et al. (1983a)

    Infrared        5         245-352        0.1-1           4.1-6.2          4.9-6.2         Traynor et al. (1984b)
    Convective      4         335-626        17.8-28.7       10-18.3          40.1-57.5

    Infrared        5         264-334        0.005-1.7       1.6-4.8          2.7-5.7         Zawacki et al. (1986)
    Convective      5         176-703        5.3-44.4        7.6-23.3         27.1-76.4
                                                                                                                         

    Table 14.  Average emission factors for nitric oxide (NO), nitrogen dioxide (NO2) and nitrogen oxides (NOx) from kerosene heaters
                                                                                                                                           

    Type of heater           Heat input rate     Emission factor       Emission factor       Emission factor      Reference
                             (kJ/min)            for NO (µg/kJ)        for NO2 (µg/kJ)       for NOx (µg/kJ)
                                                                                                                                           

    Radiant, new             144                 0.45 ± 0.05           4.4 ± 0.2             5.1 ± 0.2            Leaderer (1982)
    Radiant, new             113                 0.08 ± 0.05           5.0 ± 0.2             5.1 ± 0.2
    Radiant, new             84.4                0                     5.9 ± 0.3             5.9 ± 0.3

    Convective, new          158                 17 ± 0.3              7.0 ± 0.4             33 ± 0.6
    Convective, new          97.9                12 ± 0.6              15 ± 0.3              33 ± 1.0
    Convective, new          37.3                11 ± 0.9              17 ± 1.0              34 ± 1.7

    Radiant, new             137                 1.3 ± 0.7             4.6 ± 0.8             6.6 ± 1.3            Traynor et al. (1983b)

    Radiant, 1 year old      111                 2.1                   5.1                   8.3

    Convective, new          131                 25 ± 0.7              13 ± 0.8              51 ± 1.3

    Convective, 5 years old  94.8                11 ± 0.1              32 ± 2.8              49 ± 2.8

    Radiant                  110-200             -                     -                     13 ± 1.8             Yamanaka et al. (1979)

    Convective               110-200             -                     -                     70 ± 6.8
                                                                                                                                           
             Traynor et al. (1984a) have studied wood stoves (three airtight
    and one non-airtight) used in a house.  For each experiment, airborne
    concentrations of several pollutants were measured inside and outside
    the house during operation of one of the stoves.  The results showed
    that all indoor and outdoor concentrations of NO and NO2 were
    below 0.02 ppm.  Moreover, indoor air concentrations of some other
    pollutants were high during use of the non-airtight stove.  The
    airtight stoves had little influence on indoor concentrations of any
    pollutants.  In another study, Traynor et al. (1982a) found elevated
    airborne concentrations of NO and NO2 in three occupied houses during
    operation of wood stoves and a wood furnace.  The concentrations were
    highly variable.

         Because of the limited data, it is difficult to reach
    quantitative conclusions regarding the importance of wood stoves. 
    However, the limited information available suggests that wood stoves
    are not a major contributor to indoor nitrogen oxide exposures.  This
    is consistent with the small NO emission rates expected from the low
    temperature combustion processes characteristic of wood stoves.

    3.4.1.5  Tobacco products

         A number of studies have compared concentrations of NOx and
    other pollutants in houses with smokers and houses without smokers. 
    In general, these studies have shown that concentrations are somewhat
    greater in the homes of smokers.

         A few studies have reported emissions of NOx from cigarettes
    while sampling both sidestream and mainstream smoke together. 
    Woods (1983) reported 0.079 mg NOx/cigarette, while Moschandreas
    et al. (1985) listed emissions of 2.78 mg/cigarette for NO and
    0.73 mg/cigarette for NO2.  The National Research Council (1986)
    reported total NOx emissions of 100 to 600 µg/cigarette for
    mainstream smoke, with values 4 to 10 times greater for sidestream
    smoke.  According to the report, virtually all of the emitted NOx is
    in the form of NO; once emitted, the NO is gradually oxidized to NO2. 
    Thus environments containing cigarette smoke may have higher
    concentrations of both NO and NO2 than environments without such
    smoke.  The NO2 concentration on trains travelling between Changchun
    and Harbin, China, was found to be related to the amount of cigarette
    smoking, which was greater on daytime trains than on night-time ones. 
    On a one-way daytime train the average NO2 concentration was 54 ppb
    (range, 37-84 ppb), whereas on a two-way night-time train it was
    40.6 ppb (range, 30-59 ppb) (Du et al., 1992).

    3.4.2  Removal of nitrogen oxides from indoor environments

         A number of field studies of NO2 levels in residences have
    reported that NO2 is removed more rapidly than can be accounted for
    by infiltration alone (Wade et al., 1975; Macriss & Elkins, 1977;

    Oezkaynak et al., 1982; Traynor et al., 1982a; Ryan et al.,
    1983; Leaderer et al., 1986).  Indoors, NO2 is removed by
    infiltration/ventilation and by interior surfaces and furnishings. 
    The removal of NO2 by interior surfaces and furnishings and reactions
    occurring in air is often referred to as the reactive decay rate of
    NO2, and it can be a significant factor in the actual NO2 levels
    measured in residences.  Failure to account for the reactive decay
    rate can lead to a serious underestimation of emission rate
    measurements in chamber and test house studies and a serious
    overestimation of indoor concentrations when using emission rates to
    model indoor levels.  The NO2 reactive decay rate is typically
    determined by subtracting the decay of NO2, after a source is shut
    off, from that of a relatively non-reactive gas (e.g., CO, CO2, SF6,
    NO), which can be related to ventilation rates, expressed in room
    air changes per hour.  The measured reactive decay rates in the
    above-mentioned field studies ranged from 0.1 to 1.6 air change
    times/hour.  All studies noted that the reactive decay of NO2 is as
    important and in some cases more important than infiltration in
    removing NO2 indoors.  Leaderer et al. (1986) monitored NO2, NO, CO
    and CO2 continuously in seven houses over periods ranging from 2 to
    8 days.  They reported that the NO2 decay rate was always greater
    than that due to infiltration alone and was highly variable among
    houses and among time periods within a house.

         In an effort to identify the factors that control the NO2
    reactive decay rate, a number of small chamber (Miyazaki, 1984; Spicer
    et al., 1986), large chamber (Moschandreas et al., 1985; Leaderer et
    al., 1986) and test house studies (Yamanaka, 1984; Borrazzo et al.,
    1987b; Fortmann et al., 1987) have been conducted.  The most extensive
    small chamber work was reported by Spicer et al. (1986), where 35
    residential materials were screened for NO2 reactivity in a 1.64-m3
    chamber and a limited number of the materials were tested for the
    impact of relative humidity on the reactivity rate.  Fig. 12 shows the
    relative rates of NO2 removal for the materials screened.  The figure
    indicates that many of the materials used for building construction
    and furnishings are significant sinks for NO2 and that their removal
    rate is highly variable.  Many of the materials were found to reduce a
    significant proportion of the removed NO2 to NO.  In no cases was
    NO2 re-emitted, although some materials emitted NO.  The authors
    noted that the materials that removed NO2 most rapidly fall in two
    categories: (1) porous mineral materials of high surface area; and (2)
    cellulosic material derived from plant matter.  Higher relative
    humidities were found to enhance the removal rate for some materials
    (e.g., wool carpet), reduce the removal rate for some (e.g., cement
    block), and have little effect on others (e.g., wallboard).  In a
    series of small (0.69 m3) chamber studies (Miyazaki, 1984) reactive
    decay rates for NO2 were found to vary as a function of material type
    and to increase with increasing surface area of the material, degree
    of stirring in the chamber, temperature and relative humidity.  A
    saturation effect was noted on some of the carpets tested.

    FIGURE 12

         In a series of large chamber studies (34-m3 chamber), Leaderer
    et al. (1986) evaluated the reactive decay rate of NO2 as a function
    of material type, surface area of material, relative humidity and air
    mixing.  The reactive decay rate was found to vary as a function of
    material surface roughness and surface area.  Carpeting was found to
    be most effective in removing NO2, whereas painted wallboard was
    least effective.  Increases in relative humidity were associated with
    increases in removal rates for all materials tested, but the slope was
    a shallow one.  Of particular interest is the finding in this study
    that the degree of air mixing and turbulence was a dominant variable
    in determining the reactive decay rate for NO2.  Moschandreas et al.
    (1985) evaluated six materials in a 14.5-m3 chamber and found
    variations in decay rates according to material types and a positive
    impact of relative humidity on NO2 decay rates in an empty chamber.

         Yamanaka (1984), in assessing NO2 reactive decay rates in a
    Japanese living room, found the decay to consist of both homogeneous
    and heterogeneous processes.  The rates were found to vary as a
    function of surface property and sharply as a function of relative
    humidity.  NO production during the decay was noted.  In a test house
    study, Fortmann et al. (1987) noted that the NO2 decay rate tends to
    decrease as the concentration increases.  It is not clear whether this
    is due to surface saturation or second-order kinetics.  This study
    also noted a sharp increase in NO levels during the NO2 decay,
    indicating NO production as a result of the NO2 decay. In a test
    house study conducted over a 7-month period, Borrazzo et al. (1987b)
    found that reaction rates for NO2 in the test house were sensitive to
    the location in the house where they were measured.  This indicates
    that reaction losses during transport of NO2 from room to room in a
    house may be important.

         Reactive decay of NO2 associated with interior surface materials
    and furnishings is an important mechanism for removing NO2 from the
    air within homes. Reactive decay rates for NO2 vary as a function of
    the type and surface area of the material.  The impact of relative
    humidity on the decay rate is unclear, with some studies showing a
    pronounced impact (Yamanaka, 1984), while others show only moderate or
    little impact (e.g., Spicer et al., 1986; Leaderer et al., 1986).  The
    degree of air mixing or turbulence can have an important effect on the
    reactive decay rate.  A by-product of NO2 removal by materials may be
    NO production, and a saturation effect may occur for some materials. 
    Reactive decay of NO2 in residences is highly variable between
    residences, within rooms in a residence, and on a temporal basis
    within a residence.  The large number of variables controlling the
    reactive decay rate make it very difficult to assess in large field
    studies through questionnaire or integrated air sampling.

    3.5  Indoor concentrations of nitrogen oxides

         Indoor concentrations of NO2 are a function of outdoor
    concentrations, indoor sources (source type, condition of source,
    source use, etc.), infiltration/ventilation, air mixing within and
    between rooms, reactive decay by interior surfaces, and air cleaning
    or source venting.

    3.5.1  Homes without indoor combustion sources

         Typical studies in homes without indoor sources of NO2,
    summarized in Table 15, have reported concentrations lower than
    outdoor levels due to removal from the air of NOx by the building
    envelope and interior surfaces. Thus indoor/outdoor concentration
    ratios are consistently less than unity.  These homes provide some
    degree of protection from outdoor concentrations.  Indoor/outdoor
    ratios vary considerably according to the season of the year, the
    lowest ratios occurring in the winter and highest occurring during the
    summer.  Although urban concentrations are often highest in winter,
    this pattern in the indoor/outdoor ratio, attributed to seasonal
    differences in infiltration rates, NO2 reactivity rates, the
    penetration factor and outdoor concentrations, can result in higher
    indoor concentrations in summer than in winter.  The indoor-to-outdoor
    ratio for these homes does not appear to depend on geographical area,
    housing type or outdoor concentration.   Results of monitoring in
    Portage, Wisconsin, USA, show that the presence of a gas stove
    contributes dramatically to the indoor NO2 levels.  Table 16, taken
    from the report of Quackenboss et al. (1986) and based on data
    collected in 1981 and 1982, clearly shows that gas stoves increase not
    only indoor concentrations but also the personal exposure of children.

    3.5.2  Homes with combustion appliances

         It is estimated that gas (natural gas and liquid propane) is used
    for cooking, heating water or drying clothes in about 45% of all homes
    in the USA (US Bureau of the Census, 1982) and in nearly 100% of homes
    in some other countries (e.g., the Netherlands).  Gas appliances
    (gas cooker/oven, water heater, etc.) are the major indoor source
    category for indoor residential NO2 by virtue of the number of homes
    with such sources.  NO2 concentrations in homes with gas appliances
    are higher than those without such appliances.  Within this category,
    the gas cooker/oven and unvented heaters are by far the major
    contributors.  Cookers and ovens are especially important sources when
    used inappropriately as a supplementary room heater.  Average indoor
    concentrations (based on a 1- to 2-week measurement period) in excess
    of 100 µg/m3 have been measured in some homes with gas cookers
    (Table 17).  Homes where gas cookers with pilot lights are used have

    higher NO2 levels than homes that have gas cookers without pilot
    lights.  Average NO2 concentrations in homes with gas cookers/ovens
    exhibit a spatial gradient within and between rooms. Kitchen
    concentrations of NO2 are higher than other rooms and a steep
    vertical concentration gradient in the kitchen has been observed in
    some homes, concentrations being highest nearest the ceiling.  Average
    NO2 concentrations are highest during the winter months and lowest
    during the summer months. This seasonal temporal gradient is
    attributed to differences in infiltration, appliance use, NO2
    reactivity rates and indoors and outdoor concentrations.  The impact
    of gas appliance use on indoor NO2 levels may be superimposed upon
    the background level resulting from outdoor concentrations.  Only very
    limited data exist on short-term average (3 h or less) indoor
    concentrations of NO2 associated with gas appliance use.  These data
    suggest that short-term average concentrations of NO2 are several
    times the longer-term average concentrations measured.

         A wide variety of fuel types can be used for cooking and heating
    in different localities. These can produce various effects on indoor
    air quality.  As an example, Table 18 gives data for indoor NOx
    concentrations measured at Lanzhou City, China, where coal and
    liquified gas were used in apartments and houses (Duan et al., 1992).

        Table 15.  Average outdoor concentrations of nitrogen dioxide (NO2) and average indoor/outdoor ratios in homes without gas appliances or
               unvented space heatersa
                                                                                                                                              

    Location             Housing         Averaging  Seasons         Number   Average NO2     Indoor/outdoor ratios    Reference
                         typeb           time                       of       outdoor                              
                                                                    homes    concentration
                                                                             (µg/m3)         Kitchen      Bedroom
                                                                                                                                              

    Southern California  Mixed           7 days     Summer          70       71.9            0.80         0.75        Southern California
                                                    Spring          100      43.5            0.72         0.60        Gas Company (1986)
                                                    Winter          69       91.2            0.56         0.47

    New Haven, CT        Single family   14 days    Winter          60       13.2            0.56         0.55        Leaderer et al. (1986)
                         unattached

    Albuquerque, NM      Mixed           14 days    Winter 1        60       14.1            -            0.50        Marbury et al. (1988)
                                                    Winter 2        56       19.6            -            0.32

    California           Mobile homes    7 days     Summer          46       25.9            0.61         0.54        Petreas et al. (1988)
                                                    Winter          23       44.6            0.27         0.26

    Portage, WI          Mixed           7 days     Summer          47       15.2            0.91         0.72        Quackenboss et al. (1986)
                                                    Winter          47       17.2            0.65         0.45

    Tucson, AZ           Mixed           14 days    Summer          56       19.9            0.86         0.76        Quackenboss et al. (1986)
                                                    Spring/Autumn   41       25.6            0.71         0.55
                                                    Winter          23       36.8            0.64         0.52

    Boston, MA           Mixed           14 days    Summer          117      31.7            0.76         0.75        Ryan et al. (1988)
                                                    Autumn          117      37.8            0.43         0.40
                                                    Winter/Spring   124      33.5            0.53         0.47
                                                                                                                                              

    Table 15.  (Con't)
                                                                                                                                              

    Location             Housing         Averaging  Seasons         Number   Average NO2     Indoor/outdoor ratios    Reference
                         typeb           time                       of       outdoor                              
                                                                    homes    concentration
                                                                             (µg/m3)         Kitchen      Bedroom
                                                                                                                                              

    Northern Central     Single family   5 days     Winter          9        53.8                                     Koontz et al. (1986)
    Texas                unattached

    Suffolk County,      Single family   7 days     Winter          49       35.5            0.47         -           Research Triangle
    NY                   unattached                                                                                   Institute (1990)

    Onondago County,     Single family   7 days     Winter          66       21.7            0.70         -
    NY                   unattached      

    Portage, WI          Single family   7 days     Average over    25       12.8            0.65         0.51        Spengler et al. (1983)
                         unattached                 all seasons

    Watertown, MA        Not given       3-4 days   November        18       37.0            0.65         0.51        Clausing et al. (1984)
                                                    December        10       46.0            0.39         0.30

    Middlesbrough, UK    Not given       7 days     Winter          87       35.0            0.97         0.75        Goldstein et al. (1979)

    Middlesbrough, UK    Not given       7 days     Winter          15       34.7            -            0.75        Melia et al. (1982a,b)
                                                                                                                                              

    a  Data from field studies of private residences in the USA and United Kingdom
    b  "Mixed" indicates a single family in an attached or unattached dwelling, condominium or apartment
            Table 16.  Nitrogen dioxide concentrations (ppm) according to season and 
               stove type in Portage, Wisconsin, USAa
                                                                                      

    Season       Stove          Indoor              Outdoor             Personal
                                                                                      
                             Mean      SD        Mean      SD        Mean      SD
                                                                                      

    Summer       Gas         0.016     0.006     0.006     0.003     0.014     0.004
                 Electric    0.007     0.003     0.008     0.003     0.009     0.003

    Winter       Gas         0.027     0.013     0.008     0.003     0.023     0.009
                 Electric    0.005     0.003     0.009     0.003     0.008     0.003
                                                                                      

    a  From: Quackenboss et al. (1986); SD = standard deviation
            Table 17.  Indoor and outdoor concentrations of nitrogen dioxide (NO2) in homes with gas appliances, and the calculated average
               contribution of those appliances to indoor residential NO2 levels
                                                                                                                                               

    Location    Housing   Averaging  Type of     Season    No. of  Average measured NO2              Indoor NO2 due to source       Reference
                typea     time       appliance             homes   (µg/m3)                           (µg/m3)
                          (days)                                                                                                    
                                                                   Outdoors Kitchen Bedroom  Other   Kitchen Bedroom Other Commentb
                                                                                                                                               

    USA

    Southern    Mixed      7         Oven/range,  Summer   147     75.3     91.6    68.4     -       31      12      -     1,2      Southern
    California                       ± pilot      Spring   202     49.2     79.2    51.3     -       35      22      -     1,2      California
                                                  Winter   141     104      101.5   69       -       48      20      -     1,2      Gas Company
                                                                                                                                   (1986)
                                     Oven/range,  Winter   98      107      113     76       -       53      26      -     1,2
                                     pilot

                                     Oven/range,  Winter   38      97       74      53       -       20      7       -     1,2
                                     no pilot

                                     Water heater Winter   21      92       59      50       -       11      11      -     1,2,3
                                     in home

                                     Wall furnace Winter   90      121      161     113      -       49      36      -     1,4

                                     Floor        Summer   42      119      177     126      -       66      52      -     1,4
                                     furnace
                                                                                                                                               

    Table 17.  (Con't)
                                                                                                                                               

    Location      Housing  Averaging  Type of     Season   No. of  Average measured NO2              Indoor NO2 due to source       Reference
                  typea    time       appliance            homes   (µg/m3)                           (µg/m3)
                           (days)                                                                                                   
                                                                   Outdoors Kitchen Bedroom  Other   Kitchen Bedroom Other Commentb
                                                                                                                                               

    New Haven,    Single      14      Oven/range,  Winter    42    14.8     44.7    27.6     30.4    36      20      22    1,5      Leaderer
    CT            family,             ± pilot                                                                                       et al.
                  unattached                                                                                                        (1986)

    Albuquerque,  Mixed       14      Oven/range,  Winter    82    19.1     -       33.1     41.9    -       24      31    1,5,6    Marbury et
    NM                                ± pilot      Winter    75    20.3     -       30.9     39.3    -       24      32             al. (1988)

    California    Mobile      7       Oven/range,  Summer    265   21.1     43.1    30.2     -       30      19      -              Petreas et
                  homes               ± pilot      Winter    231   42.1     53.7    37.5     -       42      27      -     1,7      al. (1988)

    Portage,      Mixed       7       Oven/range,  Summer    36    11.5     38.9    21.1     29.6    29      13      20             Quackenboss
    WI                                ± pilot      Winter    34    15.4     69.6    31.2     50.7    60      15      42    1,8      et al.
                                                                                                                                    (1986)

    Tucson,       Mixed       14      Oven/range,  Summer    13    23.1     39.1    26.3     30.7    19      8       11             Quackenboss
    AZ                                ± pilot      Spring/   11    36.3     45.8    31.9     42.4    20      12      17             et al.
                                                   Autumn                                                                           (1986)
                                                   Winter    10    45.2     60.6    43.4     50.7    32      20      25    1,9

    Boston,       Mixed       14      Oven/range,  Summer    301   41.6     65.9    45.6     50.9    33      15      19             Ryan et al.
    MA                                ± pilot      Autumn    277   43.7     74.3    47.5     52.8    56      30      34             (1988)
                                                   Winter/   298   40.5     73.5    48.6     55.1    52      30      34    1,9
                                                   Spring
                                                                                                                                               

    Table 17.  (Con't)
                                                                                                                                               

    Location      Housing  Averaging  Type of     Season   No. of  Average measured NO2              Indoor NO2 due to source       Reference
                  typea    time       appliance            homes   (µg/m3)                           (µg/m3)
                           (days)                                                                                                   
                                                                   Outdoors Kitchen Bedroom  Other   Kitchen Bedroom Other Commentb
                                                                                                                                               

    Central       Single      5       Oven/range,  Winter    22    34.6     -       -        54.1    -       -       37    1,10     Koontz et
    Texas         family,             ± pilot                                                                                       al. (1986)
                  unattached

    Suffolk Co.,  Single      7       Oven/range,  Winter    42    37.6     77.5    -        52.4    60      -       37             Research
    NY            family,             ± pilot                                                                                       Triangle
                  unattached                                                                                                        Institute
                                                                                                                                    (1990)

    Onondago      Single      7       Oven/range,  Winter    56    30.6     62.6    0        50      41      -       27    1,9
    Co., NY       family,             ± pilot
                  unattached

    New York,     Apartments  2       Oven/range   Summer    14    109      122     98       106     30      6       13             Goldstein
    NY                                             Autumn 1  15    61       96      65       71      53      22      18             et al. 
                                                   Autumn 2  9     73       108     66       76      45      15      25             (1985)
                                      ± pilot      Winter 1  8     100      121     76       95      61      16      35
                                                   Winter 2  18    75       126     63       82      81      18      37    9,11,12
                                                   Spring    13    95       121     82       99      55      16      33

    Portage, WI   Single      7       Natural gas  All       36    15.8     65.5    36.7     -       55      29      -              Spengler et
                  family,             Oven/range,  seasons                                                                          al. (1983)
                  unattached          no pilot

                                      Liquified    All       76    11.6     65.6    37.6     -       58      31      -     1,13
                                      petroleum    seasons
                                      gas
                                      Oven/range,
                                      no pilot
                                                                                                                                               

    Table 17.  (Con't)
                                                                                                                                               

    Location      Housing  Averaging  Type of     Season   No. of  Average measured NO2              Indoor NO2 due to source       Reference
                  typea    time       appliance            homes   (µg/m3)                           (µg/m3)
                           (days)                                                                                                   
                                                                   Outdoors Kitchen Bedroom  Other   Kitchen Bedroom Other Commentb
                                                                                                                                               

    Watertown,    Not given   3-4     Gas cooking  Novemb.   60    37       74      45       51      50      26      33    1,9,14   Clausing et
    MA                                             Decemb.   51    46       86      46       60      68      32      44             al. (1984)

    Netherlands

    Arnet         Not given   7       Gas cooking  Autumn/   294   35       118     -        97      97      -       37             Noy et al.
    Enschede                          no pilot     Winter                                                                           (1984)
                                      Water
                                      heaters

    Ede           Not given   7       Gas cooking  Autumn/   173   44       113     43       54      89      17      28             Noy et al.
                                      no pilot     Winter                                                                           (1984)
                                      Water
                                      heaters

    Vlagttwedde   Rural area  7       Gas cooking  Autumn/   162   28       107     24       51      90      7       34
                                      no pilot     Winter                                      Water
                                      heaters

    Rotterdam I,  Inner city  7       Gas cooking  Autumn/   228   45       144     51       80      117     24      53
                                      no pilot     Winter
                                      Water
                                      heaters

    Rotterdam II, Inner city  7       Gas cooking  Autumn/   102   45       143     64       73      117     37      46    9,17
                                      no pilot     Winter                                       Water
                                      heaters
                                                                                                                                               

    Table 17.  (Con't)
                                                                                                                                               

    Location      Housing  Averaging  Type of     Season   No. of  Average measured NO2              Indoor NO2 due to source       Reference
                  typea    time       appliance            homes   (µg/m3)                           (µg/m3)
                           (days)                                                                                                   
                                                                   Outdoors Kitchen Bedroom  Other   Kitchen Bedroom Other Commentb
                                                                                                                                               

    United Kingdom

    Middlesbrough Not given   7       Gas cooking  Winter    428   35       213     58       -       179     24      -     1,15     Goldstein 
                                      no pilot                                                                                      et al. 
                                                                                                                                    (1979)

    Middlesbrough Not given   7       Gas cooking  Winter    183   34.7     -       60       82.7    -       39      61    1,16     Melia et 
                                                                                                                                    al.
                                                                                                                                    (1982a,b)
                                                                                                                                              

    a  "Mixed" indicates a single family in an attached or unattached dwelling, condominium or apartment
    b  1.   Background correction determined by multiplying: (a) the indoor/outdoor ratio for homes in the study with no indoor NO2 sources
            for a given season; by (b) the outdoor NO2 concentration measured for the home with sources; and subtracting the product from
            the indoor level measured in the house.
       2.   Homes containing forced air gas furnace.  These homes are thought not to contribute significantly to indoor levels for this
            sample.
       3.   Homes with electric cooker/oven, forced air gas furnace, and gas water heater in home.  Comparison is made with electric
            cooker/oven, forced air gas furnace, and gas water heater located outside home.
       4.   Homes have gas cooker/oven with source contribution calculated after correction of a gas cooker/oven.  Values are background
            corrected with gas stove.
       5.   Living room or activity room.
       6.   Sampling was done over two different periods for the same houses within the same winter period.
       7.   Outdoor values were obtained from five locations, housing type, mobile home.
       8.   Other location in home; bedroom refers to average of levels in one or more bedrooms in house.
       9.   Other location in the main living room.
       10.  Other location is point nearest centre of home.

    Table 17  (Con't)

       11.  48-h samples over 30 consecutive days.
       12.  Indoor/outdoor (I/O) ratio is assessed to be 0.6, 0.7, and 0.85 for the Winter, Spring/Autumn and Summer periods,
            respectively, for all locations, because no control home (no gas appliances) mean measurements were available.  Using these
            I/O ratios, the impact of sources was calculated as footnote 1.
       13.  Each home was sampled six times over a 1-year period.
       14.  Outdoor levels are average for homes with or without gas appliances.
       15.  Outdoor levels were recorded at 75 locations in the general sampling area and were not home-specific.  Bedroom levels were
            obtained for 107 of the 428 homes.
       16.  Outdoor levels were recorded at 82 locations in the general sampling areas and were not home-specific.  Outdoor levels were
            recorded at the beginning and end of the study.
       17.  Indoor/outdoor (I/O) ratio is assumed to be 0.6 for all locations, because no control home (no gas appliances) measurements
            were available.  Using I/O ratio of 0.6, the impact of sources was calculated as in footnote 1.
        Table 18.  Indoor concentration of NOx in Lanzhou city, Chinaa
                                                                        

    Type of residence                                 Average NOx
                                                 concentration (mg/m3)
                                                                        

                                                 Winter         Summer
                                                                        

    Apartment building with central              0.141          0.059
    heating, liquified gas for cooking

    Apartment building without central           0.136          0.059
    heating, coal for cooking and heating

    One-storey house, coal for cooking           0.106          0.046
    and heating
                                                                        

    a  From: Duan et al. (1992)

    3.5.3  Homes with combustion space heaters

         Unvented kerosene and gas space heaters are important sources of
    NO and NO2 in homes, both because of the NO and NO2 production rates
    of the heaters and the long periods of time that they are in use.  The
    concentrations of NO emitted are usually several times higher than
    those of NO2. However, in the literature, indoor air concentrations
    of NO are frequently not reported.

         Field studies indicate that average residential concentrations
    (1- or 2-week average levels) exhibit a wide variation, depending
    primarily on the amount of heater use and the type of heater
    (Table 19).  Under similar operating conditions, unvented gas space
    heaters appear to be associated with higher indoor NO2 concentrations
    than kerosene heaters.  Average concentrations in homes using unvented
    kerosene heaters have been found to be well in excess of 100 µg/m3. 
    In one study (Leaderer et al., 1986), calculations of NO2
    concentrations in residences during kerosene heater use (in homes
    without gas appliances) indicate that approximately 50% of the homes
    have NO2 concentrations above 100 µg/m3 and 8% above 480 µg/m3. A
    peak NO2 concentration of 847 µg/m3 was measured over a 1-h period
    in a home during use of a kerosene heater.

    Table 19.  Two-week average nitrogen dioxide (NO2) levels for homes
               in New Haven, Connecticut, USA, during winter, 1983a
                                                                    

    Source category;                   NO2 (µg/m3)
      location                                                      

                             n      Mean        SDb   % above
                                                        100 µg/m3
                                                                    

    No kerosene heater
     or gas stove
       Outdoors              144     13.2       5.3       0
       House average         145      7.4       4.2       0
       Kitchen               147      7.6       3.7       0
       Living room           146      7.3       3.4       0
       Bedroom               145      7.3       8.6       0

    One kerosene heater,
     no gas stove
       Outdoors              95      12.9       4.6       0
       House average         95      36.8      32.8       2.1
       Kitchen               96      39.1      35.5       4.2
       Living room           96      38.5      36.6       5.2
       Bedroom               95      31.9      30.8       5.3

    No kerosene heater,
     one gas stove
       Outdoors              42      14.8       4.2       0
       House average         42      34.3      26.2       4.8
       Kitchen               42      44.7      31.4       4.8
       Living room           42      30.4      24.8       4.8
       Bedroom               42      27.8      25.1       4.8

    One kerosene heater,
     one gas stove
       Outdoors              18      14.5       5.2       0
       House average         18      66.8      43.9      16.7
       Kitchen               18      74.5      52.0      22.2
       Living room           18      57.4      38.6      11.1
       Bedroom               18      68.5      56.5      16.7

    Two kerosene heaters,
     no gas stove
       Outdoors              13      16.5       9.4       0
       House average         13      69.5      38.0      23.0
       Kitchen               13      73.0      31.7      23.0
       Living room           13      73.6      44.3      38.5
       Bedroom               13      67.8      44.9      23.1
                                                                    

    Table 19.  (Con't)
                                                                    

    Source category;                   NO2 (µg/m3)
      location                                                      

                             n      Mean        SDb   % above
                                                        100 µg/m3
                                                                    

    Two kerosene heaters,
     one gas stove
       Outdoors              3       22.1       6.2       0
       House average         3       85.8      24.4      33.3
       Kitchen               3       94.0      22.7      66.6
       Living room           3       77.6      38.4      33.3
       Bedroom               3       85.8      19.5      33.3

                                                                    

    a  From: Leaderer et al. (1986); repeat monitoring data (n = 19)
         are included
    b  SD = standard deviation

         A large field study (Koontz et al., 1986) of indoor NO2
    concentrations in Texas homes using unvented gas space heaters (most
    also had gas cookers) found that approximately 70% of the homes had
    average concentrations in excess of 100 µg/m3 and 20% had average
    concentrations in excess of 480 µg/m3.  This study found that the
    indoor/outdoor temperature difference was the best indicator of
    average indoor NO2 levels during the colder winter periods when
    heating demands are greatest.

         Only limited data have so far been published on short-term peak
    indoor concentrations for homes with unvented gas space heaters, and
    no data are available on spatial variations or concentrations solely
    during the hours of heater operation.

         No spatial gradient of NO2 was found in homes with unvented
    kerosene space heaters, contrary to the strong spatial gradient noted
    for homes with gas appliances.  This is probably due to the strong
    convective heat output and the long operating hours of the heaters,
    which result in rapid mixing within the homes.

         Ferrari et al. (1988) conducted a study of air quality in
    homes with unvented space heaters in Sydney, Australia, over
    two winters.  NO2 concentrations were measured by both continuous
    (chemiluminescence with O3 method) and passive monitors (badges and

    Palmes tubes).  Concentrations of NO2 exceeded 0.10 ppm (average
    concentration) in 85% of homes tested, and 0.16 ppm in 44% of homes.
    More than 10% of homes had average NO2 concentrations exceeding
    0.32 ppm, and the maximum recorded was greater than 0.5 ppm.  Unvented
    gas space heaters are common in Sydney, and average use is about 3 h
    per night during the winter.  As a result, an estimated 0.5 million
    residents are exposed to NO2 concentrations exceeding 0.16 ppm for
    several hours per night during the colder months of the year.

         Improper use of gas appliances (e.g., using a gas oven or
    stove to heat a living space) and improperly operating gas appliances
    or vented heating systems (e.g., out-of-repair gas cooker or
    improper operation of a gas wall or floor furnace) can be important
    contributors to indoor NO2 concentrations, but few data are available
    to assess the magnitude of that contribution.  Little or no field data
    exist that would allow for an assessment of the contributions of wood-
    or coal-burning stoves or fireplaces to indoor NO2 concentrations,
    but such a contribution would be expected to be small.  Cigarette
    smoking is expected to add relatively small amounts of NO2 to homes
    (see also Tables 15 and 18).

         In developing countries, biomass fuels (e.g., wood, biogas,
    animal dung, etc.) are much more widely used for home heating and/or
    cooking than in developed countries, these fuels often being burnt in
    open hearth fires or poorly vented appliances within indoor spaces of
    residential dwellings (WHO, 1992).  As noted by Sims & Kjellström
    (1991), a very conservatively estimated 400 million people are
    affected by biomass smoke problems worldwide, mostly in rural areas of
    developing countries.  A disproportionate number of women and young
    children are exposed, owing to the greater numbers of hours typically
    spent by them indoors and their involvement in cooking and other
    household chores.  Increased NOx concentrations, as well as greater
    concentrations of carbon monoxide, suspended particulate matter (SPM)
    and volatile organic compounds (VOCs) are normally found in biomass
    smoke (Chen et al., 1990).  Reviews of field studies in rural areas of
    developing countries indicate that exposure levels to biomass smoke
    components are usually rather high.  Indoor concentrations for NO2,
    for example, were found to fall within the range of 0.1 to 0.3 mg/m3
    in India, Nepal, Nigeria, Kenya, Guatemala and Papua New Guinea, as
    reported in reviews by WHO (1984) and Smith (1986, 1987).  Similarly,
    Hong (1991) reported NO2 concentrations in the range of 0.01 to
    0.22 mg/m3 resulting from indoor combustion of biogas in homes in
    Chengdu, Szechuan Province, China.  Hong (1991) also reported NOx
    concentrations in the range of 0.02 to 0.16 mg/m3 in other houses in
    Gansu Province, China, where dried cow dung was used as a fuel.  The
    above NO2 indoor air concentrations from biomass smoke should be
    compared with the WHO Air Quality Guidelines recommendation of
    0.15 mg/m3 for daily exposures to NO2 (WHO, 1987).

    3.5.4  Indoor nitrous acid concentrations

         Recent studies have demonstrated that substantial concentrations
    of HNO2 can be present inside residential buildings, especially when
    unvented combustion sources are used.  HNO2 is formed by the reaction
    of NO2 with water on surfaces and the reaction is promoted by high
    humidity.  HNO2 may also be produced by other mechanisms, and this is
    the subject of active research.  Brauer et al. (1993) found that HNO2
    can represent over 10% of the concentrations usually reported as NO2.

    3.5.5  Predictive models for indoor NO2 concentration

         Efforts to model indoor NO2 levels have employed two distinct
    approaches: physical/chemical and empirical/statistical models.

         The physical/chemical modelling approach has been used by
    numerous investigators in chamber, test house and small field studies
    (involving a small number of homes) to estimate emission rates of NO2
    from combustion sources (e.g., Traynor et al., 1982a; Leaderer, 1982;
    Moschandreas et al., 1984), to estimate reactive decay rates (e.g.,
    Oezkaynak et al., 1982; Yamanaka, 1984; Leaderer et al., 1986; Spicer
    et al., 1986; Borrazzo et al., 1987a), to estimate the impact of
    ventilation and mixing on the spatial and temporal distribution of
    NO2 (e.g., Oezkaynak et al., 1982; Traynor et al., 1982b; Borrazzo
    et al., 1987a), and to evaluate the applicability of emission
    rates determined under controlled conditions in estimating indoor
    concentrations of NO2 (e.g., Traynor et al., 1982b; Borrazzo et al.,
    1987a).  More recently, studies have reported the use of distributions
    of the input variables to the mass balance equation (emission rates,
    source use, decay rates, ventilation rates, etc.), determined from the
    published literature, to estimate distributions of indoor NO2 levels
    for specific sources and combinations of sources (Traynor et al.,
    1987; Hemphill et al., 1987).

         Prediction of indoor concentrations or concentration
    distributions of NO2 in homes with combustion sources using
    physical/chemical (mass-balance) models requires accurate information
    for input parameters (e.g., emission rates).  Although data are
    available for some of the input parameters under controlled
    experimental conditions, there are very limited data available
    concerning either the variability of such input parameters in actual
    homes or the factors that control variability (e.g., variability of
    emission or decay rates).  Obtaining field measurements or estimates
    of the inputs in large numbers of homes would be expensive and
    time-consuming.  Such modelling efforts do, however, help to identify
    the potential range of indoor NO2 concentrations, factors that may
    result in high levels, and the potential effectiveness of mitigation
    efforts.

         Empirical/statistical models have been developed from large field
    studies that have measured NO2 concentrations in residences and
    associated outdoor levels for time periods of a week or more.  These
    have typically used questionnaires to obtain information on sources in
    the residences, source use, building characteristics (house volume,
    number of rooms, etc.), building use, and meteorological conditions.
    In some cases, additional measurements, including temperature have
    been recorded.  Several investigators have attempted to fit simple
    regression models to their field study databases in an effort to
    determine whether the variations in NO2 levels seen among houses can
    be explained by variations in questionnaire responses.  The goal has
    been to see how well questionnaire information or easily available
    information (meteorological data) can predict indoor NO2 levels.  In
    most cases a linear model has been used, but several investigators
    have used log transformations of variables.  These employ
    questionnaire responses and measured physical data (house volume,
    etc.) as independent variables and have met with moderate success. 
    Linear regression models, with the exception of the Petreas et al.
    (1988) model, explain from 40 to 70% of the variations in residential
    NO2 levels and typically have large standard errors associated with
    their estimates.  Although log transformations of variables have
    always produced a higher percentage of explained variation due to the
    skewed distribution of the original variables, interpretation of the
    coefficients in a nonlinear model can require special attention.

         Regression models developed from field studies employing
    questionnaires to explain variations in indoor levels of NO2 have met
    with only moderate success.

         Better information, through additional measurements and better
    questionnaire design, is needed on a range of factors if the
    statistical/empirical models are to be used to estimate indoor
    concentrations of NO2 in homes without measurements.  Factors include
    source type and condition, source use, contaminant removal
    (infiltration and reactive decay) and between and within room mixing.

    3.6  Human exposure

         To assess the health impact of exposure to nitrogen oxides, it is
    essential to conduct an accurate exposure assessment.  Such data are
    of paramount importance for the definition of dose-effect and
    dose-response relationships.  In fact, the risk to human health is not
    simply determined by indoor and outdoor concentrations of nitrogen
    oxides, but rather by the personal exposure of every individual.  The
    integrated exposure is the sum of the individual exposures to oxides
    of nitrogen over all possible time intervals for all settings or
    environments.  It requires, thus, the consideration of long-term

    average concentration level, variations and short-term exposures, as
    well as the activity patterns and personal and home characteristics of
    individuals (Berglund et al., 1994).

         Exposure is a function of concentration and time.  People spend
    various periods in different types of micro-environments with various
    concentration levels.  On average, people spend about 90% of their
    time indoors (at home, work, school, etc.), about 5% in transit
    (Chapin, 1974), and 7% (range 3-12%) near smokers (Quackenboss et al.,
    1982).  These values vary with the season, day of the week, age,
    occupation and other factors but it is decidedly important to predict
    indoor pollutant levels when total exposure is being estimated.

         Adequate exposure assessment for NO2 is particularly critical in
    conducting and evaluating epidemiological studies.  Failure to measure
    or estimate exposures adequately and address the uncertainty in the
    measured exposures can lead to misclassification errors (Shy et al.,
    1978; Gladen & Rogan, 1979; Oezkaynak et al., 1986; Willett, 1989;
    Dosemeci et al., 1990; Lebret, 1990).  Early studies comparing the
    incidence of respiratory illness in children living in homes with and
    without gas stoves used a simple categorical variable of NO2
    exposure; the presence or absence of a gas cooker.  Such a dichotomous
    grouping can result in a severe non-differential misclassification
    error in assigning exposure categories.  This type of error is likely
    to underestimate the true relationship and could possibly result in a
    null finding.

         In assessing human exposure to NO2 (and other oxides of
    nitrogen), averaging times chosen should account for the type of
    effect to be expected.  With regard to NO2, the principal biological
    responses include (a) relatively transient effects on respiratory
    function associated with acute, short-term (< 1 h) exposures, and (b)
    the likelihood of increased risk for respiratory disease in children
    associated with frequently repeated short-term peak exposures and/or
    lower level long-term exposures.

         Indirect and direct methods for personal exposure assessment are
    available.  Indirect methods combine measures of concentrations at
    fixed sites in various types of micro-environments with information
    on where people have spent their time (time-activity patterns). 
    Time-weighted average (TWA) exposure models have been developed to
    estimate total personal exposure (Fugas, 1975; Duan, 1982; Duan,
    1991).  The NO2 exposure levels predicted from TWA exposure models
    have been shown to correlate closely with the exposure levels obtained
    by direct measurements of personal exposure (Nitta & Maeda, 1982;
    Quackenboss et al., 1986; Sega & Fugas, 1991).  However, the large
    variation in NO2 concentrations (distribution) within each type of

    micro-environment (because of variability in, for example, stove use,
    emission rates, ventilation frequencies, and the day-to-day and
    person-to-person variations in the use of time) decreases the accuracy
    of the predicted exposure and increases the risk for misclassification
    of the exposure.

         Direct measurements of concentrations in the breathing zone
    of a person using personal passive exposure monitors provide
    time-integrated measurements of exposure for a certain period across
    the various micro-environments where a person spends time.  It is
    important to collect exposure data over time intervals consistent with
    the expected effects.  Effects from long-term, low-level exposure may
    be different from effects from short periods of high concentration
    (intermittent peak exposure).  Intermittent peak exposure, which
    occurs during cooking on a gas stove, may be significant to total
    exposure and adverse health effects.  If effects from peak exposure
    are to be considered in the exposure assessment, the sampling time
    must be short enough to detect these peak exposures. Such a short
    sampling time is possible with the more sensitive passive samplers and
    with conventional air monitors, such as chemiluminescence NOx
    monitors.  However, direct methods of measuring personal exposure
    are relatively costly and time-consuming.  Within-person and
    between-person variability, both in personal exposure and personal
    use of time, can be large.

         Hence a sufficient number of personal exposure measurements must
    be collected for each person (repeated measurements), and a sufficient
    number of individuals must be sampled before the measurements can be
    considered to be representative.  Personal daily exposures have been
    shown to vary between individuals on the same day by a factor of up to
    about 15 in the urban area of Stockholm and between days for the same
    individual by a factor of up to 10 (Berglund et al., 1993).

         A comparison of personal NO2 exposures, as measured by Palmes
    diffusion tubes, and NO2 exposures measured in residences had a
    correlation of 0.94 for a subsample of 23 individuals (Leaderer et
    al., 1986).  Results of this comparison are depicted in Fig. 13 and
    show an excellent correlation between average household exposure and
    measured personal exposure.

         It is important to note that indoor concentrations are strong
    predictors of personal exposure.  In the case of homes with gas or
    electric stoves, personal exposure has been shown to be closely
    related to the household indoor average concentrations (Quackenboss
    et al., 1986; Harlos et al., 1987a).

    FIGURE 13

         Results of monitoring in Portage, Wisconsin, verify that the
    presence of a gas stove contributes dramatically to personal NO2
    exposure levels. Table 16, derived from the reports of Quackenboss et
    al. (1986) and based on data collected in 1981 and 1982, clearly shows
    that gas stoves increase not only indoor concentrations but also the
    personal exposure of children.

         On the other hand, outdoor concentrations, even if measured
    outside each residence, have been found to be relatively poor
    predictors of personal exposure (Quackenboss et al., 1986; Leaderer et
    al., 1986).  The association between personal exposure and outdoor
    levels of NO2 is weakest during the winter for both gas and electric
    stove groups.

         The only route of NO2 exposure is inhalation.  The dose is
    dependent on the inhalation volume and thus on physical activity, age,
    etc.  Lung absorption of NO2 is about 80-90% during rest and over 90%
    during physical activity (WHO, 1987).

         Efforts have been made to find a sufficient biological marker for
    NO2 exposure and dose.  Increased urinary excretion of collagen and
    elastin (pulmonary connective tissue) breakdown products (including
    hydroxyproline, hydroxylysine and desmosine) has been suggested as a
    marker of diffuse pulmonary injury related to inhaled NO2.  A
    significant relationship between urinary hydroxyproline excretion and
    daily NO2 exposure was found among housewives in Japan, but the
    hydroxyproline excretion fell within the normal distribution for
    healthy people (Yanagisawa et al., 1986).  The majority of the
    housewives were exposed to active or passive cigarette smoke, and this
    exposure was independently related to the excretion of hydroxyproline. 
    Other investigators have not been able to substantiate the
    relationship between urinary hydroxyproline excretion and NO2
    exposure (Muelenaer et al., 1987; Adgate et al., 1992).

         Measurements of the NO-haem protein complex in bronchoalveolar
    lavage (Maples et al., 1991) and of 3-nitrotyrosine in urine (Oshima
    et al., 1990) have been suggested as biological markers for NO2
    exposure.  The work in progress to find a suitable biological marker
    for NO2 exposure at levels found in the general environment is
    promising; however, no metabolite has yet proved to be sensitive or
    specific enough.

         Personal exposure to air pollutants can be assessed by direct or
    indirect measures.  Direct measures include biomarkers and use of
    personal monitors.  No validated biomarkers for exposure are presently
    available for NO2.

         Studies using passive monitors to measure NO2 exposures lasting
    one day to one week have been conducted in the USA (Dockery et al.,
    1981; Clausing et al., 1986; Leaderer et al., 1986; Quackenboss et
    al., 1986; Harlos et al., 1987; Schwab et al., 1990), in the
    Netherlands (Hoek et al., 1984), in Japan (Nitta & Maeda, 1982;
    Yanagisawa et al., 1984), and in Hong Kong (Koo et al., 1990).  These
    studies generally indicate that outdoor levels of NO2, although
    related to both personal levels and indoor concentrations, are poor
    predictors of personal exposures for most populations.  Average indoor
    air residential concentrations (e.g., whole-house average or bedroom
    level) tend to be the best predictor of personal exposure, typically
    explaining 50 to 80% of the variation in personal exposures.

         Indirect measures of personal exposure to NO2 employ various
    degrees of micro-environmental monitoring and questionnaires to
    estimate an individual's or population's total exposure.  One such
    model (Billick et al., 1991), developed from an extensive monitoring
    and questionnaire database, can estimate population exposure
    distributions from easily obtained data on outdoor NO2 concentrations,
    housing characteristics and time-activity patterns.  This model is
    proposed for use in evaluating the impact of various NO2 mitigation
    measures.  The model is promising, but has not yet been validated nor
    has associated uncertainty been characterized.

    3.7  Exposure of plants and ecosystems

         The sensitivity of plants to nitrogen oxides is determined both
    by their genetic characteristics and by environmental conditions.

         The relation between exposure and uptake by plants depends on
    aerodynamic and stomatal resistance and thus increases with increasing
    light intensity, wind velocity and air humidity.  After uptake, the
    response of a plant depends on its metabolic activity, and thus
    increases with poorer nutritional supply and lower temperature.

         Moreover, the sensitivity of plants depends on the general air
    pollution situation.  Emission of SO2 is often combined with NOx,
    and these compounds act synergistically.  Therefore, the impact of
    NOx may be higher in regions with elevated SO2 concentrations.  NOx
    forms part of photochemical smog.  Although ozone is the most
    phytotoxic, the contribution of NOx to adverse effects on plants is
    not negligible.

         For vegetation and ecosystems the impact of NOx is through its
    contribution to total nitrogen disposition rather than its direct
    toxicity.  Thus, other nitrogen-containing pollutants have to be taken
    into consideration.

         The dependencies of sensitivity, as summarized above, mean that
    wide variation exists in the vulnerability of different regions of the
    world.

    4.  EFFECTS OF ATMOSPHERIC NITROGEN COMPOUNDS (PARTICULARLY NITROGEN
        OXIDES) ON PLANTS

         Effects of nitrogen on ecosystems are caused through deposition
    onto soil and foliar uptake of nitrogen in various forms. Total
    effects are often difficult to separate into component effects.  This
    section, therefore, covers nitrogen inputs in all forms to ecosystems. 
    Much of the research focuses on European ecosystems, where the
    majority of the research has been conducted.  Here NHy deposition
    either dominates or is a major constituent of total nitrogen input. 
    However, this is not true for other parts of the world.  All effects
    of atmospheric nitrogen input, in whatever form, are included as
    indicators of more globally relevant effects on ecosystems but the
    reader should bear in mind local conditions of nitrogen input when
    assessing likely local consequences.

         NOx, as used in this chapter, refers to the total nitrogen
    measured by chemiluminescence detectors; this is NO2 following
    conversion to NO, and NO itself. Other nitrogen oxides may interfere
    to some extent in this method.

         Elemental nitrogen (N2) forms 80% of the atmosphere of the
    earth.  This is equivalent to about 75 × 106 kg above each hectare of
    the earth's surface. In unpolluted conditions a small fraction
    (1-15 kg nitrogen per ha per year) is converted by nitrogen-fixing
    microorganisms to biologically more active forms of nitrogen: NH4+
    and NO3-. The natural deposition of nitrogen-containing atmospheric
    compounds other than N2 is much less.  The soil contains 5 times more
    nitrogen than the atmosphere, but weathering of rock is a negligible
    source of biologically active nitrogen.  By denitrification (reduction
    of NO3- to N2 and to a lesser extent N2O, NO and NH3), 1-30 kg
    nitrogen per ha per year is recycled from the earth to the atmosphere.

         Human activities, both industrial and agricultural, have greatly
    increased the amount of biologically active nitrogen compounds,
    thereby disturbing the natural nitrogen cycle.  Various forms of
    nitrogen pollute the air, mainly NO, NO2 and NH3 as dry deposition
    and NO3- and NH4+ as wet deposition.  Another contribution
    is from occult deposition (fog and clouds).  There are many more
    nitrogen-containing air pollutants (e.g., N2O5, PAN, N2O, amines)
    but these have not been considered in this chapter, either because
    their contribution to the total nitrogen deposition is considered to
    be small or because their concentrations are probably far below the
    effect thresholds.

         Transformations of nitrogen, as it moves from the atmosphere,
    through ecosystems and back to the atmosphere, form the nitrogen
    cycle.  This is illustrated, together with anthropogenic sources of
    nitrogen, in Fig. 14.  The component processes affected by chronic
    nitrogen deposition are indicated in Fig. 15.

    FIGURE 14

    FIGURE 15

         Nitrogen-containing air pollutants can affect vegetation
    indirectly, via chemical reactions in the atmosphere, or directly
    after being deposited on vegetation, soil or water surfaces.  The
    indirect pathway is largely neglected in this chapter, although it
    includes very relevant processes, and should be taken into account
    when evaluating the entire impact of nitrogen-containing air
    pollutants: NO and NO2 are precursors for tropospheric ozone (O3),
    which acts both as a phytotoxin and a greenhouse gas.  The effects of
    O3 will be discussed in a forthcoming Environmental Health Criteria
    monograph.  N2O contributes to the depletion of stratospheric O3,
    resulting in increasing ultraviolet radiation.  This and other aspects
    of global climate change will be evaluated in a WHO/WMO/UNEP document
    entitled "Climate and Health: potential impacts of climate change". 
    The direct impact of airborne nitrogen is due to toxic effects,
    eutrophication and soil acidification.  One effect of soil
    acidification is that aluminum enters into solution, hence increasing
    its bioavailability.  This result in root damage.  Aluminum toxicity
    will be discussed in a further Environmental Health Criteria
    monograph.

         Most biodiversity is found in (semi-)natural ecosystems, both
    aquatic and terrestrial.  Nitrogen is the limiting nutrient for plant
    growth in many (semi-)natural ecosystems.  Most of the plant species
    from these (semi-)natural habitats are adapted to nutrient-poor
    conditions, and can only compete successfully in soils with low
    nitrogen levels (Chapin, 1980; Tamm, 1991). Ellenberg (1988b) surveyed
    the nitrogen requirements of 1805 plant species from Germany and
    concluded that 50% can compete successfully only in habitats that are
    deficient in nitrogen.  Furthermore, of the plants threatened by
    increased nitrogen deposition, 75-80% are indicator species for
    low-nitrogen habitats.  When stratified by ecosystem type, it is also
    clear that the trend of rare species occurring with greater frequency
    in nitrogen-poor habitats is a common phenomenon across many
    ecosystems (Fig. 16 and Fig. 17).  Plant species threatened by high
    nitrogen deposition are common across many ecosystem types (Ellenberg,
    1988b).  The critical loads for nitrogen depend on (i) the type of
    ecosystem; (ii) the land use and management in the past and present;
    and (iii) the abiotic conditions (especially those which influence the
    nitrification potential and immobilization rate in the soil).  The
    impact of increased nitrogen deposition upon biological systems can be
    the result of direct uptake by the foliage or uptake via the soil. 
    The most relevant effects at the level of individual plants are injury
    to the tissue, changes in biomass production and increased
    susceptibility to secondary stress factors.  At the vegetation level,
    this results in changes in competitive relationships between species
    and loss of biodiversity.

    FIGURE 16

    FIGURE 17

         Effects on individual plants are discussed in section 4.1.  The
    following natural ecosystems are treated in detail in section 4.2:
    freshwater ecosystems (shallow soft-water bodies, lakes and streams)
    and terrestrial ecosystems (wetlands and bogs, species-rich
    grasslands, heathlands and forests).  Estuarine and marine systems
    are also considered.

         Air quality guidelines refer to thresholds for adverse effects. 
    Two different types of effect thresholds exist: critical levels and
    critical loads.

         The critical level is defined as:

         the concentration in the atmosphere above which direct adverse
         effects on receptors, such as plants, ecosystems or materials,
         may occur according to present knowledge.

         The critical load is defined as:

         a quantitative estimate of an exposure (deposition) to one or
         more pollutants below which significant harmful effects on
         specified sensitive elements of the environment do not occur
         according to present knowledge.

         Generally, critical levels for nitrogen-containing air pollutants
    are expressed in terms of exposure (µg/m3 and exposure duration),
    while critical loads are expressed in terms of deposition (kg nitrogen
    per ha per year).  Both critical level and load are intended to be
    set so as to protect vegetation, and can be converted into each
    other knowing the deposition velocity.  Thus, it might seem to be
    superfluous to assess both critical levels and loads.  However, with
    the currently accepted approach, critical levels and loads are more or
    less complementary: critical levels focus on effect thresholds for
    short-term exposure (1 year or less), while critical loads focus on
    safe deposition quantities for long-term exposure (10-100 years):
    critical levels are not aimed to protect plants completely against
    adverse effects.  No-observed-effect concentrations (NOECs) are
    usually lower than critical levels.  For instance, a critical level
    can be set at 5% yield reduction. Thus, owing simply to differences in
    definition, the critical level is generally higher than the critical
    load (Fig. 18b).

         In current practice there are other differences between critical
    levels and loads: critical levels give details on individual compounds
    and focus on responses on plant level, while critical loads cover all
    nitrogen-containing compounds and focus on the vegetation or ecosystem
    level.  In other words: critical loads focus on functioning of the
    ecosystem, while critical levels focus on protection of the relatively
    sensitive plant species.

         In the critical level concept, the different nitrogen-containing
    compounds are evaluated separately, because of their differences in
    phytotoxic properties, even when their load in terms of kg nitrogen
    per ha per year is the same (Ashenden et al., 1993).  Another
    difference between critical level and critical load is that critical
    level considers the possibility of more- or less-than-additive effects
    (Wellburn, 1990), while in the critical load concept additivity of
    nitrogen-containing or acidifying compounds is presumed.  Moreover,
    nitrogen-containing air pollutants have their impact not only because
    of their contribution to the nitrogen supply.  Sometimes other effects
    seem to dominate.  For instance, although occult deposition is
    generally small in terms of nitrogen deposition, it may be of great
    significance because of its ability to affect plant surfaces.

         It was concluded for these reasons that both critical levels and
    loads are necessary within the scope of air quality guidelines for
    nitrogen-containing compounds.

         Assessing effect thresholds is relatively simple in the case of
    toxic compounds with an exposure/response relationship which follows
    the usual sigmoid curve: the lowest exposure level that results in a
    response that is significantly different from the control treatment is
    the effect threshold.  However, this approach is essentially invalid
    for exposure of nitrogen-limited vegetation to nitrogen-containing air
    pollutants.  Nitrogen is a macro-nutrient and so each addition of
    nitrogen can result in a physiological response: growth stimulation
    gradually increases with higher exposure levels and changes in growth
    inhibition at higher levels (Fig. 18a).  Moreover, depending on the
    definition of adverse effects, the status of the vegetation may not be
    optimal at background levels (Fig. 18b).  These features complicate
    the assessment of effect thresholds for nitrogen-containing compounds. 
    Nevertheless, in this chapter effect thresholds are presented,
    according to current practice.

    4.1  Properties of NOx and NHy

         In this section general information is initially presented on
    uptake, detoxification, metabolism and growth aspects.  In the
    following subsections the data determining the critical levels for
    individual compounds and mixtures are discussed. The relevance of this
    information and possibilities for generalization are discussed in
    sections 4.1.8 and 4.1.9, where the critical levels are estimated. 
    Deposition on and emission from soils and vegetation is discussed in
    chapter 3.

    FIGURE 18

    4.1.1  Adsorption and uptake

         The impact of a pollutant on plants is determined by its
    adsorption, rate of uptake (flux) and the reaction of the plants. 
    Foliar uptake is probably dominant for NO, NO2 (Wellburn, 1990) and
    NH3 (Pérez-Soba & Van der Eerden, 1993), while the pathway via soil
    and roots is the major route for nitrogen-containing pollutants in wet
    deposition.

         The flux of the compounds from the atmosphere into the plant is a
    complicated process, which is highly dependent on the properties of
    both plant and compound and on environmental conditions.  This is why
    deposition velocities proved to be highly variable (chapter 3).

         The flux from the atmosphere to the leaf surface (and soil)
    depends on the aerodynamic and boundary layer resistances, which
    are determined by meteorological conditions and plant and leaf
    architecture.  The flux from the leaf surface to the final site of
    reaction in the cell is determined by stomatal, cuticular and
    mesophyll resistance.  The reaction of the plant to the nitrogen that
    arrives at the target side is dependent on the intrinsic properties of
    the plant and on its nutritional status, and again on environmental
    conditions.

         The flux of atmospheric nitrogen through the soil is conditioned
    by properties of soil and vegetation and by meteorological conditions. 
    The chemical composition of soil water, the rate of nitrification
    (NH4+ -> NO3-), the preference of the plant for either NH4+ or
    NO3-, the root architecture and the metabolic activity of the plants
    play major roles in this uptake (Schulze et al., 1989).

         Adsorption on the outer surface of leaves certainly takes
    place.  Exposure to relatively high concentrations of gaseous NH3
    (180 µg/m3) or NH4+ in rainwater (5 mmol/litre) damages the
    crystalline structure of the epicuticular wax layer of the needles of
     Pseudotsuga menziesii (Van der Eerden & Pérez-Soba, 1992).  NO2
    (Fowler et al., 1980) and NH4+ and NO3- in wet and occult
    deposition can disturb leaf surfaces in several ways (Jacobson, 1991). 
    The quantitative relevance of this effect for the field situations has
    not yet been shown in detail.

         Uptake of NH3 and NOx is driven by the concentration gradient
    between atmosphere and mesophyll.  It is generally directly determined
    by stomatal conductance and thus depends on factors influencing
    stomatal aperture.  Although in higher plants uptake through the
    stomata strongly dominates, there are indications that penetration
    through the cuticle is not completely negligible.  This has been
    demonstrated for NO and NO2 (Wellburn, 1990).  While stomata greatly
    influence the foliar uptake of aerial nitrogen compounds, many of
    these compounds subsequently alter stomatal aperture and the extent of

    further uptake.  The nitrogen status of plants is also known to affect
    stomatal behaviour towards other environmental conditions such as
    drought (Ghashghaie & Saugier, 1989).

         The flux of NH3 into a plant appeared to be linearly related to
    the atmospheric concentration (Van der Eerden et al., 1991), there
    being no mesophyll resistance (Van Hove et al., 1989).  This relation
    can become less then linear with high concentrations, e.g., some
    hundreds of µg/m3 (Wollenheber & Raven, 1993).  Mesophyll resistance
    is, however, probably more significant for NO and NO2 (Capron et al.,
    1994).

         There is increasing evidence that foliar uptake of nitrogen
    reduces the uptake of nitrogen by the roots (Srivastava & Ormrod,
    1986; Pérez-Soba & van der Eerden, 1993), although the driving
    mechanism is not yet clear.

         In the presence of low concentrations plants can emit NH3,
    rather than absorb it (chapter 3).  NO and N2O are emitted in
    significant quantities by the soil (chapter 3).

         Rain, clouds, fog and aerosols always contain significant amounts
    of ions including NH4+ and NO3-.  In the past, foliar uptake of
    nitrogen from wet deposition was considered to be negligible, but
    recent research using 15N and throughfall analysis shows that this
    path can contribute a high proportion of the total plant uptake (see
    Pearson & Stewart, 1993, and section 2.4).  In general, cations (e.g.,
    NH4+) are more easily taken up through the cuticle than anions
    (e.g., NO3-).  A substantial foliar uptake of NH4+ from rainwater
    has been measured in several tree species (Garten & Hanson, 1989). 
    Lower plants, such as bryophytes and lichens do not have stomata and a
    waxy waterproof cuticle, and thus the potential for direct uptake of
    pollutants in the form of wet or dry deposition is greater.  Epiphytic
    lichens are active absorbers of both NH4+ and NO3- (Reiners &
    Olson, 1984).  Uptake and exchange of ions through the leaf surface is
    a relatively slow process, and thus is only relevant if the surface
    remains wet for long periods.

    4.1.2  Toxicity, detoxification and assimilation

         One would expect a positive relationship between the solubility
    of a compound and its biological impact.  NO is only slightly soluble
    in water, but the presence of other substances can alter its
    solubility.  NO2 has a higher solubility, while that of NH3 is much
    higher.

         Much information exists on mechanisms of toxicity, although it is
    sometimes confusing.  NO2, NO, HNO2 and HNO3 can be incorporated
    into nitrogen metabolism using the pathway: NO3- -> NO2- ->

    (NH3 <--> NH4+) <--> glutamate -> glutamine -> other amino
    acids, amides, proteins, polyamines, etc.  The enzymes involved
    include nitrate reductase (NR), nitrite reductase (NiR) and glutamine
    synthetase (GS).  Glutamate dehydrogenase (GDH) plays a role in the
    internal cycling of NH4+.

         After exposure to NO2, nitrate can accumulate for some weeks;
    accumulation of nitrite is rarely reported, although it is certainly
    an intermediate.  Nitrite levels can be elevated for some hours due to
    the fact that NR activity is induced faster than that of NiR.  In many
    cases storage of excess nitrogen has been found to be in the form of
    arginine (Van Dijk & Roelofs, 1988), which could last months or
    longer.

         NO2-, NH3 and NH4+ are highly phytotoxic, and could well be
    the cause of adverse effects of nitrogen-containing air pollutants. 
    Wellburn (1990) suggested that the free radical *N=O plays a role in
    the phytotoxicity of NOx.

         High concentrations can cause visible injury via lipid breakdown
    and cellular plasmolysis.  At lower concentrations inhibition of lipid
    biosynthesis may dominate, rather than damage of existing lipids
    (Wellburn, 1990).

         Raven (1988) assumed that the adverse effects of nitrogen-
    containing compounds are due to their interference with cellular
    acid/base regulation.  They can influence cellular pH both before
    and after assimilation.  Assimilation of most air pollutants,
    including NH3, has been shown to result in production of protons
    (Wollenheber & Raven, 1993).  Assimilation of nitrate and a high
    buffer capacity can prevent the plant from being damaged by this
    acidification (Pearson & Stewart, 1993).  If these adverse effects can
    effectively be counteracted, assimilation of nitrogen-containing
    compounds will result in growth stimulation.

         Synergistic effects have been found in nearly all studies
    concerning SO2 and NO2 (Wellburn et al., 1981).  Inhibition of NiR
    by SO2, resulting in the inability of the plant to detoxify nitrite,
    might be the cause of this interaction.

    4.1.3  Physiology and growth aspects

         When climatic conditions and nutrient supply allow biomass
    production, both NOx and NHy result in growth stimulation at low
    concentrations and growth reduction at higher concentrations. 
    However, the exposure level at which growth stimulation turns into
    growth inhibition is much lower for NOx than for NHy (see Fig. 18a).

         Foliar uptake of NH3 generally results in an increase in
    photosynthesis and dark respiration, and in the amount of RUBISCO
    (ribulose 1,5-biphosphate carboxylase oxygenase) and chlorophyll. 
    Some authors have shown that stomatal conductance increases and the
    stomata remain open in the dark, resulting in enhanced transpiration
    and drought sensitivity (Van der Eerden & Pérez-Soba, 1992).  Most
    experiments with NO and NO2 have been conducted with relatively high
    concentration levels (> 200 µg/m3).  These experiments show
    inhibition of photosynthesis by both NO and NO2, possibly additively
    (Capron & Mansfield, 1976).  Inhibition by NO may be stronger than
    that of NO2 (Saxe, 1986).  The threshold for this response is well
    below the threshold for visible injury (Wellburn, 1990) and
    transpiration (Saxe, 1986).  With lower (nearer to ambient) NOx
    concentrations, stimulation of photosynthesis may well occur.  Both
    NOx and NHy generally cause an increase in shoot/root ratio.  The
    specific root length and the amount of mycorrhizal infection can be
    reduced by both compounds. However, these alterations in root
    properties resemble general responses to increased nitrogen nutrient
    supply.

    4.1.4  Interactions with climatic conditions

         Evidence suggests that exposure of vegetation to NH3 and to
    mixtures of NO2 and SO2 can influence the subsequent response to
    drought and frost stress.  There is also evidence that environmental
    conditions can affect the response to NOx and to NH3.

         The foliar uptake of nitrogenous compounds in the form of wet and
    occult deposition is largely via the cuticle.  Uptake and exchange of
    ions through the leaf surface is a relatively slow process, and thus
    is especially relevant if the surface remains wet for longer periods,
    e.g., in regions where exposure to mist and clouds is frequent.

         The solubility of most gases, including NO, NO2 and NH3, rises
    as temperature falls, while the metabolic activity of plants and thus
    their detoxification capacity is lower.  On the other hand, stomatal
    conductivity and thus the influx of gases generally falls as
    temperature falls.

         Guderian (1988) proposed a lower critical level in winter than
    for the whole year, in acknowledgement of several results that
    indicate greater toxicity of NO2 during winter conditions.  For
    example, exposure of  Poa pratensis in outdoor chambers to 120 µg/m3
    inhibited growth during winter but had little effect or even
    stimulated growth in summer and autumn (Whitmore & Freer-Smith, 1982). 
    Mortensen (1986) found that low light and non-injurious low
    temperature conditions can enhance the toxicity of NOx.  Capron et
    al. (1991) found that the depression relative to the control of net
    photosynthesis by 1250 µg NO/m3 plus 575 µg NO2/m3 at 10°C was
    three times, and at 5°C was almost five times, that recorded at 20°C.

    An interaction between NOx and temperature may also occur at lower
    realistic concentrations.  This is suggested by the observation of
    nitrite accumulation at low temperatures during fumigation of
     Picea rubra with 38 µg NO2/m3 plus 54 µg SO2/m3 (Wolfenden et
    al., 1991).  This temperature effect may play a role in combination
    with elevated concentrations of CO2 as well: the stimulating effect
    of CO2 on net photosynthesis was inhibited by NOx to a larger extent
    when applied at lower temperature (Capron et al., 1994).  Observation
    of NH3 injury to plants also indicates that this is greatest in
    winter (Van der Eerden, 1982).

         In contrast with the view that NOx (and NH3) injury is greater
    at low temperatures, Srivastava et al. (1975) found that inhibition by
    NOx of photosynthesis was greatest under optimal temperature and high
    light conditions, when stomatal conductance to the gas would be
    highest.

         The exposure of plants to NOx and NH3 may reduce their ability
    to withstand drought stress, owing to loss of control of transpiration
    by stomata and to an increase in the shoot/root ratio (Lucas, 1990;
    Atkinson et al., 1991; Fangmeijer et al., 1994).

    4.1.5  Interactions with the habitat

         Whether the atmospheric input of nitrogen has a positive or
    negative impact depends on the plant species and habitat.  Based on
    experimental evidence, Pearson & Stewart (1993) hypothesized that
    species which are part of a climax vegetation on nutrient-poor acidic
    soils are often relatively sensitive to NOx and NHy.  Morgan et al.
    (1992) found that NOx disrupted the NR activity to a greater extent
    in calcifuge than calcicole moss species.  Ombrotrophic mires and
    other strongly nitrogen-limited systems may be especially prone to
    detrimental effects from input of nitrogen-containing air pollutants.

         The assimilation of low concentrations of NO2 and the
    incorporation into amino acids by NR (Morgan et al., 1992; Table 20)
    are indicators that this pollutant can contribute to the nitrogen
    budget of plants (Pérez-Soba et al., 1994).  The contribution of NOx
    to the nitrogen supply increases as root-available nitrogen is lowered
    (Okano & Totsuka, 1986; Rowland et al., 1987).  Srivastava & Ormrod
    (1986) observed reduced ability to respond to a supply of nitrate to
    the roots when  Hordeum vulgare was fumigated with NO2.  Similarly,
    Pérez-Soba & Van der Eerden (1993) found reduced uptake of NH4+ from
    the soil when  Pinus sylvestris was fumigated with NH3.  Although
    there is much evidence that nitrogen-containing air pollutants play a
    role in the nitrogen demand and nitrogen metabolism of the plant,
    Ashenden et al. (1993) found no obvious relationship between
    sensitivity to NO2 and nitrogen preference, as indicated by Ellenberg
    (1985).

    4.1.6  Increasing pest incidence

         Any change in chemical composition of plants due to the uptake of
    nitrogenous air pollutants could alter the behaviour of pests and
    pathogens.  Evidence indicates that, in general, NOx and NHy
    increase the growth rate of herbivorous insects (Dohmen et al., 1984;
    Flückiger & Braun, 1986; Houlden et al., 1990; Van der Eerden et al.,
    1991).  This may also apply to fungal pathogens (van Dijk et al.,
    1992).

    4.1.7  Conclusions for various atmospheric nitrogen species and
           mixtures

    4.1.7.1  NO2

         In Table 20 the lowest effective exposure levels for NO2 are
    given.  Three different types of effects are considered:

    *    (bio)chemical: e.g., enzyme activity, consumption quality
    *    physiological: e.g., CO2 assimilation, stomatal conductivity
    *    growth aspects: e.g., biomass, reproduction, stress sensitivity

         Four exposure durations are used in this table.  These are
    (including an indication of the exposure durations and the margins):

    *    short term (hours): < 8 h
    *    air pollution episodes (days): 8 h to 2 weeks
    *    growing season or winter season (months): 2 weeks to 6 months
    *    long term (years): > 6 months

         To avoid the information being too selective, in each cell in
    this table a species is used only once.  For each cell the three
    lowest effective concentrations and exposure durations are given;
    species and references are mentioned in footnotes.  Exposure levels
    far higher than current levels measured in the field situation have
    not been considered.

         The fact that not all cells in Table 20 are filled with
    information is because many of the experiments have been conducted
    with unrealistically high concentrations. The majority of observations
    mentioned in the table are on crops; several of these show growth
    stimulation.  Most of the responses on a biochemical level deal with
    enhanced NR activity, which shows that the plants are capable of
    assimilating the NO2.  A general effect threshold as derived from
    Table 20 would be substantially higher if enhanced NR and biomass
    production of crops were not assumed to be an adverse effect. 
    However, growth stimulation is often considered an adverse effect in
    most types of natural vegetation.  Moreover, Pearson & Stewart (1993)

    assumed detoxification of NHy and NOx to be a potentially adverse
    effect, because it contributes to cellular acidification, which can
    not always be counteracted.

    4.1.7.2  NO

         In Table 21 the lowest effective exposure levels for NO are
    given.

         Most research into the effects of nitric oxide has been based on
    glasshouse crops, particularly the tomato  (Lycopersicon esculentum).
    Virtually all experiments deal with photosynthesis or enzymatic
    reactions and a few growth aspects were measured.  The existing data
    are rather difficult to interpret since controlled fumigation with NO
    inevitably results in some oxidation to NO2.  Thus atmospheres will
    contain a mixture of the oxides.  There is growing interest in the
    distinct properties and effects of NO and NO2, and the mechanisms of
    their cellular action probably differ (Wellburn, 1990).  The exchange
    properties of NO and NO2 over vegetation (personal communication by
    D. Fowler to the IPCS) and single plants (Saxe, 1986) appear quite
    different.  Their effects are also contrasting, and there is clearly
    some dispute over which oxide is the most toxic.  Earlier studies of
    the inhibition of photosynthesis found NO to act more rapidly than
    NO2 (at several ppm) but to cause less overall depression of the
    photosynthetic rate (Hill & Bennet, 1970).  More recent photosynthetic
    studies by Saxe (1986), using similar concentrations, found NO to be
    considerably more toxic than NO2.  There is very little information
    on contrasting effects of the two oxides at low concentrations, but
    this also adds weight to the suggestion that NO is biologically more
    toxic.  In her studies of NR in bryophytes, Morgan et al. (1992)
    discovered that exposure to NO initially inhibited NR while NO2
    induced activity.  At present, however, there is insufficient
    knowledge across a range of species to establish separate critical
    levels for NO and NO2, and studies using a wider variety of
    vegetation are urgently required.

    4.1.7.3  NH3

         The lowest effective exposure levels for NH3 are given in Table
    22.

         The toxicity of NH3 during very short exposure periods has been
    tested for the purpose of evaluating accidental releases during
    transport or industrial processes.  The estimated critical level for
    10 min is (100 ppm) (personal communication by Lee & Davison to the
    IPCS).  This type of exposure is out of the context of this monograph. 


        Table 20.  Lowest exposure levels (in µg/m3) and durations at which NO2
               caused significant effectsa
                                                                                      

                     (Bio)chemical     Physiological            Growth aspects
                                                                                      

    Long term                                                   200 (130); 104 h/week;
                                                                7 monthsr
                                                                120-500; 9.5 monthss
                                                                122; 37 weekst

    Growing season   50; 39 daysb      120; 22 daysj            10-43; 130 daysu
    or winter        125; 140 daysc    190 (65); 105 h          55-75; 62 daysv
                     940; 19 daysd     in 15 daysk              150-190 (28-33);
                                                                120 h in 40 daysw

    Air pollution    140; 1 daye       375 (165); 35 h in       375; 2 weeksx
    episodes         160; 7 daysf      5 daysl  190; 3 daysm    100 (25);
                     65; 1 dayg        375 (165); 35 h          20 h in 5 daysy
                                       in 5 daysn

    Short term       7500, 6 hh        940; 1 ho                2000-3000; 3.5 hz
                     7500; 4 hi        850; 7 hp
                                       1100; 1.5 hq
                                                                                      

    a  If the fumigation was not continuous an average value has been estimated
       and quoted in parentheses (calculated assuming 10 µg/m3 during the periods
       in which the fumigation was switched off).
    b   Pinus sylvestris; changes in amino acid composition, with no physiological
       changes (Näsholm et al., 1991)
    c   Lolium perenne; increase in GDH activity (Wellburn et al., 1981)
    d   Lycopersicum esculentum; decrease in nitrate content of the leaves (Taylor
       & Eaton, 1966)
    e   Picea rubens, increase in NR activity (Norby et al., 1989)

    Table 20  (Con't)

    f   Pinus sylvestris, increase in NR activity (Wingsle et al., 1987)
    g  Several bryophyte species; increase in NR activity (Morgan et al., 1992)
    h   Zea mais; increase in NiR activity (Yoneyama et al., 1979)
    i   Vicia faba; change in amino acid composition (Ito et al., 1984)
    j   Betula sp; increased water loss (Neighbour et al., 1988)
    k   Phaseolus vulgaris; reversible increase in dark respiration (Sandhu & Gupta, 1989)
    l   Glycine max; increase in photosynthesis (Sabarathnam et al., 1988a,b)
    m   Phaseolus vulgaris; increase in transpiration (Ashenden, 1979)
    n   Glycine max; enhanced dark respiration (Sabarathnam et al., 1988b)
    o   Vicia faba; reversible structural damage on cellular level (Wellburn et al., 1972)
    p   Pisum sativum; emission of stress ethylene (Mehlhorn & Wellburn, 1987)
    q   Medicago sativa, Avena sativa; inhibition of photosynthesis (Hill & Bennet, 1970)
    r  Several grass species; reduction in shoot growth (Whitmore & Mansfield, 1983)
    s   Citrus sinensis; increased fruit drop (Thompson et al., 1970)
    t   Polytrichum formosum and 3 fern species; injury and changes in growth (Ashenden
       et al., 1990; Bell et al., 1992)
    u   Brassica napus and  Hordeum vulgare; growth stimulation (resp.: Adaros et al.,
       1991a,b)
    v   Phaseolus vulgaris; increase in total dry matter, not in yield (Bender et al.,
       1991)
    w   Raphanus sativus; growth stimulation (Runeckles & Palmer, 1987)
    x   Helianthus annuus; reduction in net assimilation rate (Okano et al., 1985b)
    y   Pinus strobus; slight needle necrosis in 2 of 8 clones (Yang et al., 1983)
    z   Nicotiana tabacum; leaf necrosis (Bush et al., 1962)
        
    Table 21.  Lowest exposure levels (in µg/m3) at which NO caused
               significant effectsa
                                                                         

                      (Bio)chemical     Physiological     Growth aspects
                                                                         

    Growing season    44; 21 daysb                        625; 16 daysn
                      500; 28 daysc                       500;o

    Air pollution     375; 8 daysd      1250; 4 daysi     1250; 5 daysp
    episodes          44; 8-24 he       125; 20 hj
                      1875; 18 hf

    Short term        188; 7 hg         750; 1 hk
                      500; 3 hh         2500; 10 minl
                                        1875; 20 minm
                                                                         

    a  If the fumigation was not continuous an average value has been
       estimated and quoted in parentheses (calculated assuming 10 µg/m3
       during the periods in which the fumigation was switched off).
    b  Four bryophyte species; inhibition of nitrate-induction of NR
       (Morgan et al., 1992)
    c   Lycopersicon esculentum; induction of NiR (Wellburn et al., 1980)
    d   Lactuca sativa; induction of NiR (Besford & Hand, 1989)
    e   Ctenidium molluscum (bryophyte); inhibition of NR (Morgan et al., 1992)
    f   Capsicum annum; reduction in NiR activity (Murray & Wellburn, 1980)
    g   Pisum sativum; increase in ethylene release (Mehlhorn & Wellburn, 1987)
    h   Lycopersicon esculentum; induction of NiR (Wellburn et al., 1980)
    i  Eight indoor ornamental species; inhibition of photosynthesis
       (Saxe, 1986)
    j   Lycopersicon esculentum; inhibition of photosynthesis (Capron &
       Mansfield, 1989)
    k   Avena sativa &  Medicago sativa; inhibition of photosynthesis (Hill &
       Bennet, 1970)
    l   Lactuca sativa; inhibition of photosynthesis (Capron, 1989)
    m   Lycopersicon esculentum; inhibition of photosynthesis (Mortensen, 1986)
    n   Lactuca sativa; reduction in plant mass (Capron et al., 1991)
    o   Lycopersicon esculentum; reduction in plant mass (Anderson &
       Mansfield, 1979)
    p   Lycopersicon esculentum; reduction in plant mass (Bruggink et al., 1988)

    Table 22.  Lowest exposure levels (in µg/m3) at which NH3 caused 
               significant effectsa
                                                                        

                     (Bio)chemical    Physiological    Growth aspects
                                                                        

    Long term        50; 8 monthsb    53; 9 monthsh    25; 1 yeark
                                                       53; 8 monthsl
                                                       35; 16 monthsm

    Growing season   100; 6 weeksc    50; 6 weeksi     60; 2 monthsn
    or winter        60; 14 weeksd                     20; 90 dayso
                     180; 13 weekse                    30; 23 daysp

    Air pollution    2000; 24 hf      213; 5 daysj     120; 11 daysq
    episodes         213; 5 daysg                      1000; 2 weeksr
                                                       300; 3 dayss

    Short term                                         30 000; 1 ht
                                                       2000 2 hu
                                                       2000 6 hv
                                                                        

    a  If the fumigation was not continuous an average value has been
       estimated and quoted in parentheses (calculated assuming
       10 µg/m3 during the periods in which the fumigation was
       switched off).
    b  Species of  Violion caninea alliance; imbalanced nutrient
       status (Dueck & Elderson, 1992)
    c   Deschampsia flexuosa; change in amino acid composition (Van
       der Eerden et al., 1990)
    d   Pinus sylvestris; increased GS activity (Pérez-Soba et al.,
       1990)
    e   Pseudotsuga menziesii; imbalanced nutrient status (Van der
       Eerden et al., 1992)
    f   Lycopersicum esculentum; increase of NH4+ in the dark
       (Van der Eerden, 1982)
    g   Lolium perenne; 30% of N in the plant is derived from
       foliar uptake (Wollenheber & Raven, 1993)
    h   Pinus sylvestris; increased loss of water after two weeks
       of desiccation (Dueck et al., 1990)
    i   Populus sp.; increase in stomatal conductance in leaves;
       increase in mesophyll conductance and maximum photosynthetic
       rate at a slightly higher exposure level (Van Hove et al., 1989)
    j   Lolium perenne; significant impact acid/base regulation and
       nutrients status

    Table 22  (Con't)

    k   Pseudotsuga menziesii; erosion of wax layer (Thijse & Baas,
       1990; the authors have some doubts about the causality of this
       effect (personal communication)
    l   Calluna vulgaris; reduction in survival rate after winter
       (Dueck, 1990)
    m   Arnica montana; reduced survival after winter and flowering
       (Van der Eerden et al., 1991)
    n  Field exposure during winter; median concentration; severe
       injury of several conifer species (Van der Eerden, 1982)
    o   Viola canina, Agrostis capillaris; 50% growth stimulation
       of the shoot (not of the roots) (Van der Eerden et al., 1991)
    p   Racomitrium lanuginosum; chlorosis (Van der Eerden et al.,
       1991)
    q   Hypnum jutlandicum; chlorosis (Van der Eerden et al., 1991)
    r   Lepidium sativum; reduction in dry weight (Van Haut &
       Prinz, 1979)
    s  Several horticultural crops; leaf injury
    t  Various deciduous trees; leaf injury (Ewert, 1979)
    u   Brassica sp., Helianthus sp.; leaf injury (Benedict & Breen,
       1955)
    v   Rosa sp.; leaf injury rose (Garber, 1935)


    Several cells in Table 22 could not be filled; the majority of quoted
    effects are on biomass production and tissue injury.  It is clear that
    the data in this table are not random; nearly all of the information
    originating from one Dutch research group.  Only a few pollution
    climates were considered.  The results may be representative for
    mild oceanic climates, but probably not for cold climates with dark
    winters: toxicity of NH3 increases with lower temperature and lower
    light intensity.  The effects of NH3 need to be studied with more
    plant species and under more climatic conditions in order to obtain
    critical levels with sufficient potential for generalization.

    4.1.7.4  NH4+ and NO3- in wet and occult deposition 

         NH4+, NO3- and H+ make up about half of the ionic
    composition of rain, clouds, fog and aerosols.  The impact of the
    acidity of rain and clouds has received much attention in recent years
    (Jacobson, 1991). This is not the case with other compounds in wet
    deposition, although their relevance is recognized. At the same pH,
    Cape et al. (1991) found a much greater effect of sulfuric acid than
    of nitric acid, indicating that the impact of acid rain is not only
    through protons, but also through anions.

         There is an abundance of information on the effects of NH4+ in
    soil solution.  However, threshold concentrations for NH4+ in the
    soil (e.g. Schenk & Wehrman, 1979) can not be considered a critical
    level for rain water quality, because the type of exposure and
    response is completely different.

         Wet deposition containing NH4+ can reduce frost tolerance (Cape
    et al., 1990) and induce leaching of K+ and other cations (Roelofs
    et al., 1985).  It is not yet clear whether this type of ion exchange
    can have deleterious effects on its own in the field situation.

         Currently, too few quantitative data on the effects of nitrogen-
    containing wet and occult deposition are available for critical levels
    for this group of compounds to be derived.

    4.1.7.5  Mixtures

         A polluted atmosphere generally consists of a cocktail of
    compounds, but certain combinations are more frequent.  Because of
    their role in the formation of tropospheric O3, simultaneous
    co-occurrence of relatively high levels of O3 and NO are rarely
    observed, while sequential co-occurrences are much more frequent
    (Kosta-Rick & Manning, 1993).  If burning of fossil fuels results in
    emission of SO2, this is often combined with emission of NOx.

    a)  SO2 plus NO2

         Synergism has been found in nearly all studies concerning this
    combination, with only few exceptions (Kuppers & Klump 1988; Murray et
    al., 1992). Based on data presented by Whitmore (1985), for  Poa
     pratensis the effect threshold for combinations of SO2 and NO2, in
    equal concentrations when expressed in ppm, is in the range of 1.2-2.0
    ppm.days (Fig. 19).  This threshold applies to effects by combinations
    of SO2 and NO2; the effects of single exposures were not assessed in
    this study.  However, it is reasonable from other references to expect
    synergism, and thus to include this threshold in Table 23, in which
    combined effects are summarized.  Another threshold for combinations
    of SO2 and NO2  was defined by Van der Eerden & Duym (1988) (Fig.
    20; Table 23).

    b)  SO2 plus NH3

         Adsorption of either NH3 or SO2 on leaf surfaces is enhanced by
    the presence of the other compound (Van Hove et al., 1989). 
    Interactive physiological effects have been found as well (Dueck,
    1990; Dueck et al., 1990; Dueck & Elderson, 1992).  Currently, there
    is far too little information on this combination to quantify this
    interaction.

    FIGURE 19

    FIGURE 20

    Table 23.  Lowest exposure levels at which NO2 increases the
               effect of SO2, O3, or SO2 plus O3
                                                                        

                    (Bio)chemical    Physiological  Growth aspects
                                                                        

    Long term                                       150-190; 9 monthsf
                                                    220; 60 weeksg
                                                    19; 10-41 weeksh

    Growing season  55-75; 34 daysb  135; 28 daysd  30; 38 daysi
    or winter       135; 28 daysc                   10-43; 130 daysj
                                                    30; 43 daysk

    Air pollution                                   80; 2 weeksl
    episodes                                        75; 1 daym

    Short term                       153; 1 he      325; 1 hm
                                                    400; 1 hn
                                                                        

    a  If the fumigation was not continuous an average value has
       been estimated and quoted in parentheses (calculated
       assuming 10 µg/m3 during the periods in which the
       fumigation was switched off).
    b   Phaseolus vulgaris; inhibition of parts of nitrogen
       metabolism, when exposed sequentially with O3
       (100-120 µg/m3; 8 h/day)
    c   Lolium perenne; decrease in proline content during winter
       hardening when applied in combination with SO2 at
       188 µg/m3 (Davison et al., 1987)
    d   Lolium perenne; less negative osmotic potential during
       winter hardening when applied in combination with SO2 at
       188 µg/m3 (Davison et al., 1987)
    e   Phaseolus vulgaris; Inhibition of photosynthesis when in
       combination with SO2 (215 µg/m3); without SO2
       inhibition at 760 µg/m3 (Bennet et al., 1990)
    f  Several crops; growth stimulation by NO2 turns into a
       reduction in synergism with sequential treatment with O3
       (160-200 µg/m3; 6 h/day) (Runeckles & Palmer, 1987)
    g  Six tree species; reduced plant growth in combination with
       SO2 (280 µg/m3), both antagonism and synergism
       (Freer-Smith, 1984)
    h  10 grass species were tested in combination with SO2
       (27 µg/m3). Three species showed growth stimulation.
       Reduced growth was found at higher concentrations.
       Interactions with acidic mist and with O3 were found
       (Ashenden et al., 1993).

    Table 23  (Con't)

    i   Poa pratensis; inhibition of biomass production; in
       combination with SO2 (42 µg/m3) for 38 days; the longest
       exposure period used in the experiments. Calculated from
       data from Whitmore (1985), assuming synergism and a critical
       level for SO2 plus NO2 of 1.2 ppm.days (Whitmore,     
       1985).
    j   Brassica napus and  Hordeum vulgare; antagonism (and
       rarely synergism) with O3 (6-44 µg/m3; 8 h/day) and       
       SO2 (9-33 µg/m3, continuously): enhanced yield turns into
       reduction (Adaros et al., 1991a,b) 
    k   Plantago mayor; reduced shoot dry weight synergism with
       SO2 (60 µg/m3) and O3 (60 µg/m3, 8 h/day)
       (Mooi, 1984)
    l   Poa pratensis; inhibition of biomass production; in
       combination with SO2 (110 µg/m3) for 2 weeks (the upper
       margin of the exposure period of this cell in the table; the
       shortest fumigation in this survey was 20 days. Calculated
       from data from Whitmore (1985), assuming synergism and a
       critical level for SO2 plus NO2 of 1.2 ppm.days
       (Whitmore, 1985).
    m  Critical level for NO2 assuming SO2 to be present at
       70 µg/m3; considered to be a critical level for a 24-h mean
       (UNECE, 1994) (Van der Eerden & Duym, 1988)
    n   Lycopersicon esculentum; reduction in plant mass if in
       combination or preceded by O3 (160 µg/m3; 1 h)
       (Goodyear & Ormrod, 1988).

    c)  NO plus NO2

         When activated charcoal has been used as filter material in NO2
    fumigation experiments, NO must have been present as well, because
    activated charcoal has virtually no capacity to absorb NO.  In those
    studies, responses must have been due to NO2 plus NO.  Although the
    toxicity of NO was often considered to be much less than that of NO2,
    currently the two compounds are assumed to be equally toxic and to
    act additively.  However, Wellburn (1990) and others have stated
    that NO is more toxic, and Saxe (1994) showed that the variation in
    sensitivity amongst species is different for the two compounds.  This
    supports the suggestion of Wellburn that the mechanism of toxicity is
    different.

         For the purpose of deriving critical levels, the assumption of
    additivity may result in an underestimation.  However, there are not
    enough data to quantify this.

    d)  Mixtures with O3

         The combination NH3 plus O3 has rarely been studied.  No
    statistically significant interactions have been found as yet, but in
    one study the threshold for leaf injury was higher in the presence
    of NH3 (Van der Eerden et al., 1994).  The combination NO2 plus O3
    has been studied more frequently, but the responses differed
    considerably between experiments and species.  Additivity or
    antagonism was found by  Runeckles & Palmer (1987), Adaros et al.
    (1991a,b), and Bender et al. (1991).  Synergism was reported by Ito et
    al. (1984), Runeckles & Palmer (1987) and Kosta-Rick & Manning (1993).

         The combination of SO2 plus O3 plus NO2 has also been studied. 
    Again the responses varied between plant species and experiment. 
    Antagonism, additivity and synergism have all been found (Kosta-Rick &
    Manning, 1993).

    e)  Mixtures with elevated CO2

         Generally, an increased supply of CO2  to crops results in an
    enhanced biomass production.  The responses of native species are more
    variable but are also frequently positive.  This growth stimulation is
    limited by a deficiency of other nutrients.  Nitrogen can be one such
    limiting factor, and for this reason a nitrogen fertilizer such as
    NHy and possibly low NOx concentrations could be hypothesized to
    have a more-than-additive relationship with CO2.  However, as yet
    there is no experimental evidence for this.  Van der Eerden
    et al. (1994) and Pérez-Soba et al. (1994) found stimulation of
    photosynthesis and growth by both NH3 and CO2, but not by a
    combination of these two compounds.

         Effects of the combination of NOx and CO2 have not yet been
    studied within the scope of global climate change.  But some relevant
    information could be gained from the literature dealing with CO2
    enrichment in glasshouses.  When the flue gases of the heating system
    are used as a CO2 source, NOx (in which NO is dominant) becomes a
    major contaminant.  The fertilizing effect of elevated CO2 can
    largely disappear in the presence of NOx (Anderson & Mansfield, 1979;
    Saxe & Voight Christensen, 1984; Mortensen, 1985; Bruggink et al.,
    1988; Capron et al., 1994).

         The CO2, NH3 and NOx concentrations used in combination in
    these experiments were relatively high and therefore cannot be used
    in the critical level assessment.  More experiments with lower
    concentrations are required.

         Table 23 indicates, surprisingly, that the effective long-term
    exposures are generally higher than those of shorter duration. 
    However, long-term responses (more than half a year) have rarely been
    studied.  Therefore, the information on effects over growing season
    periods may be much more representative of long-term effects.

         A study included in a report by UNECE (1994) used 21 µg SO2/m3
    and 11 µg NO2/m3, over the entire growing season and found synergism
    in reducing biomass production of  Pisum sativum and  Spinacea
     oleracea.  Similar results were found for  Hordeum vulgare and
     Brassica oleracea, when fumigation was conducted for 120-190 days
    with 30-40 µg SO2/m3 and 30-50 µg NO2/m3.  This study cannot be
    used for the assessment of critical levels because it has not yet been
    published, but it indicates that lower levels of the two pollutants
    than those quoted in Table 23 can influence plant responses.

    4.1.8  Appraisal

         Table 24 shows the former air quality guidelines for NO2 and
    some other critical levels assessed in the same period.  Fig. 21
    summarizes the results given in Tables 20 to 23.  In this figure
    curves are drawn to estimate critical levels according to current
    practice, known as the "envelope" approach.  After having plotted all
    effective exposure levels in a graph of concentration and exposure
    time, a curve is drawn just below the lowest effective exposures. 
    Critical levels can be derived from this curve.  Fig. 21 shows that
    more experiments with exposure periods of 0.5 to 5 days are required
    to give a more solid basis for the estimation of critical levels of
    24 h.

    Table 24.  Critical levels for NO2

    Concentration       Exposure time       Reference
    (µg/m3)
                                                                        

    95                  4 h                 WHO (1987)
    30a                 annual mean         WHO (1987)
    800                 1 h                 Guderian (1988)
    60                  growing season      Guderian (1988)
    40                  winter              Guderian (1988)
                                                                        


    a    SO2 and O3 not higher than 30 µg/m3 and 60 µg/m3, respectively

    FIGURE 21

         A second approach to arrive at critical levels is the statistical
    model of Kooijman (1987).  Based on the variation in sensitivity
    between species, critical levels are calculated taking into account
    the number of tested species and the level of uncertainty (Van der
    Eerden et al., 1991).  The second approach is better, but only part of
    the available data is suitable for this approach.

         Tables 20 to 23 show that some new relevant information has
    appeared.  Comparing the data of Table 20 with those of Table 21
    (Fig. 21a and 21b), it appears that NO2 has slightly higher effect
    thresholds than NO.  However, this probably reflects the separate
    attention paid to these compounds, rather than any difference in
    toxicity.  It is now obvious that the toxicity of NO cannot be
    ignored, and it should be included in the guidance values.  The
    consideration of NO and NO2 together (leading to a guidance value for
    NOx) seems the best way of evaluating the impact of NO.  However,
    future research should evaluate the specific phytotoxic properties of
    the individual compounds and their combinations.

         It is not yet possible to discriminate in the critical level
    assessment between separate types of vegetation, such as crops,
    plantation forests, natural forests and other natural vegetation.  A
    1-h average for NO2 of 800 µg/m3 to prevent acute damage
    (Table 24) is probably too high.  A critical level for NOx of around
    300 µg/m3 would be better.  A critical level of 95 µg/m3 as a 4-h
    mean, as proposed in the previous WHO guidelines (WHO, 1987), is still
    realistic, but not very practical.  If critical levels for short
    periods (e.g., 1 or 8 h) should be defined, it is probably necessary
    to separate day- and night-time exposures.  A critical level for a
    24-h mean is more practical, as this is relevant for both peak
    concentrations of several hours and air pollution episodes of several
    days.

         For the growing season and winter half year, Guderian (1988)
    suggested critical levels of 60 and 40 µg/m3, respectively.  From
    Table 20 it can be seen that the critical level of 60 µg/m3 can cause
    substantial growth stimulation rather than reduction. Within the
    context of air quality guidelines, this type of response must be
    regarded as potentially adverse because, for instance, of its
    influence on competition within the natural vegetation.  From current
    knowledge it is evident that 60 µg/m3 is too high to prevent growth
    stimulation. In addition, the critical level of 30 µg/m3 for an
    annual mean, given in the 1987 WHO guidelines, will almost certainly
    not protect all plant species.  However, for crops, where growth
    stimulation is rarely an adverse effect, this could be acceptable if
    secondary effects are negligible.  The experimental basis for the WHO
    air quality guidelines of 1987 was relatively poor, but evidence is
    increasing that these are certainly not unrealistically low.  Not even

    all direct adverse effects are eliminated by these levels (Adaros et
    al., 1991a,b; Bender et al., 1991; Ashenden et al., 1993).  Thus, the
    updated guidelines/guidance values should be lower than the ones of
    1987.

         A long-term critical level for NO2 of 10 µg/m3, especially to
    avoid eutrophication of nutrient-poor vegetation, was proposed by
    Guderian (1988) and Zierock et al. (1986).  The basis for this
    proposal was the work of Lee et al. (1985) and Press et al. (1986),
    who found reduced growth of  Sphagnum cuspidatum in regions with an
    annual mean concentration of 38 and 11 µg/m3, respectively, compared
    to the growth in another region with 4 µg/m3 after 140 days of
    exposure.  However, Lee et al. (1985) also showed that the poor growth
    of  Sphagnum was more closely related to the excessively high
    concentrations of nitrate and ammonium ions in bog water rather than
    to the concentration of NO2 alone.  Thus, this information could well
    be used to assess water quality guidelines, but is not very useful as
    a basis for air quality guidelines.

    4.1.8.1  Representativity of the data

         Critical levels for adverse effects of NH3 on plants were
    estimated using the model of Kooijman (Van der Eerden et al., 1991). 
    To protect 95% of the species at P < 0.05, a 24-h critical level of
    270 and an annual mean critical level of 8 µg/m3 were estimated. 
    With the graphical approach the 24-h average was a little lower and
    the annual mean somewhat higher (13 and 200 µg/m3, respectively;
    Fig. 21).

         On the basis of a review by Cape (1994), critical levels for H+
    and NH4+ were adopted for locations where ground-level cloud is
    present for more than 10% of the time and where calcium and magnesium
    concentrations in rain or cloud do not exceed H+ and NH4+
    concentrations (mainly high elevation areas in cold climate zones):
    300 µmol NH4+/litre as an annual mean (UNECE, 1994).

         There remains considerable deficiency in the amount and scope of
    experimentally derived information on which to base air quality
    guidelines.  This results from the fact that most experiments have
    been performed under conditions that cannot directly be compared to
    outdoor circumstances.  In most experiments, only primary effects such
    as photosynthesis and biomass production were evaluated, and rarely
    secondary effects such as alteration of stress tolerance or
    competitive ability.  The plant species chosen in most experiments
    were crops, although evidence suggests that some native species are
    relatively more sensitive.  For instance, lower plants such as
    bryophytes and lichens are not protected by a waxy waterproof cuticle
    and do not have the potential to close stomata.  Furthermore, Pearson
    & Stewart (1993) suggested that plants species from nutrient-poor
    acidic soils are more sensitive.

         Further work, employing low concentrations of NHy and NOx
    (especially NO) in different climates, is urgently required.  It is
    not realistic to screen for all likely growth and physico-chemical
    effects in the majority of species in order to arrive at general
    effect thresholds.  Selections must be made on the basis of an
    understanding of differences in sensitivity between species.  However,
    an obvious mechanistic explanation for sensitivity differences is not
    yet available.  For instance, there appears to be no relationship
    between the sensitivity to NO2 and the nitrogen preference
    (Ellenberg, 1985; Ashenden et al., 1993).  Sensitivity classifications
    for some tens of species have been made for NO2 and NH3 (e.g. US
    EPA, 1978; Taylor et al., 1987), but it appears difficult to extend
    predicitions very far beyond those examined.  The hypotheses of Raven
    (1988) and Pearson & Stewart (1993) should be studied in more detail
    in laboratory experiments and field studies, as they could result in
    an efficient selection criterium for further screening.

         An attempt to determine the global situation regarding
    nitrogen-containing compounds is being made.  The assumption that all
    deposited nitrogen-containing compounds (which is part of the critical
    load concept) act additionally in their impact on vegetation is poorly
    based on experimental results and is probably not valid for the short
    term.

         Generalizations and simplifications have to be made to arrive at
    conclusions that are applicable in environmental policy making, but
    this should be done with great care.  Mechanistic simulation models
    can become a powerful tool for making general predictions on the basis
    of various air pollution experiments (Van de Geijn et al., 1993). 
    However, sufficient knowledge of biochemical and physiological
    mechanisms to incorporate the impact of air pollution on vegetation
    into these models is still lacking.  This applies especially to
    natural vegetation where stress sensitivity and competition are key
    factors.

         Many gaps in understanding the impact of nitrogen-containing air
    pollution on vegetation still exist, and this is a good reason to use
    a safety factor in determining critical levels and loads.  However,
    currently there is still no broadly accepted approach to quantify such
    a safety factor.

    4.1.9  General conclusions

         The sum of information on gaseous NH3 and on NH4+ in wet and
    occult deposition is still too limited to arrive at air quality
    guidelines, as they should have a broad applicability.  The critical
    levels for NH3 and NH4+ are probably only valid for temperate
    oceanic climatic zones (see sections 4.1.7.3, 4.1.7.4 and 4.1.8).

         In most studies with NO and NO2, no significant effects were
    found at levels below 100 µg/m3, but several relevant exceptions
    exist.  NO2 altered the response to O3 generally with a
    less-than-additive interaction.  In combination with SO2, NO2 acted
    more-than-additively in most cases.  With CO2 and with NO, no
    interaction and thus additivity were generally found.  The lowest
    effective concentration levels of NO2 are about equal for NO2 alone
    and in combination with O3 or SO2, although, generally, at
    concentrations near to its effect threshold NO2 causes growth
    stimulation if it is the only pollutant, while in combination with
    SO2 and/or O3 it results in growth inhibition.

         To include the impact of NO, a critical level for NOx instead of
    one for NO2 is proposed, assuming that NO and NO2 act in an additive
    manner.  A strong case can be made for the provision of critical
    levels for short-term exposures, but currently there are insufficient
    data to provide these with sufficient confidence.  Current evidence
    exists for a critical level of around 75 µg/m3 for NOx as a 24-h
    mean.

         The critical level for NOx (NO and NO2, added in ppb and
    expressed as NO2 in µg/m3) is 30 µg/m3 as an annual mean.  At
    concentrations slightly above this critical level, growth stimulation
    will be the dominant effect if the ambient conditions allow growth and
    NOx is the only pollutant.  This stimulation may be combined with a
    minor increase in sensitivity to biotic and abiotic stresses.  In
    cases where biomass production is a positive effect, e.g., in
    agriculture and plantation forests, the critical level can be higher. 
    Current knowledge is insufficient to arrive at critical levels for
    these systems.

         The critical level can be converted into deposition quantities. 
    With deposition velocities of 3 and 0.3 mm/second for NO2 and NO,
    respectively (see section 3.2.2 and Table 5), the annual deposition
    corresponding to a NOx concentration of 30 µg/m3 is 4.8 kg/ha when
    half of the NOx is NO2 and 8.3 kg/ha when all is NO2.  This
    indicates that at a concentration near to its critical level the
    contribution of NOx to the nitrogen demand is negligible for
    fertilized crops but not for natural vegetation (see section 4.2).

    4.2  Effects on natural and semi-natural ecosystems

    4.2.1  Effects on freshwater and intertidal ecosystems

         In this section the effects of atmospheric nitrogen deposition
    on freshwater and intertidal ecosystems are evaluated.  The
    effects of increased emissions of nitrogen compounds with respect to
    eutrophication are examined in order to establish ecosystem guidelines

    for nitrogen deposition.  The ecological effects of nitrogen
    deposition are reviewed for (i) shallow softwater lakes and (ii) lakes
    and streams.

    4.2.1.1  Effects of nitrogen deposition on shallow softwater lakes

         In the lowlands of western Europe, soft water is often found on
    sandy soil which is poor in calcium carbonate or almost devoid of it. 
    The water is poorly buffered and the concentrations of calcium in the
    water layer are very low.  The water bodies are shallow and fully
    mixed, with periodically fluctuating water levels.  They are mainly
    fed by rain water and thus are oligotrophic. These softwater
    ecosystems are characterized by plant communities from the
    phytosociological alliance Littorellion (Schoof-van Pelt, 1973;
    Wittig, 1982; Roelofs, 1986; Vöge, 1988; Arts, 1990).  The stands of
    these communities are characterized by the presence of rare and
    endangered isoetids, such as  Littorella uniflora, Lobelia dortmanna,
     Isoetes lacustris, I. echinospora, Echinodorus species,  Luronium
     natans and many other softwater macrophytes.  These softwater bodies
    are now almost all within nature reserves and have become very rare in
    western Europe.  A strong decline in amphibians has also been observed
    in these water bodies (Leuven et al., 1986).

         The effects of nitrogen pollutants on these softwater bodies have
    been intensively studied in the Netherlands both in field surveys and
    experimental studies.  Field observations on about 70 softwater bodies
    (with well-developed isoetid vegetation in the 1950s) showed that the
    water, in which these macrophytes were still abundant in the early
    1980s, was poorly buffered (alkalinity of 50-500 µeq/litre), slightly
    acidic (pH=5-6) and very poor in nitrogen (Roelofs, 1983; Arts et al.,
    1990).  The softwater sites where these plant species had disappeared
    could be divided into two groups.  In 12 of the 53 softwater sites,
    eutrophication, resulting from nutrient-enriched water, seemed to be
    the cause of the decline.  In this group of non-acidified water
    bodies, plant species, such as  Myriophyllum alterniflorum, Lemna
     minor or  Riccia fluitans had become dominant.  High concentrations
    of phosphate and ammonium ions were measured in the sediment. In some
    of the larger water bodies no macrophytes were found, as a result of
    dense plankton bloom.  In the second group of lakes and pools (41 out
    of 53) another development had taken place: the isoetid species were
    replaced by dense stands of  Juncus bulbosus or aquatic mosses such
    as  Sphagnum cuspidatum or  Drepanocladus fluitans.  This clearly
    indicates acidification of the water in recent decades, probably
    caused by enhanced atmospheric deposition.  In the same field study it
    was shown that the nitrogen levels in the water were higher in
    ecosystems where the natural vegetation had disappeared, compared with
    ecosystems where the Littorellion stands were still present (Roelofs,
    1983).  This strongly suggests the detrimental effects of atmospheric
    nitrogen deposition in these softwater lakes.

         Several ecophysiological studies have revealed the importance of
    (i) inorganic carbon status of the water as a result of intermediate
    levels of alkalinity, and (ii) low nitrogen concentrations for the
    growth of the endangered isoetid macrophytes.  Furthermore, almost all
    of the typical softwater plants had a relatively low potential growth
    rate.  Increased acidity and higher concentrations of ammonium ion in
    the water clearly stimulated the development of  Juncus bulbosus and
    submerged mosses such as  Sphagnum and  Drepanocladus species
    (Roelofs et al., 1984; Den Hartog, 1986).  Cultivation experiments
    confirmed that the nitrogen species involved (ammonium or nitrate
    ions) differentially influenced the growth of the studied species of
    water plants.  Almost all of the characteristic softwater isoetids
    developed better when nitrate was added instead of ammonium, whereas
     Juncus bulbosus and aquatic mosses  (Sphagnum & Drepanocladus) were
    clearly stimulated by ammonium (Schuurkes et al., 1986).  The
    importance of ammonium for the growth of these aquatic mosses was also
    reported by Glime (1992).

         The effects of atmospheric deposition have been studied in
    small-scale softwater systems during a 2-year treatment with different
    artificial rainwaters.  Acidification, without airborne nitrogen input
    (using sulfuric acid), did not result in a mass growth of  Juncus
     bulbosus, and a diverse isoetid vegetation remained present. 
    However, after increasing the nitrogen concentration in the
    precipitation (as ammonium sulfate), similar changes to those seen in
    field conditions were observed, i.e. a dramatic increase in the
    dominance of  Juncus bulbosus, of submerged aquatic mosses and of
     Agrostic canina (Schuurkes et al., 1987).  It became obvious that
    the observed changes occurred because of the effects of ammonium
    sulfate deposition, leading to both eutrophication and acidification. 
    The increased levels of ammonium in the system directly stimulated the
    growth of plants such as  Juncus bulbosus, whereas the surplus
    ammonium would be nitrified in this water (pH > 4.0).  During this
    nitrification process, H+ ions are produced, which increases the
    acidity of the system.  The results of this study clearly demonstrated
    that the changes in composition of the vegetation had already occurred
    after a 2-year treatment with > 19 kg nitrogen per ha per year.  A
    reliable critical load for nitrogen deposition in these shallow
    softwater lakes is thus most likely to be below 19 kg nitrogen per ha
    per year and probably between 5 to 10 kg nitrogen per ha per year. 
    This value is supported by the observation that the greatest decline
    in the species composition of the Dutch Litorellion communities has
    coincided with nitrogen loads of around 10-13 kg nitrogen per ha per
    year (Arts, 1990).

    4.2.1.2  Effects of nitrogen deposition on lakes and streams

         There is ample evidence that an increase of acidic and
    acidifying compounds in atmospheric deposition had resulted in recent
    acidification of lakes and streams in geologically sensitive regions

    of Scandinavia, western Europe, Canada and the USA (Hultberg, 1988;
    Muniz, 1991).  This acidification is characterized by a decrease in pH
    and acid neutralizing capacity and by increases in concentrations of
    sulfate, aluminium, and sometimes nitrate and ammonium.  It has been
    shown since the 1970s, using various approaches (field surveys,
    laboratory studies, whole-lake experiments), that these changes have
    had major consequences for plant and animal species (macrofauna,
    fishes) and for the functioning of these aquatic ecosystems.

         The critical loads of acidity (from SOy and NOy) for aquatic
    ecosystems, based on steady-state water chemistry models, were
    published by the UN Economic Commission for Europe (UNECE) in 1988 and
    1992.  These models incorporate both sulfur and nitrogen acidity, and
    critical loads are calculated on the basis of: (i) base cation
    deposition; (ii) internal alkalinity production or base cation
    concentrations; and (iii) nitrate leaching from the water system.  The
    calculated critical loads are thus site-specific (sensitive areas or
    not) and also depend on the local hydrology and precipitation (for
    full details, see Hultberg (1988), Henriksen (1988) and Kämäri et al.
    (1992)).  The critical loads of nitrogen acidity (kg nitrogen per ha
    per year) for the most sensitive lakes and streams are:

    Scandinavian        1.4-4.2        (Hultberg, 1988; Henriksen,
    waters                             1988; Kämäri et al., 1992)

    Alpine lakes        3.5-6.1        (Marchetto et al., 1994)
 
    Humic moorland      3.5-4.5        (Schuurkes et al., 1987;
    pools                              van Dam & Buskens, 1993)

         In many areas with high water alkalinity and/or high base cation
    deposition, the values of the critical load for nitrogen acidity are
    much higher than those for sensitive freshwaters.  At present, the
    possible effects of nitrogen eutrophication by ammonia/ammonium or
    nitrate deposition are not incorporated in the establishment of
    critical loads for nitrogen, except for shallow softwater lakes (see
    section 4.2.1.1).  This is because primary production in almost all
    aquatic ecosystems is limited by phosphorus availability, and thus
    nitrogen enrichment has been considered unimportant in this respect
    (Moss, 1988). This certainly holds for those aquatic ecosystems
    considered above, where the critical load with respect to acidifying
    effects are certainly more relevant than the effects of nitrogen
    eutrophication. It is, however, to be expected that the following
    aquatic ecosystems are sensitive to nitrogen enrichment: (i) alpine
    lakes; (ii) water with low background nitrogen; and (iii) humic lakes
    (Kämäri et al., 1992).  The effects of nitrogen eutrophication
    (including ammonia/ammonium) in these water bodies need further
    research and should be taken into account in future critical loads
    determinations for nitrogen.  At present it is not possible to present

    reliable critical loads for nitrogen eutrophication in these aquatic
    ecosystems.  An overview of critical loads for nitrogen in aquatic
    ecosystems is given in section 8.2.2.

    4.2.2  Effects on ombrotrophic bogs and wetlands

         In this section the effects of atmospheric nitrogen deposition in
    (semi-)natural wetlands are evaluated.  The effects of enhanced
    nitrogen inputs are considered for: (i) ombrotrophic (raised) bogs;
    (ii) fens; and (iii) intertidal fresh- and saltwater marshes.  A
    common feature of all these systems is the anaerobic nature of their
    waterlogged soils, characterized by low redox potential, high
    concentrations of toxic reduced substances and high rates of
    denitrification (Gambrell & Patrick, 1978; Schlesinger, 1991).

    4.2.2.1  Effects on ombrotrophic (raised) bogs

         Ombrotrophic ("rain-nourished") bogs, which receive all their
    nutrients from the atmosphere, are particularly sensitive to airborne
    nitrogen loads.  These bogs are systems of acidic wet areas and are
    very common in the boreal and temperate parts of Europe.  Because of
    the anaerobic conditions, decomposition rates are slow, favouring the
    development of peat.  In western Europe and high northern latitudes,
    typical plant species include bog-mosses ( Sphagnum species), sedges
     (Carex; Eriophorum) and heathers ( Andromeda, Calluna and  Erica). 
    Insectivorous plant species (e.g.,  Drosera) are especially
    characteristic of these bogs; they compensate for low nitrogen
    concentrations by trapping and digesting insects (Ellenberg, 1988b).

         Clear effects of nitrogen eutrophication have been observed in
    Dutch ombrotrophic bogs. The composition of the moss layer in the
    small remnants of the formerly large bog areas has markedly changed in
    recent decades as nitrogen loads have increased to 20-40 kg nitrogen
    per ha per year (especially as NH4+/NH3).  The most characteristic
    species  (Sphagnum) are replaced by the more nitrophilous mosses
    (Greven, 1992).  A national survey of Danish ombrotrophic bogs has
    shown a decline of the original bog vegetation together with an
    increase of more nitrogen-dependent species in areas with high ammonia
    deposition (> 25 kg ammonium nitrogen per ha per year (Aaby, 1990).

         The effects of atmospheric nitrogen deposition on ombrotrophic
    bogs have also been intensively studied in the United Kingdom (Lee et
    al., 1989; Lee & Studholme, 1992).  Many characteristic  Sphagnum
    species are now largely absent from affected ombrotrophic bog areas
    in the United Kingdom, such as the southern Pennines in England.
    Atmospheric nitrogen deposition has more than doubled in these areas
    to around 30 kg nitrogen per ha per year, compared with areas of
    healthy  Sphagnum growth.  In contrast to the situation in
    continental western Europe, most of the nitrogen deposition in the
    United Kingdom is of nitrogen oxides, although the importance of

    ammonia/ammonium deposition is also increasing in the United Kingdom
    (Fowler et al., 1980; Sutton et al., 1993).  Several studies on bogs
    in the United Kingdom have shown that increased supplies of nitrogen
    are rapidly absorbed and utilized by bog-mosses  (Sphagnum),
    reflecting the importance of nitrogen as a nutrient and its scarcity
    in unpolluted regions (Woodin et al., 1985; Woodin & Lee, 1987).  The
    high nitrogen loadings are, however, supraoptimal for the growth of
    many characteristic  Sphagnum species, as demonstrated by restricted
    development in growth experiments and transplantation studies between
    clean and polluted locations.  In areas with high nitrogen loads, such
    as the Pennines, the growth of  Sphagnum is in general less than in
    unpolluted areas (Lee & Studholme, 1992).  After transplantation of
     Sphagnum from an unpolluted site to a bog in the southern Pennines,
    a rapid increase in plant nitrogen content from around 12 to 20 mg/g
    dry weight was observed (Press et al., 1986).  A large increase in
    arginine in the shoots of these bog-mosses was also found after
    application of nitrogen.  In field experiments in northern and
    southern parts of Sweden, nitrogen was found to be the limiting factor
    for the growth of  Sphagnum.  However, other nutrients, especially
    phosphorus, may become secondarily limiting to plant growth when
    nitrogen inputs reach a threshold (Aerts et al., 1992).

         A further possible effect of the increased nitrogen content of
     Sphagnum is an increased decay rate of the peat, as nitrogen content
    strongly influences decomposition rates (Swift et al., 1979).  The
    decay rate of  Sphagnum peat in Swedish ombrotrophic bogs has been
    studied along a gradient of nitrogen deposition (Hogg et al., 1994). 
    The results of this short-term decay experiment indicated that the
    decomposition rate of  Sphagnum peat is more influenced by the
    phosphorus content of the material than by its nitrogen content,
    although some relation with nitrogen supply has been observed. 
    Further evidence is necessary to evaluate the long-term effects of
    enhanced nitrogen supply on the decay of peat.

         All these studies strongly indicate the detrimental effects of
    atmospheric nitrogen on the development of the bog-forming  Sphagnum
    species.  However, enhanced nitrogen deposition can influence the
    competitive relationships in nutrient-deficient vegetation such as
    bogs.  The effects of the supply of extra nitrogen on the population
    ecology of  Drosera rotundifolia has been recently studied in a
    4-year experiment in Swedish ombrotrophic bogs (Redbo-Torstensson,
    1994).  It was demonstrated that experimental applications of more
    than 10 kg nitrogen (as NH4NO3) per ha per year clearly affected the
    population of this insectivorous species: the establishment of new
    individuals and the survival of the plants significantly decreased in
    the vegetation treated with extra nitrogen.  This decrease in the
    total population density of the characteristic bog species  Drosera
    was not caused by toxic effects of nitrogen, but by enhanced

    competition for light with tall species such as  Eriophorum and
     Andromeda, which responded positively to the increased nitrogen
    inputs.

         On the basis of the United Kingdom and Scandinavian studies, it
    has become clear that increased nitrogen loads strongly affect
    ombrotrophic bog ecosystems, especially because of the high nitrogen
    retention capacity and closed nitrogen cycling.  The growth of
    bog-mosses is reduced, directly by nitrogen and indirectly by a
    changed competitive relationship between the prostrate dominants
    (e.g.  Eriophorum) and the subordinate plant species.  A reliable
    critical load for nitrogen in these ombrotrophic bogs is 5-10 kg
    nitrogen per ha per year, although additional long-term studies with
    enhanced nitrogen (both nitrogen oxides and ammonia/ammonium) are
    necessary to validate this figure.

    4.2.2.2  Effects on mesotrophic fens

         Fens are wetland ecosystems that are typical of alkaline to
    slightly acidic habitats in many countries.  The alkalinity is due to
    groundwater draining from surrounding rocks or sediments that are
    relatively rich in calcium carbonate.  Most of these fen ecosystems
    are characterized by rare and endangered plants species.  The effects
    of nitrogen enrichment upon mesotrophic fens have been intensively
    studied in the Netherlands (Verhoeven & Schmitz 1991; Koerselman &
    Verhoeven, 1992).  These fen ecosystems are characterised by many
     Carex species and are rich in forbs (e.g.,  Pedicularis palustris;
    orchids).  Most of these Dutch fen ecosystems are managed as hay
    meadows, with removal of the plant material further restricting
    nutrient levels, and are now nature reserves.

         A considerable increase of tall graminoids (grass or  Carex
    species), with a somewhat higher potential growth rate, was observed
    after experimentally adding nitrogen to three Dutch fen ecosystems
    (Vermeer, 1986; Verhoeven & Schmitz, 1991).  This increase caused a
    significant decrease in the diversity of subordinate plant species. 
    In one of the Dutch fen sites investigated, which had a long history
    of hay making, it has been shown that phosphorus deficiency was also a
    major factor in the productivity of the system, since there was a high
    output of phosphorus from the ecosystem with the hay (Verhoeven &
    Schmitz, 1991; Koerselman & Verhoeven, 1992).  Using the results of
    fertilization trials and nutrient budget studies in these fen
    ecosystems (Koerselman et al., 1990; Koerselman & Verhoeven, 1992),
    with their relatively closed nitrogen cycle, it seems reasonable to
    establish a critical load of 20-35 kg nitrogen per ha per year, based
    upon the output of the nitrogen from the fen system via normal
    management.  In some fen ecosystems, the critical nitrogen load based
    on the change in diversity may be substantially higher, because of the
    limitation of productivity by phosphorus (Egloff, 1987; Verhoeven &
    Schmitz, 1991).  In this situation, however, the risks of nitrogen

    losses to surface water or groundwater will increase because of
    phosphorus limitation, and this effect should be taken into account in
    critical load evaluation.  High rates of denitrification could also
    influence the establishment of critical loads for these fen
    ecosystems, and this aspect needs further investigation.

    4.2.2.3  Effects on fresh- and saltwater marshes

         In the wetland ecosystems discussed above, the nitrogen cycle is
    more closed than that of intertidal marshes.  The data on atmospheric
    nitrogen inputs to the nitrogen cycling in intertidal fresh- and
    saltwater marshes (with large prostrate graminoids as species of
     Spartina, Typha and  Carex) have been reviewed by Morris (1991). 
    It has become evident that nitrogen inputs to these marsh ecosystems
    via surface water (well above 100 kg nitrogen per ha per year) are
    much higher than the atmospheric loading.  In non-tidal freshwater
    marshes, it has been demonstrated in many studies that denitrification
    is very high with a very large output of nitrogen from the ecosystem
    (Morris, 1991).  Because of the combined effect of these processes,
    atmospheric nitrogen deposition is of only minor importance for these
    marshes, and it is not useful to establish a critical load for
    airborne nitrogen to these systems.  In his review Morris (1991)
    formulated a critical load for atmospheric nitrogen in wetland
    ecosystems of around 20 kg nitrogen per ha per year.  It is more
    appropriate to make a distinction for different types of wetlands, as
    shown above.  An overview of the critical loads for wetlands is given
    in chapter 8.

    4.2.3  Effects on species-rich grasslands

         Almost all of the research on the effects of atmospheric
    deposition on terrestrial vegetation has focused on ecosystems
    (e.g., forest, heathland and bogs) involving poorly buffered acidic
    soils.  Semi-natural grasslands with traditional agricultural use have
    also been an important part of the landscape in western and central
    Europe, and contain, or used to contain, many rare and endangered
    plant and animal species. A number of these grasslands have been set
    aside as nature reserves in several European countries (Ellenberg,
    1988b; Woodin & Farmer, 1993).  These semi-natural grasslands, which
    are of conservation interest, are generally nutrient-poor because of
    long agricultural use with low levels of manure and the removal
    of plant growth by grazing or hay making.  The vegetation is
    characterized by many low growing species and is of nutrient-poor soil
    status (Ellenberg, 1988b).  Although these grasslands are nowadays
    rare, the proportion of endangered plant and animal species in these
    communities is very high (Van Dijk, 1992).  Many experiments have
    shown that application of artificial fertilizer (nitrogen, phosphorus
    and potassium) changes these grasslands into tall, species-poor
    stands, dominated by a few highly productive crop grasses (Van Den
    Bergh, 1979; Willems, 1980; Van Hecke et al., 1981).  To maintain a

    large diversity of species, addition of fertilizer has to be avoided. 
    It is thus to be expected that these species-rich grasslands will be
    affected by increased atmospheric nitrogen input (Wellburn, 1988;
    Liljelund & Torstensson, 1988; Ellenberg, 1988b).

         Many semi-natural grassland types are present in western and
    central Europe.  Most of these grasslands belong to the so-called
    neutral grasslands (Molinio-Arrhenateretea; moist to moderately dry),
    to the dry calcareous grasslands (Festuca-Brometea) or to the acid
    grasslands on very poor soils (Nardetalia).  Detailed descriptions
    have been given by Ellenberg (1988b) and Van Dijk (1992).  To obtain
    critical loads for nitrogen for all these grasslands, it would be
    essential to study the effects of nitrogen eutrophication in a
    representative range within these communities.  Detailed data are,
    however, scarce.  Therefore, the results of an integrated research
    programme on nitrogen eutrophication in Dutch calcareous grasslands
    are used as a target study in this chapter to obtain (i) information
    on observed changes in this type of grassland caused by enhanced
    nitrogen input, and (ii) a reliable estimation of the critical load
    for nitrogen in these well-buffered non-acidic grasslands.  The
    results of this calcareous grassland study will be discussed in
    respect to other semi-natural grasslands.

    4.2.3.1  Effects of nitrogen on calcareous grasslands

         Calcareous grasslands are communities on limestone.  The subsoils
    consist of various kinds of limestone with high contents of calcium
    carbonate (> 90%), covered by shallow well-buffered rendzina soils
    (A/C-profiles; pH of the top soil is 7-8 with a calcium carbonate
    content of around 10%).  The depth of the soil varies between 10 and
    50 cm and the availability of nitrogen and phosphorus is low.  The
    grasslands are generally found on slopes with limestone in the subsoil
    and a deep groundwater table. Plant productivity is low, and the peak
    standing crop is in general between 150 and 400 g/m2.  The canopy of
    the vegetation is open and low (10-20 cm).  Calcareous grasslands are
    among the most species-rich plant communities in Europe and contain a
    large number of rare and endangered species.  The area of these
    semi-natural grasslands has decreased substantially in Europe during
    the second half of this century (Wolkinger & Plank, 1981; Ratcliffe,
    1984).  Some remnants have become nature reserves in several European
    countries.  To maintain the characteristic calcareous vegetation a
    specific management is needed to prevent their natural succession
    towards woodland (Wells, 1974; Dierschke, 1985).  The calcareous
    grasslands in the Netherlands are mown in autumn with removal of the
    hay (Bobbink & Willems, 1987).

    a)  Nitrogen enrichment and vegetation composition

         The effects of nitrogen enrichment in Dutch calcareous grasslands
    on vegetation composition have been investigated in two field

    experiments (Bobbink et al., 1988; Bobbink, 1991).  Either potassium 
    (100 kg per ha per year), phosphorus (30 kg per ha per year) or
    nitrogen (100 kg per ha per year), as well as a complete fertilization
    (nitrogen, phosphorus and potassium), were applied for 3 years to
    study the long-term effects on vegetation composition.  Nitrogen was
    given as ammonium nitrate and was applied to two nature reserves with
    opposite aspects (north and south).  Total above-ground biomass
    increased considerably, as expected, after three years of nitrogen,
    phosphorus and potassium fertilization.  In the experiments where
    the nutrients were applied individually, a moderate increase in
    above-ground dry weight was only seen with nitrogen addition: (about
    330 g/m2 compared with about 210 g/m2 in the untreated plots). The
    dry weight distribution of the species was significantly affected by
    nutrient treatments.  In the nitrogen-treated vegetation, the dry
    weight of the grass species  Brachypodium pinnatum was about 3 times
    higher than in the control plots.  Nitrogen application also resulted
    in a drastic reduction of the biomass of forb species (including
    several Dutch Red List species) and of the total number of species.
    The observed decrease in species diversity in the nitrogen-treated
    vegetation was not caused by nitrogen toxicity, but by the change in
    vertical structure of the grassland vegetation. The growth of
     Brachypodium was strongly stimulated and its overtopping leaves
    reduced the light within the vegetation.  It overshadowed the other
    characteristic species and growth of these species declined rapidly
    (Bobbink et al., 1988; Bobbink, 1991).  Stimulation of  Brachypodium
    growth and a substantial reduction in species diversity were observed
    following application of nitrogen to a 5-year permanent plot study
    using a factorial design (Willems et al., 1993).

         Many characteristic lichens and mosses have also disappeared in
    recent years from European calcareous grasslands (During & Willen,
    1986).  This has been caused partly by the indirect effects of extra
    nitrogen inputs, as shown experimentally by Van Tooren et al. (1990). 
    Data on the effects of nitrogen eutrophication on the species-rich
    fauna of calcareous grassland are not available.  However, it is very
    likely that the diversity of animals, especially of insects, will also
    be reduced when tall grasses are strongly dominating the vegetation,
    because of the decreasing abundance of many herbaceous flowering
    species which act as host or forage plants.

    b)  Nitrogen enrichment and nutrient storage in calcareous grasslands

         The seasonal distribution of nutrients after nitrogen
    fertilization in spring (120 kg nitrogen as ammonium nitrate) has been
    studied with the repeated harvest approach (Bobbink et al., 1989). 
    It resulted in a significantly increased peak standing crop of
     Brachypodium . This grass proves to have very efficient nitrogen
    uptake and very efficient withdrawal from its senescent shoots into
    its well-developed rhizome system.  Brachypodium can benefit from the
    extra nitrogen redistributed to the below-ground rhizomes by enhanced

    growth in the next spring.  The distribution of nitrogen has also been
    quantified in 3-year fertilization experiments.   Brachypodium
    greatly monopolized (> 75%) the nitrogen storage in both the
    above-ground and below-ground compartments of the vegetation with
    increasing nitrogen availability (Bobbink et al., 1988; Bobbink,
    1991).

         Nitrogen cycling and accumulation in calcareous grassland can be
    significantly influenced by two major outputs from the system:
    (i) leaching from the soil; and (ii) removal with management regimes. 
    Nitrogen losses by denitrification in dry calcareous grasslands are
    low (< 1 kg nitrogen per ha per year), owing to the well-drained soil
    conditions (Mosier et al., 1981).  Ammonium and nitrate leaching has
    been studied in Dutch calcareous grasslands by Van Dam et al. (1992). 
    Both the water fluxes and the solute fluxes at different soil depths
    have been measured over 2 years in untreated vegetation and in
    calcareous grassland vegetation sprayed at 2-weekly intervals with
    ammonium sulfate (50 kg nitrogen per ha per year).  The nitrogen
    leaching from the untreated vegetation was very low (0.7 kg nitrogen
    per ha per year) and amounted to only 2% of the total atmospheric
    deposition of nitrogen.  After the spraying with ammonium sulfate,
    nitrogen leaching increased significantly to 3.5 kg nitrogen per ha
    per year, although this figure was also a very small proportion (4%)
    of the nitrogen input in this vegetation (Van Dam et al., 1992).  It
    is thus evident that calcareous grassland ecosystems retain nitrogen
    almost completely in the system.  This is caused by a combination of
    enhanced plant uptake (Bobbink et al., 1988; Bobbink, 1991) and
    increased immobilization in the soil organic matter (Van Dam et al.,
    1992).

    4.2.3.2  Critical loads for nitrogen in calcareous grasslands

         The most important output of nitrogen from calcareous grassland
    is via exploitation or management.  The annual nitrogen removal in the
    hay varies slightly between years and sites, but in general between
    17 and 22 kg nitrogen per ha is removed from the system under normal
    management conditions in the Netherlands (Bobbink, 1991; Bobbink &
    Willems, 1991).  The ambient nitrogen deposition in Dutch calcareous
    grasslands, as determined by Van Dam (1990), is high (35-40 kg
    nitrogen per ha per year) and is nowadays the major nitrogen input to
    the system.  Legume species  (Leguminosae) also occur in calcareous
    vegetation, and form an additional nitrogen input owing to the
    nitrogen-fixing microorganisms in their root nodules (about 5 kg
    nitrogen per ha per year).

         The nitrogen mass balance of Dutch calcareous grasslands is
    summarized in Table 25. It is obvious that calcareous grasslands now
    significantly accumulate nitrogen (16-26 kg per ha per year) in the
    Netherlands.  A critical nitrogen load has been determined with a mass
    balance model, because of the lack of long-term addition experiments

    with low nitrogen loads.  Assuming a critical long-term immobilization
    rate for nitrogen of 0-6 kg nitrogen per ha per year, the critical
    nitrogen load can be derived by adding the nitrogen fluxes due to net
    uptake, denitrification and leaching, corrected for the nitrogen input
    by fixation. In this way, 15-25 kg nitrogen per ha per year is
    considered as nitrogen critical load for this ecosystem. Nitrogen
    cycling within the system (between plants and soil) is not used for
    this calculation.

    Table 25.  Nitrogen mass balance (kg nitrogen per ha per year)
               for dry calcareous grassland in the Netherlands
                                                                        
 
    Input                              Output
                                                                        

    Atmospheric deposition   35-40     Harvest             17-22
    Nitrogen fixation        5         Denitrification     1
                                       Soil leaching       1

    Total                    40-45     Total               19-24
                                                                        

         In calcareous grassland in England, addition of nitrogen
    stimulated the dominance of grasses in most cases (Smith et al., 1971;
    Jeffrey & Pigott, 1973).  In these studies, the application of
    50-100 kg nitrogen per ha per year resulted in a strong dominance of
    the grasses  Festuca rubra, F. ovina or  Agrostis stolonifera.
    However,  Brachypodium and  Bromus erectus, the most frequent
    species in calcareous grassland in continental Europe, were absent
    from these sites.  Following a survey of data from a number of
    conservation sites in southern England, Pitcairn et al. (1991)
    concluded that  Brachypodium had expanded in the United Kingdom
    during the last 100 years.  They considered that much of the early
    spread could be attributed to a decline in grazing pressure but that
    the more recent spread had, in some cases, taken place despite grazing
    or mowing, and could be related to nitrogen inputs.  However, a study
    of chalk grassland at Parsonage Downs (United Kingdom) showed no
    substantial change in species composition over the twenty years
    between 1970 and 1990, a period when nitrogen deposition is thought to
    have increased significantly (Wells et al., 1993).   Brachypodium was
    present in the sward but had not expanded as in the Dutch grasslands. 
    In a linked experimental study, applications of nitrogen to eight
    forbs and one grass  (Brachypodium) at levels of 20, 40 and 80 kg
    nitrogen per ha per year for two years did not result in
     Brachypodium becoming dominant.

         Apart from the Dutch studies, the effects of enhanced nitrogen
    inputs have been little investigated in continental European
    calcareous grasslands.  Some data from a recent fertilization

    experiment at the alvar grasslands, a thin-soiled vegetation over flat
    limestone, on the Swedish island Öland, suggest that the vegetation
    hardly responds to applications of nitrogen or phosphorus (Sykes & Van
    der Maarel, 1991; personal communication by Van der Maarel).  Only
    irrigation in combination with nutrients has caused an increase in
    grasses.  This is probably due to the low annual precipitation in this
    area (400-500 mm).

         Increased nitrogen availability is probably of major importance
    in many European calcareous grasslands. An increased availability of
    nitrogen is indicated by enhanced growth of some tall grasses,
    especially stress-tolerant species, which have a slightly higher
    potential growth rate and efficient nitrogen utilization.  It clearly
    depends on the original species composition, as to which of the
    grass species will increase following enhanced nitrogen inputs. 
    Furthermore, the difference between the Dutch and United Kingdom
    results may reflect differences in management; the impacts of grazing
    in the United Kingdom grasslands could offset any competitive
    advantage the grasses may have obtained from additional nitrogen
    inputs.  The critical load for nitrogen in these calcareous grasslands
    could be influenced by management; long-term studies involving
    additional nitrogen input with various management schemes are needed
    to quantify these aspects.

    4.2.3.3  Comparison with other semi-natural grasslands

         Productivity in grasslands is strongly influenced by nutrients,
    as shown in many agricultural studies (e.g. Chapin, 1980).  It is also
    well-known that large amounts of fertilizer (nitrogen, phosphorus and
    potassium) alter almost all grassland types into tall, species-poor
    swards dominated by a few highly productive crop grasses (e.g. Bakelaar
    & Odum, 1978; Van Den Bergh, 1979; Willems, 1980; Van Hecke et al.,
    1981).  Most of these species-rich grasslands are deficient in
    nitrogen or phosphorous, and thus characterized by plant species of
    nutrient-poor habitats.  It is thus likely that these grasslands are
    sensitive to nitrogen eutrophication from the atmosphere (Wellburn,
    1988; Ellenberg, 1988b).  Thus, it is also important to establish
    critical loads for nitrogen in the species-rich grasslands, although
    data from experiments with nitrogen application in these semi-natural
    grasslands are scarce.

         Increased nitrogen availability can also affect other
    semi-natural grasslands, although experimental evidence is quite
    scarce.  A classical study into the effects of nutrients on neutral
    grasslands is the Park Grass experiment at Rothamsted, England, which
    has been running since 1856 (Williams, 1978). After application of
    nitrogen as ammonium sulfate or sodium nitrate (48 kg nitrogen per ha
    per year), the vegetation became very poor in species and dominated by
    grasses such as  Holcus lanatus or  Agrostis sp.  The effects of
    nutrients in dry and wet dune grasslands (1% calcium carbonate) on

    sandy soils have been studied at Braunton Burrows (Devon, England) by
    Willis (1963).  Nutrients were applied over 2 years (6 × 40 kg
    nitrogen per ha per year) using a factorial design for nitrogen and
    phosphorus.  Nitrogen proved to be the most important nutrient in
    stimulating the growth of some grass species ( Festuca rubra and  Poa
     pratensis).  This enhanced growth reduced significantly the abundance
    of many small plants such as prostrate phanerogamic species, mosses
    and lichens (Willis, 1963).  In this coastal area with low nitrogen
    deposition (currently about 10 kg nitrogen per ha per year) the
    vegetation of dune grasslands is at present still species-rich,
    whereas in many Dutch dune grasslands with higher nitrogen loading
    (20-30 kg nitrogen per ha per year) certain grasses have increased and
    it has become a problem to maintain diversity.  Recent studies of the
    response of mesothrophic grasslands in the United Kingdom have shown
    that additions as small as 25 kg per ha per year can lead to changes
    in species diversity after several years of fertilizer additions and
    that changes take place more rapidly at higher rates of addition
    (Mountford et al., 1994).  This indicates that many of these
    semi-natural grasslands are also sensitive to nitrogen eutrophication
    and that the critical loads are likely to be of the same magnitude or
    slightly higher (20-30 kg nitrogen per ha per year) than in calcareous
    grasslands.

         Many other semi-natural grassland types occur, especially in the
    montane-subalpine regions, containing a large proportion of the
    biodiversity of the area.  However, data are too scarce to establish
    reliable load for these grasslands, although it may be expected that:
    (i) most of these grassland are sensitive to nitrogen; and (ii) the
    critical load for nitrogen is probably lower than for lowland
    (calcareous) grasslands.  The presented critical loads for
    species-rich grasslands are summarized in section 8.2.2.

    4.2.4  Effects on heathlands

         Various types of plant communities have been described as heath,
    but the term is applied here to plant communities where the dominant
    vegetation is small-leaved dwarf-shrubs forming a canopy of 1 m or
    less above soil surface.  Grasses and forbs may form discontinuous
    strata, and there is frequently a ground layer of mosses or lichens
    (Gimingham et al., 1979; De Smidt, 1979).  Dwarf-shrub heathlands
    occur in various parts of the world, especially in montane habitats,
    but are widespread in the atlantic and sub-atlantic parts of Europe. 
    In these parts of the European continent, natural heathland is limited
    to a narrow coastal zone.  Inland lowland heathlands are man-made
    (semi-natural), although they have existed for several centuries. 
    Lowland healths are widely dominated by the  Ericaceae, especially
     Calluna vulgaris in the dry heathlands and  Erica tetralix in the
    wet heathlands (Gimingham et al., 1979).  In these heaths, development
    towards woodland has been prevented by mowing, burning, sheep grazing
    and sod removal.

         Until the beginning of this century, the balance of nutrient
    input and output was in equilibrium in the lowland heathlands of
    western Europe (De Smidt, 1979; Gimingham & De Smidt, 1983).  The
    original land use implied a regular, periodic removal of nutrients
    from the ecosystems via grazing and sod removal (Heil & Aerts, 1993).
    Sod removal was practised less systematically in many Scandinavian and
    Scottish heathlands (Gimingham & De Smidt, 1983).  Here  Calluna has
    been renewed by burning at regular intervals, varying from 4-6 years
    in southern Sweden to 15-20 years in western Norway (Nilsson, 1978;
    Skogen, 1979).  The original land use of the lowland heathland ceased
    in the early 1900s and the area occupied by this community decreased
    markedly all over its distribution area (Gimingham, 1972; De Smidt,
    1979; Ellenberg, 1988b).  Dwarf-shrub heathlands may be divided into
    four categories according to broad differences in habitat: (1) dry
    heathlands; (2) wet heathlands; (3) montane and (4) arctic-alpine
    heathlands.

    4.2.4.1  Effects on inland dry heathlands

         During recent decades many lowland heathlands in western Europe
    have become dominated by grass species.  An evaluation, using aerial
    photographs, has shown that more than 35% of Dutch heathland has been
    altered into grassland (Van Kootwijk & Van der Voet, 1989).  In recent
    years, similar changes have been observed in SW Norway, which has the
    largest local emission of ammonia as well as the heaviest nitrogen
    input through long-range deposition in Norway (Anonymous, 1991).  It
    has been suggested that nitrogen eutrophication might be a significant
    factor in this transition to grasslands.  Field and laboratory
    experiments confirm the importance of nutrients, especially in the
    early phase of heathland development (Heil & Diemont, 1983; Roelofs
    1986; Heil & Bruggink, 1987; Aerts et al., 1990).  However,  Calluna
    can compete successfully with the grasses, even at high nitrogen
    loading, if its canopy remains closed (Aerts et al., 1990).  Apart
    from the changes in competitive interactions between  Calluna and the
    grasses, heather beetle plagues and nitrogen accumulation in the soil
    are important factors in the changing lowland heaths.  Furthermore,
    evidence is growing that frost sensitivity of the dominant
    dwarf-shrubs may also be affected by increasing nitrogen inputs.

         Heathland canopies have a strong filtering effect on air
    pollutants, a varying canopy structure being an important factor.  For
    sulfur and nitrogen it has been shown that bulk deposition accounts
    for only about 35-40% of total atmospheric input (Heil et al., 1987;
    Bobbink et al., 1992b).  Total atmospheric deposition of nitrogen is
    30-45 kg nitrogen per ha per year in the heathland sites in the
    eastern part of the Netherlands.  More than 70% of the total nitrogen
    input is deposited as ammonium or ammonia (Bobbink et al., 1992b;
    Bobbink & Heil, 1993).  Although data for nitrogen inputs in other

    European lowland heathlands are missing, it is very likely that in
    many European agricultural regions nitrogen deposition has increased
    in recent years (Asman, 1987; Buijsman et al., 1987).

         In  Calluna heathland, outbreaks of the chrysomelid heather
    beetle  (Lochmaea suturalis) occur frequently.  These beetles feed
    exclusively on the green parts of  Calluna.  The closed  Calluna
    canopy is opened over large areas and the interception of light by
     Calluna decreases strongly (Berdowski, 1987, 1993).  Thus the growth
    of the under-storey grasses ( Deschampsia or  Molinia) is enhanced
    significantly.  In general insect grazing is affected by the nutritive
    value of the plant material, and the nitrogen content is especially
    important in this respect (Crawley, 1983).  Experimental applications
    of nitrogen to heathland vegetation cause the concentrations of this
    element in the green parts of  Calluna to increase (Heil & Bruggink,
    1987; Bobbink & Heil, 1993).  It is, therefore, very likely that the
    frequency and intensity of heather beetle outbreaks are stimulated by
    increased atmospheric nitrogen input in Dutch heathland.  This
    hypothesis is supported by the observations of Blankwaardt (1977), who
    reported that from 1915 onwards heather beetle outbreaks were observed
    in the Netherlands with an interval of about 20 years, whereas in the
    last 15 years the outbreaks have occurred with a periodicity of less
    than 8 years.  It has also been observed that during a heather beetle
    outbreak  Calluna plants are more severely damaged in nitrogen-
    fertilized vegetation (Heil & Diemont, 1983). In a rearing experiment
    with larvae of the heather beetle, Brunsting & Heil (1985)
    demonstrated that the growth of the larvae was increased by higher
    nitrogen concentrations in the leaves of  Calluna.  Van der Eerden
    et al. (1990) studied the effects of ammonium sulfate on the growth of
    heather beetle after a outbreak of the beetle in vegetation
    artificially sprayed under a cover.  They found no significant effect
    of the treatments on total number or on biomass of the first stage
    larvae.  However, the development of subsequent larval stages was
    accelerated by the application of ammonium sulfate in the artificial
    rain: the percentage of third stage larvae increased by 20%, compared
    with larvae in the control treatment.  Furthermore, heather beetle
    larvae were put on  Calluna shoots taken from plants which had been
    fumigated with ammonia in open-top chambers (12 months; 4 to
    105 µg/m3) (Van der Eerden et al., 1991).  After 7 days the larvae
    were counted and weighed.  Both the mass and the development rate of
    the larvae clearly increased with increasing concentrations of
    ammonia.  The heather beetle has recently been found in SW Norway and
    it is expanding its territory.  It is probably an important cause of
     Calluna death in this region (Hansen, 1991).  It can be concluded
    that nitrogen inputs influence outbreaks of heather beetle, although
    the exact relationship between both processes needs further research.

         In the past Dutch inland heathlands were grazed by flocks of
    sheep and sods were generally removed at intervals of 25-50 years
    (De Smidt, 1979).  Nowadays these heathlands are mostly managed by

    mechanical sod removal, which can restore the heathland vegetation if
    a seed bank of the original heathland species is still present
    (Bruggink, 1993).  The increase in organic matter and in the amounts
    of nitrogen in the system during secondary succession is well known
    (Begon et al., 1990). It was shown in the 1970s that during secondary
    heathland succession the above-ground and below-ground biomass and the
    amount of litter increase (Chapman et al., 1975; Gimingham et al.,
    1979).  It is likely that changes in nitrogen accumulation will have
    occurred due to the increase in atmospheric deposition.

         Berendse (1990) performed a detailed study on the accumulation of
    organic matter and of nitrogen during the secondary succession after
    sod removal in the Netherlands.  He found a large increase in plant
    biomass, soil organic matter and total nitrogen storage in the first
    20 to 30 years after sod removal.  Furthermore, it was demonstrated
    that nitrogen mineralization was low during the first 10 years (about
    10 kg nitrogen per ha per year), but increased considerably over the
    next 20 years to 50-110 kg nitrogen per ha per year.  Regression
    analysis of the total nitrogen storage versus the years after sod
    removal revealed an annual nitrogen increase in the system of about
    33 kg nitrogen per ha per year (probably somewhat lower in the early
    years and higher in later years) (Berendse, 1990).  These values are
    in good agreement with measured nitrogen deposition in Dutch
    heathlands in the late 1980s (Bobbink et al., 1992b).

         Thus, the organic matter in the soil increases rapidly after sod
    removal, which removes almost all of the soil organic matter. 
    However, this process is accelerated by the enhanced dry matter
    production and litter production of the dwarf shrubs caused by the
    extra nitrogen inputs.  Nitrogen accumulation in the system also
    increases. Hardly any nitrogen disappears from the system because
    nitrate leaching to deeper layers is only of minor importance in Dutch
    heathlands, as shown by De Boer (1989) and Van Der Maas (1990). 
    Nitrogen availability from atmospheric inputs, in addition to
    mineralization, is within a relatively short period (about 10 years)
    high enough to stimulate the transition of heathland to grassland,
    especially after the opening of the heather canopy by secondary
    causes.

         It has been demonstrated that frost sensitivity of some tree
    species increases with increasing concentrations of air pollutants
    (e.g. Aronsson, 1980; Dueck et al., 1991).  This increase in frost
    sensitivity is sometimes correlated with enhanced nitrogen
    concentrations in the foliage of the trees.  Long-term effects of air
    pollutants on the frost sensitivity of  Calluna and  Erica are to be
    expected because of (i) the evergreen growth form of these species and
    (ii) the increasing content of nitrogen in the leaves of  Calluna,
    associated with increased nitrogen deposition in the Netherlands and
    Norway (Heil & Bruggink, 1987; Hansen, 1991).  It has been suggested
    that damage to  Calluna shoots in the successive severe winters of
    the mid-1980s was at least partly caused by the increased frost

    sensitivity.  Investigations into the effects of air pollutants on the
    frost sensitivity of heathland species outside the Netherlands started
    in the early 1990s (Hansen, 1991; Uren, 1992).

         The effects of sulfur dioxide, ammonium sulfate and ammonia upon
    frost sensitivity in  Calluna have been studied by Van der Eerden
    et al. (1990).  After fumigation with sulfur dioxide (90 µg/m3 for
    3 months), increased frost injury in  Calluna was only found at
    temperatures that seldom occur in the Netherlands (lower than -20°C). 
    Fumigation with ammonia of  Calluna plants in open-top chambers over
    a 4-7 month period (100 µg/m3) revealed that frost sensitivity was
    not affected in autumn (September or November), whereas in February,
    just before growth started, frost injury increased significantly at
    -12°C (Van der Eerden et al., 1991).  These authors also studied the
    frost sensitivity of  Calluna vegetation sprayed with six different
    levels of ammonium sulfate (3-91 kg nitrogen per ha per year). The
    frost sensitivity increased slightly, although significantly, after
    5 months in vegetation treated with the highest level of ammonium
    sulfate (400 µmol/litre; 91 kg nitrogen per ha per year), compared
    with the control treatments.  However, frost sensitivity of  Calluna
    decreased again two months later and no significant effects of the
    ammonium sulfate application upon frost hardiness were seen at that
    time. Thus, high levels of ammonia or ammonium sulfate seem to
    increase the frost sensitivity of  Calluna plants, although the
    significance of this phenomenon is still uncertain at ambient nitrogen
    inputs.  The relation between frost sensitivity and nitrogen input has
    not yet been sufficiently quantified to use it for a precise
    assessment of critical loads in this respect.

         It has been shown above that atmospheric nitrogen is the trigger
    for changes of lowland dry heathlands into grass swards in the
    Netherlands.  Lowland dry heathlands in the United Kingdom do not show
    consistent patterns over the past 10 to 40 years.  Pitcairn et al.
    (1991) assessed changes in abundance of  Calluna in three heaths in
    East Anglia (eastern England) over recent decades.  All three heaths
    showed a decline in  Calluna and an increase in grasses.  The authors
    concluded that increases in nitrogen deposition was at least partly
    responsible for the changes, but also noted that the management had
    changed.  A wider assessment of heathlands in SE England showed that
    in some cases  Calluna had declined and subsequently been invaded by
    grasses while other areas were still dominated by dwarf shrubs (Marrs,
    1993).  This clearly stresses the importance of management for the
    maintenance of dwarf shrubs in heathlands.  A simulation model, which
    integrates processes such as atmospheric nitrogen input, heather
    beetle outbreak, soil nitrogen accumulation, sod removal and
    competition between species, has been used to establish the critical
    loads of nitrogen deposition in lowland dry heathlands (Heil &
    Bobbink, 1993a,b).  The model has been calibrated with data from field
    and laboratory experiments in the Netherlands.  As an indicator of the
    effects of atmospheric nitrogen, the proportion and increase of

    grasses in the heathland system are used.  Atmospheric nitrogen
    deposition has varied between 5 and 75 kg nitrogen per ha per year in
    steps of 5-10 kg nitrogen during different simulations.  From these
    simulations, the value for the critical load of nitrogen for the
    changes from dwarf shrubs to grasses was 15-20 kg nitrogen per ha per
    year.

    4.2.4.2  Effects of nitrogen on inland wet heathlands

         The western European lowland heathlands of wet habitats are
    dominated by the dwarf shrub  Erica tetralix (Ellenberg, 1988b) and
    are generally richer in plant species than the dry heathlands.  In
    recent decades a drastic change in species composition of Dutch wet
    heathlands has been observed.  Nowadays, many wet heathlands that were
    originally dominated by  Erica have become monospecific stands of the
    grass  Molinia.  Together with  Erica almost all of the rare plant
    species have disappeared from the system.  It has been hypothesized
    that this change has been caused by atmospheric nitrogen
    eutrophication.

         Competition experiments using micro-ecosystems have clearly shown
    that  Molinia is a better competitor than  Erica at high nitrogen
    availability.  After 2 years of application of nitrogen (150 kg per ha
    per year), the relative competitive strength of  Molinia compared
    with  Erica doubled (Berendse & Aerts, 1984).  A 3-year field
    experiment with nitrogen application in Dutch lowland wet heathland
    (around 160 kg nitrogen per ha per year) also indicated that  Molinia
    is able to outdo  Erica at high nitrogen availability (Aerts &
    Berendse, 1988).  In contrast to the competitive relations between
     Calluna and the grasses,  Molinia can outdo  Erica without opening
    of the dwarf shrub canopy.  This difference is caused by the lower
    canopy of  Erica (25-35 cm), compared with  Calluna, and the tall
    growth form of  Molinia, which can overgrow and shade  Erica if
    enough nitrogen is available.  It is in this respect also important
    that heather beetle plagues do not occur in wet heathlands and that no
    frost damage has been observed in this community.

         It has been demonstrated that in many Dutch wet heathlands the
    accumulation of litter and humus has led to increased nitrogen
    mineralization (100-130 kg nitrogen per ha per year) (Berendse et al.,
    1987).  In the first 10 years after sod removal the annual nitrogen
    mineralization is very low, but afterwards it increases rapidly.  The
    leaching of accumulated nitrogen from wet heathlands is extremely low
    (Berendse, 1990).  The observed nitrogen availabilities are high
    enough to change  Erica -dominated wet heathlands into monostands of
     Molinia.

         Berendse (1988) developed a wet heathland model to simulate
    carbon and nitrogen dynamics during secondary succession.  He
    incorporated in this model the competitive relationships between

     Erica and  Molinia, the litter production from both species, soil
    nitrogen accumulation and mineralization, leaching, atmospheric
    nitrogen deposition and sheep grazing.  He simulated the development
    of lowland wet heathland after sod removal, because almost all of the
    Dutch communities are already strongly dominated by  Molinia and it
    is impossible to expect changes in this situation without drastic
    management.  Using the biomass of  Molinia with respect to  Erica as
    an indicator, his results suggested 17-22 kg nitrogen per ha per year
    as the critical load for the transition of lowland wet heathland into
    a grass-dominated sward (Berendse, 1988).  The decrease in endangered
    wet heathland forbs is partly caused by the overshading by  Molinia,
    but some species had already disappeared from wet heathlands before
    the increase of  Molinia started.  The critical load for this decline
    is probably lower than the given values and is discussed in section
    4.2.4.4.

    4.2.4.3  Effects of nitrogen on arctic and alpine heathlands

         Semi-natural  Calluna heathlands are found in the lowlands along
    the Norwegian coast to 68°N and show distinct plant gradients in the
    south-north direction, from coast to inland and from lowland to upland
    areas (Fremstad et al., 1991).  In central parts of western Norway the
    plant composition changes at an altitude of about 400 m, above which
    alpine species occur regularly in the heaths.  At this altitude
    oceanic upland  Calluna and  Erica heaths merge into alpine heaths,
    which are naturally occurring, non-anthropogenic communities.  Some
    oligotrophic alpine heaths also contain  Calluna, but most heaths in
    Fennoscandia and in European parts of Russia are dominated by other
    ericoid species ( Vaccinium spp.,  Empetrum nigrum s. lat.,
     Arctostaphylos spp.,  Loiseleuria procumbens, Phyllodoce caerulea,
     Betula nana, Juniperus communis and  Salix spp.).  Many heath types
    have a more or less continuous layer of mosses and lichens. Related
    heaths are found in alpine regions in the British Isles, in Iceland,
    in southernmost Greenland, in northern Russia, and on siliceous rocks
    in the Alps (Grabherr, 1979; Elvebakk, 1985; Ellenberg, 1988b).

         Alpine and arctic habitats have many ecological characteristics
    in common, although the climatic conditions are more severe in the
    arctic regions than in most alpine regions.  The growing season is
    short (3-3.5 months in the low arctic zone), air and soil temperatures
    are low, winds are frequent and strong, and the distribution of plant
    communities depends on the distribution of snow during winter and
    spring.  Most alpine and all arctic zones are influenced by frost
    activity or solifluction, except for soils in the low alpine and
    hemiarctic zones, where podzolic soils are found.  Decomposition of
    organic matter and nutrient cycling are slow, and a large amount of
    the nitrogen input is stored in the soil in forms which can not be
    used by plants (Chapin, 1980).  The low nutrient availability limits

    primary production.  Most species are adapted to a strict nitrogen
    economy and their nitrogen indicator values are generally low
    (Ellenberg, 1979).

         Barsdate & Alexander (1975) investigated the nitrogen balance of
    an arctic area in Alaska. The most important sources of nitrogen were
    nitrogen fixation (75%) and ammonia in precipitation (22%).  Most of
    the nitrogen input is retained in living biomass, and very little is
    leached from the soil.  Denitrification is also low, partly due to
    nutrient deficiency.  Nitrogen metabolism as such does not seem to be
    inhibited by low soil temperatures (Haag, 1974).  Nitrogen fixation in
    arctic habitats has been studied in bacteria, soil algae, lichens and
    legume species  (Leguminosae) (Novichkova-Ivanova, 1971).  Blue-green
    algae (cyanobacteria) are especially important in this respect, either
    as free-living species, species associated with mosses or phycobionts
    in lichens (e.g.  Peltigera, Nephroma and  Stereocaulon).  The rate
    of nitrogen fixation depends on temperature and moisture, and thus
    varies through the year (Alexander & Schnell, 1973).

         It is to be expected that arctic and alpine communities are
    sensitive to increased atmospheric nitrogen input, because nitrogen
    retention is very efficient, although primary production is also
    strongly regulated by factors other than nitrogen (temperature,
    moisture) (Tamm, 1991).  The effects of increased nitrogen
    availability on alpine/tundra vegetation have been studied in several
    fertilizer experiments.  In most experiments full nitrogen, phosphorus
    and potassium fertilizer was used, although sometimes nitrogen was
    applied separately.  The following effects of nitrogen addition have
    been observed:

    *    In alpine and arctic vegetation, nitrogen is quickly absorbed by
         phanerogamic species and incorporated into their tissues.  The
         increase in nitrogen contents differs for graminoids, deciduous
         and evergreen species (Summers, 1978; Shaver & Chapin, 1980;
         Lechowicz & Shaver, 1982; Karlsson, 1987).

    *    Phanerogamic plant species respond to nitrogen application in
         different ways: increased growth and biomass, enhanced number of
         tillers, more flowers and changes in phenology (Henry et al.,
         1986).

    *    Some phanerogamic plant species are damaged or even killed at
         high doses of nitrogen fertilizer (Henry et al., 1986).

    *    Changes in species cover and composition are likely when nitrogen
         is applied for a longer period of time (5-10 years).

         All these studies concentrated on effects on phanerogamic plant
    species; little information is available on the effects of nitrogen on
    cryptogams.  Many authors, however, stress that nitrogen fixation
    probably will decrease when atmospheric deposition increases in arctic

    and alpine ecosystems.  In forest studies it has already been shown
    that  Cladonia spp. and some mosses are very sensitive to nitrogen
    addition.  The suggested critical load for arctic and alpine heaths
    (5-15 kg nitrogen per ha per year) is lower than that for lowland
    heathland (15-20 kg nitrogen per ha per year).

    4.2.4.4  Effects on herbs of matgrass swards

         In recent decades, in addition to the transition from
    dwarf-shrub-dominated to grass-dominated heathlands, a reduced species
    diversity in these ecosystems has been observed.  Species of the
    acidic Nardetalia grasslands and related dry and wet heathlands seem
    to be especially sensitive.  Many of these herbaceous species (e.g.,
     Arnica montana, Antennaria dioica, Dactylorhiza maculata, Gentiana
     pneumonanthe, Genista pilosa, Genista tinctoria, Lycopodium inundatum,
     Narthecium ossifragum, Pedicularis sylvatica, Polygala serpyllifolia
    and  Thymus serpyllum) are declining or have even become locally extinct
    in the Netherlands.  The distribution of these species is related to
    small-scale, spatial variability of the heathland soils.  It has been
    suggested that atmospheric deposition has caused such changes (Van Dam
    et al., 1986).  Dwarf shrubs as well as grass species are nowadays
    dominant in the former habitats of these endangered species.

         Enhanced nitrogen fluxes into nutrient-poor heathland soil
    leads to an increased nitrogen availability in the soil.  However,
    most of the deposited nitrogen in western Europe originates from
    ammonia/ammonium deposition and may also cause acidification as a
    result of nitrification.  Whether eutrophication or acidification or a
    combination of both processes is important depends on pH, buffer
    capacity and nitrification rates of the soil.  Roelofs et al. (1985)
    found that, in dwarf-shrub-dominated heathland soils, nitrification is
    inhibited at pH 4.0-4.2 and that ammonium accumulates while nitrate
    decreases to almost zero at these or lower pH values.  Furthermore,
    nitrification has been observed in the soils from the habitats of the
    endangered species, owing to its somewhat higher pH and higher buffer
    capacity.  In soils within the pH rage of 4.1-5.9, the acidity
    produced is buffered by cation exchange processes (Ulrich, 1983). The
    pH will drop when calcium is depleted, and this may cause the decline
    of those species that are generally found on soils with somewhat
    higher pH.  To study the pH effects on root growth and survival rate,
    hydroculture experiments have been conducted over 4-week periods with
    several of the endangered species ( Arnica, Antennaria, Viola,
     Hieracium pilosella and  Gentiana) and with the dominant species
    ( Molinia and  Deschampsia) (Van Dobben, 1991).  The dominant
    species indeed have a lower pH optimum (3.5 and 4.0, respectively)
    than the endangered species (4.2-6.0).  However, the endangered
    species could survive very low pH without visible injuries during this
    short experimental period.

         The pH decrease may indirectly result in an increased leaching of
    base cations, increased aluminium mobilization and thus enhanced
    aluminium/calcium (Al/Ca) ratios of the soil (Van Breemen et al.,
    1982).  Furthermore, the reduction of the soil pH may inhibit
    nitrification and result in ammonium accumulation and consequently
    increased NH4/NO3 ratios.  In a recent field study the
    characteristics of the soil of several of these threatened heathland
    species have been compared with the soil characteristics of the
    dominant species ( Calluna vulgaris, Erica tetralix and  Molinia
     caerulea) (Houdijk et al., 1993). Generally the endangered species
    grow on soil with higher pH, lower nitrogen content, and lower Al/Ca
    ratios than the dominant species.  The NH4+/NO3 ratios were higher
    in the dwarf-shrub-dominated soils than in the soil of the endangered
    species.  Fennema (1990, 1992) has demonstrated that soil from
    locations where  Arnica is still present had a higher pH and lower
    Al/Ca ratio than soil of former  Arnica stands.  However, he found no
    differences in total soil nitrogen or NH4/NO3 ratios.  Both these
    studies indicate that high Al/Ca ratios or even increased NH4/NO3
    ratios are associated with the decline of these species.  However, no
    significant effects of Al and Al/Ca on growth rates have been observed
    in hydroculture experiments in which the effects of Al and Al/Ca
    ratios on root growth and survival rate were studied (Van Dobben,
    1991).  Comparable experiments of Pegtel (1987) with  Arnica and
     Deschampsia and Kroeze et al. (1989) with  Antennaria, Viola,
     Filago minima, and  Deschampsia showed similar results.  However,
    results of a hydroculture experiment with  Arnica showed that this
    species is very sensitive to enhanced Al/Ca ratios at intermediate or
    low nutrient levels (De Graaf, 1994).  Pot experiments have indicated
    that increased NH4/NO3 ratios lead to decreased health of  Thymus.
    Hydroculture experiments with this plant species confirmed that
    increased NH4/NO3 ratios affected the cation uptake (Houdijk, 1993). 
    In a pot experiment  Thymus, planted on acid heathland soil and on
    artificially buffered heathland soil, was sprayed with 0, 15 and
    150 kg nitrogen (as ammonium) per ha per year during 6 months (Houdijk
    et al., 1993).  In this relatively short period, a deposition of 15 kg
    nitrogen (as ammonium) per ha per year on the acid soil did not lead
    to ammonium accumulation in the soil.  As a result of nitrification,
    soil pH decreased faster than in the absence of ammonium deposition. 
    At the highest deposition (150 kg nitrogen (as ammonium) per ha per
    year), nitrification rates in the acid heathland soils were too low to
    prevent ammonium accumulation, and increased NH4/NO3 ratios probably
    caused the decline of  Thymus.  Only in the artificially buffered
    soils with higher pH were nitrification rates high enough to balance
    ammonium and nitrate.   Thymus plants on these soils were healthy
    despite very high total nitrogen contents.

         At present, however, there is too little information available on
    these rare heathland and acidic grassland species to formulate a
    critical load for nitrogen.  The observation that these heathland

    species generally disappear before dwarf shrubs are replaced by
    grasses leads to the assumption that their critical load is lower than
    the critical load for the transition to grasses (< 15-20 kg nitrogen
    per ha per year) and probably between 10 and 15 kg nitrogen per ha per
    year.  An overview of the critical loads in heathlands is given in
    section 8.2.2.

    4.2.5  Effects of nitrogen deposition on forests

    4.2.5.1  Effects on forest tree species

         The growth of the vast majority of the forest tree species in the
    Northern hemisphere was until recently limited by nitrogen.  In
    forestry, nitrogen fertilizers were used to increase wood production
    (Tamm, 1991).  An increase in the supply of an essential nutrient,
    including nitrogen, will stimulate tree growth; the initial impact of
    enhanced nitrogen deposition will, therefore, be a fertilizer effect. 
    However, continued high inputs of nitrogen produces negative effects
    on tree growth (Chapin, 1980).  Until the mid-1980s, almost all of the
    research on forest decline focused on acidification, but it has now
    become evident that enhanced nitrogen deposition may also be important
    in recent forest decline.

         The effects of high atmospheric nitrogen input are very complex
    (Wellburn, 1988; Pitelka & Raynal, 1989; Heij et la., 1991; Pearson &
    Stewart, 1993).  Chronic nitrogen deposition may result in nitrogen
    saturation, when enhanced nitrogen inputs no longer stimulate tree
    growth, but start to disrupt ecosystem structure and function, and
    increased amounts of nitrogen are lost from the ecosystem in leachate
    (Agren, 1983; Aber et al., 1989; Tamm, 1991).  The nitrogen input at
    which saturation occurs depends on a number of factors including the
    amount of deposition, vegetation type and age (see chapter 3), soil
    type and management history.  The following indirect processes,
    besides the direct effect of gaseous pollutants on the shoots, are
    important:

    *     Soil acidification, due to nitrification of ammonium.  This
         process leads to accelerating leaching of base cations and, in
         poorly buffered soils, to increased dissolution of aluminium,
         which can damage fine roots development and mycorrhizas, and thus
         reduce nutrient uptake (Ulrich, 1983; Ritter, 1990).

    *     Eutrophication.  Whether ammonium will accumulate in soil or
         not is strongly dependent upon the nitrification rate and the
         deposition levels (Boxman et al., 1988).  In addition to an
         initial growth stimulation and changes in root/shoot ratio,
         ammonium accumulation will lead to an imbalance of the
         nutritional state of the soil and concomitantly of the trees
         (Roelofs et al., 1985; Van Dijk & Roelofs, 1988; Schulze et al.,
         1989; Boxman et al., 1991).  Accumulation of nitrates in the

         ecosystem may also lead to eutrophication.  As a consequence of
         all these processes, the health of the trees declines and their
         sensitivity to drought, frost, insect pests and to pathogens can
         increase markedly (Wellburn, 1988).  These phenomena may also
         play a secondary, but certainly not unimportant, role in the
         dieback of forest trees and have also been reviewed.

         Although many tree species occur in natural forest ecosystems,
    almost all studies on air pollution have concentrated on a few
    forestry tree species from acidic, nutrient-poor soils.  Most of these
    species are conifers ( Picea, Pinus and  Pseudotsuga spp.) and the
    following section concentrates on the long-term soil-mediated effects
    on these trees.  Available data on broad-leaved species ( Fagus,
     Quercus) are also considered.  Long-term effects of nitrogen
    eutrophication on the composition of the tree layer in natural forests
    may be expected but have not yet been quantified.  Soil acidification
     per se has only been briefly reviewed, because the critical load for
    acidity and tree growth is well established (Nilsson & Grennfelt,
    1988; Downing et al., 1993).

    a)  Soil-mediated changes in nutritional status of forest tree species

         It has been shown that in areas with high ammonia/ammonium
    deposition, ammonium accumulates in acid forest soils with little or
    no nitrification.  Van Dijk & Roelofs (1988) found ammonium ion
    accumulation in damaged  Pinus and  Pseudotsuga stands receiving
    60-100 kg nitrogen per ha per year, although the pH of the soil was
    the same as that in healthy stands.  This build-up of ammonium ion
    leads to increased ratios of ammonium to base cations (Roelofs et al.,
    1985; Boxman et al., 1988), a reduction of base cation uptake and,
    eventually, nutritional problems.  Using soil columns with different
    ammonium sulfate spraying treatments, critical ratios of excess
    ammonium to base cations have been determined (Boxman et al., 1988). 
    The nutritional problems of the coniferous species studied have been
    found above values of 5, 10 and 1, respectively, for the NH4/K,
    NH4/Mg and Al/Ca ratios in soil solution.  In soil with zero or a low
    nitrification rate, 10-15 kg nitrogen per ha per year is a reliable
    critical load to prevent critical ammonium to cation ratios, whereas
    in base-cation-rich soil with moderate to high nitrification rates the
    critical loads obtained are higher (20-30 kg nitrogen per ha per
    year).

         The nutritional status of the coniferous trees studied, after
    enhanced nitrogen inputs, is affected by both ammonium accumulation
    and soil acidification.  Base cation concentrations in the soil are
    reduced by leaching, whereas base cation uptake by plants is reduced
    by excess of ammonium and of aluminium.  Furthermore, root growth is
    decreased (see later).  Laboratory, greenhouse and field measurements
    in the Netherlands, Germany and southern Sweden (Van Dijk & Roelofs,
    1988; Van Dijk et al., 1989, 1990, 1992a; Hofmann et al., 1990;

    Schulze & Freer-Smith, 1991; Boxman et al., 1991, 1994; Ericsson et
    al., 1993) have shown that the complex of factors just noted produce
    severe deficiencies of magnesium and potassium in coniferous trees. 
    Most of these studies were in areas, or involved experiments, with
    large inputs (> 40-100 kg nitrogen per ha per year).

         The magnesium and phosphorus concentrations in leaves of oak
    trees  (Fagus sylvatica), a common deciduous tree in Europe,
    decreased significantly from 1984 to 1992 in permanent plots in NW
    Switzerland.  Furthermore, the magnesium concentrations in the leaves
    of young  Fagus sylvatica decreased significantly within a 4-year
    period of fertilizer application at > 25 kg nitrogen per ha per
    year (Flückiger & Braun, 1994).  In Sweden, suboptimal concentrations
    of magnesium and potassium in  Fagus leaves were found in areas with
    the highest nitrogen deposition (Balsberg-Pählsson, 1989) and addition
    of nitrogen enhanced nutritional imbalance in a 120-year-old  Fagus
    stand (Balsberg-Pählsson, 1992).  It is thus clear that this deciduous
    tree species is also sensitive to nutritional imbalance induced by
    enhanced nitrogen supply.

         Base cations are also lost from the canopy by increased leaching,
    linked to high amounts of atmospheric deposition (Wood & Bormann,
    1975; Roelofs et al., 1985; Bobbink et al., 1992b).  As a result of
    high nitrogen inputs, the organic nitrogen concentration in the
    needles of conifers has increased significantly to supra-optimal
    levels (Van Dijk & Roelofs, 1988; De Kam et al., 1991).  Concentrations
    of nitrogen-rich free amino acids, especially arginine, have
    significantly increased in the needles with high nitrogen concentration
    (> 1.5% nitrogen in  Picea abies) (Hällgren & Näsholm, 1988;
    Pietila et al., 1991; Van Dijk et al., 1992) and in  Fagus leaves
    (Balsberg-Pählsson, 1992).

         Although there is clear evidence that high NH3/NH4 loads
    produce adverse changes in the nutritional status and the growth of
    the investigated coniferous and broad-leaved trees, it is difficult to
    obtain a critical load for nitrogen from these studies, because of the
    complexity of the ecosystem.  A quite reliable critical load for
    nitrogen deposition on beech tree health is around 15-20 kg nitrogen
    per ha per year, as demonstrated in the Swiss studies (Flückiger &
    Braun, 1994).

         The results of the EC nitrogen saturation study (NITREX), which
    incorporates long-term experiments in both clean and nitrogen-polluted
    areas and whole ecosystem manipulation of nitrogen inputs, are
    providing important evidence on the effects of nitrogen deposition
    on tree health and ecosystem health.  Atmospheric deposition of
    nitrogen was reduced from 40 to 2 kg nitrogen per ha per year in a
    nitrogen-saturated  Pinus sylvestris stand in the Netherlands (Boxman
    et al., 1994, 1995).  Throughfall water was intercepted with a roof
    and replaced by clean throughfall water from 1989 onwards.  In the

    clean plot a quick response of the soil solution chemistry was
    observed.  The nitrogen concentrations in the upper soil and the
    fluxes of this element through the soil profile decreased.  As a
    result, base cation leaching and the ratios of ammonium to various
    cations also decreased; potassium and magnesium concentrations in the
    needles increased significantly.  The needle nitrogen concentrations
    were only slightly reduced in the "clean" situation, but they were
    significantly lower than in the needles of the control plots.  The
    concentration of arginine decreased significantly in the needles of
    the trees from the clean throughfall plot. Furthermore, tree growth
    became higher after 4 years of clean throughfall than in control plots
    with high nitrogen deposition. No changes in the mycorrhizal status or
    in the undergrowth have so far been observed (Boxman et al., 1994,
    1995).  This study clearly demonstrates the detrimental effects of
    enhanced atmospheric nitrogen deposition on the nutritional balance of
    coniferous trees.

    b)  Nitrogen deposition and tree susceptibility to frost, drought and
        pathogens

         It has been suggested by several authors that sensitivity of
    trees to secondary stress factors is increased by high nitrogen
    loading (Wellburn, 1988; Pitelka & Raynal, 1989).  In field fertilizer
    applications it is often observed that tree growth starts earlier in
    the season, which may increase damage by late frost. Furthermore, it
    has been shown, after nutrient applications, that frost damage to
     Pinus sylvestris increases considerably at needle nitrogen
    concentrations above 1.8% (Aronsson, 1980), although other fertilizer
    studies have demonstrated reverse effects, i.e. improved nitrogen
    status of the plants diminishes frost damage (De Hayes et al., 1989;
    Klein et al., 1989; Cape et al., 1991).

         Only few data are available with respect to frost damage in
    direct relation to airborne nitrogen deposition.  After exposure to
    NH3 and SO2,  Pinus sylvestris saplings became more frost sensitive
    (< -10°C) than control plants (Dueck et al., 1990).  Dueck et al.
    (1990) also determined the frost sensitivity of  Pinus sylvestris
    growing in areas with low ammonia/ammonium pollution (approximately
    4 µg NH3/m3) and in highly polluted areas (40 µg NH3/m3). 
    Surprisingly, the frost sensitivity was not higher in the polluted
    area than in the other investigated sites, and was sometimes even
    lower.  After experimental treatment with ammonia (53 µg NH3/m3) the
    growth of the trees had increased, indicating that the observed change
    in frost sensitivity might have occurred as a result of changes in
    physiology and nutrient imbalance.

         The effects of simulated acid mist containing sulfate, ammonium,
    nitrate and H+ on the frost sensitivity of  Picea rubens has been
    studied (Sheppard et al., 1993; Sheppard, 1994).  There was a strong

    correlation between the application of sulfate-containing mist and an
    increase in frost sensitivity, but no such correlation was seen after
    treatment with ammonium or nitrate ions.  Sulfur compounds clearly
    affect the frost sensitivity of coniferous trees, but this effect may
    be a consequence of the nutritional status (nitrogen, base cations) of
    the trees (Sheppard, 1994).  It is concluded that the effects of
    increased nitrogen inputs on frost sensitivity remain uncertain. 
    Insufficient research has been carried out to use the results for
    assessment of a critical load.

         The water uptake of coniferous trees species may be affected by
    increasing nitrogen deposition, owing to an increase in shoot-to-root
    ratio and a reduction in fine-root length.  Indeed, the health of many
    tree species in the regions of the Netherlands with high nitrogen
    deposition was particularly poor in the dry years in the mid-1980s,
    but improved again during the subsequent normal years (Heij et al.,
    1991).  Many authors have mentioned a negative impact of high nitrogen
    supply on the development of fine roots and mycorrhiza, although
    positive effects have also been described (Persson & Ahlstrom, 1991).

         Van Dijk et al. (1990) applied 0, 48, 480 kg nitrogen (as
    ammonium sulfate) per ha per year to young  Pinus sylvestris, Pinus
     nigra and  Pseudotsuga menziesii in a pot experiment.  After seven
    months the coarse root biomass had not changed, but the fine root
    biomass decreased by 36% at the highest nitrogen application.  In
    parallel, a 63% decrease in mycorrhizal infection at the highest
    nitrogen application was found.  In the Dutch EC nitrogen saturation
    study, the fine root biomass and the number of root tips of  Pinus
     sylvestris increased after reduction of the current nitrogen
    deposition to pre-industrial levels, indicating restricted root growth
    and nutrient uptake capacity at the ambient nitrogen load of about
    40 kg nitrogen per ha per year (Boxman et al., 1994, 1995).

         In a hydroculture experiment with  Pinus nigra at pH=4.0, Boxman
    et al. (1991) found an increase in coarse/fine root ratio after
    increasing the ammonium concentration to 5000 µM.  Furthermore, a
    clear relation was found between the nitrogen content of the fine
    roots and mycorrhizal infection (as measured as the number of
    dichotomously branched roots). In a hydroculture experiment Jentschke
    et al. (1991) found, however, that 2700 µM nitrate had hardly any
    effect on the mycorrhizal development of  Picea abies seedlings
    inoculated with  Lactarius rufus.  Ammonium at 2700 µM only had a
    slight negative effect on mycorrhizal development, whereas a reduction
    in root growth was recorded.  In a pot experiment with  Picea abies,
    Meyer (1988) found optimal mycorrhizal development when the mineral
    nitrogen content of the soil was 40 mg nitrogen/kg dry soil, while at
    350 mg nitrogen/kg dry soil a 95% reduction in mycorrhizal development
    was found.  In this study no correlation was found with the soil pH.
    Alexander & Fairly (1983) found, after fertilizer application to a
    35-year-old  Picea sitchensis stand with 300 kg nitrogen (as ammonium

    sulfate) per ha, a 15% reduction in mycorrhizal development in the
    second year after application.  Termorshuizen (1990) applied 0 to
    400 kg nitrogen ha per year either as ammonium or nitrate to young
     Pinus sylvestris inoculated with  Paxillus involutus in a pot
    experiment.  Above application rates of 10 kg nitrogen per ha per year
    there was a decrease in the amount of mycorrhizal root tips and the
    number of sclerotia.

         In addition to the above-mentioned data for coniferous trees, it
    had been shown that the shoot-to-root ratios of young  Fagus
     sylvatica trees, grown in containers with acid forest soil,
    increased significantly from about 1 to between 2 and 3 after a 4-year
    experimental application of nitrogen (25 kg nitrogen per ha per year
    or more) (Flückiger & Braun, 1994).

         It is thus likely that enhanced nitrogen inputs affect drought
    sensitivity through changes in shoot to root ratios, number of fine
    roots and the ectomycorrhizal infection of the roots.  However, the
    data are too few to use for the assessment of a critical load of
    nitrogen, based upon this aspect of reduced tree health.

         There may also be significant effects of fungal pathogens or
    insect pests associated with increasing nitrogen deposition.  The
    foliar concentrations of nitrogen increased markedly in tree needles
    or leaves in experiments with nitrogen additions, and also in forest
    sites with high atmospheric nitrogen loading (Roelofs et al., 1985;
    Van Dijk & Roelofs, 1988; Balsberg-Pählsson, 1992).  Animal grazing
    generally increases with increasing palatability of the leaves or
    shoots.  Nitrogen is of major importance for the palatability of plant
    material, and this certainly holds for insect grazing (Crawley, 1983). 
    Secondary plant chemicals, e.g., phenolics, are important for
    increased resistance of plants. The total amount of phenolics in
     Fagus leaves in a 120-year stand decreased by more than 30% after
    fertilizer application of about 45 kg nitrogen per ha per year,
    compared with the control treatment (Balsberg-Pählsson, 1992).  An
    ecologically important relation between nitrogen enrichment and insect
    pests has been quantified for lowland heathland (Brunsting & Heil,
    1985; Berdowski, 1993, see section 4.1) but not, so far, for forest
    ecosystems.

         From 1982 to 1985 an epidemic outbreak of the pathogenic fungus
     Sphaeropsis sapinea was observed in coniferous forest (mainly
     Pinus nigra) in the Netherlands.  This greatly affected whole
    stands, and was especially severe in the south-east part of the
    Netherlands, where there was high airborne nitrogen deposition
    (Roelofs et al., 1985).  Van Dijk et al. (1992) showed that there was
    a significantly higher foliar nitrogen concentration in the infected
    stands, together with higher soil ammonium levels, than in the
    uninfected stands.  Most of the additional nitrogen in the needles of
    the affected stands was stored as nitrogen-rich free amino acids,

    especially arginine.  Proline concentrations were also higher in the
    infected trees, indicting a relation with water stress (Van Dijk
    et al., 1992).

         The effects of  Sphaeropsis have also been studied by De Kam et
    al. (1991).  Two-year-old plants of  Pinus nigra were grown for
    3 years in pots and given five treatments of ammonium sulfate (very
    low to about 300 kg nitrogen per ha per year), in combination with two
    levels of potassium sulfate.  The 5-year-old plants were then
    inoculated with  Sphaeropsis.  The bark necroses were much more
    frequent in the plants treated with ammonium sulfate than in the
    controls.  Effects of ammonium sulfate upon fungal damage were even
    observed at an addition of 75 kg nitrogen per ha per year, but were
    very significant in the plants treated with 150 kg nitrogen per ha per
    year.  After potassium addition the number of necroses caused by the
    fungus was greatly reduced (De Kam et al., 1991).

         In beech forests in NW Switzerland, a significant positive
    correlation has been found between the nitrogen/potassium ratios in
    the leaves and necroses caused by the beech cancer  Nectria ditissima
    (Flückiger & Braun, 1994).  These authors also experimentally
    inoculated  Fagus sylvatica trees at different applications of
    nitrogen with this beech cancer and observed increased dieback of new
    leaves and shoots.  Furthermore, the infestation of  Fagus sylvatica
    with beech aphids  (Phyllaphis fagi) was also affected by the
    nitrogen availabilities.  The degree of infestation with the aphid
    increased significantly with enhanced leaf nitrogen/potassium ratios
    (Flückiger & Braun, 1994).  Although evidence for nitrogen-mediated
    changes in susceptibility to fungal pests and insect attacks has until
    now been based upon observations of only few species, it is obvious
    that trees became more susceptible to these attacks with increasing
    nitrogen enrichment and this may play a crucial role in the dieback of
    some forest stands.

         A critical load for nitrogen had been established at 10-15 kg
    nitrogen (at no or low nitrification) to 20-30 kg nitrogen per ha per
    year in highly nitrifying soils, based upon nutritional imbalance of
    coniferous species (Boxman et al., 1988).  Recent evidence of  Fagus
     sylvatica tree health in acidic forests indicated a critical load
    of 15-20 kg nitrogen per ha per year, based upon both field and
    experimental observations.  Elevated nitrogen deposition can
    seriously affect tree healthy via a complex web of interactions (e.g.
    susceptibility to frost and drought).  Pathogens may play an important
    role in tree decline, but at this moment it is not possible to combine
    the observed processes and effects to an overall value for a critical
    load of nitrogen for tree health.

    4.2.5.2  Effects on tree epiphytes, ground vegetation and ground fauna
             of forests

    a)  Effects on ground-living and epiphytic lichens and algae

         The effects of SOy as an acidifier on epiphytic lichens have
    been extensively studied (Insarova et al., 1992; Van Dobben, 1993). 
    SOy was previously the dominant airborne pollutant, and it has been
    shown that most (epiphytic) lichens are more negatively affected by
    acidity than by nitrogen compounds (except NOy). Most lichens have
    green algae as photobionts and are affected by acidity but not by
    nitrogen.  Some of them even react positively to nitrogen (Insarova et
    al., 1992).  However, 10% of all lichen species in the world have
    cyanobacteria (blue-green algae) as the photobiont.  These
    cyanobacterial lichens are negatively affected by acidity, and also by
    nitrogen.  Most of the NW European lichens with cyanobacteria live on
    the soil surface or are tree epiphytes.  The most pollution-sensitive
    lichens are among them and they are threatened by extinction in NW
    Europe. This is probably the result of increased nitrogen deposition,
    which inhibits the functioning of the cyanobacteria.  In the
    Netherlands, for example, all cyanobacterial lichens that were present
    at the end of the 19th century are now absent. In Denmark, 96% of the
    lichens with cyanobacteria are extinct or threatened.  Furthermore,
    the cyanobacterial lichens appear frequently on the Red List of the
    European Union countries (Hallingbäck, 1991).

         Very few data exist to establish a critical load for nitrogen for
    these lichens with blue-green algae.  Nohrstedt et al. (1988)
    investigated the effects of nitrogen application (as ammonium nitrate
    or calcium nitrate) on ground-living lichens ( Peltigera aphtosa and
     Nephroma arcticum) with blue-green algae as photobionts.  The plots
    were treated once or three or four times with 120, 240 or 360 kg
    nitrogen per ha.  After a short period all  Peltigera and  Nephroma
    lichens were eliminated and even 19 years later no recolonization had
    occurred.  However, it is impossible to transform these very high
    doses to critical loads. The effects of air pollutants on lichens are
    usually related to concentrations in the air or in the precipitation. 
    It is probably more relevant to relate the effects of nitrogen on
    cyanobacterial lichens to deposition than to concentrations.  For tree
    epiphytes stemflow is most relevant, whereas for ground-living lichens
    throughfall will be more important.  Although much research is still
    needed, it has been suggested that a load of 5-15 kg nitrogen per ha
    per year is already critical for the growth of these cyanobacterial
    lichens (Hallingbäck, 1991).  These lichens may be the most sensitive
    components of some forest ecosystems and thus determine the critical
    load for these systems.

         Free-living green algae, especially of the genus  Pleurococcus
    ( Protococcus and  Demococcus are synonyms), are strongly stimulated
    by enhanced nitrogen deposition. They cover practically all outdoor

    surfaces which are not subject to frequent desiccation in regions with
    high nitrogen deposition, such as in the Netherlands and in Denmark. 
    The thickness and the colonization rate of spruce needles by green
    algae has been investigated in the Swedish Environmental Monitoring
    Programme (Brakenhielm, 1991).  The Swedish data show that these algae
    do not colonize spruce needles in regions with a total deposition
    (throughfall) lower than about 5 kg nitrogen per ha per year.  In
    areas with deposition above 20 kg nitrogen per ha per year, the green
    algal cover of the needles is so thick and the algae colonize so early
    that they may impede the photosynthesis of the spruce trees.

    b)  Effects on forest ground vegetation

         In the Netherlands the forest vegetation of a site in the central
    part of the country was investigated in 1958 (with about 20 kg
    nitrogen per ha per year) and in 1981 (with about 40 kg nitrogen per
    ha per year).  All lichens had disappeared during this period and a
    considerable increase in  Deschampsia flexuosa and  Corydalis
     claviculata was found.  A large representative sample test (n=2000),
    covering about 90% of the Dutch forests, revealed in the mid-1980s
    that among the 40 most common forest plants were:  Galeopsis
     tetrahit, Rubus species,  Deschampsia flexuosa, Dryoptesis
     cathusiana, Molinia caerulea, Poa trivialis, and  Urtica dioica
    (Dirkse & Van Dobben, 1989; Dirkse, 1993).  In Sweden,  Quercus robur
    stands in two geographical areas with different nitrogen deposition
    were compared with special emphasis on nitrogen indicator species
    (Tyler, 1987).  The stands were quite comparable except for the
    nitrogen inputs: 6-8 kg nitrogen per ha per year and 12-15 kg nitrogen
    per ha per year, respectively.  In the stand with the highest
    deposition, the soil solution was more acidic, probably due to acidic
    deposition as well (± 10 kg sulfur per ha per year), and it was
    estimated that acidification of the soil has accelerated during the
    last 30 to 50 years.  The following species were more common in the
    most polluted site:  Urtica dioca, Epilobium augustifolium, Rubus
     idaeus, Stellaria media, Galium aparine, Aegopodium podagraria and
     Sambucus spp. Thus, both in Sweden and the Netherlands, species
    indicative of nitrogen enrichment became common (Ellenberg, 1988b).

         Comparable observations were reported by Falkengren-Grerup (1986)
    and by Falkengren-Grerup & Eriksson (1990), who examined the changes
    in soil and vegetation in  Quercus and  Fagus stands in southern
    Sweden. They concluded that the exchangeable base cations were reduced
    and that aluminium had doubled over the past 35 years.  They also
    found a decrease in soil pH, with a disappearance of several species
    when pH dropped below a threshold.  In spite of soil acidification
    some species had increased in cover, and the most plausible
    explanation seemed to be increased nitrogen deposition, which was
    about 15-20 kg nitrogen per ha per year in southern Sweden and which
    had doubled since 1955.  A marked increase in cover was found for

     Lactuca muralis, Dryopteris filix-max, Epilobium augustifolium,
     Rubus idaeus, Melica uniflora, Aegopodium podagraria, Stellaria
     holostea and  S. nemorum, some of these species being nitrogen
    indicators.  Despite soil acidification, acid-tolerant species
    ( Deschampsia flexuosa, Maianthemum bifolium and  Luzula pilosa) did
    not increase.  A distinct decrease was observed for  Dentaria
     bulbifera, Pulmonaria officinalis and  Polygonatum multiflorum.
    Furthermore, Rosen et al. (1992) found a significant positive
    correlation between the increase of  Deschampsia flexuosa cover in
    the last 20 years in the Swedish forests and the pattern of nitrogen
    deposition.

         In a large semi-natural  Fagus-Quercus forest in NE France,
    about 50 permanent vegetation plots were investigated in 1972 and
    1991.  The changes in species composition on calcareous soils and in
    moderately acidic habitats were followed.  During the study period a
    significant increase in nitrophilous ground flora was observed in the
    high-pH (6.9) stands.  This indicated that at this location (with
    ambient deposition of 15-20 kg nitrogen per ha per year) there was a
    distinct effect of increasing nitrogen availability (Thimonier et al.,
    1994).

         From 1968 to 1985, three sites in a 30-year-old  Pinus
     sylvestris forest in Lisselbo (central Sweden) were annually
    fertilized with 0, 20, 40 and 60 kg nitrogen per ha per year (as
    NH4NO3 plus ambient deposition of 10 kg nitrogen per ha per year). 
    The original ground vegetation consisted of  Calluna vulgaris,
     Vaccinium vitis-idea, V. myrtillus, Cladonia spp.,  Cladina spp.,
    and the mosses  Dicranum spp.,  Pleurozium spp. and  Hylocomium
    spp.  The first changes were observed within 8 to 15 years and after
    about 20 years the experimental plots were compared and statistically
    analysed.  The original species disappeared at nitrogen applications
    above 20 kg (plus ambient deposition) nitrogen per ha per year and
    were replaced by  Epilobium augustifolium, Rubus idaeus, Deschampsia
     flexuosa, Dryopteris carthusiana and the moss  Brachythecium
     oedipodium (Dirkse et al., 1991; Van Dobben, 1993).  In another
    experiment at Lisselbo the combined effects of acidification (addition
    of H2SO4, pH=2.0) and nitrogen addition (0 and 40 kg nitrogen per ha
    per year) were investigated.  The increased nitrogen level seemed to
    be the more important factor. Acidification was the next most
    discriminating factor: all species disappeared, except for the moss
     Pohlia nutans at high additions of acidity (Dirkse & Van Dobben,
    1989; Dirkse et al., 1991).

         In southern Sweden, Tyler et al. (1992) studied the effects of
    the application of ammonium nitrate (60-180 kg nitrogen per ha per
    year) over a 5-year period on stands of  Fagus sylvatica.  They
    observed a large reduction in biomass of the ground vegetation with
    the application of nitrogen, and the frequency of most herb layer
    species declined significantly.  Soil measurements revealed that, in

    addition to eutrophication effects, the acidification of the soil
    solution was also important for the decline of the original ground
    vegetation.  In an experiment on the effects of nitrogen fertilizer
    application on bryophytes, it appeared that  Brachythecium
     oedipodium, B. reflexum and  B. starkei increased significantly at
    levels up to 60 kg nitrogen per ha per year.  At higher doses these
    species tended to decline, however.  Hylocomium splendens and
     Pleurozium schreberi declined considerably at doses of 30 to 60 kg
    nitrogen per ha per year (Dirkse & Martaki, 1992).

    c)  Effects on macrofungi and mycorrhizas

         During the last two decades many reports have described a
    decrease in species diversity and abundance of macrofungi.  These
    changes can probably be attributed to indirect effects of air
    pollution, in particular to increases in the amount of available
    nitrogen (possibly in combination with acidification), and/or to
    decreased health of trees with concomitant reduction of transport to
    the roots (Arnolds, 1991).

         When comparing sites over time, the number of fruiting bodies of
    macrofungi showed marked differences.  Most studies in western Europe,
    however, have revealed that the number of ectomycorrhizal fungi
    species has declined (Arnolds, 1991).  In the Netherlands the average
    number of ectomycorrhizal species per foray declined significantly
    from 71 in 1912-1954 to 38 in 1973-1982.  Similar changes have been
    observed in Germany: 94 ectomycorrhizal species found in 1950-1979 in
    the Völklinger area (Saarland) have not been recorded recently.  From
    the 236 species found in 1918-1942 in the Darmstadt area (Germany),
    only 137 were recorded in the early 1970s, a loss of 99 species,
    including many mycorrhizal fungi (Arnolds, 1991).  In contrast to the
    decline in mycorrhizal fungi, the number of saprotrophic species
    remained practically unchanged, while the number of lignocolous
    species increased.  This may be related to soil acidification with a
    increase in aluminium, since the proportion of forest areas in western
    Europe with a soil pH below 4.2 increased from less than 1% in 1960 to
    15% in 1988 (Schneider & Bresser, 1988).

         Arnolds (1988, 1991) concluded that acidification has very little
    effect on the diversity of ectomycorrhizal fungi, but rather triggers
    changes in species composition.  He regarded the increased nitrogen
    flux to the forest floor as the most important factor in the decline
    of mycorrhizal fungi.  Termorshuizen & Schaffers (1987) found a
    negative correlation between the total nitrogen input in mature
     Pinus sylvestris stands and the abundance of fruit bodies of
    ectomycorrhizal fungi.  Similar results were obtained by Schlechte
    (1986) who compared two sites with  Picea abies in the Göttingen area
    of Germany.  An obvious negative relation was found between nitrogen
    input (23 versus 42 kg nitrogen per ha per year) and ectomycorrhizal
    species: 85 basidiomycetes including 21 ectomycorrhizas (25%) at the

    less polluted site compared with 55 basidiomycetes including
    3 ectomycorrhizas (5%) at the most polluted site.  Environmental
    factors other than nitrogen did not differ significantly.  The
    negative impact of nitrogen seems only to hold true for mature forests
    (Termorshuizen & Schaffers, 1987).  Jansen & de Vries (1988) found a
    maximum in fruit-body production in > 20-year-old  Pseudotsuga
     menziesii stands at about 25 kg nitrogen per ha per year.  Meyer
    (1988) found a similar optimum when  Picea abies was planted in soil
    mixed with different amounts of sawdust having a high carbon/nitrogen
    ratio.

         Experiments with nitrogen fertilizer have produced similar
    results.  In a fertilizer trial with simulated nitrogen deposition in
    a  Fagus forest in southern Sweden (ambient deposition 15-20 kg
    nitrogen per ha per year), Ruhling & Tyler (1991) found, after
    applying NH4NO3 (60 and 180 kg nitrogen per ha per year), that
    within 3 to 4 years almost all mycorrhizal species ceased fruit-body
    production. In contrast, several decomposer species increased
    fruit-body production.  Wood decomposers showed no obvious reaction to
    the treatment.  No fruit-bodies were recovered when 300 kg nitrogen
    per ha was applied to  Pinus sylvestris stands as liquid manure
    (Ritter & Tölle, 1978).  The mycorrhizal frequency of the roots,
    however, was still 55% as compared to 87% in the controls. 
    Application of 112 kg nitrogen (as NH4NO3) per ha to 11-year-old
     Pinus taeda stands revealed an 88% reduction in the number of
    fruit-bodies and a 14% decrease in the number of mycorrhizas per unit
    of soil volume (Menge & Grand, 1978).  In the Lisselbo study the
    number of fruit-bodies decreased considerably at each nitrogen
    fertilizer dose (Wasterlund, 1982).  Termorshuizen (1990) applied
    0, 30 and 60 kg nitrogen (as ammonium sulfate or nitrate) per ha per
    year to young  Pinus sylvestris stands.  In general fruit-body
    production was more negatively influenced by the higher ammonium
    levels than nitrate levels.  The mycorrhizal frequency and the number
    of mycorrhizas per unit of soil volume were not influenced.  It was
    concluded by Termorshuizen (1990) that fruit-body production is much
    more sensitive to nitrogen enrichment that mycorrhizal formation. 
    Branderud (1995) found after only 1.5 year a decrease in fruit-body
    production of mycorrhizal species at a nitrogen application of 35 kg
    nitrogen (as NH4NO3) per ha in a  Picea abies stand at the Swedish
    Nitrex stand.

         In contrast, some studies have shown an increase in the number of
    fruit-bodies of insensitive mycorrhizal fungi after nitrogen
    fertilizer application, e.g.,  Paxillus involutes (Hora, 1959),
     Laccaria bicolor (Ohenoja, 1988) and  Lactarius rufus (Hora, 1959).

    d)  Effects on soil fauna of forests

         Almost all studies of changes in faunal species composition due
    to nitrogen enrichment have been conducted in arable fields or

    agricultural grasslands using complete fertilization and thus cannot
    be used to substantiate critical loads for semi-natural forest
    ecosystems (Marshall, 1977).  The relationship between acidity and
    soil fauna has also been studied in northern coniferous forests, but
    only very few studies have incorporated the effects of nitrogenous
    compounds (Gärdenfors, 1987).  The abundance of  Nematoda,
     Oligochaeta and microarthropods (especially  Collembola) had
    increased in some studies, but decreased in others, after application
    of high doses of nitrogen fertilizers (> 150 kg nitrogen per ha per
    year) (Abrahamsen & Thompson, 1979; Huhta et al., 1983; Vilkamaa &
    Huhta, 1986).  A reduction in the nitrogen deposition in a  Pinus
     sylvestris stand (Nitrex site Ysselstein) to pre-industrial levels
    increased the species diversity of microarthropods due to a decreased
    dominance of some species (Boxman et al., 1995).  However, it is not
    possible to use these few data to formulate a critical load for
    changes in forest soil fauna due to increased nitrogen deposition.

         On the basis of the results presented in this overview, the
    critical load for changes in the ground vegetation of both coniferous
    and deciduous acidic forest may be 15 to 20 kg nitrogen per ha per
    year.  The critical load for changes in the fruit-body production of
    ectomycorrhizal fungi is probably about 30 kg nitrogen per ha per
    year, while the critical load for changes in mycorrhizal frequency of
    tree roots is hard to estimate, but certainly considerably higher. 
    There is insufficient data on the effects of enhanced nitrogen
    deposition on faunal components of forest ecosystems to allow critical
    loads to be set.  Epiphytic or ground-living lichens with
    cyanobacteria as the photobiont probably form a sensitive part of
    forest ecosystems and have an estimated critical load of 10-15 kg
    nitrogen per ha per year.  A summary of the critical loads for forests
    is given in chapter 8.

    4.2.6  Effects on estuarine and marine ecosystems

         Few topics in aquatic biology have received as much attention
    over the past decade as the debate over whether estuarine and coastal
    ecosystems are limited by nitrogen, phosphorus or some other factor
    (Hecky & Kilham, 1988).  Numerous geochemical and experimental studies
    have suggested that nitrogen limitation is much more common in
    estuarine and coastal waters than in freshwater systems.  Taken as a
    whole, the productivity of estuarine waters in the USA correlates more
    closely with supply rates of nitrogen than with those of other
    nutrients (Nixon & Pilson, 1983).

         Estimation of the contribution of nitrogen deposition to the
    eutrophication of estuarine and coastal waters is made difficult by
    the multiple direct anthropogenic sources (e.g., from agriculture and
    sewage) of nitrogen against which the importance of atmospheric
    sources must be weighed.  Estuaries and coastal areas are common
    locations for cities and ports. The crux of any assessment of the

    importance of nitrogen deposition to estuarine eutrophication lies in
    establishing the relative importance of direct anthropogenic exposure
    (e.g., sewage and agricultural run-off) and indirect effects
    (e.g., atmospheric deposition). 

         The effects of nitrogen deposition in certain estuarine systems
    have been investigated.  Complete nitrogen budgets, as well as
    information on nutrient limitation and seasonal nutrient dynamics,
    have been compiled for two large "estuaries", the Baltic Sea
    (Scandinavia) and the Chesapeake Bay (USA), and for the Mediterranean
    Sea.  In the case of the Mediterranean, Loye-Pilot et al. (1990)
    suggest that 50% of the nitrogen load originates as deposition falling
    directly on the water surface.  In the case of the Baltic and
    Chesapeake, deposition of atmospheric nitrogen has been suggested as a
    major contributor to eutrophication.  Data for other coastal and
    estuarine systems are less complete, but similarities between these
    two systems and other estuarine systems suggest that their results may
    be more widely applicable.  Discussion in this monograph is limited to
    these two case studies, with some speculation about how other
    estuaries may be related.

         The Baltic Sea is perhaps the best-documented case study of the
    effects of nitrogen additions in causing estuarine eutrophication. 
    Like many other coastal waters, the Baltic Sea has experienced a
    rapidly increasing anthropogenic nutrient load. It has been estimated
    that the supply of nitrogen has increased by a factor of 4, and
    phosphorus by a factor of 8, since the beginning of the 20th century
    (Larsson et al., 1985).  The first observable changes attributable to
    eutrophication of the Baltic were declines in the concentration of
    dissolved oxygen in the 1960s (Rosenberg et al., 1990).  Decreased
    dissolved oxygen concentrations result when decomposition in deeper
    waters is enhanced by the increased supply of sedimenting algal cells
    from the surface water layers to the sediment.  In the case of the
    Baltic, the spring algal blooms that now result from nutrient
    enrichment consist of large, rapidly sedimenting algal cells, which
    supply large amounts of organic matter to the sediment for
    decomposition (Enoksson et al., 1990).  Since the 1960s, researchers
    in the Baltic have documented increases in algal productivity,
    increased incidence of nuisance algal blooms, and periodic failures
    and unpredictability in fish and Norway Lobster catches (Fleischer &
    Stibe, 1989; Rosenberg et al., 1990).  It has now been shown by a
    number of methods that algal productivity in nearly all areas of the
    Baltic Sea is limited by nitrogen.  Nitrogen-to-phosphorus ratios
    range from 6:1 to 60:1 (Rosenberg et al., 1990), but the higher ratios
    are only found in the remote and relatively unaffected area of the
    Bothnian Bay (between Sweden and Finland).  Productivity in the spring
    (the season of highest algal biomass) is fuelled by nutrients supplied
    from deeper waters during spring overturn (Graneli et al., 1990); deep
    waters are low in nitrogen and high in phosphorus, resulting in
    nitrogen-to-phosphorus ratios near 5 (Rosenberg et al., 1990),

    suggesting potential nitrogen limitation when deep waters are mixed
    with surface waters.  Low nitrogen-to-phosphorus ratios in deep water
    result from denitrification in the deep sediments (Shaffer & Rönner,
    1984).  Primary productivity measurements in the Kattegat (the portion
    of the Baltic between Denmark and Sweden) correlate closely with
    uptake of NO3-, but not of PO43- (Rydberg et al., 1990).  Level II
    and III nutrient enrichment experiments conducted in coastal areas of
    the Baltic, as well as in the Kattegat, indicate nitrogen limitation
    at most seasons of the year (Graneli et al., 1990).  Growth
    stimulation of algae has also been produced by addition of rain water
    to experimental enclosures, in amounts as small as 10% of the total
    volume (Graneli et al., 1990); rain water in the Baltic is rich in
    nitrogen but poor in phosphorus.  In portions of the Baltic where
    freshwater inputs keep the salinity low, blooms of the nitrogen-fixing
    cyanobacterium  Aphanizomenon flos-aquae are common (Graneli et al.,
    1990); cyanobacterial blooms are common features of nitrogen-limited
    freshwater lakes but are usually absent from marine waters.

         Nitrogen budget estimates indicate that the Baltic Sea as a whole
    receives 7.6 × 1010 eq of nitrogen per year, of which 2.8 × 1010 eq
    per year (37%) comes directly from atmospheric deposition (Rosenberg
    et al., 1990).  Fleischer & Stibe (1989) reported that the nitrogen
    flux from agricultural watersheds feeding the Baltic has been
    decreasing since about 1980 but that the nitrogen contribution from
    forested watersheds is increasing. They cite both increases in
    nitrogen deposition and the spread of modern forestry practices as
    causes for the increase.  It should be noted, however, that the Baltic
    also experiences a substantial phosphorus load from agricultural and
    urban lands, and that phosphorus inputs may help to maintain
    nitrogen-limited conditions (Graneli et al., 1990). If the Baltic had
    received consistent nitrogen additions (e.g., from the atmosphere or
    from agricultural run-off) in the absence of phosphorus additions, it
    might well have evolved into a phosphorus-limited system some time
    ago.

         The physical structure of the Baltic Sea, with a shallow sill
    limiting exchange of water with the North Sea contributes to the
    eutrophication of the basin, by trapping nutrients in the basin once
    they reach the deeper waters.  Because the larger algal cells that
    result from nutrient enrichment in the basin provide more nutrients to
    the deep water through sedimentation, and because only shallow waters
    have the ability to exchange with the North Sea, it is estimated that
    less than 10% of nutrients added to the Baltic are exported over the
    sill to the North Sea (Wulff et al., 1990).  Throughout much of the
    year (i.e., especially during the dry months) productivity in the
    Baltic is maintained by nutrients recycled within the water column
    (Enoksson et al., 1990).  The trapping of nutrients within the basin
    and recycling of nutrients from deeper water by circulation patterns

    suggest that eutrophication of the Baltic is a self-accelerating
    process (Enoksson et al., 1990) and has a long time-lag between
    reductions of inputs and improvements in water quality.

         In the USA, a large effort has been made to establish the
    relative importance of sources of nitrogen to Chesapeake Bay (D'Elia
    et al., 1982; Smullen et al., 1982; Fisher et al., 1988; Tyler, 1988). 
    Estimates of the contribution of nitrogen to Chesapeake Bay from each
    individual source are very uncertain; estimating the proportion of
    nitrogen deposition exported from forested watersheds is especially
    problematic but critical to the analysis, because about 80% of the
    Chesapeake Bay basin is forested.  Nonetheless, three attempts at
    determining the proportion of the total nitrate load to the Bay
    attributable to nitrogen deposition all produce estimates in the range
    of 18 to 31%.  Supplies of nitrogen from deposition exceed supplies
    from all other non-point sources to the Bay (e.g., agricultural
    run-off, pastureland run-off, urban run-off), and only point source
    inputs represent a greater input than deposition.

         It is considered that the data from these studies are indicators
    of the impact of anthropogenic nitrogen.  Nevertheless, they are
    insufficient to estimate critical loads for estuarine/marine systems. 
    It may well by that critical loads for these systems differ for
    different climatic regions.

    4.2.7  Appraisal and conclusions

         Atmospheric deposition of nitrogen-containing and acidifying
    compounds have an impact on soil and groundwater quality and on the
    health and species composition of vegetation.  Critical loads for
    these effects are given in Table 26.  Critical loads have been derived
    using empirical data that relate loads directly to effects and
    steady-state soil models that calculate critical loads from critical
    chemical values for ion concentrations or ratios in foliage, soil
    solution and groundwater (De Vries, 1993). Information on the effects
    which occur when critical loads are exceeded is given in Table 27. 
    The values given in Tables 26 and 27 apply to forest vegetation in a
    temperate climate.  Whether they are representative of other climates
    is uncertain.  An overview of the critical loads for atmospheric
    nitrogen deposition in a range of natural and semi-natural ecosystems
    is given in chapter 8.

         Effects of nitrogen and acidifying deposition on soil and
    groundwater chemistry are most evident.  Field studies showed that
    deposited nitrogen is partly retained in the forest soil.  Even at
    high nitrogen deposition rates, as in the Netherlands, soil
    acidification (which is mainly manifested by leaching of aluminium and
    nitrate) is mainly caused by sulfur deposition.  A relatively small
    contribution of nitrogen to acidification does not imply that sulfur
    has a larger impact on the health of forests, since the relationship

    between soil acidification and forest health is not very clear.  The
    eutrophying impact of nitrogen is probably more important than the
    acidifying impact at present.

         There is substantial evidence from field surveys in several
    countries of Europe that exceeding critical loads does not imply
    dieback of the forest trees in the short term (one or two decades). 
    However, it does increase the risk of damage due to secondary stress
    factors and it affects the long-term sustainability of forests.  These
    risks increase with the extent to which present loads exceed critical
    loads and with the duration.

        Table 26.  Critical loads for acidity and nitrogen for forest ecosystems in temperate climates
               (From: De Vries, 1993)
                                                                                                              

                      Effects                       Criteriaa                           Critical loads
                                                                                    (kg per ha per year)
                                                                                        (H for acidity;
                                                                                    N for eutrophication)
                                                                                                              
                                                                                    Coniferous   Deciduous
                                                                                    forests      forests
                                                                                                              

    Acidity           root damage;                  Al < 0.2 mol/m3                 1.1b         1.4b
                      inhibition of uptake;         Al/Ca < 1.0 mol/mol             1.4b         1.1b
                      Al depletion;                 delta Al(OH)3 = 0 mmol/m3       1.2b         1.3b
                      Al pollution                  Al < 0.02 mol/m3                0.5b         0.3b

    Eutrophication    inhibition of uptake of K;    NH4/K < 5 mol/mol               17-70c
                      increased susceptibility;     N < 1.8%                        21-42d
                      vegetation changes;           NO3 < 0.1 mol/m3                7-20e        11-20e
                      nitrate pollution             NO3 < 0.4-0.8 mol/m3            13-21f       24-41f
                                                                                                              

    a    Background information on the various criteria is given in De Vries (1993).  Critical Al and NO3-
         concentrations and critical Al/Ca and NH4/K ratios related to root damage, inhibition of nutrient
         uptake and vegetation changes  refer to the soil solution. Critical Al and NO3- concentrations
         related to pollution refer to phreatic groundwater. Critical nitrogen contents related to an
         increased risk for frost damage and diseases refer to the foliage.
    b    Derived by a steady-state model. Al pollution refers to phreatic groundwater.  For groundwater
         used for the preparation of drinking-water, a critical acid load of 1600 mol/ha per year was derived
         (De Vries, 1993).
    c    Derived by a steady-state model assuming 50% nitrification in the mineral topsoil (second value).
    d    Empirical data on the relation between nitrogen deposition and foliar nitrogen contents.

    Table 26  (Con't)

    e    The first value is derived by a steady-state model (worst case) and the second value is based
         on empirical data.
    f    Derived by a steady-state model using critical NO3- concentrations of 0.4 and 0.8 mol/m3,
         respectively. NO3- pollution refers to phreatric groundwater.  For deep groundwater, the
         critical load will be higher because of denitrification.

    Table 27.  Possible and observed effects when critical loads are exceeded
                                                                                                              

    Possible effects         Average critical load         Observed effects in the field
                             (kg per ha per year)a
                                                                                                              

    Root damage              1.1-1.4 H                     critical Al concentrations
                                                           exceeded greatly

    Inhibition of            1.1-1.4 H                     critical Al/Ca ratios
    uptake                                                 exceeded greatly

                             17-70 N                       critical NH4/K ratios
                                                           exceeded slightly

    Aluminium depletion      1.2-1.3 H                     depletion of secondary Al compounds

    Groundwater              0.3-0.5 H                     critical Al concentrations
    pollution                                              exceeded greatly

                             13-21 N                       critical NO3 concentrations
                                                           exceeded substantially

    Increased                21-42 N                       critical N contents exceeded
    susceptibility                                         substantially; nutrient imbalances;
                                                           increased shoot/root ratios

    Vegetation changes       7-20 N                        strong increase in nitrophilous species
                                                                                                              

    a  H = acidity; N = total nitrogen
    
    5.  STUDIES OF THE EFFECTS OF NITROGEN OXIDES ON EXPERIMENTAL ANIMALS

    5.1  Introduction

         Most of the data reviewed in this chapter concerns the effects of
    NO2, since the bulk of the NOx literature is on NO2.  The results
    of the few comparative NOx studies suggest that NO2 is the most
    toxic species studied so far.  Most of the reports describe the
    effects of NO2 on the respiratory tract, but extrapulmonary effects
    are also briefly discussed.  A broad range of NO2 concentrations has
    been evaluated, but emphasis has been placed primarily on those
    studies with exposure concentrations of 9400 µg/m3 (5.0 ppm) or less,
    with the exception of studies on dosimetry and emphysema.  Discussions
    of available literature on the effects of other nitrogen compounds,
    e.g., NO, HNO3, and mixtures containing NO2, also are included.  WHO
    (1987), Berglund et al. (1993) and US EPA (1993) comprise other
    reviews of the animal toxicological literature concerning NOx
    effects.

    5.2  Nitrogen dioxide

    5.2.1  Dosimetry

         It is generally agreed that effects of NO2 observed in several
    laboratory animal species can be qualitatively extrapolated to humans. 
    However, to extrapolate animal data quantitatively to humans,
    knowledge of both dosimetry and species sensitivity must be
    considered.  Dosimetry refers to estimating the quantity of NO2
    absorbed by target sites within the respiratory tract.  Even when two
    species receive an identical local tissue/cellular dose, cellular
    sensitivity to that dose is likely to show interspecies variability
    due to differences in defence and repair mechanisms and other
    physiological/metabolic parameters.  Current knowledge of dosimetry is
    more advanced than that of species sensitivity, impeding quantitative
    animal-to-human extrapolation of effective NO2 concentrations. 
    Nevertheless, information on dosimetry alone can be crucial to
    interpretation of the data base.  Both theoretical (modelling) and
    experimental dosimetry studies are discussed below.

    5.2.1.1  Respiratory tract dosimetry

         The uptake of NO2 in the upper respiratory tract (above the
    larynx) has been experimentally studied in dogs, rats and rabbits. 
    The upper airways of dogs and rabbits exposed to 7520 to 77 080 µg/m3
    (4.0 to 41.0 ppm) NO2 removed 42.1% of the NO2 drawn through
    the nose (Yokoyama, 1968).  The uptake of NO2 by isolated
    upper respiratory tracts of naive and previously exposed rats
    (76 000 µg/m3, 40.4 ppm NO2) was 28% and 25%, respectively (Cavanagh
    & Morris, 1987).  Kleinman & Mautz (1987) exposed dogs to 1880 or

    9400 µg/m3 (1.0 or 5.0 ppm) NO2 and found that more NO2 was
    absorbed in the upper respiratory tract with nasal breathing than with
    oral breathing.  In addition, the percentage uptake of NO2 by the
    upper respiratory tract decreased with increasing ventilation rates. 
    As ventilation increased up to four times resting values, NO2 uptake
    during nasal breathing decreased from approximately 85% to less than
    80% and during oral breathing decreased from about 60% to approximately
    45%.  At rest, about 85% of the inhaled NO2 entering the lungs was
    absorbed by the lower respiratory tract; this increased to 100% with
    high ventilation rates.

         Miller et al. (1982) and Overton (1984) modelled NO2 uptake in
    the lower respiratory tract using the same dosimetry model described
    by Miller et al. (1978) for ozone (O3), but with the diffusion
    coefficient and Henry's law constant appropriate to NO2; however,
    values of the latter constant and reaction chemistry were considered
    uncertain.  For all species modelled (i.e., rat, guinea-pig, rabbit
    and humans), the results indicate that NO2 is absorbed throughout the
    lower respiratory tract, but the major dose to tissue is delivered in
    the centriacinar region (i.e., junction between the conducting and
    respiratory airways), findings consistent with the site of
    morphological effects (see section 5.2.2.4).

         Total respiratory tract uptake has been measured in healthy and
    diseased humans.  In healthy humans exposed to an NO/NO2 mixture
    containing 545 to 13 500 µg/m3 (0.29 to 7.2 ppm) NO2 for brief (but
    unspecified) periods, 81 to 90% of the NO2 was absorbed during normal
    respiration; this increased to 91 to 92% with maximal ventilation
    (Wagner, 1970).  Bauer et al. (1986) exposed adult asthmatics to
    564 µg/m3 (0.3 ppm) NO2 via a mouthpiece for 30 min, including
    10 min of exercise (30 litres/min) and measured inspired and expired
    NO2 concentrations.  At rest, the average uptake was 72%; during
    exercise, the average uptake was 87%, a statistically significant
    increase.  Because of the large increase in minute ventilation, the
    deposition was 3.1 µg/min at rest and 14.8 µg/min during exercise.

         As discussed above, increased ventilation increases the quantity
    of NO2 delivered to the respiratory tract and shifts the site of
    deposition.  Typically, the percentage uptake of NO2 in the upper
    respiratory tract decreases, with a consequent increase in uptake by
    the lower respiratory tract owing to the deeper penetration of the
    inspired gas with increased tidal volume. These experimental results
    are qualitatively similar to conclusions for the modelled effects of
    ventilation on O3 dosimetry (Miller et al., 1985; Overton et al.,
    1987a,b).

    5.2.1.2  Systemic dosimetry

         Once deposited, NO2 dissolves in lung fluids and various
    chemical reactions occur, giving rise to products that are found in
    the blood and other body fluids.  Labelled 13NO2 (564 to 1710 µg/m3
    (0.3 to 0.91 ppm)) inhaled for 7 to 9 min by rhesus monkeys was
    distributed throughout the lungs (Goldstein et al., 1977b).  These
    investigators also concluded that NO2 probably reacts with water in
    the fluids of the respiratory tract to form nitrous and nitric acids. 
    Saul & Archer (1983) provided support for this pathway using rats
    inhaling NO2.  This study subsequently led to the discovery of
    endogenous NO (Moncada et al., 1988, 1991).

         The current database indicates that once NO2 is absorbed in lung
    fluids, the subsequent reaction products are rapidly taken up and then
    translocated via the bloodstream.  For example, Oda et al. (1981)
    reported a concentration-dependent increase in both NO2- and NO3-
    levels in the blood of mice during 1-h exposures to 9400 to
    75 200 µg/m3 (5.0 to 40.0 ppm) NO2.  The blood levels of NO2- and
    NO3- declined rapidly after exposures ended, with decay half-times
    of a few minutes for NO2- and about 1 h for NO3-.

    5.2.2  Respiratory tract effects

    5.2.2.1  Host defence mechanisms

         Respiratory tract defences encompass many interrelated responses;
    however, for simplicity, they can be divided into physical and
    cellular defence mechanisms.  Physical defence mechanisms include the
    mucociliary system of the conducting airways.  Ciliary action moves
    particles and dissolved gases within the mucous layer towards the
    pharynx, where the mucus is swallowed or expectorated.  Both nasal
    and tracheobronchial regions are immunologically active (e.g.,
    nasal-associated lymphoid tissue and bronchial-associated lymphoid
    tissue), but this function has not been studied following NO2
    exposure.  Cellular defence mechanisms (phagocytic and immunological
    reactions) operate in the pulmonary region of the lung.  Alveolar
    macrophages (AMs) are the first line of cellular defence.  The AMs
    perform such activities as detoxifying and/or removing inhaled
    particles, maintaining sterility against inhaled microorganisms,
    interacting with lymphoid cells in a variety of immunological
    reactions, and removing damaged or dying cells from the alveoli
    through phagocytosis.  Polymorphonuclear leukocytes (PMNs), another
    group of phagocytic cells, are present in relatively small numbers
    (i.e., a small percentage of cells obtained from bronchoalveolar
    lavage (BAL) fluid) from normal lungs, but in response to a variety of
    insults, there can be an influx of PMNs from blood into the lung
    tissues and onto the surface of the airways.  Once recruited to the
    lung, PMNs then ingest and kill opsonized microbes and other foreign
    substances by mechanisms similar to those for AMs.

         The responses of PMNs and AMs are frequently studied using BAL,
    the washing of the airways and alveolar spaces with saline.  Cells and
    fluid obtained from this procedure can be used in a variety of ways to
    assess immune responses.

         Humoral and cell-mediated immunity are also active in the
    respiratory tract.  The humoral part of this system primarily involves
    the B cells that function in the synthesis and secretion of antibodies
    into the blood and body fluids.  The cell-mediated component primarily
    involves T lymphocytes, which are involved in delayed hypersensitivity
    and defences against viral, fungal, bacterial and neoplastic disease.

    a)  Mucociliary clearance

         Exposure to NO2 can cause loss of cilia and ciliated epithelial
    cells, as discussed in section 5.2.2.4 on morphological changes.  Such
    changes are reflected in the functional impairment of mucociliary
    clearance at high levels of NO2 (> 9400 µg/m3, 5.0 ppm) (Giordano
    & Morrow, 1972; Kita & Omichi, 1974).  At lower exposures (2 h/day for
    2, 7 and 14 days to 564 and 1880 µg/m3, 0.3 and 1.0 ppm NO2), the
    mucociliary clearance of inhaled tracer particles deposited in the
    tracheobronchial tree of rabbits was not altered (Schlesinger et al.,
    1987).

    b)  Alveolar macrophages

         Structural, biochemical, and functional changes in AMs observed
    in experimental animal studies to be caused by NO2 exposure are
    summarized in Table 28.  The adversity of these effects is not clearly
    understood at present, but they are taken as hallmarks of adverse
    reactions.  Studies of AMs in humans are discussed in chapter 6.

         Alveolar macrophages isolated from mice continuously exposed to
    3760 µg/m3 (2.0 ppm) NO2 or to 940 µg/m3 (0.5 ppm) NO2
    continuously with a 1-h peak to 3760 µg/m3 (2.0 ppm) for 5 days/week
    showed distinctive morphological changes after 21 weeks of exposure,
    compared to controls (Aranyi et al., 1976).  Structural changes
    included the loss of surface processes, appearance of fenestrae, bleb
    formation and denuded surface areas.  Continuous exposure to a lower
    NO2 level did not result in any significant morphological changes. 
    Numerous morphological studies have shown that NO2 exposure increases
    the number of AMs (see section 5.2.2.4).

         BAL methods have also been used to study AMs.  Mochitate et al.
    (1986) reported a significant increase in the total number of AMs
    isolated from rats during 10 days of exposure to 7520 µg/m3 (4.0 ppm)
    NO2, but the number of PMNs did not increase.  The AMs from exposed
    animals also exhibited increased metabolic activity, as measured by
    the activities of glucose-6-phosphate dehydrogenase, glutathione
    peroxidase and pyruvate kinase.  The AMs also showed an increase in

        Table 28.  Effects of nitrogen dioxide (NO2) on alveolar macrophagesa
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    564          0.3          2 h/day,              Rabbits     Increase in alveolar clearance.                             Schlesinger &
    1880         1.0          14 days                                                                                       Gearhart (1987)

    564          0.3          2 h/day, 13 days      Rabbit      Decreased AM phagocytic capacity at 564 µg/m3; increase     Schlesinger
    1880         1.0                                            at 1880 µg/m3 after 2 days of exposure.  No effect on cell  (1987a,b)
                                                                number or viability; random mobility reduced at 564 µg/m3
                                                                only. No effects from 6 days of exposure on.

    564          0.3          2 h/day, 1 or         Rabbit      Acceleration in alveolar clearance at < 1880 µg/m3.         Vollmuth et
    1880         1.0          14 days                                                                                       al. (1986)
    5640         3.0

    940 or       0.5 or       Continuous base       Mouse       No observable effects on AM morphology.                     Aranyi et al.
    188 base;    0.1 base;    with 2-h/day peak                                                                             (1976)
    1880 peak    1.0 peak     (5 days/week),
                              24 weeks

    3760 or      2.0 or 0.5   Continuous base       Mouse       Morphological changes, such as loss of surface processes,   Aranyi et al.
    940 base;    base;        with 7 h/day peak                 appearance of fenestrae, bleb formation, and denuded        (1976)
    3760 peak    2.0 peak     (5 days/week),                    surface areas.
                              21 weeks

    1880         1.0          17 h                  Mouse       Bactericidal activity significantly decreased by 6 and      Goldstein et al.
    4320         2.3                                            35% at 4320 and 12 400 µg/m3, respectively; no effect at    (1974)
    12 400       6.6                                            1880 µg/m3.
                                                                                                                                              

    Table 28.  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    1880 base;   1.0 base;    7 h/day, 5 days       Rat         Accumulation of AMs.  Superimposed spikes produced          Gregory et al.
    9400 peak    5.0 peak     per week base with                changes that may persist with continued exposures.          (1983)
                              one 1.5-h peak/day,
                              15 weeks

    2444-31 960  1.3-17.0          -                Rat         Decreased production of superoxide anion radical.           Amoruso et al.
                                                                                                                            (1981)

    3760         2.0          8 h/day,              Baboon      Impaired AM responsiveness to migration inhibitory          Greene &
                              5 days/week,                      factor.                                                     Schneider (1978)
                              6 months

    5640         3.0          3 h                   Rabbit      Increased swelling of AMs.                                  Dowell et al.
                                                                                                                            (1971)

    6768         3.6          2 h                   Rat         Enhanced AM agglutination with concanavalin A.              Goldstein et al.
                                                                                                                            (1977a)

    7520         4.0          6 h/day, 7, 14,       Rat         Changes in AM morphology; no change in numbers of AMs       Hooftman et al.
                              or 21 days                        or phagocytic capacity.                                     (1988)

    7520         4.0          10 days               Rat         Increase in number of AMs; no increase in PMNs; increased   Mochitate et al.
                                                                metabolic activity, protein and DNA synthesis; all          (1986)
                                                                responses peaked on day 4 and returned to normal on
                                                                day 10.
                                                                                                                                              

    Table 28.  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    7520         4.0          Up to 10 days         Rat         Increase in number of AMs, reaching a peak on days 3 and    Suzuki et al.
                                                                5; no increase in number of PMNs; decrease in AM viability  (1986)
                                                                throughout exposure period. Suppression of phagocytic
                                                                activity on day 7 that returned to normal value at
                                                                day 10.  Decrease in superoxide radical production on
                                                                days 3, 5 and 10.

    9400         5.0          7 days                Mouse       No effect on phagocytic activity.                           Lefkowitz et al.
                                                                                                                            (1986)

    9400         5.0          3 h                   Rabbit      No change in AM resistance to pox virus.                    Acton & Myrvik
                                                                                                                            (1972)
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  AM = alveolar macrophage; PMN = polymorphonuclear leukocyte
        the rate of synthesis of protein and DNA.  All responses peaked on day
    4 and returned to control levels by the tenth day.  Suzuki et al.
    (1986) made similar observations and, in addition, found that the
    viability of AMs was decreased on day 1 and remained depressed for the
    remainder of the exposure period.  Increased numbers and metabolic
    activity of AMs would be expected to have a positive influence on host
    defences.  However, AMs are rich in proteolytic enzymes, and increased
    numbers could result in some tissue destruction when the enzymes are
    released.  Furthermore, as discussed below, although more AMs may be
    present, they often have a decreased phagocytic ability.

         Schlesinger (1987a,b) found no significant changes in the number
    or the viability of AMs in BAL from rabbits exposed to 564 or
    1880 µg/m3 (0.3 or 1.0 ppm) NO2, 2 h/day, for 13 days.  Although
    there were no effects on the numbers of AMs that phagocytosed latex
    spheres, 2 days of exposure to 564 µg/m3 (0.3 ppm) decreased the
    phagocytic capacity (i.e., number of spheres per cell); the higher
    level of NO2 increased phagocytosis.  Longer exposures had no effect. 
    The phagocytic activity of rat AMs was significantly depressed after
    7 days of exposure to 7520 µg/m3 (4.0 ppm) but returned to the
    control value at 10 days of exposure (Suzuki et al., 1986).  There may
    be a species difference in responsiveness because Lefkowitz et al.
    (1986) did not observe a depression in phagocytosis in mice exposed
    for 7 days to 9400 µg/m3 (5.0 ppm) NO2.  Suzuki et al. (1986)
    proposed that the inhibition of phagocytosis might be due to NO2
    effects on membrane lipid peroxidation.  Studies by Dowell et al.
    (1971) and Goldstein et al. (1977a) add support to this hypothesis. 
    Acute exposure to 5640-7520 µg/m3 (3.0-4.0 ppm) caused swelling of
    AMs (Dowell et al., 1971) and increased AM agglutination with
    concanavalin A (Goldstein et al., 1977a), suggesting damage to the
    membrane function.

         Two independent studies have shown that NO2 exposure decreases
    the ability of rat AMs to produce superoxide anion involved in
    antibacterial activity.  Amoruso et al. (1981) presented evidence
    of such an effect at NO2 concentrations ranging from 2440 to
    32 000 µg/m3 (1.3 to 17.0 ppm).  The duration of the NO2 exposure
    was not given; all exposures were expressed in terms of parts per
    million × hours.  A 50% decrease of superoxide anion production began
    after exposure to 54 700 µg/m3 × h (29.1 ppm × h) NO2.  Suzuki et
    al. (1986) reported a marked decrease in the ability of rat AMs to
    produce superoxide anion following a 10-day exposure to either 7520 or
    15 000 µg/m3 (4.0 or 8.0 ppm) NO2.  At the highest concentration,
    the effect was significant each day, but at the lower concentration,
    the depression was significant only on exposure days 3, 5 and 10.

         Alveolar macrophages obtained by BAL from baboons exposed to
    3760 µg/m3 (2.0 ppm) NO2 for 8 h/day, 5 days/week, for 6 months had
    impaired responsiveness to migration inhibitory factor produced by
    sensitized lymphocytes (Greene & Schneider, 1978).  This substance

    affects the behaviour of AMs by inhibiting free migration, which, in
    turn, interferes with the functional capacity of these defence cells. 
    In addition, the random mobility of AMs was significantly depressed in
    rabbits following a 2 h/day exposure for 13 days to 564 µg/m3
    (0.3 ppm), but not to 1880 µg/m3 (1.0 ppm) (Schlesinger, 1987b).

         Vollmuth et al. (1986) studied the clearance of strontium-
    85-tagged polystyrene latex spheres from the lungs of rabbits
    following a single 2-h exposure to 564, 1880, 5640 or 18 800 µg/m3
    (0.3, 1.0, 3.0 or 10.0 ppm) NO2.  An acceleration in clearance
    occurred immediately after exposure to the two lowest NO2
    concentrations; a similar effect was found by Schlesinger & Gearhart
    (1987).  At the higher levels of NO2, an acceleration in clearance
    was not evident until midway through the 14-day post-exposure period. 
    Repeated exposure for 14 days (2 h/day) to 1880 or 18 800 µg/m3
    (1.0 or 10.0 ppm) NO2 produced a response similar to a single
    exposure at the same concentration.

    c)  Humoral and cell-mediated immunity

         Various humoral and cell-mediated effects are summarized in Table
    29.

         Exposing sheep to 9400 µg/m3 (5.0 ppm) NO2, 1.5 h/day for 10 to
    11 days showed that intermittent short-term exposure may temporarily
    alter the pulmonary immune responsiveness (Joel et al., 1982).  One
    technique commonly used in determining the production of specific
    antibody-forming cells is to measure the number of plaque-forming
    cells (PFCs) in the blood or tissues of immunized animals.  In this
    study, the authors assessed immunological response by monitoring the
    daily output of PFCs in the efferent lymph of caudal mediastinal lymph
    nodes of sheep immunized with horse erythrocytes (a T-cell dependent
    antigen).  Although the number of animals used was small and the data
    were not analysed statistically, it would appear that, in the animals
    that were immunized 2 days (but not 4 days) after NO2 exposure
    started, the output of PFC was below control values.  Blastogenic
    responses of T cells from the efferent pulmonary lymph and venous
    blood also appeared to be decreased.

         Hillam et al. (1983) examined the effects of a 24-h exposure to
    9400, 18 800 and 48 900 µg/m3 (5.0, 10.0 and 26.0 ppm) NO2 on
    cellular immunity in rats after intratracheal immunization with sheep
    erythrocytes (SRBCs).  Cellular immunity was evaluated by antigen-
    specific lymphocyte stimulation assays of pooled lymphoid cell
    suspensions from either the thoracic lymph nodes or the spleen. 
    Concentration-related elevation of cellular immunity in thoracic lymph
    nodes and spleen were reported after immunizing the lung with SRBCs.


        Table 29.  Effects of nitrogen dioxide (NO2) on the immune systema
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    188 base;    0.1 base;    Continuous base       Mouse       Suppression of splenic T and B cell responsiveness to       Maigetter et al.
    470, 940,    0.25, 0.5,   with 3-h/day peak                 mitogens variable and not related to concentration or       (1978)
    or 1880      or 1.0 peak  (5 days/week), 1, 3,              duration, except for the 940 µg/m3 continuous group,
    peak                      6, 9, 12 months                   which had a linear decrease in PHA-induced mitogenesis
                                                                with NO2 duration.
    940          0.5          Continuous

    470          0.25         7 h/day,              Mouse       Reduced percentage of total T-cell population and trend     Richters & Damji
                              5 days/week,          (AKR/cum)   towards reduced percentage of certain T-cell                (1988)
                              7 weeks                           subpopulations; no reduction of mature T cells or natural
                                                                killer cells.

    470          0.25         7 h/day,              Mouse       Reduced percentage of total T-cell population and           Richters & Damji
                              5 days/week,          (AKR/cum)   percentages of T helper/inducer cells on days 37 and 181.   (1990)
                              36 weeks

    658          0.35         7 h/day,              Mouse       Trend towards suppression in total percentage of T-cells.   Richters & Damji
                              5 days/week,          (C57BL/6J)  No effects on percentages of other T-cell subpopulations.   (1988)
                              12 weeks

    752          0.4          24 h/day              Mouse       Decrease in primary PFC response at >752 µg/m3.             Fujimaki et al.
    3010         1.6          4 weeks                           Increase in secondary PFC response at 3010 µg/m3.           (1982)
                                                                                                                                              

    Table 29.  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    940 base;    0.5 base;    22-h/day base         Rat         No effect on splenic or circulatory B or T cell response    Selgrade et al.
    2820 peak    1.5 peak     (7 days/week);                    to mitogens. After 3 weeks of exposure only, decrease in    (1991)
                              6-h ramped peak                   splenic natural killer cell activity. No histological
                              (5 days/week)                     changes in lymphoid tissues
                              1, 3, 13, 52,
                              78 weeks

    940 base,    0.5 base,    Continuous base       Mouse       Vaccination with influenza A2/Taiwan virus after exposure.  Ehrlich et al.
    3760 peak    2.0 peak     with 1 h/day                      Decrease in serum neutralizing antibody; haemagglutination  (1975)
                              (5 days/week)                     inhibition titres unchanged. Before virus challenge, NO2
    3760         2.0          peak, 3 months                    exposure decreased serum IgA and increased IgG1, IgM, and
                                                                IgG2; after virus, serum IgA unchanged and IgM increased.

    1880         1.0          493 days              Monkey      Monkeys challenged five times with monkey-adapted           Fenters et al.
                                                                influenza virus during NO2 exposure. Haemagglutination      (1973)
                                                                inhibition antibody titres not altered. Compared to
                                                                controls, NO2 caused an earlier and greater increase in
                                                                serum neutralization antibody titres to the virus.

    1880         1.0          6 months              Guinea-pig  Intranasal challenge with K. pneumoniae after exposure.     Kosmider et al.
                                                                Decreased haemolytic activity of complement; decrease in    (1973)
                                                                all immunoelectrophoretic fractions.

    2820         1.5          24 h/day, 6,          Mouse       Reduction in number of splenic PFCs; lowering               Lefkowitz et al.
    9400         5.0          14, or 21 days                    concentration to 2820 µg/m3 and extending the length to     (1986)
                                                                14 or 21 days decreased PFCs by 33 and 50%, respectively;
                                                                no effect on cell-mediated immune system or
                                                                haemagglutination titres.
                                                                                                                                              

    Table 29.  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    9400         5.0          1.5 h/day,            Sheep       Reduction in PFCs from pulmonary lymph and in mitogenesis   Joel et al.
                              10-11 days                        of T cells from pulmonary lymph and blood.                  (1982)

    9400         5.0          4 h/day,              Guinea-pig  Serum antibodies against lung tissue increased with         Balchum et al.
    28 200       15.0         5 days/week,                      concentration and duration of exposure.                     (1965)
                              5.52 months

    9400         5.0          Continuous, 169       Monkey      Initial depression in serum neutralization titres with      Fenters et al.
                              days, challenged                  return to normal by day 133; no effect on secondary         (1971)
                              4 x with mouse-                   response on haemmagglutin inhibition titre.
                              adapted influenza
                              virus

    9400         5.0          3-7 days              Mouse       No effect on serum interferon levels.                       Lefkowitz et al.
    47 000       25.0                                                                                                       (1983, 1984)

    9400         5.0          24 h                  Rat         Concentration-related elevation of cellular immunity in     Hillam et al.
    18 800       10.0                                           thoracic lymph nodes and spleen after immunizing the lung   (1983)
    48 900       26.0                                           with sheep RBCs.

                                                                                                                                              

    Table 29.  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    9400         5.0          Continuous,           Monkey      Depressed postvaccination serum neutralizing antibody       Ehrlich &
                              6 months                          formation.                                                  Fenters (1973)

    9400         5.0          12 h                  Mouse       No effect on primary and secondary splenic PFC response.    Fujimaki &
                                                                                                                            Shimizu (1981);
                                                                                                                            Fujimaki et al.
                                                                                                                            (1981)
                                                                                                                                              

    a  Source: Modified from US EPA (1993)
    b  PFC = plaque-forming cell; PHA = phytohaemagglutinin; Ig = immunoglobulin; RBCs = red blood cells
             Fujimaki et al. (1982) investigated the effect of a 4-week
    exposure to 752 and 3000 µg/m3 (0.4 and 1.6 ppm) NO2 in mice (i.e.,
    primary and secondary antibody response to SRBCs, using the splenic
    PFC response as the end-point).  The primary PFC response was
    decreased by both NO2 concentrations.  Secondary antibody response
    was not affected at 752 µg/m3 (0.4 ppm), but was slightly enhanced at
    3000 µg/m3 NO2.  Acute exposure (12 h) of mice to 9400 µg/m3
    (5.0 ppm) NO2 caused no such effects (Fujimaki & Shimizu, 1981;
    Fujimaki et al., 1981).

         The effect of exposure pattern was examined by Maigetter et al.
    (1978) by exposing mice for up to 1 year to 940 µg NO2/m3 (0.5 ppm)
    continuously or to three regimens having a continuous baseline of
    188 µg/m3 (0.1 ppm) with 3-h peaks (5 days/week) of either 470, 940
    or 1880 µg/m3 (0.25, 0.5 or 1.0 ppm).  General mitogenic responses of
    splenic lymphocytes to phytohaemagglutinin (PHA) (a T cell dependent
    mitogen) and lipopolysaccharide (a B-cell dependent mitogen)
    decreased, but this was not related to the concentration or duration
    of exposure, with a single exception. The decrease in PHA-induced
    mitogenesis was linearly related to the increased duration of NO2
    exposure to 940 µg/m3 (0.5 ppm).

         Shorter exposure (6 days) to 9400 µg/m3 (5.0 ppm) NO2 did not
    affect mitogenesis of T cells (Lefkowitz et al., 1986).  Although NO2
    did not affect haemagglutination antibody titres, it did reduce the
    number of splenic PFCs to SRBCs.  The authors stated (data were not
    shown) that mice exposed to 2820 µg/m3 (1.5 ppm) NO2 for 14 or
    21 days showed a 33 and 50% decrease, respectively, in the number of
    PFCs.

         Kosmider et al. (1973) exposed guinea-pigs to 1880 µg/m3
    (1.0 ppm) NO2 for 6 months and reported a significant reduction in
    all serum immunoglobulin (Ig) fractions and complement.  Decreased
    levels of these substances may lead to an increase in the frequency,
    duration and severity of an infectious disease.  Mice exposed to NO2
    had decreased serum levels of IgA and exhibited nonspecific increases
    in serum IgM, IgG and IgG2 (Ehrlich et al., 1975).

         These effects on lymphocyte function may reflect changes in
    lymphocyte populations.  Richters & Damji (1988) found that the
    percentage of the total T lymphocyte population was reduced in the
    spleens of AKR/cum mice exposed for 7 weeks (7 h/day, 5 days/week) to
    470 µg/m3 (0.25 ppm) NO2.  The percentages of mature helper/inducer
    T and T cytotoxic/suppressor lymphocytes were also lower in the
    spleens of exposed animals.  There were no changes in the percentages
    of natural killer cells or mature T cells.  Upon a longer (36-week)
    exposure, Richters & Damji (1990) found that the numbers of T-helper/
    inducer (CD4+) lymphocytes (spleen) were reduced, but no effects
    were observed on T cytotoxic/suppressor cells.  Spontaneously
    developing lymphomas in NO2-exposed animals progressed more

    slowly than those in control animals.  This was attributed to
    the NO2-induced reduction in the T-helper/inducer lymphocytes. 
    C57BL/6J mice exposed to 658 µg/m3 (0.35 ppm) for 7 h/day,
    5 days/week for 12 weeks, also showed a suppression in the percentage
    of total matured T lymphocytes, but no effect on any specific
    subpopulation upon longer exposure (36 weeks) to 470 µg/m3 (0.25 ppm)
    (Richters & Damji, 1988).  Selgrade et al. (1991) found that chronic
    exposure (up to 78 weeks) to an urban pattern of NO2 (baseline of
    940 µg/m3 (0.5 ppm) with a ramped 6-h peak to 2820 µg/m3 (1.5 ppm))
    had no effect on splenic or circulating B or T cell mitogenic
    response.  However, there was a transient decrease in splenic natural
    killer cell activity (at 3 weeks only).

         Few studies have been undertaken to assess the effects of NO2 on
    interferon production.  Mice exposed to either 9400 or 47 000 µg/m3
    (5.0 or 25.0 ppm) NO2 for 3 to 7 days had serum levels of interferon
    similar to those of controls (Lefkowitz et al., 1983, 1984).

         Induction of autoimmunity was suggested by the work of Balchum
    et al. (1965). Guinea-pigs exposed to 9400 µg/m3 (5.0 ppm) and
    28 200 µg/m3 (15.0 ppm) NO2 had an increase in the titre of serum
    antibodies against lung tissue, starting after 160 h of NO2 exposure. 
    These antibody titres continued to increase with NO2 concentration
    and duration of exposure.

         The impact of NO2 on the humoral immune response of squirrel
    monkeys to intratracheally delivered influenza vaccine was studied by
    Fenters et al. (1971, 1973) and Ehrlich & Fenters (1973).  In monkeys
    exposed for 493 days to 1880 µg/m3 (1.0 ppm) NO2 and immunized with
    monkey-adapted virus (A/PR/8/34), the serum neutralizing antibody
    titres were significantly increased earlier and to a greater degree
    than those of controls (Fenters et al., 1973; Ehrlich & Fenters,
    1973).  In monkeys exposed to 9400 µg/m3 (5.0 ppm) NO2 for a total
    of 169 days and immunized with mouse-adapted influenza virus (A/PR/8),
    serum neutralization titres were lower than controls initially; no
    significant difference was observed by 133 days of exposure (Fenters
    et al., 1971; Ehrlich & Fenters, 1973).  In all of these studies, the
    haemagglutination inhibition antibody titres were not affected. 
    Differences between studies might be due to the difference in the
    virus used for immunization, along with exposure differences. Also,
    exposure to 1880 µg/m3 (1.0 ppm) NO2 may have increased the
    establishment of infection and the survival of the monkey-adapted
    virus within the respiratory tract, resulting in an increase in
    antibody production.

         Mice that were vaccinated with influenza virus (A-2/Taiwan/ 1/64)
    after 3 months of continuous exposure to 3760 µg/m3 (2.0 ppm) or to
    940 µg/m3 (0.5 ppm) NO2 with a 1-h daily (5 days/week) spike
    exposure to 3760 µg/m3 (2.0 ppm) had mean serum neutralizing antibody
    titres that were four-fold lower than those of clean air controls
    (Ehrlich et al., 1975).  The haemagglutination inhibition antibody
    titres in these animals were unchanged.  This agrees with the Fenters
    et al. (1973) findings in monkeys exposed to 1880 µg/m3 (1.0 ppm) for
    over 1 year.

    d)  Interaction with infectious agents

         Various experimental approaches have been employed using animals
    in an effort to determine the overall functional efficiency of the
    host's pulmonary defences following NO2 exposure.  In the most
    commonly used infectivity model, animals are exposed to either NO2
    or filtered air.  After NO2 exposure, the treatment groups are
    combined and exposed briefly to an aerosol of a viable agent, such as
     Streptococcus sp.,  Klebsiella pneumoniae, Diplococcus pneumoniae
    or influenza virus.  The animals are then returned to clean air
    for a holding period (usually 15 days), and the mortality in the
    NO2-exposed and the control groups are compared. If host defences are
    compromised by the NO2 exposure, mortality rates will be higher
    (Ehrlich, 1966; Henry et al., 1970; Coffin & Gardner, 1972; Ehrlich et
    al., 1979; Gardner, 1982).  Although the end-point is mortality, it is
    a sensitive indicator of the depression of the defence mechanisms used
    to control infection.  Because these specific defence mechanisms are
    common to laboratory animals and humans, the increased susceptibility
    to infection can be qualitatively extrapolated to humans, even though
    mortality would not be an expected outcome in humans receiving
    appropriate medical treatment.  However, different exposure levels of
    NO2 and infectious agents may be required to produce changes in human
    host defences.  Effects of NO2 on pulmonary infectious disease in
    humans are discussed in chapters 6 and 7.  Table 30 summarizes effects
    of exposure to NO2 and infectious agents observed in animals.

         An enhancement in mortality following exposure to NO2 in
    combination with a pathogenic microorganism could be due to several
    factors. Goldstein et al. (1973) showed decreases in pulmonary
    bactericidal activity following NO2 exposure.  In their first
    experiments, mice breathed aerosols of  Staphylococcus aureus
     (S. aureus) labelled with radioactive phosphorus and were then
    exposed to NO2 for 4 h. Physical removal of the bacteria was not
    affected by any of the NO2 concentrations used up to 27 800 µg/m3
    (14.8 ppm).  Concentrations > 13 200 µg/m3 (7.0 ppm) NO2 lowered
    the bactericidal activity by > 7%.  Lower concentrations (3570 and
    7140 µg/m3 (1.9 and 3.8 ppm)) had no significant effect.  In another
    experiment (Goldstein et al., 1974), mice breathed 1800, 4320 and
    12 400 (1.0, 2.3 and 6.6 ppm) NO2 for 17 h and then were exposed
    to an aerosol of  S. aureus.  Four hours later, the animals were
    examined for the number of organisms present in the lungs.  No
    difference in the number of bacteria inhaled was found in the
    NO2-exposed animals. Concentrations of 4320 and 12 400 µg/m3
    (2.3 and 6.6 ppm) NO2 decreased pulmonary bactericidal activity by
    6 and 35%, respectively, compared to controls. Exposure to 1880 µg/m3
    (1.0 ppm) NO2 had no significant effect. Goldstein et al. (1974)
    hypothesized that the decreased bactericidal activity was due to
    defects in AM function.  Jakab (1987) confirmed these findings and
    found that the concentration of NO2 required to suppress pulmonary
    bactericidal activity in mice depended on the specific organism.  For
    example, exposure to > 7520 µg/m3 (> 4.0 ppm) NO2 for 4 h
    after bacterial challenge depressed  bactericidal activity
    against  S. aureus, but it required a concentration of 18 800 to
    37 600 µg/m3 (10.0 to 20.0 ppm) before the lung's ability to kill
    deposited  Pasteurella and  Proteus was impaired. Parker et al.
    (1989) made similar observations in mice exposed for 4 h to 9400 or
    18 800 µg/m3 (5.0 or 10.0 ppm) NO2 and infected with  Mycoplasma
     pulmonis.  The higher concentration of NO2 increased mortality.
    Both concentrations: (1) reduced lung bactericidal activity and
    increased bacterial growth, without affecting deposition or physical
    clearance; and (2) increased the incidence of lung lesions as well as
    their severity. Davis et al. (1991) found no effects of lower NO2
    concentrations on bactericidal activity using the same model system.

        Table 30.  Interaction of nitrogen dioxide (NO2) with infectious agentsa
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                              

    100 base,      0.05 base,     Continuous, with          Mouse      Streptococcus     No effect                            Gardner (1980);
    188 peak       0.1 peak       1 h peak, twice/day                  sp.                                                    Gardner et al.
                                  (5 days/week),                                                                              (1982); Graham
                                  15 days                                                                                     et al. (1987)

    940 +          0.5 +                                                                 Increased mortality
     1880 peak     1.0 peak

    2260 +         1.2 +                                                                 Increased mortality
     4700 peak     2.5 peak

    376 base,      0.2 base,      Continuous base           Mouse      Streptococcus     Spike plus baseline caused           Miller et al.
    1500 peak      0.8 peak       with 1-h peak                        sp.               significantly greater mortality      (1987)
                                  twice/day                                              than baseline.
                                  (5 days/week),
                                  1 year

    564-940        0.3-0.5        Continuous,               Mouse      A/PR/8            High incidence of adenomatous        Motomiya et al.
                                  3 months                             virus             proliferation of peripheral and      (1973)
                                                                                         bronchial epithelial cells; NO2
                                                                                         alone and virus alone caused less
                                                                                         severe alterations.
                                                                                                                                              

    Table 30  (Con't)
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                              

                                  Continuous,                                            No enhancement of effect of NO2
                                  6 months                                               and virus.

    940            0.5            3 h/day,                  Mouse      Streptococcus     Increase in mortality with           Ehrlich et al.
                                  3 months                             sp.               reduction in mean survival time.     (1979)

    940            0.5            Intermittent,             Mouse      Klebsiella        Increased mortality after 6 months   Ehrlich &
                                  6 or 18 h/day,                       pneumoniae        intermittent exposure or after       Henry (1968)
                                  up to 12 months                                        3, 6, 9 or 12 months continuous
                                                                                         exposure; following 12 months
                                                                                         exposure, increased mortality was
                                  Continuous,                                            significant only in continuously
                                  24 h/day up to                                         exposed mice.
                                  12 months

    940-1880       0.5-1.0        Continuous,               Mouse      A/PR/8            Increased susceptibility to          Ito (1971)
                                  39 days                   (female)   virus             infection
    18 800         10.0           2 h/day, 1, 3,
                                  and 5 days

    940-52 600     0.5-28.0       Varied                    Mouse      Streptococcus     Increased mortality with increased   Gardner et al.
                                                                       sp.               time and concentration;              (1977a,b); Coffin
                                                                                         concentration is more important      et al. (1977)
                                                                                         than time.

    940            0.5            24 h/day,                 Mouse      K. pneumoniae     Significant increase in mortality    McGrath &
    1880           1.0            7 days/week,                                           after 3-day exposure to 9400 µg/m3;  Oyervides (1985)
    2820           1.5            3 months                                               no effect at other concentrations,
                                                                                                                                              

    Table 30  (Con't)
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                               

    9400           5.0            3 days                                                 but control mortality was very
                                                                                         high.

    1880           1.0            17 h                      Mouse      Staphylococcus    No difference in number of bacteria  Goldstein et al.
    4320           2.3                                                 aureus after      deposited, but at 4320 and           (1974)
    12 400         6.6                                                 NO2 exposure      12 400 µg/m3, there was a decrease
                                                                                         in pulmonary bactericidal activity
                                                                                         of 6 and 35%, respectively;
                                                                                         no effect at 1880 µg/m3.

    1880-4700      1.0-2.5        4 h                       Mouse      S. aureus         Impaired bactericidal activity       Jakab (1988)
                                                                                         between 1800 and 4700 µg/m3 in
                                                                                         animals injected with
                                                                                         corticosteroids

    4320           2.3                                                                   6% decrease in bactericidal
                                                                                         activity

    12 400         6.6                                                                   35% decrease in bactericidal
                                                                                         activity

    1880           1.0            48 h                      Mouse      Streptococcus     Increase proliferation of            Sherwood et al.
                                                                       sp.; S. aureus    Streptococcus sp., but not           (1981)
                                                                                         S. aureus, in lung

    1880           1.0            3 h                       Mouse      Streptococcus     Exercise on continuously moving      Illing et al.
    5640           3.0                                                 sp.               wheels during exposure; increased    (1980)
                                                                                         mortality at 5640 µg/m3
                                                                                                                                              

    Table 30  (Con't)
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                              

    2820           1.5            Continuous or             Mouse      Streptococcus     After 1 week, mortality with         Gardner et al.
                                  intermittent                         sp.               continuous exposure was greater      (1979)
                                  (7 h/day), 7 days                                      than that for intermittent; after
                                  per week, 2 weeks                                      2 weeks, no significant difference
                                                                                         between continuous and intermittent
                                                                                         exposure.

    6580           3.5                                                                   Increased mortality with increased
                                                                                         duration of exposure; no
                                                                                         significant difference between
                                                                                         continuous and intermittent
                                                                                         exposure; with data adjusted for
                                                                                         total difference in the production
                                                                                         of concentration and time,
                                                                                         mortality essentially the same.

    2820 base,     1.5 base,      Continuous 60 h           Mouse      Streptococcus     Mortality increased with 3.5- and    Gardner (1980);
    8460 peak      4.5 peak       then peak for 1,                     sp.               7-h single spike when bacterial      Garnder et al.
                                  3.5 or 7 h, then                                       challenge was immediate, and         (1982); Graham
                                  continuous 18 h                                        18 h after the spike                 et al. (1987)

    8460           4.5            1, 3.5, or 7 h                                         Mortality proportional to duration
                                                                                         when bacterial challenge was
                                                                                         immediate, but not 18 h
                                                                                         post-exposure.

    2820           1.5            7 h/day, 4, 5,            Mouse      Streptococcus     Elevated temperature (32°C)          Gardner et al.
                                  and 7 days                           sp.               increased mortality after 7 days.    (1982)
                                                                                                                                              

    Table 30  (Con't)
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                              

    2820           1.5            2 h                       Mouse      K. pneumoniae     Increased mortality only at          Purvis &
    4700           2.5                                                                   > 6580 µg/m3.  Increase in           Ehrlich (1966);
                                                                                         mortality                  
    
    6580           3.5                                                                   K. pneumoniae challenge 1 and 6 h    Ehrlich (1979)
    9400           5.0                                                                   after 9400 or 18 800 µg/m3;
    18 800         10.0                                                                  when K. pneumoniae challenge 27 h
    28 200         15.0                                                                  following NO2 exposure, effect
                                                                                         only at 28 200 µg/m3.

    3570           1.9            4 h                       Mouse      S. aureus         Physical removal of bacteria         Goldstein et al.
    7140           3.8                                                 prior to NO2      unchanged at 3570 to 27 800 µg/m3.   (1973)
                                                                       exposure

    13 160         7.0                                                                   7% lower bactericidal activity
    17 300         9.2                                                                   14% lower bactericidal activity
    27 800         14.8                                                                  50% lower bactericidal activity

    3760           2.0            3 h                       Mouse      Streptococcus     Increased mortality                  Ehrlich et al.
                                                                       sp.                                                    (1977);
                                                                                                                              Ehrlich (1980)

    4700           2.5-30.0       4 h                       Mouse      S. aureus,        Concentration-related decrease       Jakab (1987)
    56 400                                                             Pasteurella and   in bactericidal activity, starting
                                                                       Proteus           at > 7500 µg/m3 with S. aureus when
                                                                                         NO2 exposure was after bacterial
                                                                                         challenge; when NO2 was before
                                                                                         bacterial challenge, effect at
                                                                                         18 800 µg/m3.  Higher concentration
                                                                                         required to affect other organisms.
                                                                                                                                              

    Table 30  (Con't)
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                              

    6580           3.5            2 h                       Mouse      K. pneumoniae     Increased mortality of all species   Ehrlich (1975)

    65 830         35.0           2 h                       Hamster
    94 050         50.0           2 h                       Squirrel
                                                             monkey

    9400           5.0            6 h/day,                  Mouse      Cytomegalovirus   Increase in virus susceptibility     Rose et al.
                                  6 days                                                                                      (1988)

    9400           5.0            Continuous,               Squirrel   K. pneumoniae     Increased viral-induced mortality    Henry et al.
                                  2 months                   monkey    or A/PR/8         (1/3).  Increase in Klebsiella-      (1970)
                                                                       influenza virus   induced mortality (2/7); no
                                                                       deaths.           control

    19 000         10.0           Continuous,                                            Increased virus-induced mortality
                                  1 month                                                (6/6) within 2-3 days after
                                                                                         infection; no control deaths.
                                                                                         Increase in Klebsiella-induced
                                                                                         mortality (1/4); no control deaths.

    9400           5.0            4 h                       Mouse      Mycoplasma        NO2 increased incidence and          Parker et al.
    19 000         10.0                                                pulmonis          severity of pneumonia lesions and    (1989)
                                                                                         decreased the number of organisms
                                                                                         needed to induce pneumonia; no
                                                                                         effect on physical clearance,
                                                                                         decreased mycoplasmal killing
                                                                                         and increased growth; no effect on
                                                                                         specific IgM in serum;
                                                                                                                                              

    Table 30  (Con't)
                                                                                                                                              

    NO2 concentration
                            

    µg/m3          ppm            Exposure                  Species    Infective agent   Effects                              Reference
                                                                                                                                              

                                                                                         C57Bl/6N mice generally more
                                                                                         sensitive than C3H/HeN mice.
                                                                                         At 19 000 µg/m3, one strain
                                                                                         (C57BL/6N) of mice had increased
                                                                                         mortality.

    9400           5.0            2 months                  Squirrel   K. pneumoniae     Mortality 2/7; bacteria present      Henry et al.
                                                             monkey                      in lung of survivors at autopsy.     (1969)

    65 800         35.0           1 month                                                Mortality 1/4; bacteria present
                                                                                         in lungs of survivors at autopsy.

    94 000         50.0           2 h                                                    Mortality 3/3
                                                                                                                                              

    a  Modified from US EPA (1993)
             Differences in species susceptibility to NO2 or to a pathogen
    may play a role in the enhancement of mortality seen in experimental
    animals.  An enhancement in mortality was noted in mice, hamsters and
    monkeys acutely exposed to NO2 for 2 h followed by a challenge
    of  K. pneumonia (Ehrlich, 1975).  However, differences in
    susceptibility were noted between the species.  Ehrlich found
    increased mortality occurred in monkeys only at 94 000 µg/m3
    (50.0 ppm), whereas, lower NO2 levels increased mortality in mice
    (6580 µg/m3, 3.5 ppm) and hamsters (65 800 µg/m3, 35.0 ppm). The
    mouse model was the most sensitive to NO2 exposure, as shown by
    enhanced mortality from  K. pneumoniae following exposure to
    6580 µg/m3 (3.5 ppm) but not to 2820-4700 µg/m3 (1.5-2.5 ppm) NO2
    for 2 h (Purvis & Ehrlich, 1963; Ehrlich, 1975).  With prolonged
    (2 month) exposure, Henry et al. (1969) found that lower levels of
    NO2 (9400 µg/m3, 5.0 ppm) increased  susceptibility to bacterial
    infections in monkeys than the 50.0 ppm concentration found to be
    effective by Ehrlich (1975) with acute (2 h) exposure.  The
    sensitivity is also affected by the test organism.  For example, when
     Streptococcus sp. was the infectious agent, a 3-h exposure to
    3760 µg/m3 (2.0 ppm) NO2 caused an increased in mortality in mice
    (Ehrlich et al., 1977).  Sherwood et al. (1981) illustrated that
    exposure to 1880 µg/m3 (1.0 ppm) NO2 for 48 h increased the
    propensity of virulent group-C streptococci, but not  S. aureus, to
    proliferate within mouse lungs and cause earlier mortality.

         Additional factors can influence the interaction of NO2 and
    infectious agents.  Mice placed on continuously moving exercise wheels
    during exposure to 5640 µg/m3 (3.0 ppm) NO2, but not 1880 µg/m3
    (1.0 ppm), for 3 h showed enhanced mortality over non-exercised
    NO2-exposed mice using the streptococcal infectivity model (Illing et
    al., 1980).  The presence of other environmental factors, such as O3
    (Ehrlich et al., 1977; Gardner, 1980; Gardner et al., 1982; Graham et
    al., 1987) or elevated temperatures (Gardner et al., 1982), also
    exacerbated the effects of NO2.

         The influence of a wide variety of exposure regimens has been
    evaluated using the infectivity model.  For example, Gardner et al.
    (1977b) examined the effect of varying durations of continuous
    exposure on the mortality of mice exposed to six concentrations of
    NO2 (940 to 52 600 µg/m3 (0.5 to 28.0 ppm)) for durations ranging
    from 15 min to 1 year.   Streptococcus sp. was used for all
    concentrations, except 940 µg/m3, in which case  K. pneumoniae was
    used.  Mortality increased linearly with increasing duration of
    exposure to a given concentration of NO2.  Mortality also increased
    with increasing concentration of NO2 as indicated by the steeper
    slopes with higher concentrations.  When the product of concentration
    and time (C × T) was held constant, the relationship between
    concentration and time produced significantly different mortality
    responses.  At a constant C × T of approximately 21 ppm-h, a 14-h
    exposure to 2820 µg/m3 (1.5 ppm) NO2 increased mortality by 12.5%,
    whereas a 1.5-h exposure to 27 300 µg/m3 (14.0 ppm) NO2 enhanced
    mortality by 58.5%.  These findings demonstrate that concentration is
    more important than time in determining the degree of injury induced
    by NO2 in this model, and they were confirmed at additional C × T
    values (Gardner et al., 1977a,b, 1982; Coffin et al., 1977).

         Gardner et al. (1979) also compared the effect of continuous
    versus intermittent exposure to NO2 followed by bacterial challenge
    with  Streptococcus sp.  Mice were exposed either continuously or
    intermittently (7 h/day, 7 days/week) to 2820 or 6580 µg/m3 (1.5 or
    3.5 ppm) NO2.  The continuous exposure of mice to 2820 µg/m3 NO2
    increased mortality after 24 h of exposure.  During the first week of
    exposure, the mortality was significantly higher in mice exposed
    continuously to NO2 than in those exposed intermittently.  By the
    14th day of exposure, the difference between intermittent and
    continuous exposure became indistinguishable.  At the higher
    concentration, there was essentially no difference between continuous
    and intermittent regimens.  This suggests that fluctuating levels of
    NO2 may ultimately be as toxic as sustained high levels (Gardner et
    al., 1979).

         Mice were exposed continuously or intermittently (6 or 18 h/day)
    to 940 µg/m3 (0.5 ppm) NO2 for up to 12 months (Ehrlich & Henry,
    1968).  None of the exposure regimens affected resistance to
     K. pneumoniae infection during the first month.  Those exposed
    continuously exhibited decreased resistance to the infectious agent,
    as demonstrated by a significant enhancement in mortality at 3, 6, 9
    and 12 months.  In another experiment, a significant enhancement did

    not occur at 3 months, but was observed after 6 months of exposure. 
    After 6 months, mice exposed intermittently (6 or 18 h/day) to NO2
    showed significant increases in mortality over controls (18%).  Only
    the continuously exposed animals showed increased mortality (23%) over
    controls following 12 months of exposure.  After 12 months of
    exposure, mice in the three experimental groups showed a reduced
    capacity to clear viable bacteria from their lungs.  This was first
    observed after 6 months in the continuously exposed group and after
    9 months in the intermittently exposed groups.  These changes,
    however, were not statistically tested for significance.  Although it
    is not possible to compare directly the results of the studies using
     Streptococcus sp. to those using  K. pneumoniae, the data suggest
    that, as the concentration of NO2 is decreased, a longer exposure
    time is necessary for the intermittent exposure regimen to produce a
    level of effect equivalent to that of a continuous exposure.  McGrath
    & Oyervides (1985) did not confirm these findings in mice exposed to
    940, 1880 and 2820 µg/m3 (0.5, 1.0 and 1.5 ppm) NO2 for 3 months. 
    The inconsistency may be attributed to the fact that the McGrath &
    Oyervides (1985) study had 95% mortality in the control groups, making
    it virtually impossible to detect a further enhancement in mortality
    due to NO2.

         Gardner (1980), Gardner et al. (1982) and Graham et al. (1987)
    reported extensive investigations on the response to airborne
    infections in mice breathing NO2 spike exposures superimposed on a
    lower continuous background level of NO2, which simulated the pattern
    (although not the NO2 concentrations) of exposure in the urban
    environment in the USA.  Mice were exposed to spikes of 8460 µg/m3
    (4.5 ppm) for 1, 3.5 or 7 h and then were challenged with
     Streptococcus sp. either immediately or 18 h after exposure. 
    Mortality was proportional to the duration of the spike when the mice
    were challenged with bacteria immediately after exposure, but mice had
    recovered from the exposure by 18 h.  Similar findings were reported
    by Purvis & Ehrlich (1963) using  K. pneumoniae.  When a spike of
    8460 µg/m3 (4.5 ppm) was superimposed on a continuous background of
    2820 µg/m3 (1.5 ppm) for 62 h preceding and 18 h following the spike,
    mortality was significantly enhanced by a spike lasting 3.5 or 7 h
    when the infectious agent was administered 18 h after the spike
    (Gardner, 1980; Gardner et al., 1982; Graham et al., 1987).  Possible
    explanations for these differences due to the presence or absence of a
    background exposure are that mice continuously exposed are not capable
    of recovery or that new AMs or PMNs recruited to the site of infection
    are impaired by the continuous exposure to NO2.  The effect of
    multiple spikes was examined by exposing mice for 2 weeks to two
    daily 1-h spikes (morning and afternoon, 5 days/week) of 8460 µg/m3
    (4.5 ppm) superimposed on a continuous background of 2820 µg/m3
    (1.5 ppm) NO2.  Mice were challenged with the infectious agent either
    immediately before or after the morning spike.  When the infectious
    agent was given before the morning spike, the increase in mortality

    did not closely approach that of a continuous exposure to 2820 µg/m3
    (1.5 ppm) NO2.  However, in mice challenged after the morning spike,
    by 2 weeks of exposure, the increased mortality over controls
    approached that equivalent to continuous exposure to 2820 µg/m3
    (1.5 ppm) NO2.  Thus, the magnitude of the effect of the base-plus-
    spike group, which had a higher C × T than the continuous groups, did
    not exceed the effect of the continuous group.  These findings
    demonstrate that the pattern of exposure determines the response and
    that the response is not predictable based on a simple C × T
    relationship.

         Further investigations into the effects of chronic exposure to
    NO2 spikes on murine antibacterial lung defences have been conducted
    using a spike-to-baseline ratio of 4:1, which is not uncommon in the
    urban environment in the USA (Miller et al., 1987).  For 1 year, mice
    were exposed 23 h/day, 7 days/week, to a baseline of 376 µg/m3
    (0.2 ppm) or to this baseline level on which was superimposed a 1-h
    spike of 1500 µg/m3 (0.8 ppm) NO2, twice a day, 5 days/week.  The
    animals exposed to the baseline level did not exhibit any significant
    effects; however, the streptococcal-induced mortality of the mice
    exposed to the baseline plus spike regimen was significantly greater
    than that of either the NO2-background-exposed mice or the control
    mice.  Human epidemiological studies in chapter 7 indicate increased
    risk of respiratory infection.  Data from experimental animals support
    the epidemiological responses in humans.

         Antiviral defences are also compromised by NO2.  Squirrel
    monkeys exposed to 9400 or 18 800 µg/m3 (5.0 or 10.0 ppm) NO2 for
    2 or 1 month, respectively, had an increased susceptibility to a
    laboratory-induced viral influenza infection (Henry et al., 1970). 
    All six animals exposed to the highest concentration died within 2
    to 3 days of infection with the influenza virus; at the lower
    concentration, one out of three monkeys died.

         Mice exposed continuously for 3 months to 564-940 µg/m3
    (0.3-0.5 ppm) NO2 followed by a challenge with A/PR/8 influenza virus
    exhibited significant pulmonary pathological responses (Motomiya et
    al., 1973).  A greater incidence of adenomatous proliferation of
    bronchial epithelial cells resulted from the combined exposures of
    virus plus NO2 than with either the viral or NO2 exposures alone. 
    Continuous NO2 exposure for an additional 3 months did not enhance
    the effect of NO2 or the subsequent virus challenge.

         Ito (1971) challenged mice with influenza A/PR/8 virus after
    continuous exposure to 940 to 1880 µg/m3 (0.5 to 1.0 ppm) NO2 for
    39 days and to 18 800 µg/m3 (10.0 ppm) NO2, 2 h daily for 1, 3 and
    5 days.  Acute and intermittent exposure to 18 800 µg/m3 (10.0 ppm)
    NO2 as well as continuous exposure to 940 to 1880 µg/m3 (0.5 to
    1.0 ppm) NO2 increased the susceptibility of mice to influenza virus
    as demonstrated by increased mortality.

         The lower respiratory tract of mice became significantly more
    susceptible to murine cytomegalovirus infection after 6-h exposures
    for 6 days to 9400 µg/m3 (5.0 ppm) NO2 (Rose et al., 1988).  No
    effects occurred at levels < 4700 µg/m3 (2.5 ppm).  Exposure to
    9400 µg/m3 (5.0 ppm) NO2 did not significantly alter the course of
    a parainfluenza (murine sendai virus) infection in mice as measured by
    the infectious pulmonary virus titres in the lungs.  However, this
    concentration of NO2, when combined with the virus exposure, did
    increase the severity of the pulmonary disease process (viral
    pneumonitis) (Jakab, 1988).

    5.2.2.2  Lung biochemistry

         Studies of lung biochemistry in animals exposed to NO2 have
    focused on either the putative mechanisms of toxic action of NO2 or
    on detection of indicators of tissue and cell damage.  One theory of
    the mechanism underlying NO2 toxicity is that NO2 initiates lipid
    peroxidation in unsaturated fatty acids in membranes of target cells,
    thereby causing cell injury or death (Menzel, 1976).  Another theory
    is that NO2 oxidizes water-soluble, low molecular weight reducing
    substances and proteins, resulting in a metabolic dysfunction that
    manifests itself in toxicity (Freeman & Mudd, 1981).  It is likely
    that NO2 acts by both means. Several potential biochemical mechanisms
    related to detoxification of NO2 or to responses to NO2 intoxication
    have been proposed and summarized below according to impacts on
    lipids, proteins, and antioxidant metabolism and antioxidants.  The
    following discussion focuses on inhalation studies because they are
    more interpretable for risk assessment purposes;  in vitro exposure
    studies have been reviewed elsewhere (US EPA, 1993).

    a)  Lipid peroxidation

         Animal toxicology studies evaluating effects of NO2 on lipid
    peroxidation are summarized in Table 31.

         Lipid peroxidation induced by NO2 exposure has been detected at
    exposure levels as low as 75 µg/m3 (0.04 ppm).  Lipid peroxidation,
    measured as ethane exhalation, was detected after 9 months of exposure
    of rats to 75-750 µg/m3 (0.04-0.4 ppm) (Sagai et al., 1984).  Lipid
    peroxidation has also been evaluated by measuring the content of lipid
    peroxides or substances reactive to thiobarbituric acid in alveolar
    lavage fluid and lung tissue after exposure to similar NO2
    concentrations (Ichinose & Sagai, 1982; Ichinose et al., 1983).  Acute
    or subacute exposure to higher concentrations of NO2 has also been
    shown to cause a rapid increase in lung peroxide levels in several
    species.

        Table 31.  Effects of nitrogen dioxide (NO2) on lung lipid metabolisma
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    75           0.04         Continuous, 9,        Rat         Increased TBA products at 7520 µg/m3 after 9 months and     Sagai et al.
    752          0.4          18 or 27 months                   at > 752 µg/m3 after 18 months; increased ethane            (1984)
    7520         4.0                                            exhalation at all levels. No changes in total lipid,
                                                                phospholipid, total cholesterol or triglyceride contents.

    75           0.04         Continuous, 6,                    Increased ethane exhalation after 9 and 18 months.
    225          0.12         9 and 18 months
    752          0.4

    752          0.4          2 weeks               Rat         Changes in TBA-reactive substances, exhaled ethane and      Ichinose et
    2260         1.2          1-16 weeks                        enzyme activities in lung homogenates, dependent on         al. (1983)
    7520         4.0                                            concentration and duration of exposure.
    18 800       10.0

    75           0.04         9, 18,
    752          0.4          27 months
    7520         4.0

    752          0.4          4 months              Rat         Duration-dependent increase in ethane exhalation and        Ichinose &
    2260         1.2                                            TBA-reactive substances; peak increase in early weeks of    Sagai (l982)
    7520         4.0                                            exposure, return towards control in mid-exposure, and
                                                                increase late in exposure.

    752          0.4          72 h                  Guinea-pig  No effect at 752 µg/m3; increase in lung lipid content in   Selgrade et
    1880         1.0                                            BAL of vitamin C-depleted, but not normal, animals at       al. (1981)
    5640         3.0                                            1880 µg/m3 or more.
    9400         5.0
                                                                                                                                              

    Table 31  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    9400         5.0          3 h                               Increased lung lipid content in vitamin C-depleted
                                                                guinea-pigs 18-h after exposure.

    752          0.4          1 week                            No effects in normal or vitamin C-depleted animals.

    1880         1.0          Continuous,           Rabbit      Decrease in lecithin synthesis after 1 week; less marked    Seto et al.
                              2 weeks                           depression after 2 weeks.                                   (1975)

    1880         1.0          4 h/day,              Rat         Vitamin E supplement reduced the lipid peroxidation.        Thomas et al.
                              6 days                                                                                        (1967)

    5450         2.9          Continuous,           Rat         Increase in lung wet weight (l2.7%) and decrease in total   Arner &
                              5 days/week                       lipid (8.7%); decrease in saturated fatty acid content of   Rhoades (1973)
                              9 months                          lung lavage fluid and tissue; increase in surface tension
                                                                of lung lavage fluid; and decrease in lung compliance.

    1880         1.0          2 h                   Rabbit      1800 µg/m3: elevated thromboxane B2. 5640 µg/m3:            Schlesinger
    5640         3.0                                            depressed thromboxane B2. 18 800 µg/m3: depressed           et al. (1990)
    18 800       10.0                                           6-keto-prostaglandin F1alpha and thromboxane B2.

    5640         3.0          Continuous,           Rat         Decrease in linoleic and linolenic acid content of BAL.     Menzel et al.
                              17 days                                                                                       (1972)

    5640         3.0          7 days                Rat         Increased TBA reactants with vitamin E deficiency.          Sevanian et al.
                                                                                                                            (1982)
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  TBA = Thiobarbituric acid; BAL = Bronchoalveolar lavage
             Lipid peroxidation results in an alteration in phospholipid
    composition.  Exposure of either mice or guinea-pigs to an NO2
    level of 750 µg/m3 (0.4 ppm) for a week resulted in a decreased
    concentration of phosphatidyl ethanolamine and a relative increase in
    the phosphatidyl choline concentration (Sagai et al., 1987).

         Several investigators have also demonstrated NO2-induced lipid
    peroxidation in  in vitro systems.  The cell type most commonly used
    is the endothelial cell from either pig arteries or aorta.  Studies
    using these cell types have recently attempted to relate the effect on
    lipid metabolism to functional parameters such as membrane fluidity
    and enzyme activation or inactivation.

         Membrane fluidity changes are related to lipid peroxidation. 
    NO2-induced changes in membrane fluidity have been demonstrated in
    alveolar macrophages and endothelial cells in culture.  Endothelial
    cells exposed to a NO2 level of 9400 µg/m3 (5 ppm), for instance,
    exhibit decreased membrane fluidity after 3 h.  Thus, NO2 changes the
    physical state of the membrane lipids, perhaps through initiating
    lipid peroxidation, and hence impairs membrane functions (Patel et
    al., 1988).

         Lipid peroxidation can also activate phospholipase activities. 
    Activation of phospholipase A1 in cultured endothelial cells by NO2
    has been demonstrated.  This activation, which is specific for
    phospholipase A1 occurs at an NO2 concentration of 9400 µg/m3
    (5 ppm) after 40 h of exposure and is speculated to depend on a
    specific NO2-induced increase in phosphatidyl serine in the plasma
    membranes (Sekharam et al., 1991).

         One function of phospholipases is the release of arachidonic
    acid.  The effect of NO2 on the release and metabolism of arachidonic
    acid has been studied both  in vivo and  in vitro.  Both an increase
    and a decrease in the metabolism of arachidonic acid has been observed
    in several species.   In vivo exposure of rats to 18 800 µg/m3
    (10 ppm) for 2 h resulted in decreased levels of prostaglandins E2
    and F2alpha, as well as thromboxane B2, in lavage fluid.  On the
    other hand, at an exposure level of 1880 µg/m3 (1 ppm), the
    concentrations of thromboxane B2 were increased (Schlesinger et al.,
    1990).

    b)  Effects on lung proteins and enzymes

         Nitrogen dioxide can cause lung inflammation (associated with
    concomitant infiltration of serum protein, enzymes and inflammatory
    cells) and hyperplasia of Type 2 cells. Thus, some changes in lung
    enzyme activity and protein content may reflect inflammation and/or
    changes in cell types, rather than direct effects of NO2 on lung cell
    enzymes.  Some direct effects of NO2 on enzymes are possible because

    NO2 can oxidize various reducible amino acids or side chains of
    proteins in aqueous solution (Freeman & Mudd, 1981).  These effects
    are summarized in Table 32.

         Nitrogen dioxide can increase the protein content of BAL in
    vitamin-C-deficient guinea-pigs (Sherwin & Carlson, 1973; Selgrade et
    al., 1981; Hatch et al., 1986; Slade et al., 1989).  Selgrade et al.
    (1981) found effects at NO2 levels as low as 1880 µg/m3 (1.0 ppm)
    after a 72-h exposure, but a 1-week exposure to 752 µg/m3 (0.4 ppm)
    did not increase protein levels.  The results of the 1-week exposure
    apparently conflict with those of Sherwin & Carlson (1973), who found
    increased protein content of BAL from vitamin-C-deficient guinea-pigs
    exposed to 752 µg/m3 (0.4 ppm) NO2 for 1 week.  Differences in
    exposure techniques,  protein measurement methods, and/or degree of
    vitamin C deficiencies may explain the difference between the two
    studies.  Hatch et al. (1986) found that the NO2-induced increase in
    BAL protein in vitamin-C-deficient guinea-pigs was accompanied by an
    increase in lung content of non-protein sulfhydryls and ascorbic acid
    and a decrease in vitamin E content.  The increased susceptibility to
    NO2 was observed when lung vitamin C was reduced (by diet) to levels
    below 50% of normal.  A depletion of lung non-protein sulfhydryls also
    enhances susceptibility to a high level (18 800 µg/m3, 10.0 ppm) of
    NO2 (Slade et al., 1989).

         The effects of NO2 on structural proteins of the lungs has been
    of major interest because elastic recoil is lost after exposure
    (section 5.2.2.3).  Last et al. (1983) examined collagen synthesis
    rates by lung minces from animals exposed to NO2.  In rats
    continuously exposed to 9400 to 47 000 µg/m3 (5.0 to 25.0 ppm) NO2
    for 7 days, there was a linear concentration-related increase in
    collagen synthesis rate.  In a subsequent paper, Last & Warren (1987)
    confirmed that 9400 µg/m3 (5.0 ppm) increased collagen synthesis. 
    Such biochemical changes are typically interpreted as reflecting
    increases in total lung collagen, which, if sufficient, could result
    in pulmonary fibrosis after longer periods of exposure.  However, such
    correlations have not been made directly after NO2 exposure.

        Table 32.  Effects of nitrogen dioxide (NO2) on lung proteins and enzymesa
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    75           0.04         Continuous,           Rat         NPSHs increased at the 2 higher NO2 levels after 9 or 18    Sagai et al.
    752          0.4          9 and 18 months                   months; GSH peroxidase activity decreased at 752 µg/m3      (1984)
    7520         4.0                                            after 18 months and at 7520 µg/m3 after 9 or 18 months;
                                                                GSH reductase activity increased after a 9-month exposure
                                                                to 7520 µg/m3; G-6-PD was increased after a 9- or 18-month
                                                                exposure to 7520 µg/m3; no effects on 6-phosphogluconate
                                                                dehydrogenase, superoxide dismutase, or disulfide
                                                                reductase; some GSH S-transferases had decreased
                                                                activities after an 18-month exposure to 752 or
                                                                7520 µg/m3.

    752          0.4          72 h                  Guinea-pig  No effect at 752 µg/m3; increase in BAL protein in          Selgrade et
    1880         1.0                                            vitamin-C-depleted but not normal animals at > 1880 µg/m3.  al. (1981)
    5640         3.0
    9400         5.0

    9400         5.0          3 h                               Increased BAL protein in vitamin-C-depleted guinea-pigs
                                                                15-h post-exposure.

    752          0.4          Continuous,                       No effect on BAL protein in vitamin-C-depleted guinea-pigs.
                              1 week

    752          0.4          Continuous,           Guinea-pig  Increase in BAL protein content of guinea-pigs with an      Sherwin &
                              1 week                            unquantified vitamin C deficiency.                          Carlson (1973)
                                                                                                                                              

    Table 32  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    752          0.4          1 to 14 weeks         Rat         Complex concentration and duration dependence of effects.   Takahashi et
    2260         1.2                                            Example: at 752 µg/m3, cytochrome P-450 levels decreased    al. (1986)
    7520         4.0                                            at 2 weeks, returned to control level by 5 weeks. At
                                                                2260 µg/m3, cytochrome P-450 levels decreased initially,
                                                                increased at 5 weeks, and decreased at 10 weeks. Effects
                                                                on succinate-cytochrome c reductase also.

    752          0.4          4 months              Rat         Duration-dependent pattern for increase in activities of    Ichinose &
    2260         1.2                                            antioxidant enzymes; increase, peaking at week 4, and       Sagai (1982)
    7520         4.0                                            then decreasing; concentration-dependent effects.

    752          0.4          2 weeks               Rat         No effect on TBA reactants, antioxidants or antioxidant     Ichinose &
                                                    Guinea-pig  enzyme activities.                                          Sagai (1989)

    752          0.4          7 days                Rat         Decrease in cytochrome P-450 level at > 2260 µg/m3.         Mochitate et
    2260         1.2                                                                                                        al. (1984)
    7520         4.0

    846          0.45         7 h/day               Mouse       No changes in lung serotonin levels.                        Sherwin et
                              4 weeks                                                                                       al. (1986)

    884          0.47         Continuous, 10,       Mouse       Increased content of serum proteins in homogenized whole    Sherwin &
                              12, 14 days                       lung tissue.                                                Layfield (1974)

    940          0.5          Continuous,           Mouse       Decrease in lung GSH peroxidase activity at 1880 µg/m3      Ayaz &
    1880         1.0          17 months                         in vitamin-E-deficient mice. Increased activity in          Csallany (1978)
                                                                vitamin-E-supplemented mice at > 940 µg/m3.
                                                                                                                                              

    Table 32  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    1880         1.0          Continuous,           Rat         Activities of GSH reductase and G-6-PD increased at         Chow et al.
    4320         2.3          4 days                            11 700 µg/m3 proportional to duration of exposure; no       (1974)
    11 700       6.2                                            effect on GSH peroxidase. No effects at < 4320 µg/m3.

    1880         1.0          15 weeks              Rat         Changes in BAL fluid and lung tissue levels of enzymes      Gregory et al.
    9400         5.0                                            early in exposure; resolved by 15 weeks.                    (1983)

    3760         2.0          3 days                Rat         Decreased superoxide dismutase activity.                    Azoulay-Dupuis
    18 800       10.0                               Guinea-pig                                                              et al. (1983)

    3760         2.0          Continuous,           Rat         Increased activities of several glycolytic enzymes.         Mochitate et
    7520         4.0          7, 10, 14 days                    At < 7520 µg/m3, pyruvate kinase increased on days          al. (1985)
                                                                4 and 7; recovery occurred by day 14. G-6-PD increased
                                                                at all levels; at 3760 µg/m3, 14 days of exposure needed.

    3760         2.0          1-7 days              Rat         Increased lung protein content; increase in microsomal      Mochitate et
    7520         4.0                                            succinate cytochrome c reductase activity.                  al. (1984)
    18 800       10.0

    5640         3.0          7 days                Rat         Various changes in lung homogenate protein and DNA          Elsayed &
                                                                content and enzyme activities; changes more severe in       Mustafa (1982)
                                                                vitamin-E-deficient rats.

    5640         3.0          7 days                Rat         No effects on antioxidant metabolism or oxygen              Mustafa et al.
    9400         5.0          4 days                            consumption enzymes at < 9400 µg/m3.                        (1979)

    7520         4.0          7, 14 and             Rat         Increased gamma-glutamyl transferase on days 14 and 21;     Hooftman et al.
    18 800       10.0         21 days                           no consistent effect on alkaline phosphatase, lactate       (1988)
    47 000       25.0                                           dehydrogenase or total protein.
                                                                                                                                              

    Table 32  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    9020         4.8          3 h                   Guinea-pig  Increased BAL protein content in vitamin-C-deficient        Hatch et al.
                                                                guinea-pigs.                                                (1986)

    8460         4.5          16 h                              Increased lung wet weight, alterations in lung
                                                                antioxidant levels in vitamin-C-deficient guinea-pigs.

    9020         4.8          7 days                Mouse       No significant changes in lung homogenate parameters.       Mustafa et al.
                                                                                                                            (1984)

    9400         5.0          14-72 h               Mouse       Increase in lung protein (14 to 58 h) by radioactive        Csallany (1975)
                                                                label incorporation.

    9400-47 000  5.0-25.0     Continuous,           Rat         Concentration-related increase in rate of collagen          Last et al.
                              7 days                            synthesis; 125% increase at 9400 µg/m3.                     (1983)

    9400         5.0          3 h                   Rabbit      Benzo[a]pyrene hydroxylase activity of tracheal mucosa      Palmer et al.
    37 600       20.0                                           not affected.                                               (1972)
    94 000       50.0
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  NPSHs = Non-protein sulfhydryls; GSH = Glutathione; G-6-PD = Glucose-6-phosphate dehydrogenase; BAL = Bronchoalveolar lavage
             Alterations in lung xenobiotic metabolism follow a complex
    duration of exposure pattern in rats exposed to 752, 2260 and
    7520 µg/m3 (0.4, 1.2 and 4.0 ppm) NO2 (Takahashi et al., 1986).  At
    the lowest NO2 concentration tested, cytochrome P-450 levels
    decreased initially (at 2 weeks) and then returned to control levels
    by 5 weeks, where they remained throughout exposure.  At 2260 µg/m3
    (1.2 ppm), cytochrome P-450 levels decreased initially, then increased
    after 5 weeks of exposure and decreased again by 10 weeks.  A similar
    pattern of response occurred at the highest concentration.  Only
    7520 µg/m3 (4.0 ppm) NO2 affected other microsomal electron-
    transport systems.  The activity of succinate-cytochrome c reductase
    was decreased by 14 weeks of exposure to 752 µg/m3 (0.4 ppm), but at
    the higher NO2 levels, the activity was decreased sooner.  In
    contrast, Mochitate et al. (1984) also found a decrease in levels of
    cytochrome P-450 at > 2260 µg/m3 (1.2 ppm) in rats exposed for
    7 days.

         Glycolytic pathways are also increased by NO2 exposure,
    apparently due to a concurrent increase in Type 2 cells (Mochitate et
    al., 1985).  The most sensitive enzyme was pyruvate kinase, exhibiting
    an increased activity after a 14-day exposure to 3760 µg/m3 (2.0 ppm)
    NO2.  At higher NO2 concentrations (e.g., 7520 µg/m3, 4.0 ppm),
    pyruvate kinase activity increased sooner (4 and 7 days) and then
    decreased to control levels by 14 days.

    c)  Antioxidant defence systems

         Since NO2 is an oxidant and lipid peroxidation is believed to be
    a major molecular event responsible for the toxic effects of NO2,
    much attention has been focused on the effect of the  antioxidant
    defence system in the epithelial lining fluid and in pulmonary cells. 
    Investigations with subacute and chronic NO2 exposure levels of 75
    to 62 040 µg/m3 (0.04-33 ppm) have been performed both  in vivo
    and  in vitro and focussed on effects on low molecular weight
    antioxidants such as glutathione, vitamin E and vitamin C, as well as
    on some enzymes involved in the synthesis and catabolism of
    glutathione.  Experiments made  in vitro using human plasma have
    shown a rapid depletion of vitamin C and glutathione and a loss of
    vitamin E.  This result was achieved with a concentration of
    26 320 µg/m3 (14 ppm) (Halliwel et al., 1992).

         Menzel (1970) proposed that antioxidants might protect the lung
    from NO2 damage by inhibiting lipid peroxidation.  Data related to
    this hypothesis have been reported (Thomas et al., 1968; Menzel et
    al., 1972; Fletcher & Tappel, 1973; Csallany, 1975; Ayaz & Csallany,
    1978; Slade et al., 1989).  Several laboratories have observed changes
    in the activity of enzymes in the lungs of NO2-exposed animals that
    regulate levels of glutathione (GSH), the major water-soluble
    reductant in the lung.  Chow et al. (1974) exposed rats to 1880, 4320
    or 11 700 µg/m3 (1.0, 2.3 or 6.2 ppm) NO2 continuously for 4 days to
    examine the effect on activities of GSH reductase, glucose-6-phosphate
    dehydrogenase and GSH peroxidase in the soluble fraction of exposed
    rat lungs.  Linear regression analysis of the correlation between the
    NO2 concentration and enzymatic activity showed a significant
    positive correlation coefficient of 0.63 for GSH reductase and of 0.84
    for glucose-6-phosphate dehydrogenase.  No correlation was found
    between the GSH peroxidase activity and the NO2 concentration.  The
    activities of GSH reductase and glucose-6-phosphate dehydrogenase were
    significantly increased during exposure to 11 700 µg/m3 (6.2 ppm)
    NO2; GSH peroxidase activity was not affected.  The possible role of
    oedema and cellular inflammation in these findings was not examined. 
    These researchers concluded that after a slightly longer exposure
    (14 days), 3760 µg/m3 (2.0 ppm) NO2 increased the activity of
    glucose-6-phosphate dehydrogenase in rats (Mochitate et al., 1985). 
    There is evidence from recent studies that glutathione and vitamins C
    and E are all involved in normal protection of the lung from NO2
    (Rietjens et al., 1986; Hatch et al., 1986; Slade et al., 1989).

         Sagai et al. (1984) studied the effects of prolonged (9 and 18
    months) exposure to 75, 752 and 7520 µg/m3 (0.04, 0.4 and 4.0 ppm)
    NO2 on rats.  After both exposure durations, non-protein sulfhydryl
    levels were increased at > 752 µg/m3; exposure to 7520 µg/m3
    (4.0 ppm) decreased the activity of GSH peroxidase and increased
    glucose-6-phosphate dehydrogenase activity.  Glutathione peroxidase
    activity was also decreased in rats exposed to 752 µg/m3 NO2 for
    18 months.  Three GSH  S-transferases were also studied, two of which
    (aryl  S-transferase and aralkyl  S-transferase) exhibited decreased
    activities after 18 months of exposure to > 752 µg/m3 NO2.  No
    effects were observed on the activities of 6-phosphogluconate
    dehydrogenase, superoxide dismutase or disulfide reductase.  When
    effects were observed, they followed a concentration and exposure-
    duration response function.  The decreases in antioxidant metabolism

    were inversely related to the apparent formation of lipid peroxides
    (see lipid peroxidation subsection).  Shorter exposures (4 months) to
    NO2 between 752 and 7520 µg/m3 (0.4 and 4.0 ppm) also caused
    concentration- and duration-dependent effects on antioxidant enzyme
    activities (Ichinose & Sagai, 1982).  For example, glucose-6-phosphate
    dehydrogenase increased, reaching a peak at 1 month, and then
    decreased towards the control value.  Briefer (2-week) exposures to
    752 µg/m3 (0.4 ppm) NO2 caused no such effects in rats or
    guinea-pigs (Ichinose & Sagai, 1989).

         Ayaz & Csallany (1978) exposed vitamin-E-deficient and vitamin-E-
    supplemented mice continuously for 17 months to 940 or 1880 µg/m3
    (0.5 or 1.0 ppm) NO2 and assayed them for GSH peroxidase activity. 
    Exposure to 1880 µg/m3 (1.0 ppm) NO2 decreased enzyme activity in
    the vitamin-E-deficient mice.  However, in vitamin-E-supplemented
    mice, GSH peroxidase activity increased at 940 µg/m3 (0.5 ppm) NO2.

    5.2.2.3  Pulmonary function

         Animal studies of NO2 effects on pulmonary function are
    summarized in Table 33.  NO2 concentrations in many urban areas of
    the USA and elsewhere consist of spikes superimposed on a relatively
    constant background level.  Miller et al. (1987) evaluated this urban
    pattern of NO2 exposure in mice using continuous 7 days/week,
    23 h/day exposures to 376 µg/m3 (0.2 ppm) NO2 with twice daily
    (5 days/week) 1-h spike exposures to 1500 µg/m3 (0.8 ppm) NO2 for
    32 and 52 weeks.  Mice exposed to clean air and to the constant
    background concentration of 376 µg/m3 (0.2 ppm) served as controls.
    Vital capacity tended to be lower (p = 0.054) in mice exposed to NO2
    with diurnal spikes than in mice exposed to air.  Lung distensibility,
    measured as respiratory system compliance, also tended to be lower in
    mice exposed to diurnal spikes of NO2 compared with constant NO2
    exposure or air exposure.  These changes suggest that up to 52 weeks
    of low-level NO2 exposure with diurnal spikes may produce a subtle
    decrease in lung distensibility, although part of this loss in
    compliance may be a reflection of the reduced vital capacity.  Vital
    capacity appeared to remain suppressed for at least 30 days after
    exposure.  Lung morphology in these mice was evaluated only by light
    microscopy (a relatively insensitive method) and showed no exposure-
    related lesions.  The decrease in lung distensibility suggested by
    this study is consistent with the thickening of collagen fibrils in
    monkeys (Bils, 1976) and the increase in lung collagen synthesis rates
    of rats (Last et al., 1983) after exposure to higher levels of NO2.

         Tepper et al. (1993) exposed 60-day-old rats to 940 µg/m3
    0.5 ppm) NO2, 22 h/day, 7 days/week, with a 2-h spike of 2820 µg/m3
    (1.5 ppm) NO2, 5 days/week for up to 78 weeks. There were no effects
    on pulmonary function between 1 and 52 weeks of exposure.  Following
    78 weeks of exposure, flow at 25% forced vital capacity was decreased,
    perhaps indicating airway obstruction.  A significant decrease in the
    frequency of breathing was also observed at 78 weeks that was
    paralleled by a trend toward increased expiratory resistance and
    expiratory time.  Taken together, these results suggest that few, if
    any, significant effects were seen that suggest incipient lung
    degeneration.

         The age sensitivity to exposure to diurnal spikes of NO2 was
    studied by Stevens et al. (1988), who exposed 1-day- and 7-week-old
    rats to continuous baselines of 940, 1880 and 3760 µg/m3 (0.5, 1.0
    and 2.0 ppm) NO2 with twice daily 1-h spikes at 3 times these
    baseline concentrations for 1, 3 and 7 weeks.  In neonatal rats, vital
    capacity and respiratory system compliance increased following 3
    weeks, but not 6 weeks, of exposure to the 1880 and 3760 µg/m3 NO2
    baselines with spikes.  In young adult rats, respiratory system
    compliance decreased following 6 weeks of exposure, and body weight
    decreased following 3 and 6 weeks of exposure to the 3760 µg/m3
    baseline with spike.  In the young adult rats, pulmonary function
    changes returned to normal values 3 weeks after exposure ceased.  A
    correlated morphometric study (Chang et al., 1986) is summarized in
    section 5.2.2.4.

         Lafuma et al. (1987) exposed 12-week-old hamsters with and
    without laboratory-induced (elastase) emphysema to 3760 µg/m3
    (2.0 ppm) NO2, 8 h/day, 5 days/week for 8 weeks.  Vital capacity
    and pulmonary compliance were not affected by NO2 exposure.

    5.2.2.4  Morphological studies

         Inhalation of NO2 produces morphological alterations in the
    respiratory tract, as summarized in Tables 34 and 35.  This
    discussion is generally limited to those studies using NO2 levels
    < 9400 µg/m3 (5.0 ppm), but results of studies of emphysema
    conducted at higher concentrations are also discussed. Examination of
    the tables shows variability in responses at similar exposure levels
    in different studies.  This may be due to differences in animal
    species or strain, age, diet, microbiological status of the animals,

    or aspects of experimental protocol.  The latter includes the
    methodology used to evaluate the morphological response.  For example,
    simple light microscopic examination may reveal no effect, whereas
    other techniques, such as quantitative morphological (morphometric)
    procedures with electron microscopy, can detect more subtle structural
    changes.

         There is a large degree of interspecies variability in
    responsiveness to NO2; this is clearly evident from those few studies
    where different species were exposed under identical conditions
    (Wagner et al., 1965; Furiosi et al., 1973; Azoulay-Dupuis et al.,
    1983).  Variability in response may be due to differences in effective
    dose of NO2 reaching target sites, but other species differences are
    likely to play a role.  Guinea-pigs, hamsters and monkeys all appear
    to be more severely affected morphologically by equivalent exposure to
    NO2 than are rats, the most commonly used experimental animal.
    However, in most cases, similar types of histological lesions are
    produced when similar effective concentrations are used.

    a)  Sites affected and time course of effects

         The anatomic region most sensitive to NO2 and within which
    injury is first noted is the centriacinar region.  This region
    includes the terminal conducting airways (terminal bronchioles),
    respiratory bronchioles, and adjacent alveolar ducts and alveoli. 
    Within this region, those cells that are most sensitive to
    NO2-induced injury are the ciliated cells of the bronchiolar
    epithelium and the Type 1 cells of the alveolar epithelium, which are
    then replaced with non-ciliated bronchiolar (Clara) cells and Type
    2 cells, respectively.  In addition to these dynamic changes,
    permanent alterations resembling emphysema-like disease may result
    from chronic exposure.

         The temporal progression of early events due to NO2 exposure has
    been described best in the rat (e.g., Freeman et al., 1966, 1968c,
    1972; Stephens et al., 1971a, 1972; Evans et al., 1972, 1973a,b, 1974,
    1975, 1976, 1977; Cabral-Anderson et al., 1977; Rombout et al., 1986)
    and guinea-pig (Sherwin et al., 1973).  The earliest alterations
    resulting from exposure to concentrations of > 3760 µg/m3
    (2.0 ppm) are seen within 24 to 72 h of exposure and include increased
    AM aggregation, desquamation of Type 1 cells and ciliated bronchiolar
    cells, and accumulation of fibrin in small airways.  However, repair
    of injured tissue and replacement of destroyed cells can begin within
    24 to 48 h of continuous exposure.  Hyperplasia of nonciliated
    bronchiolar (Clara) cells occurs in the bronchioli, whereas in the
    alveoli, the damaged Type 1 cells are replaced with Type 2 cells. 
    These new cells are relatively resistant to effects of continued NO2
    exposure.

        Table 33.  Effects of nitrogen dioxide (NO2) on pulmonary functiona
                                                                                                                                              
    NO2 Concentration
                      
    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    376          0.2          23 h/day base         Mouse       Decreased vital capacity following base + spike             Miller et
                              (7 days/week), 1-h                NO2 exposures compared with control and base NO2            al. (1987)
    376 base,    0.2 base,    peaks twice/day,                  exposures. Tendency toward decreased respiratory
    1500 peak    0.8 peak     32 and 52 weeks                   system compliance following spike NO2 exposures
                                                                compared and control and base NO2 exposures.

    940 base,    0.5 base,    23 h/day              Rat (1-day  Increased lung volume and compliance in neonates            Stevens et
    2820 peak    1.5 peak     (7 days/week) base,   and         following 3-week, but not 6-week, exposure to the           al. (1988)
                              1-h peaks twice/day   7-weeks     two higher exposure levels. Decreased body weight
    1880 base,   1.0 base,    (5 days/week);        old)        and lung compliance in adult rats following 6-week
    5640 peak    3.0 peak     1, 3 and 6 weeks                  exposure to 3760 µg/m3 + spike. Adults recovered
                                                                3 weeks after exposure.
    3760 base,   2.0 base,
    11 300 peak  6.0 peak

    940 base,    0.5 base,    22 h/day (7 days per  Rat         Decreased delta FEF25 and frequency of breathing            Tepper et al.
    2820 peak    1.5 peak     week), 2-h peak                   following 78-week NO2 exposure.                             (1993)
                              (5 days/week); 1,
                              3, 12, 52 and
                              78 weeks

    3760         2.0          8 h/day,              Hamster     No change in vital capacity or lung compliance              Lafuma et al.
                              5 days/week,                      following NO2 exposures in both normal and                  (1987)
                              8 weeks                           elastase-treated animals.

    10 200       5.4          3 h/day for 7, 14     Rat         Tendency toward increased lung volume at low                Yokoyama et
                              or 30 days                        inflation pressures.                                        al. (1980)
                                                                                                                                              
    a  Modified from: US EPA (1993)
    b  PaO2 = Arterial oxygen tension; delta FEF25 = Change in forced expiratory flow at 25% of forced vital capacity;
       PaCO2 = Arterial carbon dioxide tension

    Table 34.  Effects of acute and subchronic exposure to nitrogen dioxide (NO2) on lung morphologya
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    207          0.11         Continuous,           Rat (1,     Various morphometric changes, depending on age              Kyono & Kawai
    865          0.46         1 month               3, 12,      and exposure level. Multiphasic pattern (e.g.,              (1982)
    5260         2.8                                21 months   decrease in air-blood barrier thickness from 1 to
    16 500       8.8                                old)        12 months of age, and increase in 21-month-old
                                                                rats).

    639          0.34         6 h/day, 5 days       Mouse       Type 2 cell hypertrophy and hyperplasia; increase           Sherwin &
                              per week, 6 weeks                 in mean linear intercept and amount of alveolar             Richters (1982)
                                                                wall area.

    940          0.5          4 h                   Rat         Loss of cytoplasic granules in and rupture of               Thomas et al.
                                                                mast cells.                                                 (1967)

    940          0.5          Continuous, up        Rat         Increased number of mast cells in trachea as exposure       Hayashi et al.
                              to 6 days                         duration increased.                                         (1987)
                                                                                                                                              

    Table 34  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    940 base,    0.5 base,    23 h/day (7 days      Rat (1 day  In proximal alveolar region: base (940 µg/m3) + peak        Crapo et al.
    2820 peak    1.5 peak     per week) base, 1-h   and         caused Type 2 cells to become spread over more              (1984); Chang
                              peaks twice/day       6 weeks     surface area in neonates and adults; Type 2 cell            et al. (1986,
    3760 base,   2.0 base,    (5 days/week);        old)        hypertrophy and increase in number of AMs in adults;        1988)
    11 280 peak  6.0 peak     6 weeks                           Type 2 cells thinner in neonates. Base (3760 µg/m3)
                                                                + peak (only adults studied) caused similar changes
                                                                plus an increase in numbers of Type 1 cells, which
                                                                were smaller than normal Type 1 cells.
                                                                In terminal bronchiolar region: base (940 µg/m3) +
                                                                peak caused no effects on percentage distribution of
                                                                ciliated cells and Clara cells in neonates or adults,
                                                                but neonates (only) had a increase in ciliated cell
                                                                surface area and mean luminal surface area of Clara
                                                                cells. Base (3760 µg/m3) + peak (only adults studied)
                                                                had fewer ciliated cells with a reduced surface area
                                                                and alterations in the shape of Clara cells.

    1000         0.53         Continuous            Rat         At < 2500 µg/m3: no pathology. At 5000 µg/m3: focal         Rombout et al.
    2500         1.33         (24 h/day)                        thickening of centriacinar septa by 2 days; progressive     (1986)
    5000         2.66         28 days                           loss of cilia and abnormal cilia in trachea and main
                                                                bronchi at > 4 days; hypertrophy of bronchiolar
                                                                epithelium at > 8 days.
                                                                At days 16 and 28, all epithelial cells hypertrophied.
                                                                                                                                              

    Table 34  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    1000         0.53         24 h/day,             Guinea-     No pathology                                                Steadman et al.
                              90 days               pig,                                                                    (1966)
                                                    rabbit,
                                                    dog,
                                                    monkey,
                                                    rat

    1320-1500    0.7-0.8      Continuous,           Mouse       Mucous hypersecretion; focal degeneration and               Nakajima et al.
                              1 month                           desquamation of mucous membrane; terminal                   (1980)
                                                                bronchiolar epithelial hyperplasia; some alveolar
                                                                enlargement; shortening of cilia.

    1880         1-1.5        Continuous,           Mouse       Terminal bronchiolar epithelial hyperplasia; some           Nakajima et al.
    2820                      1 month                           alveolar enlargement.                                       (1980)

    1880         1.0          1 h                   Rat         Degranulation and decreased number of mast cells.           Thomas et al.
                                                                                                                            (1967)

    3760         2.0          3 days                Rat         No historical changes                                       Azoulay-Dupuis
                                                                                                                            et al. (1983)

    3760         2.0          3 days                Guinea-pig  Thickening of alveolar walls; oedema; increase in           Azoulay-Dupuis
                                                                AM numbers; loss of bronchiolar cilia; inflammation.        et al. (1983)
                                                                                                                                              

    Table 34  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    3760         2.0          8 h/day,              Hamster     Moderate alveolar enlargement, primarily at                 Lafuma et al.
                              5 days/week,                      bronchiolar-alveolar duct junction; increase in mean        (1987)
                              8 weeks                           linear intercept; decrease internal surface area of
                                                                lung; no lesions in bronchial, bronchiolar, alveolar
                                                                duct, or alveolar epithelium; no change in
                                                                macrophage number.

    3760         2.0          Continuous,           Guinea-pig  Type 2 cell hypertrophy at 7 or 21 days.                    Sherwin et al.
                              7-21 days                                                                                     (1973)

    3760         2.0          Continuous,           Guinea-pig  Increase in number of LDH-positive cells with time          Sherwin et al.
                              1-3 weeks                         of exposure. Correlated to increase in Type 2 cells         (1973)
                                                                (LDH positive).

    3760         2.0          Continuous,           Rat         Minimal effect: some cilia loss in terminal bronchioles;    Azoulay et al.
                              6 weeks                           some distended or disrupted alveolar walls.                 (1978)

    9400         5.0          Continuous,           Cynomolgus  Bronchiolar epithelia hyperplasia; some focal               Busey et al.
    18 800       10.0         90 days               monkey      pulmonary odema.                                            (1974)
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  AMs = Alveolar macrophages; LDH = Lactate dehydrogenase

    Table 35.  Effects of chronic exposure to nitrogen dioxide (NO2) on lung morphologya
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    75           0.04         Continuous,           Rat         At 75 µg/m3: no significant change, but some tendency       Kubota et al.
    752          0.4          9-27 months                       towards increase in arithmetic mean thickness of air-blood  (1987)
    7520         4.0                                            barrier.  At 752 µg/m3: slight increase in arithmetic
                                                                mean thickness of air-blood barrier by 18 months, becoming
                                                                significant by 27 months; some interstitial oedema and
                                                                slight change in bronchiolar and alveolar epithelium by
                                                                27 months.  At 7520 µg/m3: hypertrophy and hyperplasia of
                                                                bronchiolar epithelium and increase in arithmetic mean
                                                                thickness of air-blood barrier by 9 months, which became
                                                                significant at 18 months and decreased slightly by
                                                                27 months; Clara cell hyperplasia. By 27 months:
                                                                interstitial fibrosis and hypertrophy of Type 1 and
                                                                Type 2 cells.

    188 base;    0.1 base;    Continuous            Mouse       Dilated airspaces and aveolar wall destruction (small       Port et al.
    1880 peak    1.0 peak     baseline; 2-h                     sample size).                                               (1977)
                              daily peak;
                              6 months
                                                                                                                                              

    Table 35  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    940          0.5          Continuous,           Rat         At 940 µg/m3: swelling of terminal bronchiolar cilia and    Yamamoto &
    1880         1.0          7 months                          hyperplasia of Type 2 cells.                                Takahashi
    7520         4.0                                            At 1880 µg/m3: cilia loss in terminal bronchioles;          (1984)
                                                                hyperplasia of Type 2 cells; and interstitial oedema.
                                                                At 7520 µg/m3: cilia loss in terminal bronchioles;
                                                                hyperplasia of Type 2 cells, interstitial oedema; decrease
                                                                in number of lamellar bodies in Type 2 cells; lysosomes
                                                                with osmiophilic lamellar structure in ciliated cells of
                                                                terminal bronchioles.

    940          0.5          Continuous, up        Rat         Type 2 cell hypertrophy and interstitial oedema by          Hayashi et al.
                              to 19 months                      4 months; increased thickness of alveolar septa by          (1987)
                                                                6 months; fibrous pleural thickening by 19 months.

    940          0.5          6-24 h/day,           Mouse       3 months: pneumonitis and alveolar size increase; loss of   Blair et al.
                              3-12 months                       cilia in respiratory bronchioles and bronchiolar            (1969)
                                                                inflammation with 24 h/day.
                                                                6-12 months: pneumonitis; cilia loss; bronchial and
                                                                bronchiolar inflammation; alveolar size increase.

    1500         0.8          Continuous,           Rat         Minimal changes: slight enlargement of alveoli and          Freeman et
                              lifetime (up                      alveolar ducts; some rounding of bronchial and bronchiolar  al. (1966)
                              to 33 months)                     epithelial cells; increase in elastic fibers around
                                                                alveolar ducts.

    1880         1.0          Continuous,           Squirrel    No pathology                                                Fenters et
                              16 months             monkey                                                                  al. (1973)
                                                                                                                                              

    Table 35  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    1880         1.0          6 h/day,              Dog         At 1880 µg/m3 - 6 months: no pathology; 12 months:          Wagner et
                              5 days/week, up                   dilated alveoli and alveolar ducts; 18 months:              al. (1965)
                              to 18 months                      dilated alveoli, oedema, thickening alveolar septa
                                                                due to inflammation.
    9400         5.0                                            At 9400 µg/m3 - 6 months: no pathology; 12 months:
                                                                dilated alveolar ducts; 18 months: oedema, congestion,
                                                                and thickened alveolar septa due to inflammatory cells.

    1880         1.0          6 h/day               Guinea-pig  Mild thickening of alveolar septa due to inflammation;      Wagner et
                              5 days/week,                      some alveolar dilatation.                                   al. (1965)
                              18 months

    1880         1.0          7 h/day,              Rat         No pathology                                                Gregory et
                              5 days/week,                                                                                  al. (1983)
                              15 weeks

    3760         2.0          Continuous,           Rat         Loss of cilia in terminal bronchioles; abnormal             Stephens et
                              2 years                           ciliogenesis; crystalloid inclusions in bronchiolar         al. (1971a,b)
                                                                epithelial cells; increased thickness of collagen fibrils
                                                                and basement membrane in terminal bronchioles.

    3760         2.0          Continuous, up        Rat         Hypertrophy of ciliated cells and cilia loss by 72 h;       Stephens et
                              to 12 months                      decreased number of ciliated cells by 7 days; normal        al. (1972)
                                                                ciliated cells from 21 days-12 months.

    3760         2.0          Continuous, up        Rat         No change in turnover of terminal bronchiolar epithelial    Evans et al.
                              to 360 days                       cells; increase in turnover of Type 2 cells in peripheral   (1972)
                                                                alveoli by 1 day, but normal by 7 days.
                                                                                                                                              

    Table 35  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    3760         2.0          Continuous,           Monkey      Bronchiolar epithelial hypertrophy, especially adjacent     Furiosi et al.
                              14 months             (Macaca     to alveolar ducts; change to cuboidal cells in proximal     (1973)
                                                    peciosa)    bronchiolar epithelium.

    3760         2.0          Continuous,           Rat         Minimal effect: some terminal bronchiolar epithelial        Furiosi et al.
                              14 months                         hypertrophy.                                                (1973)

    3760         2.0          Continuous,           Rat         Alveolar distension, especially near alveolar duct level;   Freeman et
                              lifetime (up to                   increased variability in alveolar size; loss of cilia and   al. (1968b)
                              763 days); 1500                   hypertrophy in terminal bronchiolar cells; no
                              µg/m3 for 1st                     inflammation.
                              69 days, then
                              3760 µg/m3

    7520         4.0          Continuous,           Rat         Bronchial epithelial hyperplasia                            Haydon et al.
                              16 months                                                                                     (1965)

    9400         5.0          6 h/day,              Mouse       No pathology                                                Wagner et al.
                              5 days/week,                                                                                  (1965)
                              14 months

    9400         5.0          4-7.5 h/day,          Guinea-pig  Some dilatation of terminal bronchioles; tracheal           Balchum et al.
                              5 days/week,                      inflammation; pneumonitis.                                  (1965)
                              5.5 months
                                                                                                                                              

    Table 35  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    9400         5.0          7 h/day,              Rat         Focal hyperinflation and areas of subpleural accumulation   Gregory et al.
                              5 days/week,                      of macrophages.                                             (1983)
                              15 weeks
                                                                                                                                              

    a  Modified from US EPA (1993)
             The time course of alveolar lesions over a chronic exposure was
    examined by Kubota et al. (1987) in small groups of rats exposed to
    7520 µg/m3 (4.0 ppm) NO2, 24 h/day for up to 27 months.  One phase,
    which lasted for 9 to 18 months of exposure, consisted of a decrease
    in number and an increase in cell volume of Type 1 epithelium, an
    increase in the relative ratio of Type 2 to Type 1 cells, and an
    increase in the number and volume of Type 2 cells.  A second phase,
    between 18 to 27 months of exposure, showed some recovery of the
    alveolar epithelium, but the total volume of interstitial tissue
    decreased and collagen fibres in the interstitium increased.  Thus,
    some lesions resolved with continued exposure, whereas others
    progressed.  At 752 µg/m3 (0.4 ppm), Kubota et al. (1987) found that
    the lesion typically was milder and its initiation delayed, compared
    to the higher concentration.  In general, most NO2-induced lesions
    were resolved following a recovery period.  This period may be as
    short as 30 days for exposures at < 9400 µg/m3 (5.0 ppm).  With
    continuous exposure, early morphological damage may also be resolved. 
    For example, in rats exposed continuously for 7 months to 940 µg/m3
    (0.5 ppm) NO2, resolution of epithelial lesions occurred by 4 to
    6 months of exposure (Yamamoto & Takahashi, 1984).

    b)  Effects of nitrogen dioxide as a function of exposure pattern

         Several morphological studies were designed to evaluate ambient
    NO2 patterns consisting of a low baseline level with transient spikes
    of NO2.  However, in some cases, there was no group at the baseline
    exposure only, preventing evaluation of the contribution of peaks to
    the responses.  Gregory et al. (1983) exposed rats (14 to 16 weeks
    old) for 7 h/day, 5 days/week for up to 15 weeks to atmospheres
    consisting of the following concentrations of NO2: (1) 1880 µg/m3
    (1.0 ppm), (2) 9400 µg/m3 (5.0 ppm), or (3) 1880 µg/m3 (1.0 ppm)
    with two 1.5-h spikes of 9400 µg/m3 (5.0 ppm) per day.  After
    15 weeks of exposure, histopathology was minimal, with focal
    hyperinflation and areas of subpleural accumulation of macrophages
    found in some of the animals exposed either to the baseline of
    9400 µg/m3 (5.0 ppm) or to 1880 µg/m3 (1.0 ppm) with the 9400 µg/m3
    (5.0 ppm) spikes.

         Port et al. (1977) observed dilated respiratory bronchioles and
    alveolar ducts in mice exposed to 188 µg/m3 (0.1 ppm) NO2 with daily
    2-h peaks to 1880 µg/m3 (1.0 ppm), for 6 months.  Miller et al.
    (1987) found no morphological effects in mice exposed for 1 year,
    although host defence and functional changes were noted (see sections
    5.2.2.1 and 5.2.2.3).

         Crapo et al. (1984) and Chang et al. (1986) used quantitative
    morphometric analyses to examine the proximal alveolar and terminal
    bronchiolar regions of rats exposed for 6 weeks to a baseline
    concentration of 940 or 3760 µg/m3 (0.5 or 2.0 ppm) NO2, 23 h/day
    for 7 days/week, onto which were superimposed two daily 30 min spikes
    of 3 times the baseline concentration for 5 days/week.  At the lower
    exposure level, the volumes of the Type 2 epithelium, interstitial
    matrix, and AMs increased, whereas the volume of the fibroblasts
    decreased.  The surface area of Type 2 cells increased.  Most of these
    changes also occurred at the higher exposure level, and in some cases
    the change was greater than that at the lower level (i.e., increase in
    Type 1 and Type 2 epithelial volume).  At both levels of exposure, the
    volume of Type 2 cells and interstitial fibroblasts increased, with no
    significant changes in their numbers, and the number of AMs decreased. 
    The number of Type 1 cells decreased and their average surface area
    increased in the highest exposure group.  Generally, there was a
    spreading and hypertrophy of Type 2 cells.  A correlation between
    decreased compliance (Stevens et al., 1988) and thickening of the
    alveolar interstitium was found (see section 5.2.2.3 for details of
    the pulmonary function portion of the study).  Examination of the
    terminal bronchiolar region revealed no effects at the lower exposure
    level.  At the higher level, there was a 19% decrease in ciliated
    cells per unit area of the epithelial basement membrane and a
    reduction in the mean ciliated surface area.  The size of the dome
    protrusions of non-ciliated bronchiolar (Clara) cells was decreased,
    giving the bronchial epithelium a flattened appearance, but there was
    no change in the number of cells.

    c)  Factors affecting susceptibility to morphological changes

         Age-related responsiveness to an urban pattern of NO2 was
    evaluated by Chang et al. (1986, 1988) using 1-day- or 6-week-old rats
    exposed for 6 weeks to a baseline of 940 µg/m3 (0.5 ppm) NO2 for
    23 h/day, 7 days/week, with two 1-h spikes (given in the morning and
    afternoon) of 2820 µg/m3 (1.5 ppm) 5 days/week.  Electron microscopic
    morphometric procedures were used.  In the proximal alveolar region,
    only the older animals showed an increase in the surface density of
    the alveolar basement membrane.  The increase in the mean cellular
    volume of Type 2 cells was greater in the young adult animals,
    although the neonates also exhibited this effect.  Although there was
    no qualitative evidence of morphological injury in the terminal
    bronchioles of the neonatal rats, there was a 19% increase in the
    average ciliated cell surface and a 13% increase of the mean luminal

    surface area of non-ciliated bronchiolar (Clara) cells that was not
    evident in the young adult rats.  Generally, the neonatal rats were as
    sensitive or more susceptible than young adults, depending upon the
    end-point.  However, the terminal bronchioles of the neonatal rats
    were more susceptible than those of young adults (Chang et al.,
    1988).  For example, the lower exposure altered ciliated cells and
    non-ciliated bronchiolar (Clara) cells in the neonates but not the
    young adults.  Other indices were unaffected.  Pulmonary function was
    also altered in similarly exposed rats (Stevens et al., 1988) (see
    section 5.2.2.3).  Interpretation of the neonatal effects is
    difficult.  Assuming that rats prior to weaning are more resistant to
    NO2 (Stephens et al., 1978) (see below), effects observed after a
    6-week exposure from birth may have resulted from the last 3 weeks of
    exposure, as the first 3 weeks may constitute a more resistant period. 
    In contrast, effects observed in young adults probably reflect the
    impact of the entire 6-week exposure.

         In one of the more extensive studies, Kyono & Kawai (1982)
    exposed rats at 1, 3, 12, and 21 months of age continuously for
    1 month to 207 µg/m3, 865 µg/m3, 5260 µg/m3 or 16 500 µg/m3
    (0.11, 0.46, 2.8 or 8.8 ppm) NO2.  Various morphometric parameters
    were assessed, including arithmetic mean thickness of the air-blood
    barrier and the volume density of various alveolar wall components. 
    Quantitative estimations deliberately excluded the site of main damage
    (i.e., the peripheral alveolar wall was examined).  Analysis of
    individual results was complex, but depending upon the animal's age
    and the specified end-point, exposure levels as low as 207 µg/m3
    (0.11 ppm) changed specific morphometric parameters.  There was a
    trend towards a concentration-dependent increase in air-blood barrier
    thickness in all age groups, with evidence of age-related differences
    in response.  At any concentration, the response of this end-point
    decreased in rats from 1 to 12 months old, but increased again in
    21-month-old animals.  Type 1 and 2 cells showed various degrees of
    response, depending on both age at onset of exposure and exposure
    concentration.  The response of each lung component did not always
    show a simple concentration-dependent increase or decrease, but
    suggested a multiphasic reaction pattern.

         The above studies with rats may not have used the most
    susceptible animal model, as demonstrated by Azoulay-Dupuis et al.
    (1983), who exposed both rats and guinea-pigs aged 5 to > 60 days
    old to 3760 (2.0 ppm) for 3 days.  Rats at all ages and guinea-pigs
    < 45 days old were not affected.  The 45-day-old guinea-pigs showed
    thickening of alveolar walls, alveolar oedema, and inflammation,
    whereas animals older than 45 days showed similar, but more frequent,
    alterations that seemed to increase with age.  Adults also had focal
    loss of cilia in bronchioli.

         In general, it appears that neonates, prior to weaning, are
    relatively resistant to NO2, and that responsiveness then increases
    (Stephens et al., 1978).  Furthermore, the responsiveness of mature
    animals appears to decline somewhat with age, until an increase in
    responsiveness occurs at some point in senescence.  However, the
    morphological response to NO2 in animals of different ages involves
    similarities in the cell types affected and in the nature of the
    damage incurred.  Age-related differences occur in the extent of
    damage and in the time required for repair, the latter taking longer
    in older animals.  The reasons for age differences in susceptibility
    are not known, but may involve differences in doses to the target
    cells and variable sensitivity of target cells during different growth
    phases.

         The database regarding the effects of levels of NO2
    < 9400 µg/m3 (5.0 ppm) on animals with pre-existing respiratory
    disease is very limited and only includes animals with laboratory-
    induced emphysema or infections.  Lafuma et al. (1987) exposed both
    normal and elastase-induced emphysematous hamsters (2 months old) to
    3760 µg/m3 (2.0 ppm) NO2 for 8 h/day, 5 days/week, for 8 weeks. 
    Morphometric analyses indicated that emphysematous lesions were
    exacerbated by NO2 (i.e., NO2 increased pulmonary volume and
    decreased internal alveolar surface area).  The investigators
    suggested that these results may imply a role for NO2 in enhancing
    pre-existing emphysema.  Acute infectious (influenza) lung disease
    enhanced the morphological effects of NO2 in squirrel monkeys
    exposed continuously to 1880 µg/m3 (1.0 ppm) NO2 for 16 months
    (Fenters et al., 1973).

    d)  Emphysema following nitrogen dioxide exposure

         Numerous investigators have observed morphological lesions that
    led them to the diagnosis of NO2-induced emphysema.  However, to
    evaluate these reports independently, it is necessary to apply the
    current definition of emphysema, especially because the definition
    changed after several of the reports were published.  Such an
    evaluation is described in detail by the US EPA (1993), based upon the
    most recent definition of emphysema from the report of the US National
    Heart, Lung and Blood Institute (NHLBI), Division of Lung Diseases
    Workshop (National Institutes of Health, 1985).  According to this
    document, in human lungs: "Emphysema is defined as a condition of the
    lung characterized by abnormal, permanent enlargement of airspaces
    distal to the terminal bronchiole, accompanied by destruction of their
    walls, and without obvious fibrosis".  Destruction in emphysema is
    further defined as "non-uniformity in the pattern of respiratory
    airspace enlargement so that the orderly appearance of the acinus and
    its components is disturbed and may be lost".  The report further
    indicates: "Destruction...may be recognized by subgross examination of
    an inflation-fixed lung slice...".  However, emphysema in animal

    models was defined differently.  An animal model of emphysema is
    defined as "an abnormal state of the lungs in which there is
    enlargement of the airspaces distal to the terminal bronchiole. 
    Airspace enlargement should be determined qualitatively in appropriate
    specimens and quantitatively by stereologic methods".  Thus, in animal
    models of emphysema, airspace wall destruction need not be present. 
    "Appropriate specimens" presumably refers to lungs fixed in the
    inflated state.   When reports of emphysema following NO2 exposures
    of animals are to be extrapolated to potential hazards for humans, the
    definition of human emphysema, rather than that for emphysema in
    experimental animals, should be used.  The presence of airspace wall
    destruction, critical to the definition of human emphysema, can only
    be determined independently in published reports by careful review of
    the authors' description of the lesions or by examining the
    micrographs that the author selected for publication.  Because
    descriptions in some reports are inadequate for independent
    evaluation,  more evidence may exist for emphysema than is summarized
    here.  All reports reviewed are summarized in Table 36, but only those
    showing emphysema of the type seen in human lungs are discussed in the
    text that follows.

         Haydon et al. (1967) reported emphysema in rabbits exposed
    continuously (presumably 24 h/day) for 3 to 4 months to 15 000 or
    22 600 µg/m3 (8.0 or 12.0 ppm) NO2.  They reported enlarged lungs
    that failed to collapse when the thorax was opened.  The lungs were
    fixed in an expanded state via the trachea.  In 100-µm thick sections
    from formaldehyde-fixed dried lungs they reported "dilated" airspaces
    with "distorted architecture."  In those and other tissue
    preparations, they reported that the airspaces appeared "grossly
    enlarged and irregular, which appears to be due to disrupted alveoli
    ... and the absence of adjacent alveolar collapse."  Thus, in
    appropriately fixed lungs, they reported evidence of enlarged
    airspaces with destructive changes in alveolar walls.  Although no
    stereology was performed, this appears to be emphysema of the type
    seen in human lungs.

         Freeman et al. (1972) exposed rats to 37 600 µg/m3 (20.0 ppm)
    NO2, which was reduced during the exposure to 28 200 µg/m3
    (15.0 ppm) or to 18 800 µg/m3 (10.0 ppm), for varying periods up to
    33 months. Following removal at necropsy, the lungs were fixed via the
    trachea at 25 cm of fixative pressure. Morphometry of lung and
    alveolar size was performed in a suitable, although unconventional,
    manner. The morphometry indicated enlargement of alveoli and reduction
    in alveolar surface area.  The authors also both reported alveolar
    destruction and illustrated alveolar destruction in their figures. 
    They correctly concluded that they had demonstrated emphysema in their
    NO2-exposed rats.  However, it is not entirely clear whether both
    experimental groups or only the group exposed to 28 200 µg/m3
    (15.0 ppm) had emphysema.

        Table 36.  Effects of nitrogen dioxide (NO2) on the development of emphysemaa
                                                                                                                                              

    NO2 concentration
                                              

    µg/m3                         ppm                  Exposure                           Species         Emphysemab    Reference
                                                                                                                                              

    188 with 2-h peaks to 1880    0.1 with             Daily, 6 months                    Mouse           ±             Port et al. (1977)
                                  peaks to 1.0

    263 plus 2050 µg/m3 NO        0.14                 16 h/day, 68 months                Beagle dog      -             Hyde et al. (1978)
    1200 plus 310 µg/m3 NO        0.64                                                                    +

    940                           0.5                  6, 18 or 24 h/day, 1-12 months     Mouse           -             Blair et al. (1969)

    1500                          0.8                  51-813 days                        Rat             -             Haydon et al. (1965)
    7520                          4.0

    1880 (with and without viral  1.0                  16 months                          Squirrel        ±             Ehrlich & Fenters (1973)
    challenge)                                                                            monkey

    3760                          2.0                  Continuous, 112-763 days           Rat             -             Freeman et al. (1968c)

    3760                          2.0                  8 h/day, 5 days/week               Hamster         -             Lafuma et al. (1987)
                                                       for 8 weeks

    9400                          5.0                  3 months                           Squirrel        ±             Ehrlich & Fenters (1973)
    18 800                        10.0                                                    monkey
                                                                                                                                              

    Table 36  (Con't)
                                                                                                                                              

    NO2 concentration
                                              

    µg/m3                         ppm                  Exposure                           Species         Emphysemab    Reference
                                                                                                                                              

    9400                          5.0                  Up to 18 months                    Dog,            -             Wagner et al. (1965)
                                                                                          rabbit,
                                                                                          guinea-pig,
                                                                                          rat, hamster,
                                                                                          mouse

    15 000                        8.0                  3-4 months (presumably             Rabbit          +             Haydon et al. (1967)
    22 560                        12.0                 24 h/day)

    28 200                        15.0                 3-5 months                         Rat             -             Stephens et al. (1976)

    28 200                        15.0                 Continuously for 35 days then      Rat             ±             Port et al. (1977)
                                                       intermittently for at least
                                                       2 years

    33 800                        18.0                 24 h/day for 1-6 days or           Rat             ±             Freeman et al. (1968a)
                                                       4 weeks

    37 600 reduced to either      20.0 reduced to      Up to 33 months                    Rat             +             Freeman et al. (1972)
    28 200 or 18 800              15.0 or 10.0

    47 000                        25.0                 32-65 days                         Rat             -             Freeman & Haydon (1964)

    56 400                        30.0                 22 h/day, 12 months                Hamster         -             Kleinerman et al. (1985)

    56 400                        30.0                 Continuous, up to 140 days         Rat             ±             Glasgow et al. (1987)

    56 400                        30.0                 Continuous, up to 8 weeks          Rat             -             Blank et al. (1978)
                                                                                                                                              

    Table 36  (Con't)
                                                                                                                                              

    NO2 concentration
                                              

    µg/m3                         ppm                  Exposure                           Species         Emphysemab    Reference
                                                                                                                                              

    56 400 to 65 800              30.0-35.0            23 h/day for 7 days                Hamster         -             Lam et al. (1983)

    65 800                        35.0                 6 h/day for 25 days                Rat             -             Stavert et al. (1986)

    75 200                        40.0                 6 or 8 weeks                       Mouse           -             Buckley & Loosli (1969)

    94 000 to 169 200 for         50-90 reduced        2 h/day, 5 days/week,              Hamster,        ±             Gross et al. (1968)
    4 weeks, reduced to 56 400    to 30-50             12 months                          guinea-pig
    to 94 000

    84 600 to 103 400             45-55                22-23 h/day, 10 weeks              Hamster         -             Kleinerman & Cowdrey
                                                                                                                        (1968)
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  + = emphysema; - = no emphysema; ± = equivocal
       Emphysema is defined according to the 1985 US National Heart, Lung, and Blood Institute Workshop criteria for human emphysema.
       Although many of the papers reviewed (US EPA, 1993) reported finding emphysema, some of these studies were reported according
       to previous, different criteria; some reports did not fully describe the methods used; and/or the results obtained were not in
       sufficient detail to allow independent confirmation of the presence of emphysema.  Thus, a "-" (i.e. no emphysema) should only be
       interpreted as lack of proof of emphysema, because it is conceivable that if the study were repeated with current methods and
       the current criteria applied, it might be judged to be positive.
             Hyde et al. (1978) studied beagle dogs that had been exposed 16 h
    daily for 68 months to either filtered air or to 1200 µg/m3
    (0.64 ppm) NO2 with 310 µg/m3 (0.25 ppm) NO or to 263 µg/m3
    (0.14 ppm) NO2 with 2050 µg/m3 (1.67 ppm) NO.  The dogs then
    breathed clean air during a 32- to 36-month post-exposure period.  The
    right lungs were fixed via the trachea at 30-cm fixative pressure
    in a distended state.  Semiautomated image analysis was used for
    morphometry of air spaces.  The dogs exposed to 1200 µg/m3 NO2 with
    310 µg/m3 NO had significantly larger lungs with enlarged air spaces
    and evidence of destruction of alveolar walls.  These effects were not
    observed in dogs exposed to 270 µg/m3 NO2 with 2050 µg/m3 NO,
    implying a significant role of the NO2 in the production of the
    lesions.  The lesions in dogs exposed to the higher NO2 concentration
    meet the criteria of the 1985 NHLBI workshop for emphysema of the type
    seen in human lungs.

    5.2.3  Genotoxicity, potential carcinogenic or co-carcinogenic effects

         NO2 forms nitrous and nitric acids in aqueous solutions, which
    are in equilibrium with the nitrite (NO2-) and nitrate (NO3-) ions
    that constitute the main metabolites of NO2.  Nitrous acid/NO2- is
    mutagenic  in vitro, causing deamination of bases in DNA.  The
    formation of  N-nitroso compounds from secondary amines and amides is
    another mechanism for indirect mutagenic activity (Zimmermann, 1977).

          In vitro studies with NO2 have demonstrated mutations in
    bacteria (Salmonella strain TA100) (Isomura et al., 1984; Victorin &
    Stahlberg, 1988) but not in a mammalian cell culture (Isomura et al.,
    1984).  Other experiments using cell cultures were positive concerning
    chromatid-type chromosome abberations, sister chromatid exchanges
    (SCE) and DNA single strand breaks (Tsuda et al., 1981; Shiraishi &
    Bandow, 1985; Gorsdorf et al., 1990).

         NO2 did not induce recessive lethal mutations or somatic
    mutations in Drosophila (Inoue et al., 1981; Victorin et al., 1990)
    and was negative in  in vivo studies with mice concerning chromosome
    abberations in peripheral lymphocytes or spermatocytes (Gooch et al.,
    1977) and micronuclei in bone marrow cells in mice (Victorin et al.,
    1990).

         Two studies have dealt with genotoxic effects in the relevant
    target organ, i.e. the lung, and both were positive at high
    concentrations. In the first one, Isomura et al. (1984) demonstrated
    the induction of mutations and chromosome abberations in lung cells of
    rats exposed to 27 ppm (50 000 µg/m3) for 3 h.  In the other (Walles
    et al., 1995), DNA single strand breaks were induced in lung cells of
    mice exposed to 54 000 µg/m3 (30 ppm) for 16 h or 94 000 µg/m3
    (50 ppm) for 5 h.

         Several studies have evaluated the issue of carcinogenesis and
    co-carcinogenesis, but results are often unclear or conflicting
    (Table 37).  However, there do not appear to be any published reports
    on studies using classical carcinogenesis whole-animal bioassays.  An
    excellent critical review and discussion of some of the important
    theoretical issues in interpreting these types of studies has been
    published (Witschi, 1988).  Although lung epithelial hyperplasia
    (section 5.2.2.4) and enhancement of endogenous retrovirus expression
    (Roy-Burman et al., 1982) have been thought by some to suggest
    increased carcinogenic potential, such findings are not conclusive
    (see US EPA, 1993).

         Wagner et al. (1965) suggested that NO2 may accelerate the
    production of tumours in CAF1/Jax mice (a strain that has
    spontaneously high pulmonary tumour rates) after continuous exposure
    to 9400 µg/m3 (5.0 ppm) NO2.  After 12 months of exposure, 7 out of
    10 mice in the exposed group had tumours, compared to 4 of 10 in the
    controls.  No differences in tumour production were observed after 14
    and 16 months of exposure.  A statistical evaluation of the data was
    not presented.  The frequency and incidence of spontaneously
    occurring pulmonary adenomas was increased in strain A/J mice (with
    spontaneously high tumour rates) after exposure to 18 800 µg/m3
    (10.0 ppm) NO2 for 6 h/day, 5 days/week, for 6 months (Adkins et al.,
    1986).  These small, but statistically significant, increases were
    only detectable when the control response from nine groups (n = 400)
    were pooled.  Exposure to 1880 and 9400 µg/m3 (1.0 and 5.0 ppm) NO2
    had no effect.  In contrast, Richters & Damji (1990) found that an
    intermittent exposure to 470 µg/m3 (0.25 ppm) NO2 for up to 26 weeks
    decreased the progression of a spontaneous T cell lymphoma in
    AKR/ cum mice and increased survival rates.  The investigators
    attribute this effect to an NO2-induced decrease in the proliferation
    of T cell subpopulation (especially T-helper/inducer lymphocytes) that
    produce growth factors for the lymphoma.

         Whether NO2 facilitates metastases has been the subject of
    several experiments by Richters & Kuraitis (1981, 1983), Richters &
    Richters (1983) and Richters et al. (1985). Mice were exposed to
    several concentrations and durations of NO2 and were injected
    intravenously with a cultured-derived melanoma cell line (B16) after
    exposure; subsequent tumours in the lung were counted.  Although
    some of the experiments showed an increased number of lung
    tumours, statistical methods were inappropriate.  Furthermore, the 
    experimental technique used in these studies probably did not evaluate
    metastases formation, as the term is generally understood, but more
    correctly, colonization of the lung by tumour cells.

        Table 37.  Effects of nitrogen dioxide (NO2) on carcinogenesis or co-carcinogenesisa
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    188-18 800     0.1-10.0     0.5-4 h               Mouse     Mice exposed to DMA had whole-body concentration-           Iqbal et al.
                                                                related increase in DMN.                                    (1981)

    470            0.25         7 h/day,              Mouse     NO2 slowed progression of spontaneous T cell                Richters & Damji
                                5 days/week,                    lymphomas in AKR/cum mice, increased survival, and          (1990)
                                up to 26 weeks                  decreased number of splenic CD4+ T cells.

    752            0.4          7-8 h/day,            Mouse     Increased lung tumors and mortality in mice injected        Richters &
    940            0.5          5 days/week,                    with melanoma cells after NO2 exposure.                     Kuraitis (1981,
    1500           0.8          12 weeks                                                                                    1983); Richters
                                                                                                                            et al. (1985)

    940-1500       0.5-0.8      Continuous,           Mouse     Hyperplastic foci identical to that observed after          Nakajima et
                                30 days                         exposure to known carcinogens.                              al. (1972)

    1500           0.8          8 h/day,              Mouse     Enhanced retrovirus expression in two strains of            Roy-Burman
                                5 days/week,                    mice.                                                       et al. (1982)
                                18 weeks

    1880           1.0          6 h/day,              Mouse     No effect at 1880 or 9400 µg/m3. At 18 800 µg/m3,           Adkins et al.
    9400           5.0          5 days/week,                    spontaneous adenomas in strain A/J mice increased           (1986)
    18 800         10.0         6 weeks                         only when compared to pooled control group.

    2000           1.1          Continuous,           Rat       DMA plus NO2 did not produce tumors. Design and             Benemansky
    3010           1.6          lifetime                        statistical analyses not appropriate; exposure              et al. (1981)
                                                                methods not described.
                                                                                                                                              

    Table 37  (Con't)
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    9400-18 800    5.0-10.0     2 h/day,              Mouse     Mice given 4-nitroquinoline-1-oxide during NO2              Ide & Otsu
                                5 days/week,                    exposure; NO2 had no effect on tumor production.            (1973)
                                50 weeks

    18 800         10.0         2 h/day,              Mouse     Mice given 4-nitroquinoline-1-oxide and NO2.                Otsu & Ide
                                5 days/week,                    NO2 decreased incidence of lung tumors.                     (1975)
                                50 weeks

    28 200-94 000  15.0-50.0    1-4 h                 Mouse     Mice gavaged with morpholine had concentration-             Iqbal et al.
                                                                dependent increase in whole-body content of NMOR.           (1980)

    31 020-38 500  16.5-20.5    5-6 h/day,            Mouse     In vivo production of NMOR when 1 g/kg of morpholine        Van Stee et
                                4 days; plus                    was administered each day prior to exposure.                al. (1983)
                                3 h on 5th day

    84 600         45.0         2 h                   Mouse     Mice gavaged with morpholine had an in vivo increase        Norkus et al.
                                                                in NMOR production.                                         (1984)

    199 000        106.0        0.5-4 h               Rat       Rats given morpholine in their diets or by gavage had       Mirvish et al.
                                                      Mouse     no NMOR detected in their bodies.                           (1981)
                                                                In mice morpholine, by gavage, yielded no significant
                                                                in vivo NMOR production.
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  DMA = Dimethylamine; DMN = Dimethylnitrosamine; NMOR = N-nitrosomorpholine
             Ide & Otsu (1973) did not find that chronic exposure to high
    concentrations of NO2 (somewhere between 9400 and 18 800 µg/m3,
    5.0 and 10.0 ppm) enhanced tumour production in conventional mice
    receiving five weekly injections of 0.25 mg 4-nitroquinoline-1-oxide
    (a lung-tumour-specific carcinogen).  Benemansky et al. (1981) used a
    known carcinogen, nitrosodimethylamine or its precursor dimethylamine
    (DMA) to test for interactions with a chronic exposure to NO2.
    However, appropriate statistical techniques and control groups were
    not employed and the methods of exposure and monitoring of NO2 were
    not reported, thus precluding accurate evaluation.  In another study,
    rats were injected with  N-bis (2-hydroxy-propyl)nitrosamine (BHPN)
    and continuously exposed to 75, 750 or 7500 µg/m3 (0.04, 0.4 or
    4.0 ppm) NO2 for 17 months.  Although the data indicated five times
    as many lung adenomas or adenocarcinomas in the rats injected with
    BHPN and exposed to 7500 µg/m3 NO2 (5/40 compared to 1/10), the
    results failed to achieve statistical significance (Ichinose et al.,
    1991).

         Because of evidence that NO2 could produce NO2- and NO3- in
    the blood and the fact that NO2- is known to react with amines to
    produce animal carcinogens (nitrosamines), the possibility that NO2
    could produce cancer via nitrosamine formation has been investigated. 
    Iqbal et al. (1980) was the first to demonstrate a linear time-
    dependent and concentration-dependent relationship between the amount
    of  N-nitrosomorpholine (NMOR) (an animal carcinogen) found in
    whole-mouse homogenates after the mice were gavaged with 2 mg of
    morpholine (an exogenous amine that is rapidly nitrosated) and
    exposure to 28 200 to 94 000 µg/m3 (15.0 to 50.0 ppm) NO2 for
    between 1 and 4 h.  In a follow-up study, Iqbal et al. (1981) used
    DMA, an amine that is slowly nitrosated to dimethylnitrosamine (DMN). 
    They reported a concentration-related increase in biosynthesis of DMN
    at NO2 concentrations as low as 188 µg/m3 (0.1 ppm); however, the
    rate was significantly greater at concentrations above 18 800 µg/m3
    (10.0 ppm) NO2.  Increased length of exposure also increased DMN
    formation between 0.5 and 2 h, but synthesis of DMN was less after 3
    or 4 h of exposure than after 0.5 h.

         Mirvish et al. (1981) conducted analogous research and concluded
    that the results of Iqbal et al. (1980) were technically flawed, but
    found that  in vivo exposure to NO2 could produce a nitrosating
    agent (NSA) that would nitrosate morpholine only when morpholine was
    added  in vitro.  Further experiments showed that the NSA was
    localized in the skin (Mirvish et al., 1983) and that mouse skin
    cholesterol was a likely NSA (Mirvish et al., 1986).  It has also been
    reported that only very lipid-soluble amines, which can penetrate the
    skin, would be available to the NSA.  Compounds such as morpholine,
    which are not lipid-soluble, could only react with NO2 when it was
    painted directly on the skin (Mirvish et al., 1988).  Iqbal (1984),
    responding to the Mirvish et al. (1981) criticisms, verified their
    earlier studies (Iqbal et al., 1980).   In vivo nitrosation was also
    demonstrated by Norkus et al. (1984) after morpholine administration
    and a 2-h exposure to 84 600 µg/m3 (45 ppm) NO2.

         Another study (Van Stee et al., 1983) reported that mice gavaged
    with 1 g of morpholine/kg body weight per day and then exposed (5-6 h
    daily for 5 days) to 31 000 to 38 500 µg/m3 (16.5 to 20.5 ppm) NO2
    revealed that NMOR could be produced  in vivo.  The single site
    containing the greatest amount of NMOR was the gastrointestinal tract.

         Shoaf et al. (1989) studied the uptake and nitrosation of primary
    amines by NO2 in isolated ventilated rat lungs.  The rate of
    nitrosation was very low because the nitrosation of primary amines is
    a general acid/base catalysed reaction that would be at a minimum at
    pH 7.  The authors could not replicate the previous nitrosation
    studies.  At a maximum, only 0.0001% of an amine would be nitrosated. 
    Such a rate is at or below the detection limit for nitrosamine.  The
    studies reporting nitrosation may be seriously in error.  Nitrosation
    may be a very minimal reaction and of little consequence.

         Victorin (1994) reviewed the genotoxicity of nitrogen oxides and
    concluded that there is no clear evidence of a carcinogenic potential
    of NO2.  Victorin (1994) also directed attention to the possibility
    that NOx compounds in photochemical smog may contribute secondarily
    to formation of other genotoxic compounds.  For example, it was noted
    that strongly mutagenic nitro-PAH compounds are easily formed and
    mutagenic reaction products may be formed from alkenes through
    photochemical reactions.

         Overall, the above critical evaluation indicates that there is no
    evidence establishing that tumours can be directly induced by NO2
    exposure alone.  Also, the available evidence for NO2 promoting or
    enhancing the production or growth of tumours caused by other agents
    is quite limited and conflicting.  It must therefore be concluded that
    the evidence for carcinogenicity of nitrogen oxides is at present
    inadequate, but the issue should be addressed by further research.

    5.2.4  Extrapulmonary effects

         Exposure to NO2 produces a wide array of health effects beyond
    the confines of the lung.  Thus, NO2 and/or some of its reactive
    products penetrate the lung or nasal epithelial and endothelial layers
    to enter the blood and produce alterations in blood and various other
    organs (Shoaf et al., 1989).  Effects on the systemic immune system
    are discussed under section 5.2.2.1.  Information regarding the
    effects of NO2 on animal behaviour and brain enzymes is limited to a
    few studies that cannot be readily interpreted in terms of human risks
    and will not be discussed.  The summary of other systemic effects is
    quite brief because the database suggests that effects on the
    respiratory tract are of more concern.  A more detailed discussion of
    extrapulmonary responses can be found in US EPA (1993).

         Results of research on the number of erythrocytes and leukocytes,
    haemoglobin concentration, and contents of erythrocyte membranes are
    inconsistent.  In the only such study conducted below 9400 µg/m3
    (5.0 ppm) NO2, Nakajima & Kusumoto (1968) found that the amount of
    methaemoglobin was not increased when mice were exposed to 1500 µg/m3
    (0.8 ppm) NO2 for 5 days.  This topic was of interest because some
    (but not all)  in vitro studies and high concentration  in vivo NO2
    studies found methaemoglobin effects (US EPA, 1993).

         Several studies have examined hepatic function either directly or
    indirectly after NO2 exposure.  Changes in serum chemistry (e.g.,
    plasma cholinesterase, Drozdz et al., 1976; Menzel et al., 1977)
    suggest that NO2 exposure may affect the liver.  Xenobiotic
    metabolism appears to be affected by NO2.  A 3-h exposure to NO2
    concentrations as low as 470 µg/m3 (0.25 ppm) increased
    pentobarbital-induced sleeping times in female, but not male, mice
    (Miller et al., 1980; Graham et al., 1982).  Higher exposures
    (9400 µg/m3, 5.0 ppm; 3 h) did not affect the level of hepatic
    cytochrome P-450 or the activities of several mixed-function oxidases
    in mice (Graham et al., 1982).  Other authors found mixed effects
    (i.e. increase or decrease depending on exposures) on liver cytochrome
    P-450 levels in rats (Takano & Miyazaki, 1984; Takahashi et al.,
    1986).  Significant decreases in cytochrome P-450 from rat liver
    microsomes were also found after 7 days of exposure to 752 or
    7520 µg/m3 (0.4 or 4.0 ppm) NO2, but not after exposure to
    2260 µg/m3 (1.2 ppm) NO2 (Mochitate et al., 1984).  NADPH-cytochrome
    C reductase was reduced with 5 days of exposure to 7520 and
    18 800 µg/m3 (4.0 and 10.0 ppm) NO2.  Drozdz et al. (1976) found
    decreased total liver protein and sialic acid, but increased protein-
    bound hexoses in guinea-pigs exposed to 2000 µg/m3 (1.05 ppm) NO2,
    8 h/day for 180 days.  Liver alanine and aspartate aminotransferase
    activity was increased in the mitochondrial fraction but decreased in
    the cytoplasmic fraction of the liver.  Electron micrographs of the
    liver showed intracellular oedema and inflammatory and parenchymal
    degenerative changes.

         Takahashi et al. (1986) found that continuous exposure to 2260
    and 7520 µg/m3 (1.2 and 4.0 ppm) NO2 increased the amount of
    cytochrome P-450 and cytochrome b5 in the kidney after 8 weeks of
    exposure.  Continued exposure for 12 weeks resulted in less
    substantial increases in the amount and activity of the microsomal
    electron-transport enzymes.  This is in contrast to the decreased
    activity in the liver.

         Increases in urinary protein and specific gravity of the urine
    were reported by Sherwin & Layfield (1974) in guinea-pigs exposed
    continuously to 940 µg/m3 (0.5 ppm) NO2 for 14 days.  Proteinuria
    was detected in another group of animals when the exposure was reduced
    to 752 µg/m3 (0.4 ppm) NO2 for 4 h/day.  Disc electrophoresis of the

    urinary proteins demonstrated the presence of albumin and alpha-,
    beta-, and gamma-globulins.  The presence of high molecular weight
    proteins in urine is characteristic of the nephrotic syndrome. 
    Differences in water consumption or in the histology of the kidney
    were not found.

         Few studies have examined the effects of NO2 on reproduction and
    development or the heritable mutagenic potential of NO2.  Exposure to
    1800 µg/m3 (1.0 ppm) NO2 for 7 h/day (5 days/week for 21 days)
    resulted in no alterations in spermatogenesis, germinal cells or
    interstitial cells of the testes of six rats (Kripke & Sherwin, 1984). 
    Similarly, breeding studies by Shalamberidze & Tsereteli (1971) found
    that long-term NO2 exposure had no effect on fertility.  However,
    there was a statistically significant decrease in litter size and
    neonatal weight when male and female rats exposed to 2440 µg/m3
    (1.3 ppm) NO2, 12 h/day for 3 months were bred.   In utero death due
    to NO2 exposure resulted in smaller litter sizes, but no direct
    teratogenic effects were observed in the offspring.  In fact, after
    several weeks, NO2-exposed litters approached weights similar to
    those of controls.

         Inhalation exposure of pregnant Wistar rats to NO2
    concentrations of 1000 and 10 000 µg/m3 for 6 h/day throughout
    gestation (21 days) was found to have maternal toxic effects and to
    induce developmental disturbances in the progeny (Tabacova et al.,
    1984; Balabaeva & Tabacova, 1985; Tabacova & Balabaeva, 1988). The
    maternal weight gain during gestation was significantly reduced at
    10 000 µg/m3 (5.3 ppm).  Pathomorphological changes, manifested at
    the higher exposure level, were found in maternal organs, e.g.,
    desquamative bronchitis and bronchiolitis in the lung, mild
    parenchymal dystrophy and reduction of glycogen in the liver, and
    blood stasis and inflammatory reaction in the placenta.  At gross
    examination, the placentae of the dams exposed to 10 000 µg/m3 were
    smaller in size than those of control rats.  A marked increase of
    lipid peroxides was found in maternal lungs and particularly in the
    placenta at both exposure levels by the end of gestation (Balabaeva &
    Tabacova, 1985).  Disturbances in the prenatal development of the
    progeny were registered, such as two- to four-fold increase in late
    post-implantation lethality at 1000 and 10 000 µg/m3 (0.5 and
    5.3 ppm), respectively, as well as reduced fetal weight at term and
    stunted growth at 10 000 µg/m3 (Tabacova et al., 1984).  These
    effects were significantly related to the content of lipid peroxides
    in the placenta, which was suggestive of a pathogenetic role of
    placental damage (Tabacova & Balabaeva, 1988).  Teratogenic effects
    were not observed, but dose-dependent morphological signs of
    embryotoxicity and retarded intrauterine development, such as
    generalized oedema, subcutaneous haematoma, retarded ossification
    and skeletal aberrations, were found at both exposure levels.

         In the only study that has examined postnatal development, a
    significant delay in eye opening and incisor eruption was observed in
    the progeny of maternally exposed Wistar rats (Tabacova et al., 1985). 
    The dams were exposed to 50, 100, 1000 or 10 000 µg/m3 (0.03, 0.05,
    0.53 or 5.3 ppm) NO2 for 6 h/day, 7 days/week throughout gestation,
    and the offspring were studied for 2-month post-exposure.  Significant
    deficits in the onset of normal neuromotor development and reduced
    open field activity were detected in the offspring of dams exposed to
    1000 and 10 000 µg/m3 NO2.

    5.3  Effects of mixtures containing nitrogen dioxide

         Humans are exposed to pollutant mixtures in the ambient air, and,
    because pollutant interactions do occur, it is difficult to predict
    the effects of NO2 in a mixture based upon the effects of NO2 alone.
    Epidemiological studies (chapter 7), by their very nature, evaluate
    ambient air mixtures, but the presence of confounding variables makes
    it difficult to demonstrate a cause-effect relationship.  In contrast,
    controlled animal and human clinical studies can often demonstrate the
    cause of a response, but are typically limited to binary or tertiary
    mixtures, which do not truly reflect ambient air exposures.  When
    combinations of air pollutants are studied, there are a number of
    possible outcomes on human or animal responses.  The result of
    exposure to two or more pollutants may be simply the sum of the
    responses to individual pollutants; this is referred to as additive. 
    Another possibility is that the resultant response may be greater than
    the sum of the individual responses, suggesting some type of
    interaction or augmentation of the response; this is referred to as
    synergism.  Finally, responses may be less than additive; this is
    often called antagonism. Generally, such human clinical studies, which
    focused on pulmonary function, have not found that acute exposures to
    NO2 has any impact on the response to other co-occurring pollutants
    (e.g., O3) or that additive effects occur.  Animal toxicological
    studies, with a wider array of designs and end-points, have shown an
    array of interactions, including no interaction, additivity and
    synergism.  Because no clear understanding of NO2 interactions has
    yet emerged from this database, only a brief overview is provided
    here.  A more substantive review can be found in US EPA (1993).  Other
    animal studies sought to understand the effects of ambient air
    mixtures containing NO2 or vehicular combustion exhausts containing
    NOx.  Generally these studies provide useful information on the
    mixtures, but lack NO2-only groups, making it impossible to discern
    the influence of NO2.  Therefore, this class of research is not
    described here, but is reviewed elsewhere (US EPA, 1993).

         The vast majority of interaction studies have involved NO2 and
    O3.  For lung morphology end-points, NO2 had no interaction with O3
    (Freeman et al., 1974) or with sulfur dioxide (SO2) (Azouley et al.,
    1980) after a subchronic exposure. Some biochemical responses to NO2

    plus O3 display no positive interaction or synergism.  For example,
    Mustafa et al. (1984) found synergism for some end-points (e.g.,
    increased activities of O2 consumption and antioxidant enzymes),
    but no interaction for others (e.g., DNA or protein content)
    in rats exposed for 7 days.  Ichinose & Sagai (1989) observed a
    species-dependence in regard to the interaction of O3 (752 µg/m3,
    0.4 ppm) and NO2 (752 µg/m3, 0.4 ppm) after 2 weeks of exposure. 
    Guinea-pigs, but not rats, had a synergistic increase in lung lipid
    peroxides.  Rats, but not guinea-pigs, had synergistic increases in
    antioxidant factors (e.g., non-protein thiols, vitamin C, glucose-6-
    phosphate dehydrogenase, GSH peroxidase).  Schlesinger et al. (1990)
    observed a synergistic increase in prostaglandin E2 and F2 alpha in the
    lung lavage of acutely exposed rabbits; the response appeared to have
    been driven by O3.  However, with 7 or 14 days of repeated 2-h
    exposures, only prostaglandin E2 was decreased and appeared to have
    been driven by NO2; there was no synergism (Schlesinger et al.,
    1991).

         The infectivity model has been frequently used to study NO2-O3
    mixtures.  In this model, mice are exposed to O3 and NO2 alone or in
    mixtures for various durations.  The mice are then challenged with an
    aerosol of viable bacteria.  An increase in mortality indicates
    detrimental effects on lung host-defence mechanisms.  Ehrlich et al.
    (1977) found additivity after acute exposure to mixtures of NO2 and
    O3.  They reported synergism after subchronic exposures.  Exposure
    scenarios involving NO2 and O3 have also been performed using a
    continuous baseline exposure to one concentration or mixture, with
    superimposed short-term peaks to a higher level.  This body of work
    (Ehrlich et al., 1979; Gardner, 1980; Gardner et al., 1982; Graham et
    al., 1987) shows that differences in the pattern and concentrations of
    the exposure are responsible for the increased susceptibility to
    pulmonary infection, without indicating clearly the mechanism
    controlling the interaction.

         Some aerosols may potentiate response to NO2 by producing local
    changes in the lungs that enhance the toxic action of co-inhaled NO2. 
    The impacts of NO2 and H2SO4 on lung host defences have been
    examined by Schlesinger & Gearhart (1987) and Schlesinger (1987a).  In
    the former study, rabbits were exposed for 2 h/day for 14 days to
    either 564 µg/m3 (0.3 ppm) or 1880 µg/m3 (1.0 ppm) NO2, or
    500 µg/m3 H2SO4 alone, or to mixtures of the low and high NO2
    concentrations with H2SO4.  Exposure to either concentration of NO2
    accelerated alveolar clearance, whereas H2SO4 alone retarded
    clearance.  Exposure to either concentration of NO2 with H2SO4
    resulted in retardation of clearance in a similar manner to that seen
    with H2SO4 alone.

         Schlesinger (1987a) used a similar exposure design, but different
    end-points.  Exposure to 1800 µg/m3 (1.0 ppm) NO2 with acid resulted
    in an increase in the numbers of PMNs in lavage fluid at all time
    points (not seen with either pollutant alone), and an increase in
    phagocytic capacity of AMs after two or six exposures.  In contrast,
    exposure to 564 µg/m3 (0.3 ppm) NO2 with acid resulted in depressed
    phagocytic capacity and mobility.  The NO2/H2SO4 mixture was
    generally either additive or synergistic, depending on the specific
    cellular end-point being examined.

         Last et al. (1983) and Last & Warren (1987) found that exposure
    to high levels of NO2 (< 9400 µg/m3, 5.0 ppm) with very high
    concentrations of H2SO4 (1 mg/m3) caused a synergistic increase in
    collagen synthesis rate and protein content of the lavage fluid of
    rats.

         Dogs were exposed for 68 months (16 h/day) to raw or
    photochemically reactive vehicle exhaust which included mixtures of
    NOx œ one with a high NO2 level and a low NO level (1200 µg/m3,
    0.64 ppm, NO2; 310 µg/m3, 0.25 ppm, NO), and one with a low NO2
    level and a high NO level (270 µg/m3, 0.14 ppm, NO2; 2050 µg/m3,
    1.67 ppm, NO) (Stara et al., 1980).  Following the end of exposure,
    the animals were maintained for about 3 years in normal indoor air. 
    Numerous pulmonary function, haematological and histological
    end-points were examined at various times during and after exposure. 
    The lack of an NO2-only or NO-only group precludes determination of
    the nature of the interaction.  Even so, the main findings are of
    interest.  Pulmonary function changes appeared to progress after
    exposure ceased.  Dogs in the high NO2 group had morphological
    changes considered to be analogous to human centrilobular emphysema
    (see section 2.2.2.4).  Because these morphological measurements were
    made after a 2.5- to 3-year holding period in clean air, it cannot be
    determined with certainty whether these disease processes abated or
    progressed during this time.  This study suggests progression of
    damage after exposure ends.

    5.4  Effects of other nitrogen oxide compounds

    5.4.1  Nitric oxide

         The toxicological database for NO is small, relative to NO2.  It
    is often difficult to obtain pure NO in air without some contamination
    with NO2.  An excellent review on the effects of NO on animals and
    humans has been prepared by Gustafsson (1993) for the Swedish
    Environmental Protection Agency.  The following sections are based on
    the information in this review.

    5.4.1.1  Endogenous formation of NO

         Endogenous NO synthesis occurs by NO formation from physiological
    substrate (the amino acid L-arginine) in cells of many of the organ
    systems, such as nerve tissue, blood vessels and the immune system. 
    NO has been found to be produced by at least three different
    oxygen-utilizing NO synthases, for purposes such as signalling in the
    nervous system, mediating vasodilation in both systemic and pulmonary
    circulation, and mediating cytotoxicity and host defence reactions in
    the immune system (Garthwaite, 1991; Barinaga, 1991; Moncada et al.,
    1991; McCall & Valance, 1992; Snyder & Bredt, 1992; Moncada, 1992). 
    The impact of these findings for an understanding of the toxicological
    effects of NO is still difficult to assess.

         The actions of endogenous NO can be divided into two main groups. 
    The first group involves low concentrations of NO (nano- to picomolar)
    formed by constitutive enzymes in nerve and endothelial cells.  Nitric
    oxide has local discrete actions exerted via activation of an enzyme,
    guanylate cyclase, in the target cell (Ignarro, 1989).  The second
    group involves high concentrations of NO (micro- to nanomolar) formed
    by enzymes that can increase in amount through the induction of these
    enzymes upon exposure to bacterial toxins or to growth-regulating
    factors (cytokinins).  The inducible NO formation occurs especially in
    macrophages and neutrophil leukocytes and is important for the killing
    of bacteria and parasites, and possibly also for cytostasis in
    antitumour reactions (Hibbs et al., 1988; Ignarro, 1989; Moncada et
    al., 1991; Moncada, 1992).

         For effects of inhaled NO it is important to consider that 
    endogenous NO regulates pulmonary vascular resistance; it is found in
    small amounts in exhaled air and has been suggested to be necessary
    for normal oxygenation of the blood (Persson et al., 1990; Gustafsson
    et al., 1991).

    5.4.1.2  Absorption of NO

         Yoshida et al. (1981) found that < 10% of the NO "inhaled" by
    isolated perfused lungs of rabbits was absorbed.  In normally
    breathing humans, 85 to 92% of NO was absorbed at concentrations
    ranging from 400 to 6100 µg/m3 (0.33 to 5.0 ppm) (Wagner, 1970;
    Yoshida & Kasama, 1987); values for NO2 were 81 to 90% (Wagner,
    1970).  Absorption of NO with exercise was 91 to 93% in humans
    (Wagner, 1970).  Yoshida et al. (1980) found the percentage of
    absorption of NO in rats acutely exposed to 169 300 µg/m3 (138 ppm),
    331 300 µg/m3 (270 ppm) and 1 079 800 µg/m3 (880 ppm) to be 90%, 60%
    and 20%, respectively.  The progressive decrease in absorption was
    ascribed to an exposure-induced decrease in ventilation.  In dogs
    exposed to vehicle exhaust mixtures, 73% of the constituent NO was
    removed by the nasopharyngeal region; this compared to 90% removal for

    NO2 (Vaughan et al., 1969).  Thus, respiratory tract absorption of NO
    has some similarities to that for NO2, in spite of solubility
    differences.  The lower solubility of NO may, however, result in
    greater amounts reaching the pulmonary region, where it may then
    diffuse into blood and react with haemoglobin (Yoshida & Kasama,
    1987).   In vivo exposures seem to indicate that NO has a faster rate
    of diffusion through tissue than NO2 (Chiodi & Mohler, 1985).

    5.4.1.3  Effects of NO on pulmonary function, morphology and host lung
             defence function

         No change in respiratory function was found in guinea-pigs
    exposed to NO at 19 600 µg/m3 (16 ppm) or 61 300 µg/m3 (50 ppm) for
    4 h (Murphy et al., 1964).  Increased airway responsiveness to
    acetylcholine was observed in guinea-pigs exposed to 6130 µg/m3
    (5 ppm) NO for 30 min, twice a week for 7 weeks.  In sheep,
    significant reversal of vasoconstriction to an infused thromboxane
    analogue was seen with acute exposure to 6130 µg/m3 NO (Fratacci et
    al., 1991).  At the same exposure level, hypoxic vasoconstriction was
    significantly diminished and was nearly abolished at 49 000 µg/m3
    (40 ppm) NO in inhaled air (Frostell et al., 1991).

         Reversal of methacholine-induced bronchoconstriction by NO has
    been reported in guinea-pigs at 6130 µg/m3 (5 ppm) (Dupuy et
    al., 1992), while in rabbits full reversal of methacholine
    bronchoconstriction was seen at 98 100 µg/m3 (80 ppm) (Högman et
    al., 1993).  Relaxation of bronchial smooth muscle can be exerted
     in vitro by mechanisms dependent on an intact airway epithelium.  An
    endogenous muscle-relaxing factor released by the epithelium has been
    suggested, but it is not clear whether it is endogenous NO (Barnes,
    1993).

         The few studies that have examined histological response to
    non-lethal levels of NO are outlined in Table 38.  With chronic
    exposure, the morphological changes seen are similar to those with
    NO2 (see section 5.2.2.4 on morphological effects of NO2), except
    that NO levels needed to produce them are higher.  Additionally, Hugod
    (1979) noted that the absence of  NO-induced alterations in the
    alveolar epithelium suggested that the observed responses occurred
    after absorption of NO; that is, they were not caused by direct action
    of deposited NO.  Perhaps higher exposure concentrations of NO are
    needed for direct toxic action (e.g., results of Holt et al., 1979). 
    Some of the effects seen by Oda et al. (1976) with 12 270 µg/m3
    (10.0 ppm) NO may have been due to the presence of 1880 to 2820 µg/m3
    (1.0 to 1.5 ppm) NO2 in the exposure atmosphere.

         It is important to note that in all existing studies of NO
    toxicity in the lungs, histological evaluation of the lungs was
    rudimentary and no quantitative measurements were carried out to test
    for airspace enlargement or destruction.

        Table 38.  Effects of nitric oxide (NO) on respiratory tract morphologya
                                                                                                                                              

    NO2 Concentration
                      

    µg/m3        ppm          Exposure              Species     Effectsb                                                    Reference
                                                                                                                                              

    2460           2                        Continuous,         Rat       Slight emphysema-like alterations of alveoli.     Azoulay et
                   (NO2 = 0.08 ppm)b        6 weeks                                                                         al. (1977)

    2950           2.4                      Continuous,         Mouse     No difference from control.                       Oda et al.
                   (NO2 = 0.01-0.04 ppm)b   for lifetime                                                                    (1980b)
                                            (23-29 months)

    6150           5                        Continuous,         Rabbit    Oedema; thickening of alveolo-capillary           Hugod (1979)
                   (NO2 = < 0.1 ppm)b       14 days                       membrane due to fluid in interstitial space;
                                                                          fluid-filled vacuoles seen in arteriolar
                                                                          endothelial cells and at junctions of
                                                                          endothelial cells; no changes in alveolar
                                                                          epithelium; no inflammation.

    12 300         10                       2 h/day, 5 days     Mouse     Enlarged air spaces in lung periphery;            Holt et al.
                                            per week, up to               paraseptal emphysema; some haemorrhage;           (1979)
                                            30 weeks                      some congestion in alveolar septa; increased
                                                                          concentration of goblet cells in bronchi.

    12 300         10                       Continuous,         Mouse     Bronchiolar epithelial hyperplasia; hyperaemia;   Oda et al.
                   (NO2 = 1-1.5 ppm)b       6.5 months                    congestion; enlargement of alveolar septum;       (1976)
                                                                          increase in ratio of lung to body weight.
                                                                                                                                              

    a  Modified from US EPA (1993)
    b  This represents reported nitrogen dioxide (NO2) levels measured during exposure
             A recent study (Mercer et al., 1995) suggests that NO may be more
    potent than NO2 in introducing certain changes in lung morphology. 
    More specifically, male rats were exposed to either NO or NO2 at
    0.5 ppm with twice daily 1-h spikes of 1.5 ppm for 9 weeks.  The
    number of pores of Kohn and detached alveolar septa were evaluated by
    electron microscopy, using stereological procedures for the study of
    lung structure that involved morphometric analyses of electron
    micrographs.  The average number of pores per lung for the NO group
    exceeded by approx. 2.5 times the mean number for the NO2 groups,
    which was more than 10 times that for controls.  Analogously, the
    average number of detached septa per lung was significantly higher for
    the NO group (X = 117) than the NO2 group (X = 20) or the controls
    (X = 4).  There was also a statistically significant 30% reduction in
    interstitial cells in the NO group, but no significant differences in
    the other parenchymal cell types between the controls and the NO- or
    NO2-exposed groups.  Lastly, the thickness of the interstitial space
    was reduced for the NO group (X = 0.24 µm versus 0.32 µm for controls)
    but not for the NO2 group (X = 0.29 µm), and epithelial cell
    thickness did not differ between the groups.

         The effects of NO on host defence function of the lungs has been
    examined in two studies.  Holt et al. (1979) found immunological
    alterations in mice exposed to 12 270 µg/m3 (10 ppm) NO for 2 h/day
    (5 days/week for 30 weeks).  However, interpretation is complicated by
    the duration dependence of some of the responses (e.g., an enhancement
    of the humoral immune response to SRBCs was seen at 10 weeks, but this
    was not evident at the end of the exposure series).  The effects of NO
    on bacterial defences were examined by Azoulay et al. (1981).  Male
    and female mice were exposed continuously to 3760 µg/m3 (2.0 ppm) NO
    for 6 h to 4 weeks to assess the effect on resistance to infection
    induced by a bacterial aerosol administered after each NO exposure. 
    There were no statistically significant effects for either sex at any
    of the time points studied.

    5.4.1.4  Metabolic effects

         Mice exposed to NO concentrations of 12 300 to 25 800 µg/m3
    (10 to 21 ppm) for 3 h daily for 7 days showed no change in the
    levels of reduced glutathione in their lungs (Watanabe et al., 1980). 
     In vitro data indicate that NO stimulates guanylate cyclase and
    therefore leads to smooth muscle relaxation and vasodilation and
    functional effects on the nervous system (Katsuki et al., 1977;
    Ignarro, 1989; Garthwaite, 1991; Moncada et al., 1991).  These effects
    are probably responsible for vasodilation in the pulmonary circulation
    and an acute bronchodilator effect of inhaled NO.  However, it is
    unclear whether other effects might be exerted from ambient NO via
    this pathway.  Due to the rapid inactivation of NO in haemoglobin,
    internal organs other than the lungs are unlikely to be affected
    directly by cyclic GMP-mediated vasodilator influence from ambient
    concentrations of NO.

         Methaemoglobin formation, via the formation of nitrosylhaemoglobin
    (Oda et al., 1975, 1979, 1980a,b; Case et al., 1979; Nakajima et al.,
    1980) and subsequent oxidation with oxygen, is well known (Kon et al.,
    1977; Chiodi & Mohler, 1985).  During NO exposure of mice to 24 500 to
    98 100 µg/m3 (20-80 ppm), the levels of methaemoglobin were found
    to increase exponentially with the NO concentration (Oda et al.,
    1980b).  After the cessation of NO exposure, methaemoglobin decreased
    rapidly, with a half-time of only a few minutes.  In humans the ability
    to reduce methaemoglobin varies genetically and is lower in infants. 
    Of the NO reaction products with haemoglobin, methaemoglobin attains
    higher levels than nitrosylhaemoglobin (Maeda et al., 1987).  Exposure
    of mice to 2940 µg/m3 (2.4 ppm) NO for 23-29 months resulted in
    nitrosylhaemoglobin levels at 0.01%, while the maximal methaemoglobin
    level was 0.3% (Oda et al., 1980b).  At 12 300 µg/m3 for 6.5
    months the nitrosylhaemoglobin level was 0.13% and the level of
    methaemoglobin was 0.2% (Oda et al., 1976).  Rats exposed to
    2450 µg/m3 (2 ppm) continuously for six weeks showed no detectable
    methaemoglobin (Azoulay et al., 1977).

    5.4.1.5  Haematological changes

         Mice exposed to NO at 11 070 µg/m3 (9 ppm) for 16 h had
    decreased iron transferrin (Case et al., 1979), and when exposed to
    12 300 µg/m3 (10 ppm) for 6.5 months had increased leukocyte count
    and proportion of polymorphonuclear cells (Oda et al., 1976).  Red
    blood cell morphology, spleen weight and bilirubin were also affected. 
    A slight increase in haemolysis was seen in mice exposed to
    2940 µg/m3 (2.4 ppm) of NO (Oda et al., 1980a).

    5.4.1.6  Biochemical mechanisms for nitric oxide effects: reaction
             with iron and effects on enzymes and nucleic acids

         NO has an affinity for haem-bound iron which is two times higher
    than that of carbon monoxide.  This affinity leads to the formation of
    methaemoglobin and the stimulation of guanylate cyclase.  Furthermore,
    NO reacts with thiol-associated iron in enzymes and eventually
    displaces the iron.  This is a possible mechanism for the cytotoxic
    effects of NO (Hibbs et al., 1988; Weinberg, 1992).   In vitro, the
    NO donor sodium nitroprusside has been shown to mobilize iron from
    ferritin (Reif & Simmons, 1990).  NO might possibly modulate
    arachidonic acid metabolism via interference with iron (Kanner et al.,
    1991a,b).

         NO inhibits aconitase, an enzyme in the Krebs cycle, and also
    complex 1 and 2 of the respiratory chain (Hibbs et al., 1988; Persson
    et al., 1990; Stadler et al., 1991).  Permanent modification of
    haemoglobin has been found; possibly via deamination (Moriguchi
    et al., 1992).  NO can also deaminate DNA, evoke DNA chain breaks,

    and inhibit DNA polymerase and ribonucleotide reductase (Wink et al.,
    1991; Lepoivre et al., 1991; Kwon et al., 1991; Nguya et al., 1992). 
    NO might be antimitogenic and inhibit T cell proliferation in rat
    spleen cells (Fu & Blankenhorn, 1992), and NO donors inhibit DNA
    synthesis, cell proliferation, and mitogenesis in vascular tissue
    (Garg & Hassid, 1989; Nakaki et al., 1990).  ADP (adenosine
    diphosphate) ribosylation is stimulated by NO-generating agents
    (Nakaki et al., 1990).

         Substantial  in vitro evidence has recently been published
    describing other effects of NO in tissues.  These include:  inhibition
    of glyceraldehyde-3-phosphate dehydrogenase (GAPDH) via ADP
    ribosylation (Alheid et al., 1987; Dimmeler et al., 1992); macrophage
    mediated-nitric oxide dependent mechanisms which include inhibition of
    the electron transport chain (Nathan, 1992); inhibition of DNA
    synthesis (Hibbs et al., 1988); inhibition of protein synthesis
    (Curran et al., 1991) and decrease in cytosolic free calcium by a
    cGMP-independent mechanism (Garg & Hassid, 1991).

    5.4.2  Nitric acid

         There have been only a few toxicological studies of HNO3, which
    exists in ambient air generally as a highly water-soluble vapour.  A
    few investigators have examined the histological response to instilled
    HNO3 (usually 1%), a procedure used in developing models of
    bronchiolitis obliterans in various animals, namely dogs, rabbits and
    rats (Totten & Moran, 1961; Greenberg et al., 1971; Gardiner &
    Schanker, 1976; Mink et al., 1984).  However, the relevance of such
    instillation studies is questionable, except to provide information
    for the design of inhalation studies.

         Only two studies have been designed specifically to examine the
    pulmonary response to pure HNO3 vapour.  Abraham et al. (1982)
    exposed both normal sheep and allergic sheep (i.e., having airway
    responses similar to those occurring in humans with allergic airway
    disease) for 4 h to 4120 µg/m3 (1.6 ppm) HNO3 vapour.  The exposure,
    using a "head-only" chamber, decreased specific pulmonary flow
    resistance in both groups of sheep; this indicated the absence of any
    bronchoconstriction.  Allergic, but not normal, sheep showed increased
    airway reactivity to carbachol, both immediately and 24 h after HNO3
    exposure.  In another study, rats exposed for 4 h to 1000 µg/m3
    (0.38 ppm) HNO3 vapour or for 4 h/day for 4 days to 250 µg/m3
    (0.1 ppm) HNO3 showed a decrease in stimulated or unstimulated
    respiratory burst activity of alveolar macrophages (AMs) obtained by
    lavage, as well as an increase in elastase inhibitory capacity of BAL
    (Nadziejko et al., 1992).

    5.4.3  Nitrates

         Only one inhalation study conducted at levels < 1 mg/m3
    NO3- has been reported.  Busch et al. (1986) exposed rats and
    guinea-pigs with either normal lungs or elastase-induced emphysema to
    ammonium nitrate aerosols at 1 mg/m3 for 6 h/day, 5 days/week for
    4 weeks.  Using both light and electron microscopy, the investigators
    concluded that there were no significant effects of exposure on lung
    structure.

    5.5  Summary of studies of the effects of nitrogen compounds on
         experimental animals

         Responses to NO2 exposure have been observed in several
    laboratory animal species, resulting in the conclusion that these
    effects could occur in humans.  In addition, mathematical dosimetry
    models suggest that the greatest dose of NO2 is delivered to the same
    region in both animal and human lungs (i.e. the centriacinar region
    which is the junction of the conducting airway with the gas exchange
    area).  Thus, the responses of laboratory animals can be qualitatively
    extrapolated to humans.

         NO2 exposure causes lung structural alterations.  Exposure to
    3760 µg/m3 (2.0 ppm) for 3 days has resulted in centriacinar damage,
    including damaged cilia and alveolar wall oedema.  Prolonged exposures
    produce changes in the cells lining the centriacinar region, and the
    tissue in this region (i.e., alveolar interstitium) becomes thicker. 
    These effects were seen in rats exposed to 940 µg/m3 (0.5 ppm)
    baseline with brief peaks of 2800 µg/m3 (1.5 ppm) for 6 weeks or
    exposures to 940 µg/m3 (0.5 ppm) NO2 for 4 to 6 months.

         Several animal studies clearly demonstrate that chronic exposure
    to concentrations of NO2 > 9400 µg/m3 (> 5.0 ppm) can
    cause emphysema of the type seen in human lungs.  Increased lung
    distensibility was reported in mice exposed to 375 µg/m3 (0.2 ppm)
    with peaks of 1500 µg/m3 (0.8 ppm) after 1 year of exposure.

         NO2 increases susceptibility to bacterial and viral pulmonary
    infections in animals.  Reduced phagocytic activity and reduced
    mobility were observed in AMs from rabbits exposed for 13 days to
    500 µg/m3 (0.3 ppm).  The lowest observed concentration that
    increases lung susceptibility to bacterial infections after acute
    exposure is 3750 µg/m3 (2.0 ppm) NO2 (a 3-h exposure study in mice). 
    Acute (17 h) exposures to > 4250 µg/m3 (> 2.3 ppm) NO2 also
    decrease pulmonary bactericidal activity in mice.  After long-term
    exposures (e.g., 3 to 6 months) to 940 µg/m3 (0.5 ppm) NO2, mice
    have decreased resistance to lung bacterial infections. Exposure of
    mice for 1 year to 375 µg/m3 (0.2 ppm/week) with 1480 µg/m3
    (0.8 ppm) spike followed by infection with streptococcus resulted in

    increased mortality.  NO2 also increases lung susceptibility to viral
    infections in mice.  Subchronic (7-week) exposures to concentrations
    as low as 470 µg/m3 (0.25 ppm) NO2 can alter the systemic immune
    system in mice.

         NO2 exposure has been shown to cause a clear dose-related
    decrease in pulmonary antibacterial defences. Decreases in pulmonary
    antibacterial defences occurred at concentrations ranging from
    7520 µg/m3 (4 ppm) for  Staphylococcus aureus to 37 500 µg/m3
    (20 ppm) for  Proteus mirabilis.  Dose-response increases in
    bacterial-induced mortality in mice was demonstrated with continuous
    exposure to 940 µg/m3 (0.5 ppm) after 3 months.

         When the relationship of NO2 exposure concentration and duration
    was studied, concentration had more influence than duration on the
    outcome.  This conclusion is primarily based on investigations of lung
    antibacterial defences of mice, which also indicate that the exposure
    pattern (e.g., baseline level with daily peaks of NO2 or exposure
    24 h/day versus 6 to 7 h/day) has an impact on the study results.

         Structural changes in the lung become more severe as exposure
    progresses from weeks to months at a given NO2 concentration. Longer
    exposures resulted in effects at lower concentrations.

         NO2 showed positive effects in some studies with Salmonella
    strain TA100 and caused DNA strand breaks in a mammalian cell culture. 
    NO seems to be less active.  High concentrations of NO2 have induced
    mutations in lung cells  in vivo, but not in other organs. There are
    no classical chronic bioassays for carcinogenicity. Studies concerning
    enhancement of spontaneous tumours, co-carcinogenic effects, or
    facilitation of the metastases of tumours to the lung are inadequate
    to form conclusions.  Possible secondary effects concern the
     in vivo formation of nitrite and nitrosamines and atmospherically
    formed mutagenic reaction products from NOx and hydrocarbons.

         The effects of exposure to mixtures of NO2 and other pollutants
    are dependent on the exposure regimen, species and end-point measured. 
    Most mixture research involves NO2 and O3 and shows that additivity
    and synergism can occur.  A similar conclusion can be drawn from the
    more limited research with NO2 and sulfuric acid.  Findings of either
    additivity or synergism are of concern because of the ubiquitous
    co-occurrence of NO2 and O3.  Extrapolation of these findings is not
    currently possible.

         NO is a potent vasodilator and effects can be demonstrated with
    inhaled concentrations of approximately 6130 µg/m3 (5 ppm) in sheep
    and guinea-pigs.  NO also reduces resistance to bacterial infection
    via the inhalation route in female mice exposed to 2452 µg/m3
    (2 ppm).  Morphological alterations in the alveoli and thickening of

    the alveolocapillary membrane are seen in rabbits at 6130 µg/m3. 
    Methaemoglobin formation is seen at concentrations above 12 260 µg/m3
    (10 ppm).

         NO2 acts as a strong oxidant.  Unsaturated lipids are readily
    oxidized with peroxides as the dominant product.  Both ascorbic acid
    and alpha-tocopherol inhibit the peroxidation of unsaturated lipids. 
    When ascorbic acid is sealed within bi-layer liposomes, NO2 rapidly
    oxidizes the sealed ascorbic acid.  The protective effects of
    alpha-tocophernol (vitamin E) and ascorbic acid (vitamin C) in animals
    and humans are due to the inhibition of NO2 oxidation.  NO2 also
    oxidizes membrane proteins. The oxidation of either membrane lipids or
    proteins results in the loss of cell permeability control.  The lungs
    of NO2-exposed humans and experimental animals have larger amounts of
    protein within the lumen.  The recruitment of inflammatory cells and
    the remodelling of the lung are a consequence of these events.

         The oxidant properties of NO2 also induce the peroxide
    detoxification pathway of glutathione peroxidase, glutathione
    reductase, and glucose-6-phosphate dehydrogenase.  Increases in the
    peroxide detoxification pathway occur in animals in a roughly
    dose-response relationship following NO2 exposure.

         The mechanism of action of NO is less clear.  NO is readily
    oxidized to NO2 and then peroxidation occurs.  Because of concomitant
    exposure to some NO2 in NO exposures, it is difficult to discriminate
    NO effects from those of NO2.  NO is, however, a potent second
    messenger modulating a wide variety of essential cellular functions.

         Peroxyacetyl nitrate (PAN) decomposes in water generating
    hydrogen peroxide.  Little is known of the mechanism of action, but
    oxidative stress is likely for PAN and its congeners.

         Inorganic nitrates may act by alterations in intracellular pH. 
    Nitrate ion is transported into Type 2 cells, acidifying the cell. 
    Nitrate also mobilizes histamine from mast cells.  Nitrous acid could
    also act to alter intracellular pH, but this mechanism is unclear.

         The mechanisms of action of the other nitrogen oxides are unknown
    at present.

    6.  CONTROLLED HUMAN EXPOSURE STUDIES OF NITROGEN OXIDES

    6.1  Introduction

         The effects of nitrogen oxides (NOx) on human volunteers exposed
    under controlled exposure conditions are evaluated in this chapter. 
    Of the NOx species typically found in the ambient air, NO2 has been
    the most extensively studied.  Nitric oxide (NO), nitrates, nitrous
    acid and nitric acid also have been evaluated and are discussed here,
    as are investigations of mixtures of NOy and other co-occurring
    pollutants.  A more extensive detailed review of this literature can
    be found in US EPA (1993).

         Most volunteers for human clinical studies are young, healthy
    adult males, but other potentially susceptible subpopulations,
    especially asthmatics, patients with chronic obstructive pulmonary
    disease (COPD), children and the elderly have also been studied.  Many
    exposures are conducted while the volunteer performs some form of
    controlled exercise.  The exercise increases ventilation, which
    increases the mass of pollutant inhaled per unit time and may alter
    the distribution of the dose within the lung.  More information on
    NO2 dosimetry is presented in chapter 5. Important methodological and
    experimental design considerations for controlled human studies have
    been discussed in greater detail by Folinsbee (1988).

         In many human clinical studies of NO2 exposure, both pulmonary
    function and airway responsiveness to bronchoconstrictors have been
    measured.  Spirometric measurements of lung volume, as well as
    measurements of airway resistance, ventilation volume, breathing
    pattern, and other tests provide information about some of the basic
    physiological functions of the lung.  Dynamic spirometry tests (forced
    expiratory tests such as forced expiratory volume in 1 second (FEV1),
    maximal and partial flow-volume curves (including those using gases of
    different densities such as helium), peak flow measurements, etc.),
    and measurements of specific airway resistance/conductance (SRaw,
    SGaw) are also used.  Most of these tests evaluate large airway
    function.  However, since NO2 deposition occurs primarily around
    the junction of the tracheobronchial and pulmonary regions (section
    5.2.1), many of these tests may not provide the necessary information
    to evaluate fully the effects of NO2.  Other tests that may evaluate
    small airway function (e.g., multiple breath nitrogen washout tests,
    closing volume tests, aerosol deposition/distribution tests, density
    dependence of flow-volume curves, and frequency dependence of dynamic
    compliance) are less frequently used, and the extent to which they
    indicate small airways function is not clearly established.  As
    discussed below, NO2 can increase airway responsiveness to chemicals
    that cause bronchoconstriction, such as histamine or cholinergic
    agonists (i.e., acetylcholine, carbachol or methacholine).  Other
    challenge tests use allergens, exercise, hypertonic saline or cold-dry

    air.  Responses are usually measured by evaluating changes in airway
    resistance (Raw) or spirometry (e.g. FEV1) after each dose of the
    challenge is administered.  Generally, asthmatics are significantly
    more responsive than healthy normal subjects to these types of airway
    challenge (O'Connor et al., 1987).  However, there is some overlap
    between the most responsive healthy subjects and the least responsive
    (to histamine) asthmatics (Pattemore et al., 1990).

         In the following sections, the changes in pulmonary function and
    airway responsiveness after NO2 exposure in healthy subjects  are
    discussed.  Responses of asthmatics and patients with chronic
    obstructive pulmonary disease (COPD) are then evaluated.  A brief note
    regarding age-related susceptibility is followed by a review of the
    effects of NO2 on pulmonary host defences and on biochemical markers
    in lung lavage fluid or in the blood.  The effects of two other
    oxidized nitrogen compounds, NO and nitric acid vapour are also
    discussed.  Finally, the effects of mixtures of oxidized nitrogen
    compounds (NO2, NO, HNO3) with other gaseous or particulate
    pollutants are considered.  An overall summary is presented at the end
    of the chapter.

    6.2  Effects of nitrogen dioxide

    6.2.1  Nitrogen dioxide effects on pulmonary function and airway
           responsiveness to bronchoconstrictive agents

         Much research has focused on NO2-induced changes in pulmonary
    function and airway responsiveness to bronchoconstrictive agents. 
    Healthy adults do not typically respond to low levels of NO2
    (< 1880 µg/m3, 1 ppm).  However, asthmatics appear to be the most
    susceptible members of the population (section 6.2.1.2).  Asthmatics
    are generally much more sensitive to inhaled bronchoconstrictors.  The
    potential addition of an NO2-induced increase in airway response to
    the already heightened responsiveness to other substances raises the
    possibility of exacerbation of asthma by NO2.  Another potentially
    susceptible group includes patients with COPD (section 6.2.1.3).  A
    major concern with COPD patients is the absence of an adequate
    pulmonary reserve, so that even a relatively small alteration in lung
    function in these individuals could potentially cause serious
    problems.  In addition, both adolescents and the elderly have
    been evaluated, to determine whether differential age-related
    susceptibility exists (section 6.2.1.4).

        Table 39.  Effects of nitrogen dioxide (NO2) on lung function and airway responsiveness of healthy subjectsa
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    188          0.1          60                                     15 M           23-29 years,       No symptoms; no odour   Hazucha et al.
                                                                                    NS                 detection; no effect    (1982, 1983)
                                                                                                       on SRaw.

    188          0.1          240                                    6              Normal adults      No effects of NO2       Sackner et al.
    564          0.3                                                                                                           (1980)
    940          0.5
    1880         1.0

    226          0.12         60                                     4 M/6 F        13-18 years        No effects on lung      Koenig et al.
                                                                                                       function.               (1985)

    226          0.12         40         10           32.5           3 M/7 F        14-19 years        No effects on Rtau or   Koenig et al.
    338          0.18         40                                     4 M/6 F        15-19 years        spirometry.             (1987a,b)

    230          0.12         20                                     5 M/4 F        20-36 years,       Suggestion of change    Bylin et al.
    460          0.24                                                               NS                 in SRaw in normals:     (1985)
    910          0.48                                                                                  SRaw tended to increase
                                                                                                       at 476 µg/m3 and
                                                                                                       tended to decrease at
                                                                                                       910 µg/m3.  Analysis of
                                                                                                       variance indicates no
                                                                                                       significance.  No effects
                                                                                                       on bronchial  reactivity.
                                                                                                       Median odour threshold
                                                                                                       75 µg/m3.
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    282          0.15         120        60           50 W           6 M            19-24 years        No symptoms; no         Kagawa & Tsuru
                                                                                                       pulmonary function      (1979); Johnson
                                                                                                       effects.  Suggested     et al. (1990)
                                                                                                       individual changes
                                                                                                       in SGaw.

    338          0.18         30         10 (L)       L approx. 25   9 M            18-23 years,       No change in lung       Kim et al.
    564          0.3                     16 (H)       H approx. 72                  "collegiate        function.               (1991)
                                                                                    athletes"

    508          0.27         60                                                    Healthy,           Possible small          Rehn et al.
    1993         1.06                                                               young M            increase in Raw at      (1982)
                                                                                                       508 µg/m3 (0.27 ppm).

    564          0.3          120        60           50 W           6              19-25 years        No effect on SGaw.      Kagawa (1986)

    564          0.3          225        30           approx. 40     10 M/10 F      20-48 years        No symptom, lung        Morrow & Utell
                                         (3 × 10)                                   (FEV1/FVC          function or airway      (1989)
                                                                                    76-95%)            reactivity responses
                                                                                                       to carbachol for either
                                                                                                       of the 20-48 year or
                                                                                                       the 49-69 year age
                                                                                                       groups.

    564          0.3          225        21           30-40          10 M/10 F      49-69 years,
                                         (3 × 7)                                    (FEV1/FVC
                                                                                    72-84%)
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    940          0.5          120        15           Light/         10             Healthy, three     Decreased quasistatic   Kerr et al.
                                                      moderate                      ex-smokers in      compliance. Non-random  (1979)
                                                                                    group              exposure sequence air-
                                                                                                       NO2. No change in
                                                                                                       spirometry or resistance.
                                                                                                       Apparent compliance
                                                                                                       change may be due to
                                                                                                       exposure order.

    940          0.5          120        15                          10             Normal adults      Decreased static lung   Kulle (1982)
                                                                                                       compliance.

    940          0.5          240        30           55             10 M           26.4 years         No significant effects  Stacy et al.
                                                                                                       on spirometry or Raw.   (1983)

    1128         0.6          120        60           25             8 M/8 F        51-76 years        No statistically        Drechsler-Parks
                                                                                                       significant changes     et al. (1987)
                                                                                                       in lung function due
                                                                                                       to NO2 exposure in
                                                                                                       either age group.

                                                                     8 M/8 F        18-26 years,
                                                                                    NS

    1128         0.6          180        60           approx. 40     7 M/2 F        Healthy, NS        No change in            Frampton et al.
                                         (6 × 10)                                                      spirometry, Raw or      (1989a)
                                                                                                       carbachol reactivity.
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    94 with      0.05 with    135        60                          11 M/4 F       Non-reactive
    3760         2.0 spikes   3 × 15     (6 × 10)                                   (carbachol)
    spikes

    (1) 1128     (1) 0.6      180        60           39             6 M/2 F        30.3 ± 1.4         There were no changes   Frampton et al.
                                                                                    years, NS          in airway mechanics     (1991)
                                                                                                       (FVC, FEV1, SGaw).
                                                                                                       Responsiveness to
    (2) Var.     (2) Var.     180        60           43             11 M/4 F       25.3 ± 1.2         carbachol was
    (94          (0.05                                                              years, NS          significantly increased
    background   background                                                                            after 2820 µg/m3 NO2
    with         with         180        60           approx. 40     5 M/3 F        32.6 ± 1.6         (Group 3) but not after
    3 × 15 min   3 × 15                                                             years, NS          the other exposures
    at 3760)     min at                                                                                (Groups 1 and 2). Degree
                 2.0 ppm)                                                                              of baseline
                                                                                                       responsiveness to
                                                                                                       carbachol was not
                                                                                                       related to response after
                                                                                                       2820 µg/m3.

    (3) 2820     (3) 1.5      180        60           39             12 M/3 F       23.5 ± 0.7
                                                                                    years, NS
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    1128         0.6          120/day    60           approx. 30-40  4 M/1 F        NS, 21-36          No effects of repeated  Boushey et al.
                              for 4 days                                            years, FEV1/       NO2 exposure on         (1988) (Part 2)
                                                                                    FVC% range         respiratory function
                                                                                    73-83%,            (SRaw, FVC, FEV1) or
                                                                                    "normal"           symptoms.
                                                                                    methacholine
                                                                                    responsiveness

    1128         0.6          60         60           70             20 M           Healthy            No effect of NO2 on     Adams et al.
                                                      50             20 F                              spirometry or airway    (1987)
                                                                                                       resistance.

    1166         0.62         120        15           33             5 M            Healthy            No significant          Folinsbee et al.
                                         30           33             5 M                               pulmonary function      (1978)
                                                                                                       responses attributed
                                                                                                       to NO2 exposure.

    1316-3760    0.7-2.0      10                                     10                                Increased resistance    Suzuki &
                                                                                                       10 min after exposure.  Ishikawa (1965)

    1316         0.7          60                                     5              19-22 years,       No effects on airway    Toyama et al.
                                                                                    3 of 5 were        conductance.            (1981)
                                                                                    investigators
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    1880         1.0          120 (2     60           Light          16             Healthy            Air-NO2-NO2 fixed       Hackney et al.
                              consecutive                                                              exposure sequence.      (1978)
                              days)                                                                    1.5% decrease in FVC
                                                                                                       after second day of
                                                                                                       NO2. Not clear that
                                                                                                       the decreased FVC is
                                                                                                       an NO2 effect or an
                                                                                                       order effect. No other
                                                                                                       effects.

    1880         1.0          120/day,                               22             Healthy, NS,       Overall trend for a     Goings et al.
    3760         2.0          3 days                                 21, 22         seronegative       slight decrement in     (1989)
    5640         3.0                                                 22                                FEV1 with NO2 exposure
                                                                                                       (< 1%).  No change in
                                                                                                       methacholine
                                                                                                       responsiveness as a
                                                                                                       result of NO2 exposure
                                                                                                       or viral infection
                                                                                                       status.

    1880         1.0          120                                    16             11 S               After 14 100 µg/m3      Beil & Ulmer
    4700         2.5          120                                    16             5 NS               (120 min) and           (1976)
    9400         5.0          120                                    16                                9400 µg/m3 (14 h),
    14 100       7.5          120                                    16             8 S                responsiveness to
    9400         5.0          840                                    8                                 acetylcholine increased.
                                                                                                       Resistance increased
                                                                                                       after all but the
                                                                                                       1880 µg/m3 exposure.
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    3760         2.0          60                                     8 M/3 F        18-36 years,       Vitamin C blocked       Mohsenin (1987b)
                                                                                    NS                 NO2-induced increase
                                                                                                       in airway reactivity
                                                                                                       to methacholine.

    3760         2.0          120                                    13 M/5 F       Normal, NS,        No symptoms; no lung    Mohsenin (1988)
                                                                                    18-33 years        function changes.
                                                                                                       Increased methacholine
                                                                                                       reactivity.

    7520-9400    4.0-5.0      10                                                                       Bag exposure            Abe (1967)
                                                                                                       technique. Airway
                                                                                                       resistance increased
                                                                                                       30 min after end of
                                                                                                       exposure. No change in
                                                                                                       spirometry.

    7520         4.0          75         15 (L)       L 20-29        16 M/ 9 F      18-45 years,       No change in SRaw       Linn & Hackney
                                         15 (H)       H 44-57                       NS                 associated with NO2.    (1983); Linn
                                                                                                       Small but significant   et al. (1985b)
                                                                                                       decrease in blood
                                                                                                       pressure; some mild
                                                                                                       increase in symptoms.

    9400         5.0          15                                     16             Healthy            Decreased DLCO 18%.     Von Nieding et
                                                                                                                               al. (1973a)
                                                                                                                                              

    Table 39  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristice
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    9400         5.0          120        Intermittent Light          11 M           Healthy            Increased resistance    Von Nieding et
                                                                                                       60%. Remained elevated  al. (1977)
                                                                                                       for 60 min. Possible
                                                                                                       decrease in PaO2.

    9400         5.0          120        60           220            11 M           Healthy            Resistance increased    Von Nieding et
                                         (4 × 15)                                                      60%. Remained elevated  al. (1979)
                                                                                                       60 min after exposure.
                                                                                                       Possible decrease in
                                                                                                       earlobe PO2.
                                                                                                                                              

    a  Modified from US EPA (1993)
       Abbreviations:
       M = Male; F = Female; S = Active smoker; NS = Non-smoker; FEV1 = Forced expiratory volume in 1 second; FVC = Forced vital capacity;
       SRaw = Specific airway resistance; Var = Variable; Raw = Airway resistance; SGaw = Specific airway conductance; W = Watts; L = Light;
       H = Heavy; RT = Total respiratory resistance; DLCO = Diffusing capacity for carbon monoxide; PaO2 = Arterial partial pressure of oxygen;
       PO2 = Partial pressure of oxygen
        6.2.1.1  Nitrogen dioxide effects in healthy subjects

         The effects of NO2 levels greater than 1880 µg/m3 (1.0 ppm) on
    respiratory function in healthy subjects have been examined in several
    studies (Table 39). Early work indicated that NO2 increased Raw
    or total respiratory resistance (RT) at concentrations above
    2820 µg/m3 (1.5 ppm) in healthy volunteers (Abe, 1967; Von Nieding et
    al., 1970, 1973a, 1979; Von Nieding & Wagner, 1977).  Although Beil &
    Ulmer (1976) found a small but statistically significant increase in
    RT after a 2-h exposure to > 4700 µg/m3 (> 2.5 ppm) NO2, the
    response was not appreciably increased by raising the NO2
    concentration to 9400 or 14 100 µg/m3 (5.0 or 7.5 ppm).  Also, airway
    responsiveness to acetylcholine was increased after exposure to
    14 100 µg/m3 for 2 h or to 9400 µg/m3 for 14 h, but not after the
    2-h exposures to < 9400 µg/m3.

         In contrast, some investigators found no effects at high
    concentrations.  For example, a 75-min exposure with light and heavy
    exercise to 7520 µg/m3 (4.0 ppm) NO2 did not affect Raw (Linn et
    al., 1985b), and a 1-h resting exposure to 3760 µg/m3 (2 ppm) did not
    cause a change in lung volume, flow-volume characteristics on either
    full or partial expiratory flow-volume (PEFV) curves, or SGaw
    (Mohsenin, 1987b, 1988).  However, NO2 did increase airway
    responsiveness to methacholine (Mohsenin, 1987b, 1988).

         Goings et al. (1989) found no effects of exposure to NO2 at
    1880, 3760 or 5640 µg/m3 (1, 2 or 3 ppm; for 2 h/day on 3 consecutive
    days) on respiratory symptoms, lung function or airway reactivity to
    methacholine.  Laboratory-induced influenza virus infection did not
    alter airway responsiveness in either sham (clean air) or NO2
    exposure groups.  The infectivity portion of this study is discussed
    in section 6.2.2.

         The influence of exposure pattern was examined by Frampton et al.
    (1991), using healthy subjects exposed for 3 h to either 1128 µg/m3
    (0.60 ppm), 2820 µg/m3 (1.5 ppm) or a variable concentration protocol
    where three 15 min peaks of 3760 µg/m3 (2.0 ppm) were added to a
    background level of 94 µg/m3 (0.05 ppm).  Nitrogen dioxide did not
    affect airway mechanics (forced vital capacity (FVC), FEV1, SGaw). 
    However, after exposure to 2820 µg/m3, but not to the other
    concentrations, there was a small but statistically significant
    increase in airway responsiveness to carbachol.  This study supported
    the earlier observations by Mohsenin (1987b, 1988) of increased airway
    responsiveness after a 1-h exposure to 3760 µg/m3.  Mohsenin (1987b)
    further observed that the NO2-induced increase in airway
    responsiveness could be blocked by elevation of serum ascorbate level
    through pretreatment with the antioxidant ascorbic acid (vitamin C).

         At concentrations below 1880 µg/m3 (1.0 ppm) NO2, pulmonary
    function and airway responsiveness have generally not been found to be
    affected in healthy adult subjects (Beil & Ulmer, 1976; Folinsbee et
    al., 1978; Hackney et al., 1978; Kerr et al., 1979; Sackner et al.,
    1980; Toyama et al., 1981; Kulle, 1982; Hazucha et al., 1982, 1983;
    Stacy et al., 1983; Kagawa, 1986; Adams et al., 1987; Drechsler-Parks
    et al., 1987; Drechsler-Parks, 1987; Boushey et al., 1988; Morrow &
    Utell, 1989; Frampton et al., 1989a, 1991; Kim et al., 1991).
    Although some investigators have at times reported statistically
    significant effects, there does not appear to be a consistent pattern
    of acute responses in healthy subjects at these low NO2
    concentrations.

         Kagawa & Tsuru (1979) reported the lowest NO2 exposure
    concentration that appeared to cause an effect.  Healthy men were
    exposed to 282 µg/m3 (0.15 ppm) NO2 for 2 h while performing light,
    intermittent exercise.  The authors suggested that NO2 caused some
    statistically significant changes, i.e. a 0.5% decrease in vital
    capacity (VC) and a 16% decrease in an index of small airway function
    (i.e. FEF75HeO2: FEF75AIR; the ratio of forced expiratory flow at 75%
    FVC expired while breathing a helium-oxygen mixture compared to FEF75
    while breathing air).  These findings should be interpreted with the
    consideration that multiple t-tests were used in the statistical
    analysis of these data.  Rehn et al. (1982) reported a small (17%)
    increase in SRaw in men exposed to 500 µg/m3 (0.27 ppm) for 1 h, but
    a higher concentration (2000 µg/m3, 1.06 ppm) did not cause an
    effect.

         Bylin et al. (1985) reported that the SRaw of normal subjects
    exposed to 230, 460 and 910 µg/m3 (0.12, 0.24 and 0.48 ppm) for
    20 min was unaffected. Specific comparisons revealed a significant 11%
    increase in SRaw at 460 µg/m3 (0.24 ppm) and a 9% decrease in SRaw
    at 910 µg/m3.  Bronchial responsiveness to histamine was increased by
    910 µg/m3 NO2.

         Symptomatic responses of subjects exposed to NO2 were evaluated
    in several of the above studies.  None of these studies, including
    exposures for as long as 75 min to 7520 µg/m3 (4.0 ppm) NO2 (Linn &
    Hackney, 1983; Linn et al., 1985b), resulted in a significant increase
    in respiratory symptoms.  In studies of sensory effects, subjects were
    unable to detect the odour of 188 µg/m3 (0.1 ppm) NO2 (Hazucha et
    al., 1983), but Bylin et al. (1985) observed an odour threshold of
    75 µg/m3 (0.04 ppm) for normal subjects and 150 µg/m3 (0.08 ppm) for
    asthmatics.

    6.2.1.2  Nitrogen dioxide effects on asthmatics

         Studies of the effects of exposures to NO2 on respiratory
    function and airway responsiveness of asthmatics are summarized in
    Table 40.  Asthmatics are generally more responsive than healthy
    subjects to NO2.  However, as can be seen in Table 40, there is
    substantial variability in observed responses between and even within
    laboratories.  This variability is illustrated in Fig. 22 and 23, in
    which changes in airway resistance and FEV1 are related to the
    "exposure dose" of NO2 (calculated as ppm × litres of air breathed
    over the duration of exposure) (US EPA, 1993).  The individual
    investigations that yielded the data used to develop these
    illustrations will be discussed in more detail below.  Other studies,
    not discussed separately, are also summarized in Table 40.  The review
    by the US EPA (1993) provides more detail on many of these studies. 
    Although differences in exposure protocols may explain some of the
    differences between studies, the explanation most often invoked is
    that there may be differences in the severity of asthma among the
    subject groups tested.  There are numerous definitions of "asthma
    severity" (see, for example, National Institutes of Health, 1991). 
    Those applied to the key asthma studies discussed here (based on the
    data available) are:  (1) mild: controlled by bronchodilators and
    avoidance of known precipitating factors, does not interfere with
    normal activities; and (2) moderate: often requires periodic use of
    inhaled steroids in treatment and may interfere with work or school
    activities.  Those with severe asthma are seldom used as subjects for
    NO2 studies because their disease can include life-threatening
    episodes.  Typical volunteers for the studies described here had mild
    allergic asthma.

         Avol et al. (1988) studied a group of moderate-to-severe
    asthmatics exposed to 564 and 1128 µg/m3 (0.3 and 0.6 ppm) NO2 for
    2 h with moderate intermittent exercise. NO2 did not cause
    significant changes in SRaw or FEV1.  Results of tests of airway
    responsiveness to cold air suggested a slightly increased response
    after exposure to 564 µg/m3, but not after 1128 µg/m3. A post hoc
    analysis of a subgroup of subjects with the most abnormal lung
    function (i.e., FEV1/FVC ratios < 0.65) did not find enhanced
    susceptibility. In a subsequent study using 564 µg/m3 NO2, Avol et
    al. (1989) found decreases in FEV1, FVC and peak expiratory flow rate
    (PEFR), but no change in responsiveness to cold air challenge.

        Table 40.  Effects of nitrogen dioxide (NO2) on lung function and airway responsiveness of asthmaticsa
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    188          0.1          60                                     9              20-51 years,       No effect of NO2 on     Ahmed et al.
                                                                                    "history of        FEV1, SGaw or on        (1983a)
                                                                                    bronchial asthma"  ronchial reactivity to
                                                                                                       ragweed antigen, either
                                                                                                       immediately or 24 h
                                                                                                       after exposure.

    188          0.1          60                                     20 M/34 F      18-39 years        No significant effect   Ahmed et al.
                                                                                                       on SGaw, FEV1, VISOV;  (1983b)
                                                                                                       variable effect on
                                                                                                       carbachol reactivity.
                                                                                                       No information on
                                                                                                       controlled exposure.

    188          0.1          60                                     15 M           21-46 years,       No significant          Hazucha et al.
                                                                                    mild or inactive   changes in RT or        (1982, 1983)
                                                                                    disease            responsiveness to
                                                                                                       methacholine associated
                                                                                                       with NO2 exposure.

    207          0.11         60                                     6 M/1 F        1 Smoker,          No change in SRaw or    Orehek et al.
    (132-301)    (0.07-0.16)                                                        3 asthmatic,       in responsiveness to    (1981)
                                                                                    4 allergic         grass pollen in 3
                                                                                                       allergic asthmatics
                                                                                                       and 4 allergic
                                                                                                       subjects.
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    210          0.11         60                                     13 M/7 F       15-44 years,       13/20 subjects had      Orehek et al.
    (169-244)    (0.09-0.13)                                                        13 mild/7 mod      enhanced responses to   (1976)
                 (n = 20)                                                           asthmatics;        carbachol after
                                                                                                       210 µg/m3 NO2. Post
                                                                                                       hoc statistical
                                                                                                       analysis questionable.

    489          0.26                                                               65 years           1/4 subjects had        Orehek et al.
                 (n = 4)                                                                               enhanced responses      (1976)
                                                                                                       to carbachol after
                                                                                                       489 µg/m3 NO2.

    226          0.12         60                                     4 M/6 F        12-18 years,       No significant effects  Koenig et al.
                                                                                    asympt.,           on pulmonary function   (1985)
                                                                                    extrinsic          due to NO2. Increased
                                                                                    allergic           symptoms after NO2
                                                                                    asthmatics         exposures.

    226          0.12         60                                     4 M/6 F        12-18 years        No change in FEV1,      Koenig et al.
    226          0.12         40         10           33             4 M/6 F        11-19 years        RT increased 10.4%      (1987a,b)
    338          0.18         40         10           39             7 M/3 F        12-18 years,       (NS), 3% decrease
                                                                                    asympt.,           in FEV1 (p < 0.06).
                                                                                    extrinsic
                                                                                    allergic
                                                                                    asthmatics
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    230          0.12         20                                     6 M/2 F        17-45 years,       No significant change   Bylin et al.
    460          0.24         20                                                    very mild          in SRaw at any NO2      (1985)
    910          0.48         20                                                    asympt.            levels. Histamine
                                                                                                       reactivity tended to
                                                                                                       increase.

    260          0.14         30                                     8 M/12 F       17-56 years,       Overall trend for SRaw  Bylin et al.
    510          0.27                                                               very mild          to decline during       (1988)
    1000         0.53                                                               asympt.            exposure period, not
                                                                                                       related to NO2
                                                                                                       concentration.
                                                                                                       Histamine bronchial
                                                                                                       reactivity tended to
                                                                                                       increase after 260 and
                                                                                                       510 µg/m3 NO2 exposure.

    376          0.2          120        60           approx. 20     12 M/19 F      18-55 years,       No effects on           Kleinman et al.
                                                                                    wide range of      spirometry or airway    (1983)
                                                                                    asthma severity    resistance. Airway
                                                                                                       reactivity to
                                                                                                       methacholine results
                                                                                                       variable-tended to
                                                                                                       increase with exposure.
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    470          0.25         30         10           30             9 M/2 F        18-55 years,       Mouthpiece exposure     Joerres &
                                                                                    mild asympt.       system. No changes in   Magnussen
                                                                                                       methacholine            (1991)
                                                                                                       responsiveness were
                                                                                                       observed after NO2
                                                                                                       exposure.

    470          0.25         30                                     10 M/4 F       20-55 years,       After NO2 exposure,     Joerres &
                                                                                    mild asthma,       responsiveness to       Magnussen
                                                                                    most asympt.       inhaled SO2 was         (1990)
                                                                                                       increased. No effect
                                                                                                       of NO2 alone on SRaw.

    564          0.3          30         20           approx. 30     5 M/4 F        23-34 years        No changes in SRaw,     Rubinstein et
                                                                                                       FVC, FEV1, SBN2 or      al. (1990)
                                                                                                       symptoms after NO2
                                                                                                       exposure. NO2 exposure
                                                                                                       did not increase airway
                                                                                                       responsiveness to SO2.
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    564          0.3          30         10           30             15             20-45 years,       Resting 20 min          Bauer et al.
                                                                                    mild asympt.       exposures produced no   (1986)
                                                                                                       effects. Slight excess
                                                                                                       decrease in FEV1 and
                                                                                                       PEFR in NO2 plus
                                                                                                       exercise above that
                                                                                                       caused by exercise
                                                                                                       alone. PEFR, -16% (air).
                                                                                                       -28% (NO2); FEV1 -5.5%
                                                                                                       (air), -9.3% (NO2).
                                                                                                       Significantly increased
                                                                                                       response to cold air
                                                                                                       after NO2 exposure.

    564          0.3          225        30           30-40          10 M/10 F      19-54 years        Group findings          Morrow & Utell
                                         (3 × 10)                                                      indicated no            (1989)
                                                                                                       significant responses.
                                                                                                       No change in lung
                                                                                                       function, symptoms,
                                                                                                       carbachol reactivity.
                                                                                                       Subjects studied
                                                                                                       previously (Bauer et
                                                                                                       al., 1986) showed
                                                                                                       possible responses to
                                                                                                       NO2. New subject
                                                                                                       subgroup showed
                                                                                                       significantly greater
                                                                                                       response in air
                                                                                                       exposures.
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    564          A. 03        110        60           42             A. 13 M        19-35 years,       FEV1 decreased 11%      Roger et al.
                                                                                    mild asthmatics    in NO2 but only 7% in   (1990)
                                                                                                       air, after first 10 min
                                                                                                       of exercise. Smaller
                                                                                                       changes later in
                                                                                                       exposure.

    282          B. 0.15      75         30           42             B. 21                             No increase in airway
    564          0.3                                                                                   reactivity to
    1128         0.6                                                                                   methacholine 2 h after
                                                                                                       exposure. Nochange in
                                                                                                       FEV1 or SRaw as a
                                                                                                       result at NO2
                                                                                                       exposure.

    564          0.3          180        90           30             24 M/10 F      10-16 years        After 60 min of         Avol et al.
                                                                                                       exposure, FEV1, FVC     (1989)
                                                                                                       and PEFR (-3.4, -4.0
                                                                                                       and -5.6%,
                                                                                                       respectively) were
                                                                                                       significantly reduced.
                                                                                                       No change in airways
                                                                                                       responsiveness to cold
                                                                                                       air challenge. SRaw
                                                                                                       increased 17% after NO2
                                                                                                       exposure. After 180 min
                                                                                                       of exposure, the
                                                                                                       responses had returned
                                                                                                       to baseline levels.
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    564          0.3          120        60           40             27 M/32 F      18-50 years,       Exercise-related        Avol et al.
                                                                                    some moderate      increases in symptoms.  (1988)
                                                                                    asthmatics         Possible NO2-related
                                                                                                       decrease in FEV1,
                                                                                                       PEFR. Increased cold
                                                                                                       air response after
                                                                                                       564 µg/m3.

    1128         0.6          120        60           41                                               More consistent
                                                                                                       increases in SRaw at
                                                                                                       1128 µg/m3 but not
                                                                                                       significantly different
                                                                                                       from air and 564 µg/m3.

    564          0.3          60         30           41             15 M/6 F       20-34 years,       No effect of NO2.       Linn et al.
    1880         1.0          60         30           41                            mild asthmatics    Exercise-related        (1986)
    5640         3.0          60         30           41                                               increase in SRaw under
                                                                                                       all conditions.

    940          0.5          120        15                          9 M/4 F        19-50 years,       Increased respiratory   Kulle (1982)
                                                                                    3 Smokers          symptoms in 4/13
                                                                                                       subjects. Also,
                                                                                                       increased static lung
                                                                                                       compliance. Impossible
                                                                                                       to determine amount of
                                                                                                       effect due to NO2.
                                                                                                                                              

    Table 40  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    940          0.5          60                                     10             22-44 years,       No change in symptoms.  Mohsenin (1987b)
                                                                                    mild asthmatics    Significant group mean
                                                                                                       increase in
                                                                                                       responsiveness to
                                                                                                       methacholine after
                                                                                                       NO2 exposure. No other
                                                                                                       function changes.

    940 +        0.5 +        120        60           approx. 20     6 M/12 F       33 years,          No significant effect   Linn et al.
    857          0.3 ppm                                             6 ex-smokers   physician-         on spirometry, RT.      (1980a)
    SO2          SO2                                                 asthma         diagnosed

    7520         4.0          75         a. 15        a. 25          12 M/11 F      18-34 years,       No NO2 effects on       Linn & Hackney
                                         b. 15        b. 49                         physician-         SRaw, symptoms, heart   (1984); Linn et
                                                                                    diagnosed          rate, skin              al. (1985b)
                                                                                    asthma             conductance. Small
                                                                                                       decrease in
                                                                                                       systolic blood
                                                                                                       pressure.
                                                                                                                                              

    a  Modified from US EPA (1993)
       M = male; F = female; SGaw = specific airway conductance; FEV1 = forced expiratory volume in 1 second; VISOV = volume of isoflow;
       PEFR = peak expiratory flow; SRaw = specific airway resistance; FVC = forced vital capacity; Asympt. = asymptomatic;
       RT = total respiratory resistance; NS = not significant; SO2 = sulfur dioxide; SBN2 = single breath nitrogen washout

    FIGURE 22

    FIGURE 23

         Roger et al. (1990) reported the effects of NO2 exposure on mild
    asthmatics.  Their first study was a pilot study of 12 mild asthmatics
    exposed to 564 µg/m3 (0.3 ppm) for 110 min, including three 10-min
    periods of exercise.  After the first 10 min of exercise in NO2,
    there was a decrease in FEV1 that persisted for the remainder of the
    exposure period, although the overall responses were progressively
    less with successive periods of exercise, as is common with
    exercise-induced asthma when the exercise is intermittent.  Their
    subsequent concentration-response study of twenty-one subjects 
    included six responsive subjects from the pilot study; volunteers were
    exposed to 282, 564 and 1128 µg/m3 (0.15, 0.30 and 0.60 ppm) NO2 for
    75 min, with three 10-min exercise periods.  In contrast to the pilot
    study, there were no effects of NO2 on pulmonary function or airway
    responsiveness to methacholine, tested 2 h after exposure ceased.  The
    authors suggested that the differences between the pilot and the main
    study may have been due to more reactive airways in the pilot study
    asthmatics.  Because the studies were conducted during different
    seasons, seasonal differences in temperature, air pollution, ambient
    aeroallergens or other factors may have contributed to some of the
    variability in response.

         Asthmatics exposed to 230, 460 and 910 µg/m3 (0.12, 0.24 and
    0.48 ppm) NO2 for 20 min were studied by Bylin et al. (1985). 
    Changes in SRaw during the four exposures averaged +3% after air and
    +9%, -2% and -14% after the three levels of NO2, respectively; these
    changes were not significantly different.  There was a tendency for an
    increase in thoracic gas volume (TGV) after NO2 exposures (9 to
    10%), but differences in pre-exposure values for TGV were probably
    responsible, rather than NO2.  There were no significant changes in
    tidal volume or respiratory rate.  At the highest concentration tested
    (910 µg/m3, 0.48 ppm), histamine bronchial responsiveness was
    increased.

         In mild asthmatics exposed for 30 min to 260, 510 and 1000 µg/m3
    (0.14, 0.27 and 0.53 ppm), there were no significant changes in SRaw,
    although there was a general trend for SRaw to fall throughout the
    period of exposure at all NO2 concentrations (Bylin et al., 1988). 
    There was, however, a significant increase (p = 0.03) in airway
    responsiveness to histamine after 30 min of exposure to 510 µg/m3
    (0.27 ppm) only. The absence of a concentration-related increase in
    responsiveness is not inconsistent with other studies.  This
    observation contrasts with earlier results (Bylin et al., 1985) that

    suggested a possible increased responsiveness after exposure to
    910 µg/m3 (0.48 ppm). Because of the use of a non-parametric pair
    comparison test that was not adjusted for multiple comparisons, the
    raw data presented in the paper were subjected to reanalysis (US EPA,
    1993) using a Friedman non-parametric analogue of an F test, which
    is probably more appropriate for these data than a series of
    Wilcoxon matched pairs signed rank tests.  This analysis showed no
    statistically significant change in histamine responsiveness due to
    NO2 exposure.

         Asthmatics exposed to 564 µg/m3 (0.3 ppm) NO2 by mouthpiece
    for 20 min at rest followed by 10 min of exercise (30 litres/min)
    experienced a statistically significant spirometric response to NO2
    (Bauer et al., 1986).  After NO2 exposure, 9 out of 15 asthmatics had
    a decrease in FEV1; both the pre-post exposure difference on the NO2
    day (10.1%) and the pre-post NO2 minus the pre-post air (i.e.,
    delta-delta) differences (6%) were significant using a paired t-test. 
    Maximum expiratory flow at 60% total lung capacity (PEFV curve) was
    also decreased, but FVC and SGaw were not altered.  Nine out of
    twelve subjects experienced an increase in airway responsiveness to
    cold air.  The mouthpiece exposure system used in this study contained
    relatively dry air (relative humidity, RH, of 9 to 14% at 20°C) and
    airway drying may have interacted with NO2 to cause greater
    responses.  However, Bauer et al. (1986) controlled for the airway
    drying effect by exposing subjects to clean air at the same
    temperature and RH.  Nevertheless, air temperature and humidity
    effects may be an important consideration for NO2 effects in winter
    in the temperate regions of the world.
 
        Linn et al. (1985b) and Linn & Hackney (1984) exposed mild
    asthmatics to 7520 µg/m3 (4.0 ppm) NO2 for 75 min, with two 15-min
    exercise periods.  There was no significant difference in lung
    function that could be attributed to NO2; if anything, SRaw tended
    to be slightly lower with the NO2 exposures.

         The reasons for the differences between the group of asthmatics
    exposed to 7520 µg/m3 (4 ppm) for 75 min (with exercise) (Linn et
    al., 1985b) and the group exposed to 564 µg/m3 (0.30 ppm) for 30 min
    with exercise studied by Bauer et al. (1986) are not clear.  The
    subjects of Bauer et al. were exposed to NO2 in dry air through a
    mouthpiece which could have caused some drying of the upper airways;
    Linn et al. (1985b) used a chamber exposure.  Second, the subjects in
    the Linn et al. (1985b) study tended to have milder asthma than the
    subjects in the Bauer et al. (1986) study.  There were differences in
    the season in which the two studies were conducted, and there may have
    been a difference in background exposure to NO2 (outdoors and/or
    indoors).  In addition, increased bronchial reactivity to cold air was
    an important finding in the Bauer et al. (1986) study, but it was not
    measured by Linn et al. (1985b).

 

         Further research was conducted by Linn et al. (1986) on mild
    asthmatics exposed to 564, 1880 and 5640 µg/m3 (0.30, 1.0 and
    3.0 ppm) NO2 for 1 h.  The exposures included intermittent, moderate
    exercise.  As in the previous study with 7520 µg/m3 (4.0 ppm) NO2,
    there were no significant effects of NO2 on spirometry, SRaw or
    symptoms.  Furthermore, there was no significant effect on airway
    responsiveness to cold air.  In order to examine the suggestion that
    the severity of response to NO2 may be related to the clinical
    severity of asthma, the authors selected three subjects characterized
    as having more severe illness.  Although they experienced markedly
    larger changes in resistance than other milder asthmatics under all
    exposure conditions, there was no indication that the responses of
    these subjects were related to NO2 exposure.

         Mohsenin (1987a) found no changes in symptoms, spirometry, or
    plethysmography in mild asthmatics exposed to 940 µg/m3 (0.5 ppm)
    NO2 for 1 h at rest.  However, airway responsiveness to methacholine
    increased after the NO2 exposure.

         The effects of previous NO2 exposure on SO2-induced
    bronchoconstriction has been  examined by Joerres & Magnussen (1990)
    and Rubinstein et al. (1990).  Neither study found changes in
    pulmonary function after NO2 exposure.  Joerres & Magnussen (1990)
    exposed mild-to-moderate asthmatic subjects to 470 µg/m3 (0.25 ppm)
    NO2 for 30 min while breathing through a mouthpiece at rest.  After
    the NO2 exposure, airway responsiveness to 1965 µg/m3 (0.75 ppm)
    SO2 was increased.  Rubinstein et al. (1990) exposed asthmatics to
    564 µg/m3 (0.30 ppm) NO2 for 30 min (including 20 min light
    exercise).  No mean change in responsiveness to SO2 occurred, but one
    subject showed a tendency toward increased responsiveness.  The
    reasons for the different findings in these two studies is not clear,
    especially as the subjects of Rubinstein et al. (1990) were exposed to
    a higher NO2 concentration and exercised during exposure.  However,
    Joerres & Magnussen's subjects appeared to have had slightly more
    severe asthma and were somewhat older.  The modest increase in SRaw
    caused by exercise in the Rubinstein et al. (1990) study may have
    induced a refractory state to SO2.  Finally, the different method of
    administering the SO2 bronchoprovocation test may have had an
    influence.  Joerres & Magnussen (1990) increased minute ventilation
    (V.E) at a constant SO2 concentration, whereas Rubinstein et al.
    (1990) increased SO2 concentration at constant VE.

         A number of studies of the effects of NO2 exposure in asthmatics
    on changes in airway responsiveness to bronchoconstrictors have been
    presented in Table 40, but not evaluated in the text.  Various types
    of inhalation challenge tests have been used (methacholine, histamine,
    cold air, etc.).  Some exposures were conducted at rest and others
    while performing some exercise.  For twenty studies for which
    individual data were available, a meta analysis (Folinsbee, 1992) was
    performed to assess the changes in airway responsiveness in asthmatics
    exposed to NO2.  The aim of the meta analysis was to examine the
    diversity of response seen in the various studies and to examine
    factors such as NO2 concentration, exercise, and airway challenge
    method that could help explain some of the variability in response. 
    Such questions could not be adequately addressed using individual
    studies.  The analysis provides only a qualitative examination of
    concentration-response relationships.  For this analysis, the
    directional change (i.e., increased or decreased) in airway
    responsiveness after NO2 exposure was determined for each subject. 
    The data were then organized by exposure concentration range and
    whether or not exposures included exercise.  Within each exposure
    category the fraction of subjects with increased airway responsiveness
    was determined (see Table 41).  For the total of 355 individual NO2
    exposures, 59% of the asthmatics had increased responsiveness.  If the
    response was not associated with NO2 exposure, the fraction would be
    expected to approach 50%.  The excess increase in responsiveness can
    be attributed primarily to the NO2 exposures conducted at rest
    (fraction was 69%).  There was a larger fraction of increased
    responsiveness during the resting exposures in all three concentration
    ranges (see Table 41).  In the exercising studies, however, there was
    no effect because only 51% had an increase in airway responsiveness. 
    There was a trend for a slightly larger percentage (approx. 75%) of
    subjects to have increased airway responsiveness after NO2 exposures
    above 376 µg/m3 (0.20 ppm) and under resting conditions.  Of those
    six studies independently reporting a statistically significant
    response (Kleinman et al., 1983; Bylin et al., 1985, 1988; Bauer et
    al., 1986; Mohsenin, 1987a; Joerres & Magnussen, 1990), four were
    resting exposures, and in four the exposure duration was 30 min or
    less.  Although the authors offered various hypotheses for this

    apparent effect of low-level NO2 resting exposures, the mechanisms
    are unknown.  Changes in responsiveness were seen with relatively
    brief exposures.  One possible explanation for the absence of
    response in the exercising exposures is that exercise-induced
    bronchoconstriction may interfere with the NO2-induced response or
    that prior exercise may cause the airways to become refractory to the
    effects of NO2.  Possible confounding influences of nitric oxide, not
    measured in most studies, cannot be determined.

    Table 41.  Fraction of nitrogen dioxide-exposed subjects with
               increased airway responsivenessa
                                                                        

    Nitrogen dioxide   All             Exposures      Exposure
    concentration      exposures       with exercise  at rest
    (ppm)
                                                                        

    Asthmatics

    0.05-0.20          0.64 (105)b     0.59 (17)      0.65 (88)b
    0.20-0.30          0.57 (169)      0.52 (136)     0.76 (33)b
    > 0.30             0.59 (81)       0.49 (48)      0.73 (33)c
    All NO2            0.59 (355)b     0.51 (202)     0.69 (154)b
      concentrations

    Healthy

    < 1.0              0.47 (36)                      0.47 (36)
    < 1.0              0.79 (29)b    0.73 (15)        0.86 (14)c
                                                                        

    a  Data are fraction of subjects with an increase in airways
       responsiveness above the value for clean air. Numbers in
       parenthesis indicate actual number of subjects in each category.
       Total number = 355.  Ties (i.e. no change) were excluded.
    b  p < 0.01 two-tailed sign test
    c  p < 0.05 two-tailed sign test

         A similar meta analysis for healthy subjects indicated increased
    airway responsiveness after exposure to NO2 concentrations greater
    than 1880 µg/m3 (1 ppm). Exercise during exposure did not appear to
    influence the responses as much in the healthy subjects as in the
    asthmatics, but a similar trend was evident.

    6.2.1.3  Nitrogen dioxide effects on patients with chronic obstructive
             pulmonary disease

         Patients with COPD represent an important potentially sensitive
    population group. Studies evaluating NO2 effects on respiratory
    function in COPD subjects are summarized in Table 42.  The results of
    two NO2 exposure studies (9400 to 15 040 µg/m3, 5 to 8 ppm NO2 for
    up to 5 min) were discussed by Von Nieding et al. (1980), who found
    that the responses of bronchitics were generally similar to those of
    healthy subjects.  There was a tendency for the response to NO2 to be
    greater in the subjects with the highest baseline Raw.  Percentage
    changes ranged from approximately 25 to 50%.  In a review of their
    studies, Von Nieding & Wagner (1979) showed that Raw increased in
    chronic bronchitics exposed to > 3760 µg/m3 (2.0 ppm) NO2.

         The responses of COPD patients were affected by exposure (with
    mild exercise) to 564 µg/m3 (0.3 ppm) NO2 for 3.75 h (Morrow &
    Utell, 1989).  Forced vital capacity showed progressive and
    significant decreases during and following NO2 exposure, the largest
    change of -9.6% occurring after 3.75 h of exposure.  Smaller
    decrements in FEV1 (-5.2%) occurred at the end of exposure. There was
    no effect of NO2 on SGaw or diffusing capacity.  The severity of
    disease (based on impairment of lung function: FEV1 < 60% predicted
    vs. > 60% predicted) generally did not influence the magnitude of
    response to NO2.  The COPD patients showed a decrement in FEV1
    compared to the healthy, elderly non-smokers who experienced an
    improvement in FEV1.  In contrast, Linn et al. (1985a) found no
    effects from a 1-h exposure (with exercise) to 940, 1880 and
    3760 µg/m3 (0.5, 1.0 and 2.0 ppm) NO2 in a diverse group of COPD
    patients.  The reasons for the marked difference in responses between
    the two studies are not known.  Ambient exposure to air pollution in
    general and NO2 in particular was probably much higher for the
    subjects in the Linn et al. (1985a) study.  Thus, attenuation of
    physiological responses may have been a factor.

         Hackney et al. (1992) studied effects of field exposure to
    ambient air and chamber exposure to 564 µg/m3 (0.3 ppm) NO2 in       
    older adults with evidence of COPD and a history of heavy smoking. 
    They reported only slight adverse effects of NO2.  The study did not
    strongly confirm the findings  of Morrow & Utell (1989) and Morrow et
    al. (1992), and the authors speculated that ambient exposure history
    may have been responsible for differences between these studies.

    6.2.1.4  Age-related differential susceptibility

         Studies evaluating possible age-related differences in
    susceptibility to NO2 effects on respiratory function in healthy
    subjects are summarized in Table 39.

         Research on asthmatics is summarized in Table 40.  Spirometry
    measurements of young (18 to 26 years old) and older (51 to 76 years
    old) men and women were not affected by exposure to 1128 µg/m3
    (0.6 ppm) NO2 with light intermittent exercise (Drechsler-Parks et
    al., 1987; Drechsler-Parks, 1987).  In addition, Morrow & Utell (1989)
    did not observe any pulmonary function or airway responsiveness
    effects due to a lower level of NO2 (564 µg/m3, 0.3 ppm) in young or
    elderly healthy subjects.

         Koenig et al. (1985) found no "consistent significant changes in
    pulmonary functional parameters" after 1-h resting exposures of
    asthmatic adolescents to 226 µg/m3 (0.12 ppm) NO2. Subsequent
    mouthpiece exposures to 226 µg/m3 NO2, with exercise, caused
    increases in RT and decreases in FEV1 after both air and NO2
    exposure, which were apparently due to exercise alone (Koenig et al.,
    1987a,b).  When subjects were exposed to a higher level of NO2
    (338 µg/m3, 0.18 ppm), no differences in RT occurred.  Decreases in
    FEV1 were -1.3 and -3.3% for air and NO2, respectively; this
    difference (p = 0.06) may indicate a possible response trend.

    6.2.2  Nitrogen dioxide effects on pulmonary host defences and
           bronchoalveolar lavage fluid biomarkers

         Nitrogen dioxide can enhance susceptibility to infectious
    pulmonary disease, as clearly demonstrated in the animal toxicological
    literature (chapter 5).  Epidemiological studies (chapter 7) suggest
    similar effects.  Human clinical studies of NO2 effects on host
    defences are summarized in Table 43.

         Kulle & Clements (1988) and Goings et al. (1989) (two reports of
    the same study) examined the effect of NO2 exposure on susceptibility
    to attenuated influenza virus.  Healthy adults were exposed for
    2 h/day for 3 days to either clean air or 1880, 3760 or 5640 µg/m3
    (0, 1.0, 2.0 or 3.0 ppm) NO2.  The virus was administered
    intranasally after the second day of exposure, and infectivity was
    defined as the presence of virus in nasal washes, a rise in either
    nasal wash or serum antibody titres to the virus, or both.  Although
    the rates of infection were elevated after NO2 exposure in some of
    the NO2-exposed groups (91% of subjects exposed to 1880 or
    3760 µg/m3 (1 or 2 ppm) infected vs. 71% of controls), the changes
    were not significant.  The investigators concluded that the results of
    the study were inconclusive, rather than negative, because the
    experimental design had a low power to detect a 20% difference in
    infection rate, decreasing the possibility of statistical
    significance.

        Table 42.  Effects of nitrogen dioxide on lung function and airway responsiveness of chronic obstructive pulmonary disease patientsa
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    564          0.3          225        21           25             13 M/7 F       47-70 years,       Total NO2 inhaled       Morrow &
                                         (3 × 7)                                    8 mild,            dose 1.215 mg.          Utell (1989)
                                                                                    12 moderate        Decrease in FVC after
                                                                                                       exposure-9.6%. 5.2%
                                                                                                       decline in FEV1
                                                                                                       significant after
                                                                                                       approx. 4-h exposure.

    564          0.3          240        28           25             15 M/11 F      47-69              No significant change   Hackney et al.
                                         (4 × 7)                                                       in FVC or FEV1 with     (1992)
                                                                                                       NO2 exposure

    940          0.5          120        15           25             7              24-53 years,       No effects in           Kerr et al.
                                                                                    daily cough        bronchitics alone.      (1979)
                                                                                    for 3 months       Possible decrease in
                                                                                                       quasistatic compliance.

    940          0.5          60         30           16             13 M/9 F       48-69 years,       No change in FVC,       Linn et al.
                                                                                    some with          FEV1, etc. at any NO2   (1985a)
    1880         1.0                                                                emphysema,         level. SRaw tended to
                                                                                    some with          increase after first
    3760         2.0                                                                chronic            exercise period.
                                                                                    bronchitis         Possible decrease in
                                                                                                       peak flow at
                                                                                                       3760 µg/m3. No symptom
                                                                                                       changes. No change
                                                                                                       in SaO2.
                                                                                                                                              

    Table 42  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    940-9400     0.5-5        15                                     88                                Decrease in earlobe     Von Nieding et
                                                                                                       blood PO2 at            al. (1971, 1970)
                                                                                                       > 7520 µg/m3.
                                                                                                       Increased Raw at
                                                                                                       > 3008 µg/m3.

    1880-9400    1-5          30 breaths                             84 M           30-72 years,       Increase in Raw         Von Nieding et
                              (15 min)                                              chronic non-       related to NO2          al. (1973a)
                                                                                    specific disease   concentration. No
                                                                                                       effect on Raw below
                                                                                                       2820 µg/m3.

    9400         5            60                                                                       Changes in PO2 of
                                                                                                       earlobe capillary
                                                                                                       blood. Change occurred
                                                                                                       in first 15 min,
                                                                                                       effect did not
                                                                                                       increase with further
                                                                                                       exposure.
                                                                                                                                              

    Table 42  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    1880-15 040  1-8 ppm      5-60                                   116            25-74 years        At 7520-9400 µg/m3      Von Nieding &
                                                                                                       for 15 min, PaO2        Wagner (1979)
                                                                                                       decreased
                                                                                                       (arterialized capillary
                                                                                                       blood). Raw increased
                                                                                                       with exposure to
                                                                                                       > 3008 µg/m3.
                                                                                                                                              

    a  Modified from US EPA (1993)
       Abbreviations: FVC = Forced vital capacity; FEV1 = Forced expiratory volume in 1 second; PaO2 = Arterial partial pressure of oxygen;
       PO2 = Partial pressure of oxygen; Raw = Airway resistance; SRaw = Specific airway resistance; SaO2 = Arterial oxygen saturation

    Table 43.  Effects of nitrogen dioxide on host defences of humansa
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    508          0.27         60                                     M              Healthy, young     No change in nasal or   Rehn et al.
    1993         1.06                                                                                  tracheobronchial        (1982)
                                                                                                       clearance.

    (1) 1128     (1) 0.6      180        60           39             6 M/2 F        30.3 ± 1.4         Total NO2 uptake (1)    Frampton et al.
                                                                                    years, healthy,    3.4 mg (2) 5.6 mg, (3)  (1989b)
                                                                                    NS                 approx.3.3 mg (4)
                                                                                                       8.1 mg. BAL fluid
                                                                                                       analysis showed no
                                                                                                       significant effect on
                                                                                                       total protein or
                                                                                                       albumin
    (2) Var      (2) Var      180        60           43             11 M/4 F       25.3 ± 1.2         Apparent increase in
    (94          (0.05                                                              years, healthy,    alpha-2-macro-globulin
    background   background                                                         NS                 3.5 h after exposure
    with 3 ×     with 3 × 15                                                                           to 0.6 ppm (Group 1)
    15 min at    min at                                                                                but not after the
    3760)        2.0 ppm)                                                                              other protocols. No
                                                                                                       changes in percentage
                                                                                                       of lymphocytes or
                                                                                                       neutrophils. Concluded
                                                                                                       that NO2 at these
                                                                                                       concentrations neither
    (3) 1128     (3) 0.6      180        60           approx. 40     5 M/3 F        32.6 ± 1.6         altered epithelial
                                                                                    years, healthy,    permeability nor
                                                                                    NS                 caused inflammatory
                                                                                                       cell influx.
                                                                                                                                              

    Table 43  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    (4) 2820     (4) 1.5      180        60           39             12 M/3 F       23.5 ± 0.7
                                                                                    years, healthy,
                                                                                    NS

    1128         0.6          120/day    60           approx. 30-40  4 M/1 F        21-36 years,       Slight increase in      Boushey et al.
                              for 4 days                                            Healthy, NS.       circulating (venous)    (1988) (Part 2)
                                                                                    FEV1/FVC%          lymphocytes:
                                                                                    range 73-83%,      1792 ± 544 per mm3
                                                                                    "normal"           (post-NO2) vs.
                                                                                    methacholine       1598 ± 549 per mm3
                                                                                    responsiveness     (baseline). No change
                                                                                                       in BAL lymphocytes
                                                                                                       except an increase in
                                                                                                       natural killer cells:
                                                                                                       7.2 ± 3.1% (post-NO2)
                                                                                                       vs. 4.2 ± 2.4%
                                                                                                       (baseline). No change
                                                                                                       observed in IL-1
                                                                                                       or TNF.

    1128         0.6          180        60           approx. 40     7 M/2 F        Healthy, NS        No change in cell       Frampton et al.
                                         (6 × 10)                                                      recovery or             (1989a)
                                                                                                       differential counts.
                                                                                                       Possible decrease in
                                                                                                       macrophage
                                                                                                       inactivation of virus
                                                                                                                                              

    Table 43  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    94 with      0.05 with    135        60                          11 M/4 F       Nonreactive        in vitro. Possible
    3760         2.0 spikes   3 × 15     (6 × 10)                                   (carbachol), no    sensitive subgroup.
    spikes                                                                          recent upper
                                                                                    resp. infection

    1880         1.0          180        Intermittent                3 M/5 F        Healthy            No responses.           Jorres et al.
                                                                                                                               (1992)

    1880         1.0          120/day                                22             Healthy, NS,       Study conducted over    Goings et al.
    3760         2.0          3 days                                 21, 22         seronegative       3-year period. NO2 did  (1989)
    5640         3.0                                                 22                                not significantly
                                                                                                       increase viral
                                                                                                       infectivity, although
                                                                                                       a trend was observed.
                                                                                                       This study had a low
                                                                                                       power to detect small
                                                                                                       differences in
                                                                                                       infection rate.

    3760         2.0          240        120          50             10             Healthy, NS        Increased bronchial     Devlin et al.
                                                                                                       PMN's and decreased     (1992); Becker
                                                                                                       macrophage phagocytosiset al. (1993)

    3760         2.0          360        Intermittent                12             Healthy, NS        Immediate and 18-h      Frampton et al.
                                                                                                       post-BAL increase       (1992)
                                                                                                       in PMN.
                                                                                                                                              

    Table 43  (Con't)
                                                                                                                                              

    NO2 concentration         Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    4230         2.25         20         20           approx. 35     8              Healthy, NS        Increased levels of     Sandstroem et
    7520         4.0                                                 8                                 mast cells in BAL       al. (1989)
    10 340       5.5                                                 8                                 fluid at all
                                                                     Total n = 18                      concentrations.
                                                                                                       Increased numbers of
                                                                                                       lymphocytes at
                                                                                                       > 7520 µg/m3
                                                                                                       (BAL 24-h
                                                                                                       post-exposure).

    7520         4.0          20 min-    20           approx. 35     8              Healthy, NS        Total cell counts       Sandstroem et
                              alternate                                                                were reduced. Alveolar  al. (1990a)
                              days for                                                                 macrophages had
                              12 days                                                                  enhanced phagocytic
                                                                                                       activity but fewer
                                                                                                       were present.
                                                                                                       Decreased numbers
                                                                                                       of mast cells, T and
                                                                                                       B lymphocytes, and
                                                                                                       natural killer cells
                                                                                                       (BAL 24-h
                                                                                                       post-exposure).
                                                                                                                                              

    a  Modified from US EPA (1991)

       Abbreviations: M = Male; F = Female; NS = Non-smoker; FEV1 = Forced expiratory volume in 1 second; FVC = Forced vital capacity;
       BAL = Bronchoalveolar lavage; IL-1 = Interleukin-1; TNF = Tumour necrosis factor; VAR = Variable
             Others investigated the effects of NO2 on cells and fluids in
    bronchoalveolar lavage (BAL) of healthy adults.  Frampton et al.
    (1989a) used two different exposure protocols that had the same
    concentration × time product.  One group was exposed for 3 h to
    1128 µg/m3 (0.6 ppm), whereas the other was exposed to a background
    level of 94 µg/m3 (0.05 ppm) with three 15-min spikes of 3760 µg/m3
    (2.0 ppm).  Both exposures included exercise.  Pulmonary function and
    airway responsiveness were not affected.  Alveolar macrophages (AM)
    obtained by BAL after exposure to 1128 µg/m3 NO2 tended to
    inactivate virus less effectively than AM collected after air
    exposure.  The AMs that showed the impairment of virus inactivation
    also showed an increase in interleukin-1 production, not seen in the
    AMs from other subjects.  Interleukin-1 is a proinflammatory protein
    produced by AMs, which performs a number of immunoregulatory
    functions, including induction of fibroblast proliferation, activation
    of lymphocytes, and chemotaxis for monocytes.  The study had
    relatively low statistical power to detect an effect.  Becker et al.
    (1993) reported no change in virus inactivation properties of alveolar
    macrophages lavaged from subjects exposed to 3760 µg/m3 (2 ppm) for
    4 h.

         Using exposures similar to the above, with the addition of two
    groups exposed to 2820 µg/m3 (1.5 ppm) NO2 for 3 h, one with BAL at
    3.5 h post-exposure and the other with BAL at 18 h post-exposure,
    Frampton et al. (1989b) examined changes in protein in BAL fluid.  The
    total protein and albumin content of BAL fluid obtained at either 3.5-
    or 18-h post-exposure was not changed.  In BAL fluid obtained 3.5 h
    after exposure to 1128 µg/m3 (0.60 ppm) there was an increase in
    alpha-2-macroglobulin, a regulatory protein that has antiprotease
    activity and immunoregulatory effects.  This response was not seen in
    the group lavaged at 18 h post-exposure and no such effect occurred at
    a higher NO2 concentration (2820 µg/m3).

         Sandstroem et al. (1989) exposed healthy subjects to 4230, 7520
    and 10 340 µg/m3 (2.25, 4.0 and 5.5 ppm) for 20 min (with moderate
    exercise) and performed BAL 24 h after exposure.  Increased numbers of
    mast cells were observed at all NO2 concentrations; numbers of
    lymphocytes were increased only at > 7520 µg/m3.  In order to
    determine the time course of this response, Sandstroem et al. (1990a)
    exposed four groups of healthy subjects to 7520 µg/m3 NO2 for 20 min
    (mild exercise) and then performed BAL 4, 8, 24 or 72 h after
    exposure. Increased numbers of mast cells and lymphocytes were
    observed at 4, 8 and 24 h but not at 72 h.  There was no change in the
    numbers of AMs, eosinophils, polymorphonuclear leukocytes, T cells or
    epithelial cells, or in the albumin concentration of lavage fluid. 
    The authors interpreted the increased numbers of mast cells and
    lymphocytes as a nonspecific inflammatory response.

         Sandstroem et al. (1990b) also evaluated responses to repeated
    NO2 exposures.  Healthy subjects were exposed to 7520 µg/m3
    (4.0 ppm) NO2 for 20 min/day (with moderate exercise) on alternate
    days over a 12-day period (seven exposures in all); BAL was performed

    24 h after the last exposure.  The first 20 ml of BAL fluid was
    treated separately and presumed to represent primarily bronchial cells
    and secretions; subsequent fractions presumably were from the alveolar
    region.  In the first fraction, there was a reduction in the numbers
    of mast cells and AMs; AM phagocytic activity (on a per cell
    basis) was increased.  In addition, there were reduced numbers of
    T-suppressor cells, B cells and natural killer (NK) cells in the
    alveolar portion of the BAL.  This pattern of cellular response
    contrasts with that after single NO2 exposure (Sandstroem et al.,
    1990a).

         Rubinstein et al. (1991) studied five healthy volunteers exposed
    for 2 h/day for 4 days to 1128 µg/m3 (0.60 ppm) NO2 with
    intermittent exercise.  A slight increase in circulating (venous
    blood) lymphocytes was observed.  The only change observed in BAL
    cells was a modest increase in the percentage of NK cells, suggesting
    a possible increase in immune surveillance.

         Three recent studies examined the effects of longer exposures to
    1880 or 3760 µg/m3 (1.0 to 2.0 ppm) NO2 on lavaged cells and
    mediators.  Devlin et al. (1992) (also Becker et al., 1993) studied
    healthy subjects exposed to 3760 µg/m3 NO2 for 4 h with alternating
    15-min periods of rest and moderate exercise.  One of the main
    findings after NO2 exposure was that there was a three-fold increase
    in PMNs in the first lavage sample, representing predominantly
    bronchial cells and fluid.  In addition, macrophages recovered from
    the predominantly alveolar fraction showed a 42% decrease in ability
    to phagocytose  Candida albicans  and a 72% decrease in release of
    superoxide anion.  In another study, Frampton et al. (1992) exposed
    exercising subjects to 3760 µg/m3 NO2 for 6 h.  Bronchoalveolar
    lavage was performed either immediately or 18 h after exposure.
    There was a modest increase in lavage fluid PMN levels (< two-fold
    increase) but no change in lymphocytes.  Alveolar macrophage
    production of superoxide anion was not altered in these subjects. 
    These two studies suggest that NO2 exposure may induce a mild
    bronchial inflammation and may also lead to impaired macrophage
    function.  On the other hand, Joerres et al. (1992) examined both
    healthy and asthmatic subjects exposed to 1880 µg/m3 NO2 for 3 h,
    but observed no changes in cells or mediators in BAL fluid or in the
    appearance of bronchial mucosal biopsies after this exposure.  Neither
    macrophage function nor a specific bronchial washing were examined in
    this study.

         Rehn et al. (1982) reported that a 1-h exposure to either 500 or
    2000 µg/m3 (0.27 or 1.06 ppm) NO2 did not alter nasal or
    tracheobronchial mucociliary clearance rates.

    6.2.3  Other classes of nitrogen dioxide effects

         There have been isolated reports that higher levels of NO2
    (> 7520 µg/m3, 4.0 ppm) can decrease arterial oxygen partial
    pressure (PaO2) (Von Nieding & Wagner, 1977; Von Nieding et al.,
    1979) and cause a small decrease in systemic blood pressure (Linn et
    al., 1985b).  However, the impact of such changes is not clear,
    especially considering the high concentrations of NO2 required.

         The effects of NO2 on the constituents of BAL fluid, blood and
    urine have been examined in very few studies and are reviewed in more
    detail elsewhere (US EPA, 1993).  The general purpose of this research
    was to examine mechanisms of pulmonary effects or to determine whether
    NO2 exposure could result in systemic effects.  Investigations of the
    effects of NO2 on levels of serum enzymes and antioxidants have been
    conducted, but few effects were found and they cannot be interpreted
    (Posin et al., 1978; Chaney et al., 1981).  For example, Chaney et al.
    (1981) found an increase in glutathione levels, but Posin et al.
    (1978), using a higher NO2 concentration, did not find such an
    effect.  Studies of exposure to NO2 concentrations between 2820 and
    7520 µg/m3 (1.5 and 4.0 ppm) found either slight or no changes in BAL
    levels of alpha-1-antitrypsin, which inhibits protease activity
    (Mohsenin & Gee, 1987; Johnson et al., 1990; Mohsenin, 1991). Healthy
    subjects exposed to 7520 µg/m3 NO2 (Mohsenin, 1991) at rest for 3 h
    showed increased lipid peroxidation products in BAL fluid obtained
    immediately after exposure. In addition, the activity or the elastase
    inhibitory capacity (EIC) of alpha-1-protease inhibitor (alpha-1-PI)
    was decreased after NO2 exposure.  However, vitamin C supplementation
    for 4 weeks prior to NO2 exposure markedly attenuated the EIC
    response and resulted in a lower level of lipid peroxidation products. 
    The author suggested that the reduced activity of alpha-1-PI may have
    implications for the pathogenesis of emphysema, especially in smokers. 
    At a lower NO2 concentration (3760 µg/m3, 2.0 ppm, for 4 h), Becker
    et al. (1993) reported no change in alpha-1-antitrypsin.  Potential
    effects of NO2 on collagen metabolism have been investigated by
    examining urinary excretion of collagen metabolites after a 3-day
    (4 h/day) exposure to 1128 µg/m3 (0.6 ppm) NO2, but no effects were
    found (Muelenaer et al., 1987).

    6.3  Effects of other nitrogen oxide compounds

         Relatively few controlled human exposure studies have been
    conducted that evaluate NOx species other than NO2.  Such studies
    are summarized in Table 44 and concisely discussed here.

        Table 44.  Effects of other nitrogen oxide (NOx) compounds on humansa
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    HNO2         0.004        210                                    15             Healthy            A dose-dependent        Kjaergaard et
                 0.077                                               (11 M/4 F)     22-57 years        vasodilation in         al. (1993)
                 0.395                                                                                 bulbar conjunctiva.
                                                                                                       Significant increase
                                                                                                       of polymorphonuclear
                                                                                                       neutrophils, cuboidal
                                                                                                       and squamous epithelium
                                                                                                       cell counts in the
                                                                                                       tear fluid

    HNO3

    129          0.050        40         10           approx. 25-30  5 M/4 F        12-17 years,       FEV1 decreased -4.4%    Koenig et al.
                                                                                    asthmatic          after HNO3 and -1.7%    (1989a)
                                                                                                       after HNO3 plus air
                                                                                                       exposure. RT increased
                                                                                                       +22.5% after HNO3 and
                                                                                                       +7.4% after air
                                                                                                       exposure.

    200          0.078        120        100          Mod.           4 M/1 F        Healthy            In BAL, increase in     Becker et al.
                                                                                                       AM phagocytosis and     (1991)
                                                                                                       AM infection
                                                                                                       resistance.

    500          0.194        240        240          40             10             Healthy            No effect on FEV1,      Aris et al.
                                                                                                       FVC, SRaw or BAL        (1991)
                                                                                                       cells.
                                                                                                                                              

    Table 44  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    NO

    1230         1.0          120        60           50 W           8 M            19-24 years        Suggested change in     Kagawa (1982)
                                                                                                       density dependance of
                                                                                                       expired flow.

    12 300-      10-39        15                                     191            Healthy,           Increase in total       Von Nieding et
    47 970                                                                          20-50 years        respiratory resistance  al. (1973b)
                                                                                                       at > 24 600 µg/m3 and
                                                                                                       a decrease in PaO2 at
                                                                                                       > 18 450 µg/m3.

    NH4NO3

    200          (1.1 MMAD)   120        60           approx. 20     20             Normal             No significant changes  Kleinman et al.
                                                                     19             Asthmatic          due to NH4NO3 in        (1980)
                                                                                                       normals or asthmatics
                                                                                                       except possible
                                                                                                       decrease in RT.
                                                                                                       No symptoms and
                                                                                                       effects.

    80 + 940     (0.55 MMAD)  240        30           55             12             Normal             No effects.             Stacy et al.
    µg/m3        +0.5 ppm                                                                                                      (1983)
    NO2          NO2
                                                                                                                                              

    Table 44  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    NaNO3

    10, 100,     (0.2 MMAD)   10                                     5              Normal             No effects.             Sackner et al.
    1000                                                             5              Asthmatic                                  (1979)

    1000                                                             6              Normal
                                                                     6              Asthmatic

    7000         (0.46        16 (× 2)                               10             Normal             No effects.             Utell et al.
                 MMAD)        32 (total)                             11             Mild asthmatics                            (1979)

    7000         (0.49        16 (× 2)                               11             Influenza          No symptoms. SGaw       Utell et al.
                 MMAD)        32 (total)                                            patients           decrease 17% and VE    (1980)
                                                                                                       max 40% TLC decreased
                                                                                                       by 12% after nitrate,
                                                                                                       within 2 days of onset
                                                                                                       of illness. Similar
                                                                                                       effect 1 week later
                                                                                                       but not 3 weeks later.
                                                                                                                                              

    a  Modified from US EPA (1993)
       Abbreviations:

       W = Watt; M = Male; PaO2 = Arterial partial pressure of oxygen; HNO3 = Nitric acid; FEV1 = Forced expiratory volume in 1 second;
       FVC = Forced vital capacity; SRaw = Specific airway resistance; BAL = Bronchoalveolar lavage; AM = Alveolar macrophage; F = Female;
       RT = Total respiratory resistance; NS = Not significant; MMAD = Mass median aerodynamic diameter; SGaw = Specific airway conductance;
       VE max 40% TLC = Maximum expiratory flow at 40% of total lung capacity on a partial expiratory flow-volume curve
             Von Nieding et al. (1973b) exposed healthy subjects and smokers
    to 12 300 to 47 970 µg/m3 (10 to 39 ppm) NO for 15 min.  Total
    respiratory resistance increased significantly (approx. 10-12%) after
    exposure to > 24 600 µg/m3 (> 20 ppm) NO.  Diffusing capacity
    was not changed, but a small decrease (7 to 8 torr) in PaO2 was noted
    between 18 450 and 36 900 µg/m3 (15 to 30 ppm).  Kagawa (1982)
    examined the effects of a 1230 µg/m3 (1 ppm) NO exposure for 2 h in
    normal subjects.  A few individuals had increases in SGaw, and a few
    had decreases.  Analysis of the group mean data produced only one
    apparently statistically significant change: an 11% decrease in flow
    at 50% FVC in a helium-air mixture compared to this flow in air. 
    However, because the data were analysed by multiple t-tests the
    results should be interpreted with this in mind.

         NO is naturally formed in the body from the amino acid L-arginine
    and performs a second messenger function in several organ systems.  It
    has been measured in expired air (Gustafsson et al., 1991) and causes
    vasodilation in the pulmonary circulation.  Recently, NO has been used
    clinically to treat pulmonary hypertension in COPD patients and in
    infants with persistent pulmonary hypertension of the newborn (Zapol
    et al., 1994).

         In healthy volunteers made hypoxic by breathing 12% oxygen in
    nitrogen, the inhalation of 49 403 µg/m3 (40 ppm) NO prevented the
    hypoxia-induced increase in pulmonary artery pressure (Frostell et
    al., 1993).  Systemic arterial pressure was not changed.  No
    evaluation of effects on lung function were performed.  Adnot et al.
    (1993) studied a group of COPD patients who had pulmonary artery
    pressures averaging 32 mmHg.  They breathed 6130 to 49 403 µg/m3
    (5 to 40 ppm) NO for successive 10-min periods.  There was a
    dose-dependant decrease in pulmonary artery pressure during NO
    inhalation and no alteration of systemic arterial pressure.  Moinard
    et al. (1994) observed a 20% drop in pulmonary artery pressure in COPD
    patients after breathing 18 391 µg/m3 (15 ppm) NO for 10 min.  Based
    on an improvement in alveolar ventilation in some segments of the
    lung, the authors postulated that NO may also act as a bronchodilator. 
    Hoegman et al. (1993) suggested a modest bronchodilator effect of
    98 080 µg/m3 (80 ppm) NO.  Based on findings in animals, which are
    summarized in chapter 5, NO does cause bronchodilation at similar
    concentrations (Barnes, 1993).

         Nitrous acid and nitric acid may be formed from the reaction of
    NO2 with water.  Nitrous acid may also be produced directly in the
    combustion process.

         Koenig et al. (1989a) examined the responses of adolescent
    asthmatics to a 40-min exposure to 129 µg/m3 (0.05 ppm) HNO3 vapour
    via a mouthpiece exposure system.  After 30 min of rest and 10 min of
    exercise while breathing HNO3, there was a 4.4% decrease in FEV1
    compared to a 1.7% decrease after air  breathing.  A 22.5% increase in
    total respiratory resistance was also observed after HNO3 exposure,
    compared to a 7.4% increase after air breathing.

         The effects of HNO3 on BAL endpoints have been reported.  Becker
    et al. (1992) exposed healthy subjects to 200 µg/m3 (0.078 ppm) HNO3
    for 120 min, including 100 min of moderate exercise.  Bronchoalveolar
    lavage performed 18-h after exposure indicated increased phagocytic
    activity of AMs and increased resistance to respiratory syncytial
    virus infection.  There were no changes in markers of tissue damage. 
    Aris et al. (1991) exposed healthy subjects to 500 µg/m3 (0.194 ppm)
    HNO3 for 4 h, including moderate exercise.  No change in lactate
    dehydrogenase levels, lavage fluid protein or differential cell counts
    in the BAL were observed.  Pulmonary function (FEV1, FVC and SRaw)
    was not significantly affected.

         Kjaergaard et al. (1993) studied the effects of nitrous acid on
    the eyes of 15 healthy non-smokers exposed to 8, 148 or 758 µg/m3
    (4, 77 or 395 ppb) for 3.5 h.  There was an increase in trigeminal
    sensitivity (CO2 induced eye irritation) related to the concentration
    of nitrous acid.  Eye inflammation was increased, as indicated by
    increased PMNs and epithelial cells in tear fluid.

         Neither sodium nitrate (NaNO3) nor ammonium nitrate caused
    effects on pulmonary function of normal or asthmatic subjects (Sackner
    et al., 1979; Utell et al., 1979; Kleinman et al., 1980; Stacy et al.,
    1983).  However, there was a decrease in airway conductance and in
    PEFV curves in normal subjects with acute influenza exposed to
    7 mg/m3 of NaNO3 aerosol (Utell et al., 1980).  This is several
    orders of magnitude above the nitrate concentrations found in most
    ambient air.

    6.4  Effects of nitrogen dioxide/gas or gas/aerosol mixtures on lung
         function

         Table 45 summarizes studies of human subjects exposed to
    NO2-containing pollutant mixtures.  Most of the studies have been
    limited primarily to spirometry and plethysmography.  More extensive
    discussion can be found in US EPA (1993).

         With a few exceptions (to be discussed below), most research on
    interactions either showed no effects of the individual pollutants or
    the mixture, or it indicated that NO2 did not enhance the effects of
    the other pollutant(s) in the mixture (Table 45).  Most attention has
    focussed on NO2 mixtures with ozone (O3), although combinations with
    SO2, NO, particles, and a mixture of SO2 plus O3 have also been
    tested.  Due to the varied exposure protocols in the database, no
    consistent physiological trends are evident.  The generally negative
    responses could either reflect a true lack of interaction or other
    important design considerations.  For example, asthmatics were not
    studied.  Because pulmonary function studies of NO2 alone  cause
    variable effects with no clear concentration-responses, detecting
    interactions would be expected to be difficult unless there was
    significant synergism.

        Table 45.  Effects of nitrogen dioxide mixtures on healthy subjectsa
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    75 NO2       0.04 NO2     60         60           56             42 M/8 F       Healthy            No apparent effect      Avol et al.
                 (Amb)                                                                                 over and above that     (1983)
                                                                                                       of O3 alone.

    75 NO2       0.04 NO2     60         60           22.4           33 M/33 F      Children,          No effects of ambient   Avol et al.
                 (Amb)                                                              8-11 years         air exposures.          (1985a, 1987)

    103 NO2      0.055 NO2    60         60           32             46 M/13 F      Adolescents,       Ambient air exposures   Avol et al.
                 (Amb)                                                              12-15 years        effect  attributed      (1985b)
                                                                                                       to O3.

    132 NO2      0.07 NO2     120        60           approx. 20     14 M/20 F      29 years           Small decreases in      Linn et al.
                 (Amb)                                                                                 FVC, FEV1, in ambient   (1980b)
                                                                                                       air mostly attributable
                                                                                                       to O3. No association
                                                                                                       of NO2 levels with lung
                                                                                                       function change.

    545 NO2      (a) 0.29 NO2 240 (2     120          approx. 20     4              Healthy            With each group,        Hackney et al.
    +980 O3      +0.50 O3     consecutive                                                              minimal alterations     (1975b)
                              days of                                                                  in pulmonary function
    545 NO2      (b) 0.29 NO2 exposure                                                                 caused by O3 exposure.
    +980 O3      +0.50 O3     to each                                                                  Effects were not
    +34 350      +30.0 CO     mixture)                                                                 increased by addition
    CO                                                                                                 of NO2 or NO2 plus CO
                                                                                                       to test atmospheres.
                                                                                                                                              

    Table 45  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    545 NO2      (a) 0.29 NO2 120 (2     60           approx. 20     7              Healthy            Little or no change     Hackney et al.
    +490 O3      +0.25 O3     consecutive                                                              in pulmonary function   (1975b)
                              days of                                                                  found with O3 alone.
    545 NO2      (b) 0.29 NO2 exposure)                                                                Addition of NO2 or of
    +490 O3      +0.25 O3                                                                              NO2 plus CO did not
    +34 350      +30.0 CO                                                                              noticeably increase
    CO                                                                                                 the effect. Seven
                                                                                                       subjects included;
                                                                                                       some believed to be
                                                                                                       unusually reactive
                                                                                                       to respiratory
                                                                                                       irritants.

    940 NO2      0.50 NO2     120        30           40             10 M           Young adults,      FEV1, decreased 8-14%.  Folinsbee
    +980 O3      +0.5 O3                                                            NS                 No differences between  et al. (1981)
                                                                                                       O3 plus NO2 and O3
                                                                                                       alone.

    1128 NO2     0.60 NO2     120        60           25             8 M/8 F        18-26 years,       No significant          Drechsler-Parks
    +882 O3      +0.45 O3                                                           NS                 changes attributable    (1987)
                                                                                                       to NO2.

                                                                     8 M/8 F        51-76 years        Tendency (p > 0.05)
                                                                     8 M/8 F        51-76 years        for NO2 plus O3 to be
                                                                                                       greater than O3 alone.

    1128 NO2     0.60 NO2     60         60           70             20 M           Healthy            No additional effect    Adams et al.
    +588 O3      + 0.30 O3                            50             20 F                              of NO2 over and above   (1987)
                                                                                                       effect of O3.
                                                                                                                                              

    Table 45  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    1128         0.60 ppm NO2 120        60           40             21 F           Healthy, NS        NO2 exposure increased  Hazucha et al.
    NO2                                                                                                airway responses to     (1994)
                 0.3 ppm O3   120        60           40                                               methacholine after a
                 (3 h later)                                                                           subsequent O3 exposure.

    282 NO2      0.15 NO2     120        60           approx. 25     6 M            Some               Possible small          Kagawa (1986)
    +294 O3      + 0.15 O3                                                          smokers            decrease in SGaw.
    +200         + H2SO4
    H2SO4

    282 NO2      0.15 NO2     120        60           approx. 25     3 M            Some               Possible small
    +294 O3      + 0.15 O3                                                          smokers            decrease in FEV1.
    +393 SO2     + 0.15 SO2
    +200         + H2SO4
    H2SO4

    564 NO2      0.30 NO2     120        20           approx. 25     6 M            Some               Possible small
    +588 O3      +0.30 O3                                                           smokers            decrease in SGaw.
    +200         + H2SO4
    H2SO4

    282 NO2      0.15 NO2     120        60           approx. 25     7 M            19-23 years        No significant          Kagawa
    +294 O3      +0.15 O3                                                                              enhancement of the      (1983a,b)
    +393 SO2     +0.15 SO2                                                                             effects of O3 and/or
                                                                                                       SO2 by presence of
                                                                                                       NO2.
                                                                                                                                              

    Table 45  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    301 NO2      0.16 NO2     480        0                           15             16-26 years        No change in FVC,       Islam &
    +157 O3      +0.08 O3                                                                              acetylcholine airway    Ulmer (1979b)
    +891 SO2     +0.34 SO2                                                                             reactivity.

    564 NO2      0.3 NO2      120        60                          6 F            19-25 years        No significant effects  Kagawa (1990)
    +738 NO      +0.6 NO                                                            NS                 on pulmonary function
                                                                                                       or airway
                                                                                                       responsiveness to
                                                                                                       acetylcholine.

    940 NO2      0.50 NO2     135        60           approx. 20     11 M/9 F       20-53 years        No effects on           Kleinman
    1310 SO2     + 0.5 SO2                                                                             function; possible      et al. (1985)
    +26          +                                                                                     symptom responses.
    Zn(NH4)2     Zn(NH4)2                                                                              NO2 effects not
    (SO4)2       (SO4)2                                                                                discernible from
    +330 NaCl    + NaCl                                                                                mixture.

    940 NO2      0.50 NO2     120        60           approx. 20     10 M/14 F      26 ± 4             No significant effect   Linn et al.
    1310 SO2     + 0.50 SO2                                                         years, 21 NS,      on lung function in     (1980a)
                                                                                    3 S                normals. Trend for a
                                                                                                       slight decrease in
                                                                                                       FVC after combined
                                                                                                       exposure.
                                                                                                                                              

    Table 45  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    7520-9400    4-5 NO2      10                                     5 M            21-40 years,       Time course of          Abe (1967)
    NO2          +4-5 SO2                                                           4 NS, 1 S          response different.
    +4920-6150                                                                                         SO2 alone had immediate
    SO2                                                                                                increase in resistance;
                                                                                                       NO2 had delayed
                                                                                                       increase. Mixture had
                                                                                                       intermediate effects
                                                                                                       on resistance.

    9400 NO2     5.0 NO2      120        60           ?              8 M            < 30 years         FVC (-5%), FEV1.0       Islam &
    +1960 O3     +0.1 O3                                             8 M            30-40 years        (-11.7%), decreased     Ulmer (1979a)
    +13 100 SO2  +5.0 SO2                                            8 M            > 49 years         with exercise exposure
                                                                                                       to this mixture
                                                                                                       in < 30 years group.

    9400 NO2     5.0 NO2      120        intermittent                9 M            Healthy,           No interaction on PaO2  Von Nieding
    +196 O3      +0.1 O3                                                            20-38 years        or RT                   et al. (1977)
    +13 100 SO2  +5.0 SO2

    9400 NO2     5.0 NO2      120        intermittent                11 M           Healthy,           No interaction on PaO2
    +196 O3      +0.1 O3                                                            20-38 years,       or Rt
                                                                                    25 S, 9 NS

    9400 NO2     5.0 NO2      120        60           approx. 20                    23-38 years,       RT increased from 1.5   Von Nieding
    +196 O3      + 0.1 O3                             (70 W)                        two atopic         to 2.4 (p < 0.01);      et al. (1979)
    +13 100      + 5.0 SO2                                                                             questionable decrease
    SO2                                                                                                in PaO2 (8 torr).
                                                                                                                                              

    Table 45  (Con't)
                                                                                                                                              

    Concentrations            Exposure   Exercise     Exercise       Number of      Subject            Effects                 Reference
                              duration   duration     ventilation    subjects/      characteristics
                              (min)      (min)        (litres/min)   gender
    µg/m3        ppm
                                                                                                                                              

    188 NO2      0.1 NO2      120        60           approx. 20                    23-38 years,       No effects.
    +786 SO2     +0.3 SO2                                                           two atopic
                                                                                                                                              

    a  Modified from US EPA (1993)
       Abbreviations:

       Amb = Ambient air; CO = Carbon monoxide; F = Female; FEV1 = Forced expiratory volume in 1 second; FEV1.0 = Forced expiratory volume
       in 1 second; FVC = Forced vital capacity; H2SO4 = Sulfuric acid; M = Male; NaCl = Sodium chloride; (NH4)2SO4 = Ammonium sulfate;
       NO = Nitric oxide; NS = Non-smoker; O3 = Ozone; PaO2 = Arterial partial pressure of oxygen; RT = Total respiratory resistance;
       S = Active smoker; SGaw = Specific airway conductance; SO2 = Sulfur dioxide; W = Watts; ZnSO4 = Zinc sulfate
             Abe (1967) studied brief exposures to NO2-SO2 mixtures.  Both
    gases were at 4 to 5 ppm (i.e., 7520 to 9400 µg/m3 NO2 and 4920 to
    6150 µg/m3 SO2).  The effects were additive, with both gases causing
    bronchoconstriction.  Independently, the effect of SO2 was immediate
    and short-lasting, whereas the effect of NO2 was delayed and more
    persistent.  The effect of the mixed gases was intermediate between
    the two independent responses.  Kagawa (1983a,b) reported that the
    interaction of 282 µg/m3 (0.15 ppm) NO2 plus 393 µg/m3 (0.15 ppm)
    SO2 in normal  subjects exposed for 2 h with light intermittent
    exercise caused an increase in SGaw.  However, because a large number
    of repeated t-tests with an alpha level of 0.05 were used, it is
    possible that the responses were due to chance.

         The Rancho Los Amigos group (Linn et al., 1980b; Linn & Hackney,
    1983; Avol et al., 1983, 1985a, 1987) conducted several studies of
    NO2-containing ambient air mixtures.  The mean NO2 level in the
    ambient air (from the Los Angeles Air Basin) ranged from 75 to
    132 µg/m3 (0.04 to 0.07 ppm).  Normal and  asthmatic adults,
    adolescents and children were exposed for approximately 2 h during the
    summer smog seasons of 1978 to 1984.  The various pulmonary function
    effects observed (see Table 45) were attributed to O3.  However, in
    another study, Hazucha et al. (1994) found that ozone-induced
    increases in airway responsiveness to methacholine were enhanced by
    prior (3 h earlier) exposure to 1128 µg/m3 (0.60 ppm) NO2.  There
    was also a slightly greater FEV1 decrement after the NO2-O3
    sequence.

         There has been one study on the effects of HNO3 vapour in
    combination with O3 (Aris et al., 1991).  Ten healthy men were
    exposed (with moderate exercise) to 430 µg/m3 HNO3 for 2 h and then,
    after 1 h, to 392 µg/m3 (0.20 ppm) O3 for 3 h.  No changes were
    observed in FVC, FEV1 or SRaw after HNO3 exposure.  Ozone exposure
    caused increased SRaw and decreased FVC and FEV1.  Prior exposure to
    HNO3 vapour rather than air resulted in somewhat smaller changes in
    lung function after ozone exposure.  Clearly HNO3 did not potentiate
    responses to ozone.

    6.5  Summary of controlled human exposure studies of oxides of
         nitrogen

         Human responses to a variety of oxidized nitrogen compounds have
    been evaluated.  By far, the largest database and the one most
    suitable for risk assessment is that available for controlled
    exposures to NO2.  The database on human responses to NO, nitric acid
    vapour, nitrous acid vapour and inorganic nitrate aerosols is not as
    extensive.  A number of sensitive or potentially sensitive subgroups
    have been examined, including adolescent and adult asthmatics, older
    adults, and patients with chronic obstructive pulmonary disease and
    pulmonary hypertension.  Exercise increases the total uptake and

    alters the distribution of the inhaled material within the lung.  The
    proportion of NO2 deposited in the lower respiratory tract is also
    increased by exercise.  This may increase the effects of the above
    compounds in people who exercise during exposure.

         As is typical with human biological response to inhaled particles
    and gases, there is variability in the biological response to NO2. 
    Healthy individuals tend to be less responsive to the effects of NO2
    than  individuals with lung disease.  Asthmatics are clearly the most
    responsive group to NO2 that has been studied to date.  Individuals
    with chronic obstructive pulmonary disease may be more responsive than
    healthy individuals, but they have limited capacity to respond to NO2
    and thus quantitative differences between COPD patients and others are
    difficult to assess.  There is not sufficient information available at
    present to evaluate whether age or gender should be considered in the
    risk evaluation.

         NO2 causes decrements in lung function, particularly increased
    airway resistance in resting healthy subjects at 2-h concentrations as
    low as 4700 µg/m3 (approx. 2.5 ppm).  Available data are insufficient
    to determine the nature of the concentration-response relationship.

         NO2 exposure results in increased airway responsiveness to
    broncoconstrictive agents in exercising healthy, non-smoking subjects
    exposed to concentrations as low as 2800 µg/m3 (approx. 1.5 ppm) for
    exposure durations of 1 h or longer.

         Exposure of asthmatics to NO2 causes, in some subjects,
    increased airway responsiveness to a variety of provocative mediators,
    including cholinergic and histaminergic chemicals, SO2 and cold air. 
    The presence of these responses appears to be influenced by the
    exposure protocol, particularly whether or not the exposure includes
    exercise.  These responses may begin at concentrations as low as
    380 µg/m3 (0.20 ppm).  A meta analysis suggests that effects may
    occur at even lower concentrations.  However, no concentration-
    response relationship is observed between 350 and 1150 µg/m3
    (approx. 0.2 and 0.6 ppm).

         Modest increases in airway resistance may occur in patients with
    COPD from brief exposure (15-60 min) to concentrations of NO2 as low
    as 2800 µg/m3 (approx. 1.5 ppm) and decrements in spirometric
    measures of lung function (3 to 8%) change in FEV1 may also be
    observed with longer exposures (3 h) to concentrations as low as
    600 µg/m3 (approx. 0.3 ppm).

         Exposure to NO2 at levels above 2800 µg/m3 (approx. 1.5 ppm)
    may alter numbers and types of inflammatory cells in the distal
    airways or alveoli.  NO2 may alter the function of cells within the
    lung and production of mediators that may be important in lung host

    defences. The constellation of changes in host defences, alterations
    in lung cells and their activities, and changes in biochemical
    mediators is consistent with the epidemiological findings of increased
    host susceptibility associated with NO2 exposure.

         In studies of mixtures of NO2 with other pollutants, NO2 has
    not been observed to increase responses to other co-occurring
    pollutant(s) beyond what would be observed for the other pollutant(s)
    alone.  A notable exception is the observation that pre-exposure to
    NO2 enhances the ozone-induced change in airway-responsiveness in
    healthy, exercising subjects during a subsequent ozone exposure.  This
    observation suggests the possibility of delayed or persistent
    responses to NO2.

         Within an NO2 concentration range that may be of interest with
    regard to risk evaluation (i.e., 100-600 µg/m3), the characteristics
    of the concentration-response relationship for acute changes in lung
    function, airway responsiveness to bronchoconstricting agents, or
    symptoms cannot be determined from the available data.

         NO is acknowledged as an important endogenous second messenger
    within several organ systems.  Inhaled NO concentrations above
    6000 µg/m3 (approx. 5 ppm) can cause vasodilation in the pulmonary
    circulation without affecting the systemic circulation.  The lowest
    effective concentration is not established.  Information on pulmonary
    function and lung host defences consequent to NO exposure are too
    limited for any conclusions to be drawn at this time.  Relatively high
    concentrations (> 40 000 µg/m3) have been used in clinical
    applications for brief periods (< 1 h) without reported adverse
    reactions.

         Nitric acid levels in the range of 250-500 µg/m3 (100-200 ppb)
    may cause some pulmonary function responses in adolescent asthmatics,
    but not in healthy adults.

         Limited information on nitrous acid suggests that it may cause
    eye inflammation at 760 µg/m3 (0.40 ppm).  There are currently no
    published data on human pulmonary responses to nitrous acid.

         Limited data on inorganic nitrates suggest that there are no
    lung function effects of nitrate aerosols with concentrations of
    7000 µg/m3 or less.

    7.  EPIDEMIOLOGICAL STUDIES OF NITROGEN OXIDES

    7.1  Introduction

         This chapter discusses epidemiological evidence regarding effects
    of NOx on human health.  Primary emphasis is placed on assessment of
    the effects of NO2 because it is the oxide of nitrogen measured in
    most epidemiological studies and the one of greatest concern from a
    public health perspective.  Human health effects associated with
    exposure to NO2 have been the subject of several literature reviews
    since 1970 (National Research Council, 1971, 1977; US EPA, 1982a,
    1993; Samet et al., 1987, 1988).  Oxides of nitrogen have also been
    reviewed previously by the World Health Organization (WHO, 1977),
    which presented a comprehensive review of studies conducted up to
    1977.  This chapter focuses on studies conducted since 1977, while
    also using some key information from earlier literature, as reviewed
    in more detail by US EPA (1993).

         The studies discussed in this chapter are those that provide
    useful quantitative information on exposure-effect relationships for
    health effects associated with levels of NO2 likely to be encountered
    in the ambient air.  In addition, some studies that do not provide
    quantitative information are briefly discussed in the text in order to
    help elucidate particular points concerning the health effects of
    NO2.

    7.2  Methodological considerations

         Key epidemiological studies on NO2 health effects are evaluated
    below for several factors of importance for interpreting their results
    (US EPA, 1982a,c). Such factors include: (1) exposure measurement
    error; (2) misclassification of the health outcome; (3) adjustment
    for covariates; (4) selection bias; (5) internal consistency; and
    (6) plausibility of the effect based on other evidence.

    7.2.1  Measurement error

         Measurement error regarding exposure may be a major problem in
    epidemiological studies of NO2.  Ideally, personal monitors should be
    placed on all subjects for the entire period of a study, but this is
    often not feasible.  Moreover, personal monitoring may not overcome
    measurement error altogether.  For example, the monitors that are
    presently available do not accurately measure short-term peaks or
    long-term chronic exposures.  Other means of estimating NO2 exposure
    include source description, in-home monitors and fixed-site outdoor
    monitors.  These approaches are generally cheaper than personal
    monitors but may be subject to greater measurement error, both random
    (non-systematic) and systematic.

         In general, a measurement error in estimation of exposure that is
    independent of the health outcome will result in underestimation of
    associations between exposure and dichotomous health outcomes (Samet &
    Utell, 1990).  Whittlemore & Keller (1988) examined the data of Melia
    et al. (1980) and showed that a 20% misclassification rate of the
    exposure category could result in an underestimate of the logistic
    regression coefficient by as much as 50%.  Even when exposure
    measurement error is not independent of the outcome, measures of
    association are biased towards the null, unless the probability of the
    health outcome is very close to 0 or 1 (Stefanski & Carroll, 1985).

         At present, there is little information on the relative
    importance of peak and average NO2 levels as causes of respiratory
    effects in humans. In most homes and outdoor settings, peak values may
    be related to average values, and reduction of peaks may lower
    time-weighted averages.  However, if health effects are largely
    associated with the peak levels of NO2, then the use of averages as
    the sole guide to exposures will increase measurement error.

         NO2 may act as a precursor for other biologically active
    substances (such as nitrous acid).  If these agents are responsible
    for some or all of the observed respiratory effects, then measurement
    of NO2 will provide an imprecise estimate of the effective dose.

    7.2.2  Misclassification of the health outcome

         Misclassification of the health outcome can occur whether the
    outcome is continuous, (such as a measure of pulmonary function) or
    dichotomous (such as the presence or absence of respiratory symptoms). 
    Lung function is typically measured with spirometry, a well-
    standardized technique (Ferris, 1978).  The measurement errors of the
    instruments collecting the data have also been carefully estimated,
    and random errors will simply add to the error variance.  On the other
    hand, respiratory symptoms and health status are usually measured by a
    questionnaire.  Responses to symptom questions will be correlated and
    will depend on the interpretation of the respondent.  As noted below,
    a specific respiratory disease is likely to be reflected by a
    constellation of symptoms.  Therefore, it is appropriate to consider
    aggregate, as well as single, specific symptom reports.  Obviously,
    questionnaire measurements involving recent recall are better than
    those based on recall of events occurring several years earlier. 
    Questionnaires for cough and phlegm production have been standardized,
    e.g., the British Medical Research Council (BMRC) questionnaire
    (American Thoracic Society, 1969) and revisions of that questionnaire
    (Ferris, 1978; Samet, 1978).  These questionnaires and modifications
    of them have been used extensively.

    7.2.3  Adjustment for covariates

         It is common when analysing a data set to discover that one or
    more key covariates for the analysis were not measured.  Schenker et
    al. (1983) discussed socioeconomic status, passive smoking and gender
    as important covariates in childhood respiratory disease studies. 
    Other covariates often of importance are age, humidity and other
    co-occurring pollutants (e.g., particulate matter).  The concern is
    that, had missing covariates been measured, the estimate of the
    regression coefficient of a variable of interest would have been
    significantly different.  Although the problem is faced by most
    investigators, literature on the subject is sparse.  For example,
    Kupper (1984) showed that high correlations between the variables just
    described will result in "unreliable parameter estimates with large
    variances".  Gail (1986) considered the special case of omitting a
    balanced covariate from the analysis of a cohort study and concluded
    that: "In principle, the bias may be either toward or away from zero,
    though in more important examples - the bias is toward zero.  In
    important applications with additive or multiplicative regression,
    there is no bias".  Neither report provided information on how to
    attempt to correct for the bias or on approaches for investigating the
    possible bias in a given situation.

         Most studies of respiratory disease and NO2 exposure discussed
    here measured important covariates such as age, socioeconomic level of
    the parents, gender and parental smoking habits.  The estimated effect
    (regression coefficient of disease on NO2 exposure) will be
    overestimated if a missing covariate is positively or negatively
    correlated with both exposure and health outcome.  The estimated
    effect will be underestimated if positively correlated with exposure
    or outcome and negatively correlated with the other.  Ware et al.
    (1984) found that parents with some college education were more likely
    to report respiratory symptoms and less likely to use a gas stove,
    leading to an underestimate of the health effect, if education were
    omitted from the analysis.

    7.2.4  Selection bias

         The possibility of selection bias, although a concern of every
    study, seems very low for NO2 epidemiological studies.  Selection
    bias would require selection of participants based on exposure
    (e.g., use of gas stove) and also health outcome.  Because most
    epidemiological studies of these exposures are population based, there
    is little possibility of selection based on health end-points. 
    Nevertheless, the loss of subjects by attrition associated with both
    exposure and health studies must be considered.

    7.2.5  Internal consistency

         Internal consistency is also a useful check on the validity of a
    study, but authors often do not report sufficient detail to check for
    such consistency.  For example, in the case of known risk factors for
    respiratory effects, a study should find the anticipated associations
    (e.g., passive smoking with increased respiratory illness or with more
    wheeze in asthmatic children), and certain patterns of age or gender
    effects should be observed.  Consistency between studies also provides
    an indication of the overall strength of the database.

    7.2.6  Plausibility of the effect

         Health outcomes should be ones for which there are plausible
    bases to suspect that NO2 exposure could contribute to such effects. 
    Two health outcome measures have been most extensively considered
    in the epidemiological studies: lung function measurements and
    respiratory illness occurrence.  Human clinical and animal
    toxicological studies have not indicated a demonstrated effect on lung
    function at ambient levels in normal subjects.  On the other hand,
    human clinical and animal toxicological studies have shown that NO2
    exposure can impair components of the respiratory host defence system,
    resulting in increased susceptibility of the host to respiratory
    infection.  Thus, reported increases in respiratory symptoms and
    disease among children in epidemiological studies of NO2 exposure can
    be plausibly hypothesized to reflect an increase in respiratory
    infection.

         Each study is subsequently reviewed with special attention given
    to the above factors.  Those studies that address these factors most
    appropriately provide a stronger basis for the conclusions that they
    draw.  Consistency between studies indicates the level of the strength
    of the whole database.

    7.3  Studies of respiratory illness

         Respiratory illness and factors determining its occurrence and
    severity are important public health concerns.  The possible
    association of NO2 exposure with respiratory illness is of public
    health importance because both the potential for exposure to NO2 and
    childhood respiratory illness are common (Samet et al., 1983; Samet &
    Utell, 1990).  This takes on added importance because recurrent
    childhood respiratory illness (independent of NO2) may be a risk
    factor for later susceptibility to lung damage (Samet et al., 1983;
    Glezen, 1989; Gold et al., 1989).  The epidemiological studies
    relating NO2 exposure to respiratory illness are discussed in
    sections 7.3.1 and 7.3.2.

    7.3.1  Indoor air studies

         In this section, studies that meet criteria for use in a
    quantitative analysis are presented.  Firstly, studies conducted by
    Melia and colleagues in the United Kingdom are discussed.  This is
    followed by an evaluation of two large studies conducted in six cities
    in the USA.  Several other quantitative studies conducted by different
    authors in various countries and cities are then presented.  These are
    followed by discussion of some additional recent large-scale studies
    that yield useful quantitative information, e.g., a study of NO2
    relationship to respiratory disease in young children in Albuquerque,
    New Mexico, USA.  Lastly, other studies that provide information
    concerning respiratory illness are also discussed.

    7.3.1.1  St Thomas' Hospital Medical School Studies (United Kingdom)

         Results of several British studies have been reported by Melia
    et al. (1977, 1978, 1979, 1980, 1982a,b, 1985, 1988), Goldstein et al.
    (1979, 1981), and Florey et al. (1979, 1982).  Parts of these studies
    were reviewed previously (US EPA, 1982a), but their importance
    requires a more complete discussion of them.

         The initial study (Melia et al., 1977) was based on a survey of
    5658 children (excluding asthmatics, thus 100 less than the number
    reported), aged 6 to 11 years, with sufficient questionnaire
    information in 28 randomly selected areas of England and Scotland.  A
    self-administered questionnaire was completed by a parent to obtain
    information on the presence of morning cough, day or night cough,
    colds going to chest, chest sounds of wheezing or whistling, and
    attacks of bronchitis.  The questionnaire, distributed in 1973, asked
    about symptoms during the previous 12 months.  Colds going to the
    chest accounted for the majority of symptoms reported.  Information
    about cooking fuel (gas or electric), age, gender and social class
    (manual versus non-manual labour) was obtained, but there were no
    questions about parental smoking.  Melia et al. (1977) noted that
    although they could not include family smoking habits in the analysis,
    the known relation between smoking and social class (Tobacco Research
    Council, 1976) allowed them to avoid at least some of the potential
    bias from this source.  It seemed unlikely that, within the social
    class groups studied, there was a higher prevalence of smoking in
    homes where gas was used for cooking.  No measurements of NO2, either
    indoors or outdoors, were given.

         The authors presented their results in the form of a contingency
    table for non-asthmatics with complete covariate information. Table 46
    is a summary of that data for non-asthmatic children. The authors
    indicated that there was a trend for increased symptoms in homes with
    gas stoves, but the increase was only significant for girls in urban
    areas.  The authors gave no measures of increased risk. The data in

    Table 46 have been reanalysed using a multiple logistic model as shown
    in Table 47.  Because it had been suggested that gender had an effect
    on the relationship with "gas cooker", interaction terms for gender
    were included in the original model.  None of these proved to be
    significant, and they were subsequently dropped from the model.  When
    separate terms for each gender were used for the effect of "gas
    cooker", an estimated odds ratio of 1.25 was obtained for boys and an
    odds ratio of 1.39 was obtained for girls.  The combined odds ratio
    for both genders was 1.31 (95% confidence limits of 1.16 and 1.48) and
    was statistically significant (p < 0.0001).  The other main effects
    of gender, SES and age were all statistically significant.  This
    reanalysis suggests that gas stove use was associated with an
    estimated 31% increase in the odds of children having respiratory
    illness symptoms.

         Melia et al. (1979) reported further results of a national survey
    covering a new cohort of 4827 boys and girls, aged 5 to 10 years, from
    27 randomly selected areas that were examined in 1977.  The study
    collected information on the number of smokers in the home.  In the
    1977 cross-sectional study, only prevalence of day or night cough in
    boys (p approx. or equal 0.02) and colds going to the chest in girls
    (p < 0.05) were found to be significantly higher in children from
    homes where gas was used for cooking compared with children from homes
    where electricity was used.  As shown in Table 48, grouping responses
    according to the six respiratory questions into (1) none or (2) one or
    more symptoms or diseases yielded a prevalence higher in children
    from homes where gas was used for cooking than in those from homes
    where electricity was used (p approx. or equal 0.01 in boys,
    p = 0.07 in girls).  The effects of gender, social class, use of pilot
    lights and number of smokers in the house were examined.

         The reanalysis of the data in Table 48, applying a multiple
    logistic model, is given in Table 49. This model contained the same
    terms as the analysis in Table 47. As in the previous analysis, none
    of the interaction terms proved to be significant, and they were
    subsequently dropped from the model.  When separate terms for each
    gender were used for the effect of "gas cooker", an estimated odds
    ratio of 1.29 was obtained for boys and an odds ratio of 1.19 was
    obtained for girls.  The combined odds ratio for both genders was
    1.24 (95% confidence limits of 1.09 and 1.42).  This effect was
    statistically significant (p < 0.0002).  The other main effects of
    gender, SES and age were all statistically significant.  This
    reanalysis suggests that gas stove use in this study is associated
    with an estimated 24% increase in the odds of having symptoms.

    Table 46.  Symptom rates of United Kingdom children by age, gender,
               social class and type of cookera
                                                                        

                       Social classes I-IIIa     Social classes IIIb-V
                         (non-manual)                  (manual)
                       Electric   Gas            Electric       Gas
                                                                        

    Age < 8 years

    Boys               25.6%      26.1%          29.9%          37.5%
                       (203)      (88)           (375)          (309)

    Girls              22.2%      30.4%          31.8%          33.5%
                       (171)      (112)          (393)          (337)

    Age 8 to 11 years

    Boys               20.8%      23.3%          25.0%          29.0%
                       (365)      (189)          (675)          (654)

    Girls              18.1%      19.2%          17.8%          27.8%
                       (303)      (187)          (674)          (623)
                                                                        
    a  Numbers in parentheses refer to number of subjects; source:
       Melia et al. (1977)

    Table 47.  Multiple logistic analysis of data from the study of Melia
               et al. (1977)
                                                                        

    Factora                      Odds ratio   95% Confidence    p value
                                              interval
                                                                        

    SES and age by gender
      interactions (2 d.f.)                                     0.2922

    Gas by gender
      interaction (1 d.f.)                                      0.3953

    Gas cooker                   1.31         1.16-1.48         < 0.0001

    Gender (female)              0.86         0.76-0.97         0.0121

    SES (manual)                 1.31         1.14-1.51         0.0001

    Age (< 8 years)              1.47         1.30-1.66         < 0.0001
                                                                        
    a  SES = Socioeconomic status; d.f. = Degrees of freedom

    Table 48.  Unadjusted rates of one or more symptoms among United
               Kingdom children by age, gender, social class and type of
               cookera
                                                                        

                       Social classes I-IIIa     Social classes IIIb-V
                           (non-manual)                (manual)
                       Electric   Gas            Electric       Gas
                                                                        

    Age < 8 years
    Boys               27.4%      31.7%          32.8%          36.7%
                       (277)      (145)          (485)          (313)

    Girls              24.4%      27.6%          27.8%          36.3%
                       (291)      (134)          (497)          (336)

    Age 8 to 11 years
    Boys               19.2%      28.3%          23.6%          26.9%
                       (286)      (113)          (501)          (338)

    Girls              14.8%      18.6%          21.5%          18.5%
                       (243)      (118)          (437)          (313)
                                                                        

    a  Numbers in parentheses refer to number of subjects; source:
       Melia et al. (1979)

    Table 49.  Multiple logistic analysis of data from the study of Melia
               et al. (1979)
                                                                        

    Factora                      Odds ratio   95% Confidence    p value
                                              interval
                                                                        

    SES and age by gender
      interactions (2 d.f.)                                     0.5749

    Gas by gender
      interaction (1 d.f.)                                      0.5566

    Gas cooker                   1.24         1.09-1.42         < 0.0001

    Gender (female)              0.82         0.72-0.94         0.0030

    SES (manual)                 1.25         1.08-1.45         0.0034

    Age (< 8 years)              1.69         1.48-1.93         < 0.0001
                                                                        
    a  SES = Socioeconomic status; d.f. = Degrees of freedom

         In 1978, 808 schoolchildren (Melia et al., 1980), aged 6 to
    7 years, were studied in Middlesborough, an urban area of northern
    England.  Respiratory illness was defined as in the previous study. 
    Weekly indoor NO2 measurements were collected from 66% of the homes,
    the remaining 34% refusing to participate.  NO2 was measured weekly
    by triethanolamine diffusion tubes (Palmes tubes) attached to walls in
    the kitchen area and in the children's bedrooms.  In homes with gas
    stoves, weekly levels of NO2 in kitchens ranged from 10 to 596 µg/m3
    (0.005 to 0.317 ppm) with a mean of 211 µg/m3 (0.112 ppm) and levels
    in bedrooms ranged from 8 to 318 µg/m3 (0.004 to 0.169 ppm) with a
    mean of 56 µg/m3 (0.031 ppm). In homes with electric stoves, weekly
    levels of NO2 in kitchens ranged from 11 to 353 µg/m3 (0.006 to
    0.188 ppm) with a mean of 34 µg/m3 (0.018 ppm), and levels in
    bedrooms ranged from 6 to 70 µg/m3 (0.003 to 0.037 ppm) with a mean
    of 26 µg/m3 (0.014 ppm).  Outdoor levels of NO2 were determined
    using diffusion tubes systematically located throughout the area; the
    weekly average ranged from 26 to 45 µg/m3 (0.014 to 0.024 ppm).  One
    analysis by the authors was restricted to those 103 children in homes
    where gas stoves were present and where bedroom NO2 exposure was
    measured; the data are shown in Table 50.  A linear regression model
    was fit to the logistic transformation of the rates.  Cooking fuel was
    found to be associated with respiratory illness, independent of social
    class, age, gender or presence of a smoker in the house (p = 0.06). 
    However, when social class was excluded from the regression, the
    association was weaker (p = 0.11).  For the 6- and 7-year-old children
    living in homes with gas stoves, there appeared to be an increase in
    respiratory illness with increasing levels of NO2 in their bedrooms
    (p = 0.10), but no significant relationship was found between
    respiratory symptoms in those children, their siblings or parents and
    levels of NO2 in kitchens.

         Because no concentration-response estimates were given by the
    authors, a multiple logistic model was fitted to the data in Table 50
    with a linear slope for NO2 and separate intercepts for boys and
    girls.  NO2 levels for the groups were estimated by fitting a
    log-normal distribution to the grouped NO2 data, and the average
    exposures within each interval were estimated (see Hasselblad et al.,
    1980).  The estimated logistic regression coefficient for NO2
    (in µg/m3) was 0.015 with a standard error of 0.007.  The likelihood
    ratio test for NO2 gave a chi-square of 4.94 with one degree of
    freedom, with a corresponding p value of 0.03.

         The study was repeated in January to March of 1980 by Melia et
    al. (1982a,b).  This time, children aged 5 to 6 years were sampled
    from the same neighbourhood as the previous study, but only families
    with gas stoves were recruited.  Environmental measurements were made
    and covariate data were collected in a manner similar to the previous
    study (Melia et al., 1980).  Measurements of NO2 were available for
    54% of the homes.  The unadjusted rates of one or more symptoms by

    Table 50.  Unadjusted rates of one or more symptoms among United
               Kingdom boys and girls according to bedroom levels of
               nitrogen dioxidea
                                                                        

                                Bedroom levels of NO2 (ppm)
                                                                        

                   < 0.020      0.020-0.039      > 0.039      Total
                                                                        

    Boys           43.5%        57.9%            69.2%        54.5%
                   (23)         (19)             (13)         (55)

    Girls          44.0%        60.0%            75.0%        54.2%
                   (25)         (15)             (8)          (48)

    TOTAL          43.7%        58.8%            71.4%        54.4%
                   (48)         (34)             (21)         (103)
                                                                        

    a  Numbers in parentheses refer to number of subjects
       (from: Melia et al., 1980)

    gender and exposure level are shown in Table 51.  The authors
    concluded that "... no relation was found between the prevalence of
    respiratory illness and levels of NO2".  A reanalysis by Hasselblad
    et al. (1992) of the data in Table 51 was made using a multiple
    logistic model similar to the one used for the previous study (Melia
    et al., 1980).  The model included a linear slope for NO2 and
    separate intercepts for boys and girls.  Nitrogen dioxide levels for
    the groups were estimated by fitting a log-normal distribution to the
    grouped bedroom NO2 data.  The estimated logistic regression
    coefficient for NO2 (in µg/m3) was 0.0037 with a standard error of
    0.0052.  The likelihood ratio test for the effect of NO2 gave a
    chi-square of 0.51 with one degree of freedom (p = 0.48).

         Melia et al. (1983) investigated the association between gas
    cooking in the home and respiratory illness in a study of 390 infants
    born between 1975 and 1978.  When the child reached 1 year of age, the
    mother was interviewed by a trained field worker to complete a
    questionnaire.  The mother was asked whether the child usually
    experienced morning cough, day or night cough, wheeze or colds going
    to the chest, and whether the child had experienced bronchitis, asthma
    or pneumonia during the past 12 months.  No relation was found between
    type of fuel used for cooking at home and the prevalence of
    respiratory symptoms and diseases recalled by the mother after
    allowing for the effects of gender, social class and  parental
    smoking.  The authors gave prevalence rates of children having at

    least one symptom, according to gas stove use and gender.  The
    combined odds ratio for presence of symptoms according to gas stove
    use was 0.63 with 95% confidence interval of 0.36 to 1.10.

    Table 51.  Unadjusted rates of one or more symptoms among United
               Kingdom boys and girls according to bedroom levels of
               nitrogen dioxidea
                                                                         

                                 Bedroom levels of NO2 (ppm)
                                                                        

                   < 0.020       0.020-0.039       > 0.039      Total
                                                                        

    Boys             56.4%          67.6%          72.0%        64.4%
                     (39)           (37)           (25)         (101)

    Girls            60.0%          41.0%          52.2%        49.4%
                     (25)           (39)           (23)         (87)

    Total            57.8%          53.9%          62.5%        57.5%
                     (64)           (76)           (48)         (188)
                                                                        

    a  Numbers in parentheses refer to number of subjects; source:
       Melia et al. (1982a,b)

         Melia et al. (1988) studied factors affecting respiratory
    morbidity in 1964 primary school children living in 20 inner city
    areas of England in 1983 as part of a national study of health and
    growth.  Data on age, gender, respiratory illness, cooking fuels,
    mother's education and size of family were obtained by questionnaire.
    Smoking was not studied. The same respiratory questions were asked as
    in previous studies.  Melia et al. (1990) reported indoor levels of
    NO2 associated with gas stoves in inner city areas of England in
    1987.  The mean weekly NO2 level measured in 22 bedrooms of homes
    with gas stoves was 45 ± 25 µg/m3 (24.1 ± 13.2 ppb).  The mean weekly
    NO2 level measured in four bedrooms of homes without gas stoves was
    40 ± 22 µg/m3 (20.7 ± 11.8 ppb).  Melia et al. (1988) reported a
    relative risk of 1.06 (95% confidence interval of 0.94 to 1.17) for
    one or more respiratory conditions associated with exposure to gas or
    kerosene fuel used in the home after adjustment for ethnic group,
    gender, age group, mother's education, family size and single parent
    family status.

    7.3.1.2  Harvard University - Six Cities Studies (USA)

         Several authors (Spengler et al., 1979, 1986; Speizer et al.,
    1980; Ferris et al., 1983; Ware et al., 1984; Berkey et al., 1986;
    Quackenboss et al., 1986; Dockery et al., 1989a; Neas et al., 1990,
    1991) have reported on two cohorts of children studied in six
    different cities in the USA.  The six cities were selected to
    represent a range of air quality based on their historic levels of
    outdoor pollution.  They included: Watertown, Massachusetts; Kingston
    and Harriman, Tennessee; southeast St. Louis, Missouri; Steubenville,
    Ohio; Portage, Wisconsin; and Topeka, Kansas.  In each community
    during 1974-1977, approximately 1000 first- and second-grade
    schoolchildren were enrolled in the first year and an additional
    500 first-graders were enrolled in the next year (Ferris et al.,
    1979).  Families reported the number of people living in the home and
    their smoking habits, parental occupation and educational background,
    and fuels used for cooking and heating.  Outdoor pollution was
    measured at fixed sites in the communities as well as at selected
    households.  Indoor pollution including NO2 was measured in several
    rooms of selected households.

         Speizer et al. (1980) reported results from the six cities
    studies based on 8120 children, aged 6 to 10 years, who had been
    followed for 1 to 3 years.  Health end-points were measured by a
    standard respiratory questionnaire completed by the parents of the
    children.  The authors used log-linear models to estimate the effect
    of current use of gas stoves versus electric stoves on the rates of
    serious respiratory illness before age 2, yielding an odds ratio of
    1.12 (95% confidence limits of 1.00 and 1.26) for gas stove use.  The
    results were adjusted for presence of adult smokers, presence of air
    conditioning, and family SES.

         Ware et al. (1984) reported results for a larger cohort of
    10 160 white children, aged 6 to 9 years, in the same six cities over
    a longer period (1974-1979).  Directly standardized rates of reported
    illnesses and symptoms did not show any consistent pattern of
    increased risk for children from homes with gas stoves.  Logistic
    regression analyses controlling for age, gender, city and maternal
    smoking level gave estimated odds ratios for the effect of gas stoves
    ranging from 0.93 to 1.07 for bronchitis, chronic cough, persistent
    wheeze, lower respiratory illness index, and illness for the last
    year.  The lower respiratory illness index indicated the presence of
    bronchitis, restriction of activity due to lower respiratory illness,
    or chronic cough during the past year.  The 95% confidence bounds
    around all of these symptom-specific odds ratios included 1.  Only two
    odds ratios approached statistical significance: (1) history of
    bronchitis (odds ratio = 0.86, 95% confidence interval 0.74 to 1.00)
    and (2) respiratory illness before age 2 (odds ratio = 1.13, 95%
    confidence interval 0.99 to 1.28).  When the odds ratio for

    respiratory illness before age 2 was adjusted for parental education,
    the odds ratio was 1.11 with 95% confidence limits of 0.97 and 1.27
    (p = 0.14).  Thus, the study suggests an increase of about 11% in
    respiratory illness before the age of 2 years, which is about the same
    as that reported by Speizer et al. (1980), although the increase was
    not statistically significant at the 0.05 level.  The end-point in the
    Ware et al. (1984) study most similar to that of the Melia studies was
    the lower respiratory illness index.  The authors gave the unadjusted
    prevalence, and from those data, an estimated odds ratio of 1.08 with
    95% confidence limits of 0.97 and 1.19 was calculated.  Although this
    odds ratio was not adjusted for other covariates, such adjustments
    minimally affected other end-points in this study.  Analyses by Ware
    et al. (1984) on the other end-points found that effects of adjustment
    for covariates was minimal.

         During the period from 1983 to 1986, a new cohort of about
    1000 second- to fifth-grade schoolchildren in each community was
    enrolled and given an initial symptom questionnaire (Dockery et al.,
    1989a).  The authors studied reported respiratory symptoms on a
    subsequent symptom questionnaire (second annual) for 5338 white
    children aged 7 to 11 years at the time of enrolment.  The end-points
    of chronic cough, bronchitis, restriction of activity due to chest
    illness, and persistent wheeze were not associated with gas stove use
    in the home, but the health end-point of doctor-diagnosed respiratory
    illness prior to age 2 yielded an odds ratio of 1.15 with 95%
    confidence limits of 0.96 to 1.37.  The odds ratio for chronic cough
    was 1.15 with 95% confidence limits of 0.89 and 1.91.  The odds ratio
    was adjusted for age, sex, parental education, city of residence, and
    use of unvented kerosene heaters.

         Neas et al. (1990, 1991) studied the effects of measured NO2
    among a stratified one-third random sample of the children that were
    part of the Dockery et al. (1989a) analysis.  The sample was
    restricted to 1286 white children 7 to 11 years of age at enrolment
    with complete covariate information and at least one valid indoor
    measurement of both NO2 and respirable particles.  Methods for
    measuring indoor pollutants were described by Spengler et al. (1986). 
    Indoor pollutants were measured in each child's home for 2 weeks
    during the heating season and 2 weeks during the cooling season.  The
    two 2-week measurements were averaged to estimate each child's annual
    average NO2 exposure.  NO2 was measured by Palmes passive diffusion
    tubes at three locations: kitchen, activity room and the child's
    bedroom.  The three locations were averaged to create a household
    annual average NO2 exposure.

         The analysis of the Neas et al. (1990, 1991) study was based on
    the final symptom questionnaire (third annual), completed by parents
    following the indoor measurements.  The questionnaire reported
    symptoms during the previous year, including attacks of shortness of
    breath with wheeze, persistent wheeze, chronic cough, chronic phlegm

    and bronchitis.  The authors used a multiple logistic model with
    separate city intercepts, indicator variables for gender and age,
    parental history of chronic obstructive pulmonary disease, parental
    history of asthma, parental education and single parent family status. 
    Increases in symptoms were estimated for an additional NO2 exposure
    of 28.3 µg/m3 (0.015 ppm).  Table 52 shows the odds ratios for the
    five separate symptoms associated with the increase in NO2 exposure.

    Table 52.  Odds ratios and 95% confidence intervals for the effect of
               an additional load of 0.015 ppm NO2 on the symptom
               prevalence (from: Neas et al., 1991)
                                                                        

    Symptom              Odds ratio   95% Confidence interval
                                                                        

    Shortness of breath    1.23           0.93 to 1.61
    Persistent wheeze      1.16           0.89 to 1.52
    Chronic cough          1.18           0.87 to 1.60
    Chronic phlegm         1.25           0.94 to 1.66
    Bronchitis             1.05           0.75 to 1.47
                                                                        

         Neas et al. (1990, 1991) defined a combined symptom as the
    presence of any of the symptoms just reported.  A multiple logistic
    regression of this combined lower respiratory symptom, equivalent to
    the single response regression, gave an estimated odds ratio of 1.40
    with a 95% confidence interval of 1.14 to 1.72.  The odds ratio for
    the combined symptom score was slightly higher than in other studies,
    but was not inconsistent with those results.  The reference category
    for each of the symptom-specific odds ratios included some children
    with the other lower respiratory symptoms, whereas the children in the
    reference category for combined lower respiratory symptoms were free
    of any of these symptoms.  When split by gender, the odds ratio was
    higher in girls, a result similar to the gender modification reported
    by Melia et al. (1979).  When separate logistic analyses were
    performed for each community, the adjusted odds ratios ranged from
    1.26 for Topeka, Kansas, to 1.86 for Portage, Wisconsin.  When the
    cohort was restricted to the 495 children in homes with a gas stove,
    the adjusted odds ratio was 1.37 with a 95% confidence interval of
    1.02 to 1.84.  Table 53 provides the adjusted odds ratios for combined
    lower respiratory symptoms across ordered NO2 exposure categories. 
    The association is statistically significant for the upper exposure
    category and the overall results are consistent with a linear
    dose-response relationship between NO2 and lower respiratory symptoms
    in children.

    Table 53.  Odds ratios and 95% confidence intervals for the effect of
               ordered NO2 exposures on the prevalence of lower
               respiratory symptoms (from: Neas et al., 1991)
                                                                        

    NO2 level (ppm)                Number of     Odds     95% Confidence
                                   children      ratio    interval

    Range              Mean
                                                                        

    0 to 0.0049        0.0037        263          1.00
    0.005 to 0.0099    0.0073        360          1.06     0.71 to 1.58
    0.010 to 0.0199    0.0144        317          1.36     0.89 to 2.08
    0.020 to 0.0782    0.0310        346          1.65     1.03 to 2.63
                                                                        

         Neas et al. (1992) reported that the estimated effect of an
    additional load of 28.3 µg NO2/m3 (0.015 ppm) on lower respiratory
    symptoms was consistent across the seasons and sampling locations. 
    Table 54 provides the odds ratios and 95% confidence intervals for
    this association by season and sampler location.  The NO2 levels
    measured by the activity room and bedroom sampler were more strongly
    associated with lower respiratory symptoms than those in the kitchen. 
    The NO2 measurements in the kitchen were influenced more by transient
    peak levels associated with meal preparation on gas stoves, whereas
    the other sampling locations were more reflective of the child's
    long-term average exposures to NO2 in the home.  Spengler et al.
    (1992) suggested that children spend relatively little time (0.5 h per
    day) in the kitchen when the range is operating.

    7.3.1.3  University of Iowa Study (USA)

         Ekwo et al. (1983) surveyed 1355 children 6 to 12 years of age
    for respiratory symptoms and lung function in the Iowa City School
    District.  Parents of the children completed a questionnaire that was
    a modification of one developed by the American Thoracic Society.  The
    children were a random sample from those families whose parents had
    completed the questionnaire.  Eight measures of respiratory illness
    were reported by the authors, but only two were similar to the
    end-points used in the United Kingdom studies (section 7.3.1.1) and
    the Harvard Six City studies (section 7.3.1.2).  Parental smoking was
    also measured and used as a covariate in the analyses.  Results of the
    analyses, based on 1138 children, are presented in Table 55.  No
    measurements of NO2 exposure, either inside or outside the homes,
    were reported.

    Table 54.  Odds ratios and 95% confidence intervals for the effect of
               an additional 0.015 ppm NO2 on the prevalence of lower
               respiratory symptoms according to sampling location and
               season (from: Neas et al., 1992)
                                                                        

    Sampler location and        Mean difference    Odds    95% Confidence
    season                      gas vs. electric   ratio   interval
                                    (ppm)
                                                                        

    Household annual average        0.016          1.40    1.14 to 1.72
    Household winter average        0.018          1.16    1.04 to 1.29
    Household summer average        0.014          1.46    1.13 to 1.89
    Kitchen annual average          0.022          1.23    1.05 to 1.44
    Activity room annual
      average                       0.014          1.50    1.20 to 1.87
    Bedroom annual average          0.013          1.47    1.17 to 1.85
                                                                        


    Table 55.  Analysis of Iowa city school children respiratory symptoms
               according to gas stove type and parental smoking
               (from: Ekwo et al., 1983)
                                                                        

    Factor                   Hospitalization for     Chest congestion and
                             chest illness           phlegm with colds
                             before age two
                                                                      

                             Odds ratio     SEa      Odds ratio     SEa
                                                                        

    Gas stove use              2.4b        0.684        1.1        0.188
    Smoking effects
      Father alone smokes      2.3b        0.856        1.0        0.213
      Mother alone smokes      2.9b        1.239        1.3        0.363
      Both smoke               1.6         0.859        1.2        0.383
                                                                         

    a  SE = Standard error of the odds ratio
    b  Indicates statistical significance at the 0.05 probability level

    7.3.1.4  Agricultural University of Wageningen (The Netherlands)

         Houthuijs et al. (1987), Brunekreef et al. (1987), and Dijkstra
    et al. (1990) studied the effect of indoor factors on respiratory
    health in 6- to 9-year-old children from 10 primary schools in five

    non-industrial communities in the southeast region of the Netherlands. 
    Personal exposure to NO2 and home concentrations were measured.  An
    important NO2 emission and exposure source in these homes are
    geysers, which are unvented, gas-fired, hot water sources at the water
    tap.  Exposure to tobacco smoke was assessed by a questionnaire that
    also reported symptom information.  The study used Palmes diffusion
    tubes to measure a single weekly average personal NO2 exposure.  In
    January and February 1985, NO2 in the homes of 593 children who had
    not moved in the last 4 years was measured for 1 week.  Personal
    exposure was also estimated from time budgets and room monitoring. 
    Estimated and measured exposures to NO2 are given in Table 56.

    Table 56.  Estimated and measured personal NO2 exposure (µg/m3)
               for a single weekly average (from: Houthuijs et al., 1987)
                                                                         

    NO2 Source                     Estimated              Measured
                                                                         

                     Number  Arithmetic  Standard   Arithmetic  Standard
                                mean     deviation     mean     deviation
                                                                         

    No geyser        370        22          7           22          9
    Vented geyser    112        29          9           31          12
    Unvented geyser  111        40          9           42          11
                                                                         


         Three health measures were obtained from the questionnaire, a
    modified form of the WHO questionnaire.  The different items were
    combined to create three categories: cough, wheeze and asthma.  Asthma
    was defined as attacks of shortness of breath with wheezing in the
    past year.  The presence of any of the three symptoms was used as a
    combination variable.  The results are presented in Table 57.  A
    logistic regression model was used to fit the combination variable.
    Exposure was estimated by fitting a log-normal distribution to the
    grouped data, and the mean exposure values for each group were
    estimated by a maximum likelihood technique (Hasselblad et al.,
    1980).  The estimated logistic regression coefficient was œ0.002,
    corresponding to an odds ratio of 0.94 for an increase of 28.3 µg/m3
    (0.015 ppm) in NO2, with 95% confidence interval of 0.70 to 1.27. 
    Thus, these studies did not demonstrate an increase in respiratory
    disease with increasing NO2 exposure, but the range of uncertainty is
    quite large and the rates were not adjusted for covariates such as
    parental smoking and age of the child.  One potential explanation
    offered by the authors for the negative findings with respect to NO2
    exposure was the smaller sample size of the measured NO2 data
    compared to the categorical data (i.e., gas stove versus electric

    stove use).  They could not estimate whether more precision was gained
    by use of measured NO2 than was lost by the reduction in the sample
    size.  Houthuijs et al. (1987) reported earlier that the presence of
    an unvented geyser in the kitchen is associated with a higher
    prevalence of respiratory symptoms and that the NO2 difference
    between no geyser present and an unvented geyser is about 0.01 ppm.

    7.3.1.5  Ohio State University Study (USA)

         Mitchell et al. (1975) and Keller et al. (1979a) conducted a
    12-month study of respiratory illness and pulmonary function in
    families in Columbus, Ohio, prior to 1978.  The sample included 441
    families divided into two groups using either gas or electric cooking. 
    Participating households were given diaries to record respiratory
    illnesses for 2-week periods.  Respiratory illnesses included colds,
    sore throat, hoarseness, earache, phlegm and cough.  Only one incident
    of illness per person per 2-week period was recorded.  The study
    measured NO2 exposure, by both the Jacobs-Hochheiser and continuous
    chemiluminescence methods.  The electric stove users averaged
    38 µg/m3 (0.02 ppm) NO2 exposure, whereas the gas stove users
    averaged 94 µg/m3 (0.05 ppm).  The report did not indicate which
    rooms were measured in order to obtain this average.

         No differences were found in any of the illness rates for
    fathers, mothers or children.  No analyses were carried out using
    multiple logistic regression or Poisson regression (these methods were
    relatively new at the time).  No estimates were made that can be
    considered comparable to the odds ratios reported in the other
    studies.  However, the authors did show a bar graph of all respiratory
    illness for children under 12.  The rates were 389 (per 100 person-
    years) for electric stove use and 377 for gas stove use.  These rates
    were not significantly different even after adjustment for covariates,
    including family size, age, gender, length of residence and father's
    education.  No mention was made of adjustments for smoking status or
    smoking exposure for the children.

         In a second, related study (Keller et al., 1979b), 580 people
    drawn from households that participated in the earlier study were
    examined to confirm the reports and to determine the frequency
    distribution of reported symptoms among parents and children in gas or
    electric cooking homes.  A nurse-epidemiologist examined selected
    subjects who reported ill and obtained throat cultures.  The
    percentage of children having respiratory illnesses in homes with a
    gas stove was 85.1% (n = 87) versus 88.8% (n = 89) in homes with
    electric stoves.  The unadjusted proportions permit the calculation of
    an estimated odds ratio of 0.71 with 95% confidence interval of 0.30
    to 1.74.  Unfortunately the adjusted rates were not reported.

         Neas et al. (1991) commented that Keller's model controls for a
    series of variables that specify the child's prior illness history and
    that if chronic exposure to NO2 is a risk factor for prior illnesses,
    controlling for the child's illness history would substantially reduce
    the estimated effect of current NO2 exposure.

    7.3.1.6  University of Dundee (United Kingdom)

         Ogston et al. (1985) studied infant mortality and morbidity in
    the Tayside region of northern Scotland.  The subjects were 1565
    infants born to mothers who were living in Tayside in 1980.  Episodes
    of respiratory illness were recorded during the first year of life. 
    The information was supplemented by observations made by a health
    visitor and scrutinized by a paediatrician who checked diagnostic
    criteria and validity.  One health end-point assessed was defined as
    the presence of any respiratory disease during the year.  The use of
    gas cooking fuel was associated with increase respiratory illness
    (odds ratio = 1.14, 95% confidence interval 0.86 to 1.50) after
    adjustment for parental smoking, mother's age and type of home heating
    (Table 58).  The study did not give measured NO2 exposure values, but
    referenced the other studies conducted elsewhere in the United Kingdom
    for exposure estimates.

    7.3.1.7  Harvard University - Chestnut Ridge Study (USA)

         Schenker et al. (1983) reported a large respiratory disease study
    of 4071 children aged 5 to 14 in the Chestnut Ridge region of western
    Pennsylvania.  The region is predominately rural, with numerous
    underground coal mines and four large coal-fired electricity-
    generating plants in the area.  A standardized children's
    questionnaire (Ferris, 1978) was sent to parents of all children in
    grades 1 to 6 in targeted schools.  An SES scale derived from the
    parent's occupation and education was divided into quintiles to
    provide SES strata.  Important confounding factors considered in the
    analysis were gender, SES and maternal smoking.  In the multiple
    logistic model, no significant association was found between gas stove
    use and any of the respiratory or illness variables after adjusting
    for SES.  No odds ratios or other numerical data were reported.

        Table 57.  Frequency and prevalence of reported respiratory symptoms with respect to different
               categories of mean indoor NO2 concentrations in a population of 775 children
               aged 6 to 12 old (from: Dijkstra et al., 1990)
                                                                                                       

                                       Frequency and prevalence in category of indoor NO2
                                                                                                       

    Symptom                    0-20 µg/m3          21-40 µg/m3         41-60 µg/m3        > 60 µg/m3
                               (n = 336)            (n = 267)            (n = 93)          (n = 79)
                                                                                                       

    Cough                    16       4.8%       12       4.5%       7        7.5%       3        3.8%

    Wheeze                   30       8.9%       18       6.7%       3        3.2%       7        8.9%

    Asthma                   22       6.6%       12       4.5%       2        2.2%       3        3.8%

    One or more symptoms     36      10.7%       24       9.0%       8        8.6%       8       10.1%
                                                                                                       
    
    Table 58.  Regression coefficients for multiple logistic analyses of   
               respiratory illness in Tayside children (from: Ogston et
               al., 1985)
                                                                        

    Factor                   Regression    Odds ratio      95% Confidence
                             coefficient                   limits
                                                                        

    Parental smoking         0.429         1.54

    Age of mother            -0.094        not available
    (in 5-year groups)

    Presence of gas stove    0.130         1.14            0.86, 1.50
                                                                        


    7.3.1.8  University of New Mexico Study (USA)

         Samet et al. (1993) conducted a prospective cohort study between
    January 1988 and June 1990 to test the hypothesis that exposure to
    NO2 increases the incidence and severity of respiratory illness
    during the first 18 months of life.  A total of 1315 infants were
    enrolled into the study at birth in Albuquerque, New Mexico.  The
    subjects were healthy infants from homes without smokers and who spent
    less than 20 h/week in day care.  Illness experience was monitored by
    a daily diary of symptoms completed by the mother and a telephone
    interview conducted every two weeks.  For a sample of the ill
    children, a nurse practitioner made a home visit to conduct a
    standardized history and physical assessment.  Exposure to NO2 was
    estimated by a 2-week average concentration measured in the subjects'
    bedrooms with passive samplers.  Estimates of exposure based on
    bedroom concentration were tightly correlated with estimates of
    exposures calculated as time-weighted averages of the concentrations
    in the kitchen, bedroom and activity room.  The authors defined
    illness events as the occurrence on at least two consecutive days of
    any of the following: runny or stuffy nose, wet cough, dry cough,
    wheezing or trouble with breathing.  Wheezing was defined as a


    high-pitched musical sound audible during breathing, and trouble with
    breathing as the parent's perception of rapid or laboured breathing. 
    Illness events ended with two consecutive symptom-free days.

         The analysis was limited to the 1205 subjects completing at least
    1 month of observation; of these, 823 completed the full protocol. 
    Multivariate methods were used to control for potential confounding
    factors and to test for effect modification.  In analyses of
    determinants of incident illnesses, the outcome variable was the
    occurrence of illness during 2-week intervals of days at risk.  The
    independent variables considered in the multivariate analyses included
    the fixed factors of birth order, gender, ethnicity, parental asthma
    and atopic status, household income, and maternal education.  Other
    variables considered were the temporally varying factors of age,
    calendar month, day-care attendance and breast-feeding. Potential
    confounding and effect modification by cigarette smoking was
    controlled by excluding subjects from households with smokers.

         Lambert et al. (1993) reported that in this prospective cohort
    study during the winter, bedroom concentrations in homes with gas
    stoves averaged 0.021 ppm (SD = 0.022 ppm).  In bedrooms of homes with
    electric stoves, concentrations averaged 0.007 ppm (SD = 0.006 ppm). 
    Approximately 77% of the bedroom NO2 observations were less than
    0.02 ppm; only 5% were greater than 0.04 ppm.  The 90th percentile of
    the weekly measured concentrations was 0.05 ppm NO2 in bedrooms.

         Samet et al. (1993) performed the analysis using the generalized
    estimated equations described by Zeger & Liang (1986). This takes into
    account the correlation structure when estimating regression
    coefficients and their standard errors.  The multivariate models
    examined the effects of the unlagged NO2 exposures, lagged NO2
    exposures and stove type (Table 59).  None of the odds ratios was
    significantly different from unity, the value for the reference
    category of 0 to 0.02 ppm.  Additionally, the odds ratios did not tend
    to increase consistently from the middle category of exposure to the
    highest category.  Furthermore, exposure to NO2 and the durations of
    the four illness categories were not associated.  The authors added
    NO2 exposure to the model as a continuous variable, while controlling

    for the same covariates included in Table 59.  For each of the five
    illness variables, the estimated multiplier of the odds ratio per
    0.001 ppm increment of NO2 was 0.999, with confidence limits
    extending from approximately 0.995 to 1.002.

    7.3.1.9  University of Basel Study (Switzerland)

         Braun-Fahrlaender et al. (1989, 1992) and Rutishauser et al.
    (1990a,b) studied the incidence and duration of common airway symptoms
    in children up to 5 years old over a 1-year period in a rural, a
    suburban and two urban areas of Switzerland.  Parents were asked to
    record daily their child's respiratory symptoms (from a list) over a
    6-week period.  Additionally, covariates, including family size,
    parental education, living conditions, health status of the child,
    parents' respiratory health, and smoking habits of the family, were
    assessed by questionnaire.  During the same 6-week period NO2 was
    measured weekly using Palmes tubes, both inside and outside the home
    of the participants.  Meteorological data were obtained from local
    monitoring stations, but additional air quality data from fixed
    monitoring sites were only available for the two urban study areas. 
    NO2 concentrations inside the home were on average lower than in the
    outside air (Fig. 24). Indoor levels for Basel, Zurich, Wetzikon and
    Rufzerfeld were 33.8, 28.4, 20.5 and 11.2 µg/m3 (0.018, 0.015, 0.011
    and 0.006 ppm), respectively.  The indoor NO2 concentration depended
    to some extent on the concentration of the outside air.

         The analysis was restricted to 1063 Swiss nationals (from a total
    of 1225 participating families).  For all four study areas, regional
    mean incidence rates of upper respiratory illness, cough, breathing
    difficulties and total respiratory illness, adjusted for individual
    covariates and weather data, were regressed (using Poisson regression)
    against regional differences in annual mean NO2 concentrations.  All
    the relative risks were computed for a 20-µg/m3 (0.011-ppm) increase
    in pollution concentration.  The NO2 concentration measured by indoor
    passive sampler was associated with the duration of any episode
    (relative duration of 1.16, 95% confidence interval of 1.12 to 1.21),
    upper respiratory episodes (relative duration of 1.18, 95% confidence
    interval of 1.01 to 1.38), and coughing episodes (relative duration of
    1.15, 95% confidence interval of 1.03 to 1.29).  A discussion of
    associations with outdoor levels is presented in section 7.3.2.

    FIGURE 24

        Table 59.  Odds ratiosa for effect of nitrogen dioxide exposure on incidence of respiratory illness
               (from: Samet et al., 1993)
                                                                                                                                          

    NO2 exposure              All illnesses                  All lower                    Lower, with                   Lower, with
                                                                                           wet cough                      wheezing
                                                                                                                                    

                        Odds ratio     95% CIb        Odds ratio     95% CIb        Odds ratio     95% CIb        Odds ratio     95% CIb
                                                                                                                                          

    Unlaggedc              1.04       0.96-1.12          0.98       0.89-1.09          1.00       0.89-1.12          0.92       0.73-1.15
    0.02-0.06 ppm          0.94       0.81-1.08          0.93       0.76-1.13          0.94       0.77-1.16          0.88       0.56-1.37
    > 0.04 ppm

    Laggedc                1.01       0.93-1.10          0.97       0.87-1.08          0.97       0.87-1.09          0.95       0.75-1.19
    0.02-0.06 ppm          0.92       0.77-1.10          0.91       0.72-1.15          0.89       0.68-1.16          0.98       0.66-1.48
    > 0.04 ppm

    Gas Stoved             0.98       0.90-1.07          0.91       0.81-1.04          0.94       0.82-1.07          0.84       0.64-1.09
                                                                                                                                          

    a  Obtained by generalized estimating equation method. Adjusted for season, age, gender, ethnicity, birth order, day care, income,
       maternal education, breast feeding, parental atopy and asthma, and maternal history of respiratory symptoms.
    b  CI = Confidence interval
    c  Reference category is 0-0.02 ppm NO2
    d  Reference category is electric stove
        7.3.1.10  Yale University Study (USA)

         Berwick et al. (1984, 1987, 1989), Leaderer et al. (1986) and
    Berwick (1987) reported on a 12-week study (six 2-week time periods)
    of lower and upper respiratory symptoms in 159 women and 121 children
    (aged 12 or less) living in Connecticut.  Levels of NO2 were measured
    in 91% of the homes, 57 of which had kerosene heaters and 62 of which
    did not.  Ambient NO2 levels ranged from 9 to 19 µg/m3 (0.005 to
    0.01 ppm) for the six 2-week time periods.  Two-week average indoor
    NO2 levels in homes of monitored children were highest for homes with
    kerosene heaters  and  gas stoves (91 µg/m3, 0.05 ppm; n = 8),
    second highest for kerosene only (36 µg/m3, 0.02 ppm; n = 45), third
    highest for gas stoves only (32 µg/m3, 0.02 ppm; n = 13), and lowest
    for no sources (6 µg/m3, 0.003 ppm; n = 43).  Indoor levels did not
    fluctuate greatly over time, as indicated by the 2-week averages.  A
    comparison of personal NO2 exposures, as measured by Palmes diffusion
    tubes, and NO2 exposures measured in residences had a correlation of
    0.94 for a subsample of 23 individuals.  Results of this comparison
    show an excellent correlation between average household exposure and
    measured personal exposure (see section 3.6 and Fig. 13).

         The study defined lower respiratory illness as the presence of at
    least two of the following: fever, chest pain, productive cough,
    wheeze, chest cold, physician-diagnosed bronchitis, physician-
    diagnosed pneumonia and asthma.  Information on many potential
    covariates (e.g., SES, age, gender and exposure to environmental
    tobacco smoke) was obtained.  The covariates having the largest effect
    were age of child, family SES and history of respiratory illness, as
    shown by multiple logistic analysis.  When controlling for SES and
    history of respiratory illness, children under 7 years of age exposed
    to 30 µg NO2/m3 (0.016 ppm) or more were found to have a risk of
    lower respiratory symptoms 2.25 times higher than that of unexposed
    children (95% confidence limits of 1.69 and 4.79).  Older children and
    adults showed no increased risk.

         Although the Berwick study had relatively extensive information
    on exposure, several problems are evident.  Unvented kerosene
    space-heaters also release volatile organic compounds and combustion
    particles.  The 4-year age-specific relative risks for lower
    respiratory disease are very variable, and it is not clear why these
    3-year strata were collapsed into 2 strata at 7 years of age.  The

    analyses may be sensitive to the adjustment for SES, which can be
    correlated with exposure.  This is less of a problem in studies with
    larger sample sizes (e.g., Melia et al. 1977, 1979), but may be
    critical in the Berwick study.  Furthermore, Neas et al. (1991) noted
    that the Berwick study controlled for prior illnesses, as did the
    Keller study, which would reduce the estimated effect of current NO2
    exposure.

    7.3.1.11  Freiburg University Study (Germany)

         Kuehr et al. (1991) conducted a cross-sectional study on the
    prevalence of asthma in childhood in relation to NO2 levels in the
    city of Freiburg and two Black Forest communities. A study group of
    704 children (with 41 asthmatic) aged 7 to 16 years took part in a
    standardized interview and medical examination. Indoor and outdoor
    exposure information was taken into account.  Passive smoking
    exposures were assessed.  Stoves used as heating devices carried a
    4.8-fold relative risk for asthma compared to other types of heating
    (95% CI 1.95-11.8).

    7.3.1.12  McGill University Study (Canada)

         In a case-control study carried out in Montreal, Quebec, Canada,
    between 1988 and 1990, NO2 levels measured by passive NO2 monitoring
    badge were studied in relation to the incidence of asthma among 3- and
    4-year-old children (Infante-Rivard, 1993).  Multivariate
    unconditional logistic regression was carried out for the 140 subjects
    who had NO2 measurements; the analysis included NO2 and the
    variables retained in the final conditional model that includes SES
    and parental smoking.  The author reported an increase in asthma
    incidence associated with NO2 exposure levels.  However, the Task
    Group noted the exceptionally large effect estimates given the
    exposure levels.

    7.3.1.13  Health and Welfare Canada Study (Canada)

         Dekker et al. (1991) studied asthma and wheezing syndromes as
    part of a questionnaire-based study of 17 962 Canadian school
    children.  The questionnaire was developed from the 1978 American
    Thoracic Society questionnaire, which was the same as that used in the
    Harvard Six Cities Study.  For analysis, the sample was restricted to
    children aged 5 to 8 years and excluded those children with cystic
    fibrosis as well as those living in mobile homes, tents, vans,
    trailers and boats.  The authors calculated odds ratios adjusted for

    age, race, gender, parental education, gender of the respondent,
    region of residence, crowding, dampness and environmental tobacco
    smoke.  The adjusted odds ratio of asthma as a function of gas cooking
    was 1.95 with 95% confidence limits of 1.41 and 2.68.  The adjusted
    odds ratio of wheezing as a function of gas cooking was 1.04 with 95%
    confidence limits of 0.77 and 1.42.  The authors noted that this
    finding needed to be treated with caution, however, because of the few
    subjects with asthma in the study who were exposed to gas cooking
    (n = 60).

    7.3.1.14  University of North Carolina Study (USA)

         Margolis et al. (1992) studied the prevalence of persistent
    respiratory symptoms in 393 infants of different SES by analysing data
    from a community-based cohort study of respiratory illness in the
    first year of life in central North Carolina between 1986 and 1988. 
    Infants were limited to those weighing more than 2000 g and who did
    not require neonatal care outside the normal newborn nursery.  Of
    those eligible, 47% were enrolled and, of these, 77% completed the
    study and were included in the analysis.  Compared with the 1241
    infants from families refusing enrolment, the 1091 eligible study
    infants were more likely to be of high SES and were more often black. 
    Study infants were less likely to have mothers who smoked.

         The presence of persistent respiratory symptoms was measured at
    the 12-month home interview using an American Thoracic Society
    children questionnaire (modified for infants) for studies of
    respiratory illness.  Infants who were reported to "usually cough" or
    "occasionally wheeze" were classified as having persistent respiratory
    symptoms.

         Of the 393 infants that Margolis et al. (1992) included in their
    study, approximately 41 lived in homes with gas cooking.  The relative
    risk of persistent respiratory symptoms among infants exposed to gas
    cooking unadjusted for any covariates was 1.12 (95% confidence
    interval of 0.63 to 2.04).

    7.3.1.15  University of Tucson Study (USA)

         The study by Dodge (1982) was based on a cohort of 676 children
    in the third and fourth grades (about 90% aged 8-10 years) of schools
    in three Arizona communities.  Gas cooking stoves were associated with
    increased symptoms: asthma odds ratio = 1.47, wheeze odds ratio 
    = 1.24, sputum odds ratio = 2.28, and cough odds ratio = 2.21. 
    However, only 79 children (19%) had electric heat, so the numbers were
    small and only cough was significant at the 0.05 level.  After
    controlling for height and age, gas stoves were not associated with a
    decline in the growth of FEV1.

    7.3.1.16  Hong Kong Anti-Cancer Society Study (Hong Kong)

         In 1985, 362 primary school children (age 7-13 years) were
    included in a study of NO2 exposure and respiratory illness in Hong
    Kong (Koo et al., 1990).  Exposures to NO2 were estimated by use of
    personal badge monitors, worn for a single period of 24 h, and
    supplemented by monitors placed in classrooms.  NO2 exposures were
    estimated in the same manner for the mothers of the study children. 
    Mothers and children completed respiratory illness questionnaires.  No
    association was found between respiratory symptoms and NO2 exposures
    for children (mean 19 ppb).  Among the mothers (mean exposure 19 ppb)
    allergic rhinitis and chronic cough were associated with NO2.

    7.3.1.17  Recent studies

         This section includes studies that have reported preliminary
    results only or have appeared recently in the scientific literature.

         Spengler et al. (1993) reported results for evaluation of more
    than 15 000 schoolchildren in various sites in the USA and Canada, but
    found no statistically significant increases in respiratory symptoms
    to be associated with use of gas heaters or cookers.

         Goren et al. (1993) reported no association between gas heating
    and respiratory health effects among 8000 schoolchildren in Israel.

         Preliminary results reported by Peat et al. (1990) indicated no
    relationship between relatively high NO2 in Australian homes with
    gas use in Sydney and respiratory symptoms or bronchial hyper-
    responsiveness.

         Pilotto (1994) reported a prospective study of health effects of
    unflued gas heater emissions on 425 Australian schoolchildren aged
    6-11 years.  Short-term indoor monitoring by means of passive
    diffusion badge monitors placed in classrooms or worn at home was
    carried out to determine daily 6-h averages.  Children exposed to a
    level of 0.08 ppm or more, compared with a background level of
    0.02 ppm, had increased rates of respiratory illnesses and school
    absences.

    7.3.2  Outdoor studies

         Several studies have examined the relationship of estimated
    ambient NO2 levels to respiratory health outcome measures, including
    various respiratory symptomatologies.  Those that provide a
    quantitative estimate of effect are indicated in Table 60.

        Table 60.  Effects of outdoor NO2 exposure on respiratory disease
                                                                                                                                          

    Study                              Health end-point              NO2 levels (ppm)/period            Odds ratio or       95% CI
                                                                                                           estimate
                                                                                                                                          

    Dockery et al. (1989b)             Bronchitis                    0.007-0.023 annual average              1.7            0.5 to 5.5
                                       Chronic cough                                                         1.6            0.3 to 10.5
                                       Chest illness                                                         1.2            0.3 to 4.8
                                       Wheeze                                                                0.8            0.4 to 1.6
                                       Asthma                                                                0.6            0.3 to 0.9

    Braun-Fahrlaender et al. (1992)    Duration of respiratory       Change of 0.011 6-week                  1.11           1.07 to 1.16
                                       episodes                      average

    Schwartz et al. (1991)             Croup                         0.005-0.037 daily                       1.28           1.07 to 1.54

    Jaakkola et al. (1991)             Upper respiratory             Contrasted polluted versus              1.6            1.1 to 2.1
                                       infection                     less polluted areas by
                                                                     comparison of annual levels
                                                                                                                                          
        7.3.2.1  Harvard University - Six City Studies (USA)

         As part of the US Six City Studies, Dockery et al. (1989b)
    obtained respiratory illness and symptom data from questionnaires
    distributed from September 1980 to April 1981.  Indoor air aspects
    of this study (Dockery et al., 1989a) were described in the section
    on indoor studies.  The questionnaires obtained information on
    bronchitis, cough, chest illness, wheeze and asthma.  A centrally
    located air monitoring station was established in 1974 where ambient
    sulfur dioxide, NO2, ozone, total suspended particulate matter and
    meteorological variables were measured.  The authors used multiple
    logistic regression analysis in order to adjust for covariates of
    gender, age, maternal smoking, gas stove use and separate intercepts
    for each city.  Although the strongest associations were found between
    respiratory symptoms and particulate matter, there were increased odds
    ratios of respiratory symptoms with ambient NO2.  These were not
    statistically significant, but the direction for bronchitis, chronic
    cough and chest illness was consistent with the studies of indoor
    exposure.  The odds ratios for various health end-points for an
    increase in NO2 from the lowest-exposure city to the highest-exposure
    city 12 to 43 µg/m3 (0.0065 to 0.0226 ppm) are shown in Table 60.

    7.3.2.2  University of Basel Study (Switzerland)

         Braun-Fahrlaender et al. (1992) studied the incidence and
    duration of common airway symptoms in children up to 5 years old. 
    This study, also discussed in section 7.3.1.9, was conducted over
    a 1-year period in a rural, a suburban and two urban areas of
    Switzerland.  Parents were asked to record their child's respiratory
    symptoms (from a list) daily over a 6-week period.  Additionally,
    covariates including family size, parental education, living
    conditions, health status of the child, parents' respiratory health
    and smoking habits of the family were assessed by questionnaire. 
    Weekly NO2 measurements were made during the same 6-week period using
    Palmes tubes, both inside and outside the home of the participants. 
    Meteorological data were obtained from local monitoring stations, but
    additional air quality data from fixed monitoring sites were only
    available for the two urban study areas.  The analysis was restricted
    to 1063 Swiss nationals (from a total of 1225 participating families). 
    For all four study areas, regional mean incidence rates of upper
    respiratory illness, cough, breathing difficulties and total
    respiratory illness, adjusted for individual covariates and weather
    data, were regressed (using Poisson regression) against regional
    differences in annual mean NO2 concentrations.  There was no
    association between long-term differences in NO2 levels by region and
    mean annual rates of respiratory incidence.

         The adjusted annual mean symptom duration by region and the
    corresponding NO2 levels (measured by passive samplers) are shown in
    Table 61.  A second-stage regression of the adjusted natural logarithm
    of regional mean duration on NO2 levels yields significant
    associations between outdoor NO2 levels and the average duration of
    any respiratory episode (relative duration of 1.11, 95% confidence
    interval of 1.07 to 1.16) and upper respiratory episodes (relative
    duration of 1.14, 95% confidence interval of 1.03 to 1.25).  A
    positive trend for the duration of coughing episodes was also seen
    (relative duration of 1.09, 95% confidence interval of 0.97 to 1.22). 
    No association was seen with the duration of breathing difficulties. 
    All the relative risks are computed for a 20-µg/m3 (0.011-ppm)
    increase in pollution concentration.  In the suburban and rural areas,
    NO2 was the only air pollutant measured.  Correlation between
    outdoor passive NO2 sampler and total suspended particulate (TSP)
    measurements in the two urban study areas was quite high (0.52).  The
    high correlation between NO2 and TSP suggests that this NO2
    association may reflect confounding with TSP.  The lack of TSP data
    for the other two regions precludes eliminating TSP as a possible
    confounder in this analysis.  But the consistency of the NO2 findings
    are evident and, although the association with symptom duration in
    Zurich and Basel may well be due to confounding with TSP, the
    cross-sectional association across the four regions supports a
    possible NO2 role.

    7.3.2.3  University of Wuppertal Studies (Germany)

         Schwartz et al. (1991) evaluated respiratory illness in five
    German communities.  Children's hospitals, paediatric departments of
    general hospitals, and paediatricians reported daily the numbers of
    cases of croup.  A diagnosis of croup was based on symptoms of
    hoarseness and barking cough, inspiratory stridor, dyspnoea, and a
    sudden onset.  The counts were modelled using Poisson regression
    with adjustments for weather, season, temperature, humidity and
    autoregressive errors.  Statistically significant effects of both
    ambient particulate matter and NO2 were found on the counts of
    respiratory illnesses.  A relationship between short-term fluctuations
    in air pollution and short-term fluctuations in medical visits for
    croup symptoms was found in this study.  The estimated relative risk
    was 1.28 with 95% confidence limits of 1.07 and 1.54 for an increase
    from 10 to 70 µg NO2/m3 (0.005 to 0.037 ppm).

    7.3.2.4  University of Tubigen (Germany)

         Rebmann et al. (1991) studied 875 cases of croup in Baden-
    Württemberg in relation to ambient NO2 levels over a 2-year period. 
    Monthly NO2 means varied from 23 to 78 µg/m3.  Statistical
    regression methods indicated weak but statistically significant
    influences of the daily ambient NO2 mean on the occurrence of croup.

        Table 61.  Adjusted annual symptom duration (days) and NO2 levels in four regions of Switzerland
               (from: Braun-Fahrlaender et al., 1992)
                                                                                                                           

    Region         Any symptom     URI durationa      Cough       Breathing difficulty     Indoor NO2         Outdoor NO2
                     duration                        duration             duration        concentration      concentration
                                                                                               (ppm)             (ppm)
                                                                                                                           

    Basel             4.50             1.99            2.32                 1.55              0.0166             0.0272

    Zurich            4.21             1.85            2.01                 1.72              0.0118             0.0248

    Wetzikon          4.00             1.62            2.10                 3.47              0.0103             0.0173

    Rafzerfeld        3.88             1.72            2.02                 1.25              0.0059             0.0133
                                                                                                                           

    a  URI = Upper respiratory illness
        7.3.2.5  Harvard University - Chestnut Ridge Study (USA)

         In the autumn of 1980, Vedal et al. (1987) conducted a panel
    study on 351 children selected from the 1979 Chestnut Ridge study. 
    Parents and children were instructed at the beginning of the school
    year in completing daily diaries of respiratory symptoms.  Lower
    respiratory illness was defined as wheeze, pain on breathing, or
    phlegm production.  Of the 351 subjects selected for the 8 month of
    follow-up, 128 participated in the completion of diaries.  Three
    subgroups were established: one without respiratory symptoms, one with
    symptoms of persistent wheeze, and one with cough or phlegm production
    but without persistent wheeze.  Maximum hourly NO2 levels, measured
    at a single monitoring site in the study region, for each 24-h period
    were used to reflect the daily pollutant level.  During September 1980
    to April 1981, the mean NO2 maximum daily level was 40.5 µg/m3
    (0.021 ppm) with a range of 12 to 79 µg/m3 (0.006 to 0.042 ppm). 
    Regression models could not be fit for asymptomatic subjects; thus 55
    subjects were included in the analysis of lower respiratory illness,
    but NO2 levels were not predictive of any symptom outcome.

    7.3.2.6  University of Helsinki Studies (Finland)

         Jaakkola et al. (1991) studied the effects of low-level air
    pollution in three cities by comparing the frequency of upper
    respiratory infections over a 12-month period in 1982 as reported by
    parents of children aged 14 to 18 months (n = 679) and 6 years
    (n = 759).  Pollutants studied included ambient levels of NO2, the
    annual mean of which was 15 µg/m3 (0.008 ppm).  Other pollutants
    monitored were sulfur dioxide, hydrogen sulfide and particles. 
    Passive smoking and SES were taken into account.  The authors reported
    a significant association between the occurrence of upper respiratory
    infections and living in an air-polluted area for both age groups
    studied, both between and within cities.  The adjusted odds ratio was
    1.6 (95% confidence interval of 1.1 to 2.1) in the 6-year-old age
    group.  The authors concluded that the combined effect of sulfur
    dioxide, particulates, NO2, hydrogen sulfide and other pollutants may
    be a contributing factor in the study results.

    7.3.2.7  Helsinki City Health Department Study (Finland)

         Pönkä (1991) studied effects of ambient air pollution and minimum
    temperature on the number of patients admitted to hospital for asthma
    attacks in Helsinki from 1987 to 1989.  During the 3-year period,
    4209 hospitalizations for asthma occurred. The temperature ranged from
    œ37 to +26°C, with a 3-year mean of 5°C, and the number of admissions
    increased during cold weather.  After standardization for minimum
    temperature, the multiple-regression analysis indicated that NO2 and
    carbon monoxide levels were significantly related to asthma admission. 
    The NO2 levels averaged 38.6 µg/m3 (0.02 ppm) for the 3-year period,
    ranging from 4.0 to 169.6 µg/m3 (0.002 to 0.09 ppm).  During the

    period of high NO2 (mean 45.8 µg/m3, 0.024 ppm) levels, the mean
    number of all admissions was 29% greater than during the lower
    pollution period (28.1 µg/m3, 0.015 ppm).  Indoor NO2 levels and
    cooking fuel use were not reported.

    7.3.2.8  Oulu University Study (Finland)

         The number of daily attendances for asthma at the emergency room
    of the Oulu University Central Hospital, Finland, was recorded for one
    year, along with daily measures of air pollutants at four points
    around the city (Rossi et al., 1993).  Daily mean levels of NO2
    ranged up to 69 µg/m3 (maxima 0-154 µg/m3).  Asthma visits were
    reported to be significantly associated with NO2, SO2, H2S and TSP
    levels.  After adjustment for daily temperature, only NO2 was
    significantly correlated with attendances.  The association of NO2
    and asthma attacks was stronger in winter months than during the
    summer.

    7.3.2.9  Seth GS Medical College Study (India)

         A survey of air pollution and health was carried out in Bombay,
    India, in 1978 (Kamat et al., 1980).  The study included 4129 adults
    in three urban areas and one rural area.  A single monthly mean NO2
    level was reported for each study area - annual averages were 4 µg/m3
    in the rural area, and 14-16 µg/m3 in the city.  Winter levels in
    the city study were higher than at other times of the year (up to
    40 µg/m3).  It was reported that chronic cough with sputum, frequent
    colds and exertional dyspnoea were significantly associated with NO2
    levels.  These symptoms were also associated with atmospheric levels
    of SO2 and suspended particulate matter, and it was not possible to
    identify a separate influence of NO2 alone.

    7.4  Pulmonary function studies

         Pulmonary function studies are part of any comprehensive
    investigation of the possible effects of any air pollutant. 
    Measurements can be made in the field, they are non-invasive, and
    their reproducibility has been well documented.  Age, height, gender
    and presence of respiratory symptoms are important determinants of
    lung function. Furthermore, changes in pulmonary function have been
    associated with exposure to tobacco smoke, particulate matter and
    other factors.  The studies reviewed below evaluate pulmonary function
    changes in relation to indoor or outdoor NO2 exposures.  Several of
    the respiratory disease studies described earlier also included
    information on pulmonary function.

    7.4.1  Harvard University - Six City Studies (USA)

         Ware et al. (1984) described analysis of lung function values
    using multiple linear regression on the logarithm of the lung function
    measures.  Covariates included sex, height, age, weight, smoking
    status of each parent, and educational attainment of the parents. 
    Exposure to gas stoves was associated with reductions of 0.7% in mean
    FEV1 (forced expiratory volume in 1 second) and 0.6% in mean forced
    vital capacity (FVC) at the first examination (p < 0.01), and
    reductions of 0.3% at the second examination (not significant). 
    The estimated effect of exposure to gas stoves was reduced by
    approximately 30% after adjustment for parental education.  The
    authors stated that the adjustment for parental education may be an
    over-adjustment, and may partially represent gas stove use because of
    association between parental education and type of stove.

         Berkey et al. (1986) used the data from children seen at two to
    five annual visits to study factors affecting pulmonary function
    growth.  Children whose mothers smoked one pack of cigarettes per day
    had FEV1 growth rates approximately 0.17% per year lower (p = 0.05). 
    The same data provided no evidence for an effect of gas stove exposure
    on growth rate.

         Dockery et al. (1989b) obtained pulmonary function data
    during the 1980 and 1981 school year.  Only TSP concentration was
    consistently associated with estimated lower levels of pulmonary
    function.  There was little evidence for an association between lower
    pulmonary function levels and the annual mean concentration of NO2 or
    any other pollutant.

         Neas et al. (1991) also reported that indoor NO2 levels were not
    significantly associated with a deficit in children's pulmonary
    function levels in either of two examinations (FEV1 and FVC).

    7.4.2  National Health and Nutrition Examination Survey Study (USA)

         Schwartz (1989) studied air pollution effects on lung function in
    children and youths aged 6 to 24 years.  FVC, FEV1, and peak flow
    measurements taken as part of the National Health and Nutrition
    Examination Survey II (NHANES II) were examined after controlling
    for age, height, race, gender, body mass, cigarette smoking and
    respiratory symptoms.  Air pollution measurements were taken from all
    population-oriented monitors in the US EPA database.  Each person was
    assigned the average value of each air pollutant from the nearest
    monitor for the 365 days preceding the spirogram.  Highly significant
    negative regression coefficients were found for three pollutants
    (TSP, NO2 and O3) with the three lung function measurements.  For an
    increase of NO2 exposure of 28.3 µg/m3 (0.015 ppm), an estimated
    decrease of about 0.045 litres was seen in both FVC and FEV1.

    7.4.3  Harvard University - Chestnut Ridge Study (USA)

         Vedal et al. (1987) conducted a panel study on 351 children
    selected from the 1979 Chestnut Ridge cross-sectional study of
    elementary school-aged children (mean age = 9.5 years).  Peak
    expiratory flow (PEF) was measured daily in 144 children for
    9 consecutive weeks and was regressed against daily maximum hourly
    ambient concentrations of NO2, SO2 and coefficient of haze.  No air
    pollutant was strongly associated with PEF.  All pollutant levels were
    relatively low; NO2 levels ranged from 12 to 79 µg/m3 (0.006 to
    0.042 ppm).  No indoor measurements were made, nor were any surrogates
    for indoor pollution included in the analysis.

    7.4.4  Other pulmonary function studies

         Ekwo et al. (1983) obtained pulmonary function measurements from
    89 children whose parents did not smoke and 94 children whose parents
    smoked, and reported no differences in lung function associated with
    gas stove use in a cohort of children 6 to 12 years of age.

         Dijkstra et al. (1990) examined pulmonary function in Dutch
    children; lung function was measured at the schools.  There was a weak
    negative association between FEF25-75% (25 and 75% of FVC) and exposure
    to NO2.  FEV1, PEF and FEF25-75% were all negatively associated with
    exposure to tobacco smoke.  The authors concluded that the study
    failed to document clear associations between indoor exposure to NO2
    and lung function changes in 6- to 12-year-old Dutch children.

         Lebowitz et al. (1985) studied a cluster sample of 117 middle-
    class households in Tucson, Arizona, USA.  Symptom diaries and peak
    flows were obtained over a 2-year period.  Outdoor sampling of O3,
    TSP, CO and NO2 was done in or near the clusters.  Indoor sampling of
    O3, TSP, respirable suspended particles and CO was done in a
    subsample of the homes.  Information such as the presence of a gas
    stove or smoking was also obtained. The presence of a gas stove was
    used as a surrogate for indoor NOx exposure.  Children's peak flow
    was associated with gas stove use (p = 0.066) for an analysis
    excluding TSP.  In adult asthmatics, gas stove use was significantly
    associated with peak flow decrements (p < 0.001).  This was true
    across smoking groups, but the difference was greatest for smokers.

         Lung function studies were conducted in a prospective survey
    undertaken by Kamat et al. (1980) on 4129 subjects in three urban
    areas of Bombay and a rural control area during February to July,
    1977.  The survey revealed that the population in low polluted areas
    had higher lung function for FEF25-75% and PEF.  Thus, there was
    suggestive evidence that the higher values obtained from lung function
    tests in rural subjects as compared to urban subjects could be due to
    increased levels of NO2.

    7.5  Other exposure settings

         Certain recreational settings have been shown to result in NO2
    exposures that greatly exceed the chronic, low-level exposures
    described in the previous epidemiological studies.

    7.5.1  Skating rink exposures

         Hedberg et al. (1989) reported cough, shortness of breath, and
    other symptoms among players and spectators of two high school hockey
    games played at an indoor ice arena in Minnesota, USA.  These symptoms
    were related to emissions from a malfunctioning engine of the ice
    resurfacer.  Although the exact levels of NO2 were not known at the
    time of the hockey game, levels of 7500 µg/m3 (4 ppm) were detected
    2 days later with the ventilation system working, suggesting that
    levels during the games were higher.  Hedberg et al. (1989, 1990)
    reported that pulmonary function testing performed on members of one
    hockey team with a single exposure demonstrated no decrease in lung
    function parameters at either 10 days or 2 months after exposure. 
    Dewailly et al. (1988) reported another incident at a skating rink in
    Quebec, Canada, in 1988 involving referees and employees with
    respiratory symptoms such as coughing, dyspnoea and a suffocating
    feeling.  Five days after the incident, NO2 levels had come down to
    5600 µg/m3 (3 ppm), suggesting much higher levels during the
    incident.

         In another skating rink study, Smith et al. (1992) reported the
    outcome of a questionnaire administered to all students from two high
    schools on 25 February, 1992, 3 days after 11 students participating
    in a Wisconsin indoor ice hockey tournament had been treated in
    emergency rooms for acute respiratory symptoms (i.e., cough,
    haemoptysis, chest pain and dyspnoea).  The game had been attended by
    131 students, 57 of whom reported symptoms.  A simulation test on
    24 February yielded NO2 levels of 2800 µg/m3 (1.5 ppm) in the air
    over the rink after use of the ice resurfacing machine.  Higher levels
    may have been reached on the night of the game.

         Brauer & Spengler (1994) measured indoor air NO2 concentrations
    at 20 skating rinks (most of all the operating ones) in the New
    England area of the USA.  Palmes tubes were used to measure NO2 over
    a 7-day sampling period at each rink, the samplers being placed on
    the main resurfacer used in the rink, at the score keepers' bench
    around a breathing height, at the opposite side of the rink from the
    scorekeeper bench, and outdoors nearby the rink away from parking
    lots or other vehicular traffic.  In contrast to the outdoor NO2
    concentrations observed (geometric mean = 0.018 ppm, range
    0.001-0.193 ppm), those found indoors averaged about 10-fold higher
    (geometric means = 0.128, 0.169, 0.168 ppm for resurfacer, bench area,
    and second sampler opposite to bench, respectively).  These NO2
    levels may be fairly typical for the approximately 2000 operating

    rinks in North America, with some differences being found depending on
    whether the resurfacer was propane- or gasoline-powered or used a
    catalytic converter.

    7.6  Occupational exposures

         Certain occupational exposure studies have shown that NO2
    exposures in occupational settings greatly exceed the chronic,
    low-level exposure described in general population epidemiological
    studies.  Occupational exposure studies generally refer to a highly
    selected group of adult workers, usually male.  The probability of a
    healthy worker effect needs to be considered when evaluating the
    significance of such studies.

         Gamble et al. (1987) studied 232 workers in four diesel bus
    garages for the effects of NO2 on acute respiratory illness and
    pulmonary function. Response was assessed by an acute respiratory
    questionnaire and before- and after-shift spirometry.  Measurements
    over the shift of NO2 (using passive Palmes tube samplers) were made
    on each worker and collected on the same day as the pulmonary function
    tests and questionnaires.  Other irritant gases were measured and were
    well below federal standards.  Mean NO2 levels over the shift ranged
    from 0.56 (SD = 0.38) ppm NO2 in the highest garage to 0.13
    (SD = 0.06) ppm NO2 in the lowest garage.  Short-term NO2
    measurements indicated levels above 1 ppm as being common.  The
    authors reported that the prevalence of acute respiratory symptoms
    were elevated above expected in the high-exposure (> 0.3 ppm) group
    only.  No reduction in pulmonary function was associated with
    exposure.

         Gamble et al. (1983) examined chronic respiratory effects in
    259 sodium chloride miners for whom diesel emissions were the
    principal NO2 source.  The Medical Research Council respiratory
    symptom questionnaire containing smoking history was administered by
    trained interviewers.  A chest X-ray and spirometry were also
    conducted.  Personal samples of NO2 and respirable particles for jobs
    in each mine were used to estimate cumulative exposure.  Mean exposure
    ranged from a low of 0.2 (SD = 0.1) ppm NO2 to a high of 2.5
    (SD = 1.3) ppm NO2.  The author reported that although cough was
    associated with age and smoking, and dyspnoea was associated with age,
    neither symptom was associated with exposure (i.e., years worked,
    estimated cumulative NO2 or respiratory particle exposure).  Reduced
    pulmonary function showed no association with NO2 exposure.

         Robertson et al. (1984) reported on a 4-year study of lung
    function in 560 British coal miners for whom the NOx source consisted
    of diesel emissions and blasting.  Overall average NO2 levels at nine
    coal mine sites ranged from 38 to 113 µg/m3 (0.02 to 0.06 ppm), and
    nitric oxide (NO) levels ranged from 0.13 to 1.19 ppm.  No
    relationship was found between exposure and decline in FEV1 or

    respiratory symptoms.  Jacobsen et al. (1988) conducted a more
    extensive investigation on nearly 20 000 miners at the same nine
    British coal mines to examine whether long-term exposure to low
    concentration of NO2 and NO was associated with increased
    susceptibility to respiratory infections.  Shift median levels were
    0.2 ppm NO and 0.03 ppm NO2. This complete and intensive study had
    problems with misclassification of exposure and outcome that are not
    uncommon when existing data are used for purposes that were not
    foreseen when the data were collected.  The authors concluded that the
    long-term exposure to the above-mentioned levels do not detectably
    increase the chance that miners will absent themselves from work
    because of a chest infection.

         Douglas et al. (1989) reported data obtained between 1955 and
    1987 on 17 patients examined at the Mayo Clinic for silo-filler's
    disease shortly after exposure to silo gas (NO2 levels from 200 to
    2000 ppm).  Health outcomes ranged from hypoxaemia and transient
    airway obstruction to death.  Epler (1989) noted that prevention is
    essential for elimination of silo-filler's disease.  Meulenbelt &
    Sangster (1990) indicated that, after a symptom-free period
    immediately following exposure to NO2, severe respiratory failure can
    develop several hours later. Other studies also examined high
    exposures (Lowry & Schuman, 1956; Grayson, 1956; Gregory et al., 1969;
    Yockey et al., 1980).

    7.7  Synthesis of the evidence for school-age children

         The weight of the evidence does not indicate that NO2 exposures
    at levels reported in studies evaluated here have any consistent
    effect on pulmonary function of a biologically significant magnitude. 
    Many of the indoor studies, however, suggest an increase in
    respiratory morbidity in children from exposure to NO2, although the
    effects in the majority of the studies do not reach statistical
    significance (p < 0.05).  The consistency of the results across the
    indoor studies is examined and the evidence from some of the studies
    is combined in a quantitative analysis presented below.  Indoor NO2
    epidemiological studies not included in combined analysis are listed
    in Table 62.

    7.7.1  Health outcome measures

         The studies in the quantitative analysis that follows use
    health outcome measures that provide an indication of the state of
    respiratory health of the various samples of children aged up to
    12 years.  The NO2 studies utilized standard questionnaires to
    evaluate lower respiratory health in children.  Diagnoses of specific
    respiratory diseases such as bronchiolitis or asthma were not made. 
    The factor of importance here is that an attempt was made to measure

    some aspect of lower respiratory morbidity.  Table 63 lists the health
    outcome measures for each study considered.  Whereas specific measures
    such as colds going to the chest (Melia et al., 1977), chest
    congestion, and phlegm with colds (Ekwo et al., 1983) are used to
    provide measures of lower respiratory morbidity, other measures use
    indexes, grouped responses or combined indicators of lower respiratory
    morbidity, some of which include measures such as colds going to
    chest.

         Childhood lower respiratory morbidity is characterized by a
    grouping of similar symptoms and diseases that reflect changes located
    anatomically in the lower respiratory tract.  This characterization
    represents an indication of severity of the respiratory morbidity
    status of the children and is a multifaceted approach to respiratory
    health in a population living under natural conditions.  Lower
    respiratory morbidity is the combination of different respiratory
    effects that have in common an evaluation of the morbidity status of
    the lower respiratory tract.  The measure of effect on the lower
    respiratory tract varied among the studies; the indicators, however,
    are conventional symptom and illness outcomes.  The symptoms are
    tabulated from similar standardized questionnaires (Ferris, 1978) and
    are directed at eliciting the same basic dataœan indication of the
    presence of illness or infection in the lower respiratory tract.

         Although the use of identical health outcome measures would be
    most desirable, the level of similarity and the common elements
    between the outcome measures in the NO2 studies provide some
    confidence in their use in the quantitative analysis.  However, the
    symptoms and illnesses combined are to some extent different and could
    indeed reflect different underlying processes.  Thus caution is
    necessary in interpreting the analysis.  This concern is addressed
    further later in this section as part of the statistical aspects of
    the random effects model.

    7.7.2  Biologically plausible hypothesis

         The human clinical and animal toxicological studies that examined
    NO2 effects on aspects of the respiratory host defence system provide
    a biologically plausible hypothesis compatible with the relationship
    seen between respiratory symptoms and morbidity and NO2 exposure in
    epidemiological studies.  However, research gaps in both animal
    toxicological and clinical studies exist, indicating the need for
    further research efforts.  A brief discussion is presented here.
        Table 62.  Indoor NO2 epidemiological studies not included in combined
               analysis - school age children (> 5 years)
                                                                                     

    Exposure            Published result                             Reference
                                                                                     

    Gas stoves          Increased prevalence of cough                Dodge (1982)

    Gas stoves          No significant association with              Schenker et al.
                        respiratory illness                          (1983)

    Gas stoves          No association with respiratory              Melia et al.
                        illness                                      (1988)

    Gas stoves          Increased lower respiratory                  Berwick et al.
                        symptoms in children < 7 years; no           (1989)
                        increase in children aged > 7 years

    Gas stoves          Increased prevalence of asthma               Kuehr et al.
                                                                     (1991)

    Gas cookers         Increased odds ratio for asthma;             Dekker et al.
                        non-significant increase in wheeze           (1991)

    Gas heaters and     No statistically significant increase        Spengler et al.
    cookers             in overall respiratory illness in            (1993)
                        24 cities in the USA

    Gas heaters         No association with respiratory              Goren et al.
                        illness                                      (1993)

    NO2 levels          No significant association with              Hoek et al.
                        bronchitis, asthma, frequent                 (1984)
                        coughs, allergy

    NO2 levels          No association with respiratory              Peat et al.
                        symptoms or bronchial                        (1990)
                        hyper-responsiveness

    NO2 levels          No association with lower                    Koo et al.
                        respiratory symptoms                         (1990)

    NO2 level           Increased odds ratio for asthma              Infante-Rivard
                                                                     (1993)

    NO2 in school       Increased respiratory symptoms and           Pilotto (1994)
    classrooms          absences from school
                                                                                     
    
        Table 63.  Health outcome measures in selected NO2 epidemiological studies
                                                                                                                                              

    Location (date      Health outcome used                Methoda                    NO2 exposure         Age       Sample   Reference
    of study)           in meta-analysis                                              measure used         (years)   size
                                                                                      in analysis
                                                                                                                                              

    Netherlands         Respiratory illness combination    Questionnaire completed    NO2 measured with    6-12      775      Brunekreef
    (1986)              variable of presence of one or     by parent (WHO).           Palmes tubes. Gas                       et al. (1987);
                        more of cough, wheeze or asthma.                              and electric                            Dijkstra et
                                                                                      appliances.                             al. (1990)

    28 areas of         Colds going to chest showed a      Respiratory symptoms       Gas stove vs.        6-11      5658     Melia et al.
    England and         prevalence of 26.8-19.8%.          questionnaire completed    electric stove                          (1977)
    Scotland (1973)                                        by parent of child for
                                                           the last 12 months

    27 areas of         Group response to respiratory      As above                   Gas stove vs.        5-10      4827     Melia et al.
    England and         questions into none or one or                                 electric stove                          (1979)
    Scotland (1977)     more symptoms or diseases.
                        Colds going to chest
                        (26.4-19.6%) showed the highest
                        prevalence, followed by wheeze
                        (10.1-6.2%), cough and episodes
                        of asthma or bronchitis in
                        last year.

    Middlesborough,     Group response to respiratory      As above                   NO2 measured with    6-7       103      Florey et al.
    England (1978)      questions as above.                                           Palmes tubes. Gas                       (1979); Goldstein
                                                                                      stove homes only.                       et al. (1979);
                                                                                                                              Melia et al.
                                                                                                                              (1980, 1982a,b)
                                                                                                                                              

    Table 63  (Con't)
                                                                                                                                              

    Location (date      Health outcome used                Methoda                    NO2 exposure         Age       Sample   Reference
    of study)           in meta-analysis                                              measure used         (years)   size
                                                                                      in analysis
                                                                                                                                              

    Middlesborough,     As above                           As above                   NO2 measured with    5-6       188      Melia et al.
    England (1980)                                                                    Palmes tubes. Gas                       (1982a,b)
                                                                                      stove homes only.

    6 USA cities        Lower respiratory illness index    Questionnaire (Ferris,     Gas vs. electric     6-10      8240     Ware et al.
    (1974-1979)         (index of respiratory health)      1978) completed by                                                 (1984)
                        indicating during the past year    parent for symptoms
                        the presence of either             during previous
                        bronchitis, respiratory illness    12 months
                        that kept the child home 3 days
                        or more, or persistent cough for
                        3 months of the year.

    6 USA cities        Combined indicator of one or       Symptom questionnaire      NO2 measured with    7-11      1286     Neas et al.
    (1983-1986)         more lower respiratory symptoms    completed by parent        Palmes tubes. Gas                       (1990, 1991)
                        as defined. The highest            for the year during        and electric stoves.
                        prevalences were for chronic       which measurements
                        phlegm and wheeze. The other       of NO2 were taken.
                        symptoms in the index are
                        shortness of breath, chronic
                        cough and bronchitis. Chest
                        illness reflects a restriction
                        of the child's activities for
                        3 or more days.

    Iowa City,          Chest congestion and phelgm        Questionnaire completed    Gas stove vs.        6-12      1138     Ekwo et al.
    Iowa, USA           with colds.                        by parent (ATS).           electric stove.                         (1993)
                                                                                                                                              

    Table 63  (Con't)
                                                                                                                                              

    Location (date      Health outcome used                Methoda                    NO2 exposure         Age       Sample   Reference
    of study)           in meta-analysis                                              measure used         (years)   size
                                                                                      in analysis
                                                                                                                                              

    Columbus, Ohio,     Lower respiratory illness          Telephone interview by     Gas stove vs.        < 12      553      Keller et al.
    USA (1978)          syndrome characterized by cough,   nurse epidemiologist.      electric stove.                         (1979a,b)
                        wheezing, bringing up phlegm
                        and like symptoms considered
                        as "chest colds".
                                                                                                                                              

    a  ATS = American Thoracic Society
             The evidence from animal toxicological and human clinical studies
    of host defence provides a rationale for investigating the relation
    between exposure to NO2 and an increase in frequency and severity of
    respiratory symptoms and/or infections in humans.  When microorganisms
    attack a lung that has been exposed to NO2, host defence mechanisms
    altered by the NO2 exposure may result in increased severity or rate
    of respiratory illness.  Although the host defence system reacts both
    very specifically and generally to the challenge, the overall response
    in humans is expressed as a generalized demonstration of signs
    and symptoms that may be associated with a site such as the lower
    respiratory tract. It may also be reported or objectively discerned
    as a general outcome, such as a chest cold, a cough or an incident of
    asthma or bronchitis.

    7.7.3  Publication bias

         Publication bias, also known as the "file drawer problem"
    (Rosenthal, 1979), is the result of the increased likelihood of
    publication of studies that have positive results, leading to a bias
    in the literature reviewed towards positive results.  There are two
    factors that make this bias less likely for epidemiological studies of
    NO2.  Firstly, the studies are expensive, well publicized, and
    the results are usually published in order to give credit to the
    researchers involved.  Secondly, many of the studies included in this
    section did not produce statistically significant findings, indicating
    that there was not a substantial barrier in publishing negative
    studies.  However, some studies are necessarily excluded because they
    provide insufficient information.  Although, this can lead to bias,
    there is little that can be done to correct for this problem.  This
    problem is not normally referred to as a publication bias, but it is a
    similar problem.

    7.7.4  Selection of studies

         An attempt has been made to include as many studies as possible
    in the quantitative analysis.  The requirements for inclusion were:
    (1) the health end-point measured must be reasonably close to the
    standard end-point; (2) significant exposure differences between
    subjects must exist and some estimate of exposure must be available;
    and (3) an odds ratio for a specified exposure gradient must have been
    calculated, or data presented so that an odds ratio can be calculated. 
    The standard end-point chosen was the presence of lower respiratory
    symptoms and illness in children aged 5 to 12 years.  The subsequent
    analysis is based on the assumption that the relative risk of
    developing lower respiratory symptoms is similar across this age range
    and across the range of study settings as a function of NO2 exposure,
    even though the baseline rates may differ by age and study setting. 
    After a careful review of the published literature, nine studies that
    met these criteria were selected for inclusion in the quantitative
    analysis.

         The NO2 exposure gradient for the quantitative analysis of
    relative risks was selected as 28.3 µg/m3 (0.015 ppm). This is
    comparable to the reported long-term exposure difference between homes
    with gas stoves and homes with electric stoves. In the USA, Neas et
    al. (1991) reported a household annual average difference of
    32.5 µg/m3 (0.0173 ppm) between homes with electric stoves and homes
    with gas stoves with pilot lights.  In the United Kingdom, Melia et
    al. (1980, 1982a,b) reported a difference of 31.1 µg/m3 (0.0165 ppm)
    in bedroom levels between homes with electric stoves and homes with
    gas stoves.  In four studies, chronic NO2 exposures were estimated
    from direct measurements using 1- to 2-week integrated indoor samples
    by Palmes passive diffusion tubes.

         For five studies that characterized NO2 exposure according to
    differences between gas stove and electric stoves, the exposure
    gradients were estimated from the two previously sited studies with
    direct NO2 measurements.  Appropriate exposure estimates ideally
    should be country-specific, current with the studies in location and
    time, and derived from a representative sample that appropriately
    characterizes the exposure.  For three studies conducted in the USA
    (Keller et al., 1979a,b; Ekwo et al., 1983; Ware et al., 1984),
    exposure gradients were based on the studies of Neas et al. (1991). 
    For the studies of Melia et al. (1977, 1979), exposure gradients were
    based on Melia et al. (1980, 1982a,b).  The effects of exposure
    measurement error related to the use of surrogate exposure estimates
    were discussed earlier.

    7.7.4.1  Brief description of selected studies

         Melia et al. (1977) studied children aged 6 to 11 years and
    developed an indicator of the presence of at least one of a group of
    symptoms including cough, colds going to the chest, and bronchitis. 
    The symptom reported most of the time was a cold going to the chest,
    which was used as an indicator of lower respiratory morbidity.  This
    study did not measure NO2 exposure, and so the assumption was made
    that the increase in NO2 exposure from gas stove use in England was
    reasonably similar to that in the other British studies that measured
    NO2 (31.1 µg/m3, 0.0165 ppm).  The estimated odds ratio was 1.31,
    with 95% confidence limits of 1.16 and 1.48. After adjusting to a
    standard increase of 28.3 µg/m3 (0.015 ppm), the odds ratio became
    1.28 with 95% confidence limits of 1.14 and 1.43.  No adjustment was
    made for parental smoking in this study.

         The cross-sectional data reported by Melia et al. (1979) on
    children aged 5 to 10 years were also employed to estimate an odds
    ratio, although no exposure estimates were made.  The presence or
    absence of a gas stove was used as a surrogate as in the Melia et al.
    (1977) study.  The estimated odds ratio was 1.24, with 95% confidence

    limits of 1.09 and 1.42.  After adjusting to a standard increase of
    28.3 µg/m3 (0.015 ppm), the odds ratio became 1.22 with 95%
    confidence limits of 1.08 and 1.37.

         Melia et al. (1980) studied children aged 6 to 7 years and
    measured bedroom NO2 levels for the exposure estimate.  This study
    applied the same combined health end-point as the previous study.  The
    estimated odds ratio for an increase of 28.3 µg/m3 (0.015 ppm) was
    1.49 with 95% confidence limits of 1.04 and 2.14.  Melia et al.
    (1982a,b) studied children aged 5 to 6 years and also measured NO2
    exposure in the bedroom and applied the same combined health
    end-point.  The estimated odds ratio for an increase of 0.015 ppm was
    1.11, with 95% confidence limits of 0.84 and 1.46. The 10th and the
    90th percentiles of the weekly measured concentrations were 0.009 and
    0.065 ppm NO2, respectively, in bedrooms (Melia et al., 1982b).

         In the first Harvard Six Cities study cohort, Ware et al. (1984)
    reported an index of respiratory illness.  Exposure to NO2 was based
    on the presence or absence of a gas stove (32.5 µg/m3, 0.0173 ppm). 
    The estimated odds ratio was 1.08 with 95% confidence limits of
    0.97 and 1.19.  After adjusting to a standard increase of 28.3 µg/m3
    (0.015 ppm), the odds ratios became 1.07 with 95% confidence limits of
    0.98 and 1.17.

         A second cohort of subjects in the Harvard Six Cities study was
    initially reported by Dockery et al. (1989a).  This cohort of children
    aged 7 to 11 years was then reinterviewed after indoor NO2
    measurements were made, and the results of this analysis were reported
    by Neas et al. (1990, 1991).  The 10th and 90th percentiles of the
    weekly measured concentrations were 0.008 and 0.033 ppm NO2,
    respectively, in bedrooms (Neas et al., 1991).  The estimated odds
    ratio for an increase in the presence of any respiratory symptom
    resulting from an increase in exposure of 28.3 µg/m3 (0.015 ppm) was
    1.40, with 95% confidence limits of 1.14 and 1.72.

         Ekwo et al. (1983) studied several respiratory illness end-points
    from children surveyed at ages 6 to 12 years.  No exposure
    measurements were obtained, and the exposure was based on the presence
    or absence of a gas stove (32.5 µg/m3, 0.0173 ppm).  None of the
    end-points matched the end-point of interest closely.  The two most
    similar end-points were hospitalization for chest illness before
    age 2 and chest congestion and phlegm with colds.  The estimated odds
    ratio for hospitalization was 2.40.  The estimated confidence limits
    for cough and phlegm with colds was 1.09, with 95% confidence limits
    of 0.82 and 1.45.  This last symptom appears to be most similar to the
    end-point of interest, and so it was included in the synthesis.

         The data presented by Dijkstra et al. (1990) on the study in the
    Netherlands were analysed and gave an estimated odds ratio of 0.94 for
    an increase of 28.3 µg/m3 (0.015 ppm) in NO2 exposure.  The 95%

    confidence limits were 0.70 and 1.27.  The study had measured NO2
    exposure data, but the meta-analysis did not adjust for covariates
    because the covariates were not included in the tables that included
    NO2 exposure.

         Keller et al. (1979b) did not find any statistically significant
    changes in respiratory disease associated with gas stove use, but the
    unadjusted estimated odds ratio for lower respiratory illness was
    1.10, with 95% confidence limits of 0.74 and 1.54.  Assuming that the
    exposure increase was 32.5 µg/m3 (0.0173 ppm), the odds ratio was
    adjusted to an exposure of 28.3 µg/m3 (0.015 ppm).  This resulted in
    an odds ratio of 1.09 with 95% confidence limits of 0.82 and 1.46.

    7.7.4.2  Studies not selected for quantitative analysis

         Five studies with sufficient information for analysis were
    excluded from the synthesis.  Two studies (Melia et al., 1983; Ogston
    et al., 1985) were on children under 1 year of age, whereas the others
    were on children of elementary school age.  Furthermore, the end-point
    of wheeze is more predominant in children less than 1 year old as
    opposed to older children, and the outcome measure in Ogston et al.
    (1985) included upper respiratory illness, making it dissimilar to the
    others.  The Berwick et al. (1989) analysis has been criticized for
    its lack of consistency across age groups, which may have resulted
    from the very small sample sizes.  The Swiss study (Braun-Fahrlaender
    et al., 1989, 1992) examined end-points that might not be considered
    similar to those of the other studies, such as upper respiratory
    disease, breathing difficulties and duration of various respiratory
    measures.  The Melia et al. (1988, 1990) study did not demonstrate
    significant exposure differences between the two groups contrasted
    (6.4 µg/m3, 0.0034 ppm).  These differences in exposure were much
    smaller than those seen for any other study of gas stove exposure.  If
    the relative risk were adjusted for an increase of 28.3 µg/m3
    (0.015 ppm), the relative risk would be about 1.29, which is
    comparable to the odds ratios seen in the other studies.  Because the
    difference in exposure groups was so small, requiring a very large
    adjustment, it was decided not to combine this study with the other
    studies.  For these reasons, the above-mentioned studies were
    not included in the synthesis.  These studies, however, support
    qualitatively the results of the synthesis.

    7.7.5  Quantitative analysis

         Combining evidence, often referred to as meta-analysis, is not
    new, having been used as early as 1904 (Pearson, 1904).  Such analyses
    are being used more frequently as indicated by Mann (1990).  The
    National Research Council (1986) combined evidence on the effect of
    environmental tobacco smoke on lung cancer using Peto's method as
    described by Yusuf et al. (1985).  Several methods for combining
    clinical trials were discussed by Laird & Mosteller (1990).  The

    evidence to be combined in this section comes from observational
    studies.  As a result, some of the methods used for clinical trials
    are not appropriate here, and the findings must be treated cautiously
    in light of the assumptions made when combining non-experimental
    studies.

         Two basic models are employed in order to combine evidence
    (Hasselblad et al., 1992).  The first model assumes that each study
    estimates the same fixed, but unknown, parameter.  Most methods of
    combining evidence make this assumption.  One of the earliest attempts
    to combine data using a fixed-effects model was given by Birge (1932). 
    His method weights the estimates inversely by their variances and
    produces combined estimate and associated confidence limits.  Other
    methods include the Mantel-Haenszel method (Mantel & Haenszel, 1959),
    which is used to combine contingency tables.  Recently, Bayesian
    methods have been used to combine evidence, and methods particularly
    appropriate to these kinds of studies were described by Eddy (1989)
    and Eddy et al. (1990a,b).  Bayesian analyses require the choice of a
    prior distribution for the parameter of interest, which is often a
    non-informative prior.  A non-informative prior is one that, prior to
    seeing the evidence, favours no value of the parameter over any other. 
    The interesting fact about use of these methods is that, for the data
    sets considered in Table 63, the results of the computations were
    nearly identical.  This is because the (marginal) likelihood for the
    odds ratio is closely approximated by a log-normal curve.

         The second basic model assumes that the parameter of interest is
    not fixed, but is itself a random variable from a distribution.  The
    value of this random variable is different for each study, but each
    study gives some information about the mean of the distribution. 
    These models go by several names, including random-effects models,
    mixed models, two-stage models and hierarchical models. The purpose of
    a random-effects model is to relax the assumption that each study is
    estimating exactly the same parameter. This idea is not new, having
    been discussed by Cochran (1937).  A discussion of the interpretation
    of random-effects models in clinical trials and several methods of
    estimating the parameters of these models was provided by DerSimonian
    & Laird (1986).  If the studies being combined tend to estimate the
    same parameter, then the results using a random-effects model will
    approach the results using a fixed-effects model.  On the other hand,
    if the studies are estimating very different parameters, then the
    confidence limits will tend to be much broader than those obtained
    from a fixed-effects model.

    FIGURE 25

         The nine studies described earlier (Tables 63 and 64) were
    combined using both kinds of models.  Graphs of the odds ratio from
    each study are depicted in Fig. 25.  Each curve can be given one of
    three interpretations: (1) the normal approximation to the marginal
    likelihood of the logarithm of the odds ratio, (2) a distribution such
    that the 2.5 percentile and the 97.5 percentile points of the
    distribution are the 95% confidence limits of the estimated odds
    ratio, and (3) the posterior for the odds ratio for a particular study
    given a flat prior on the log odds ratio.  The results using a
    fixed-effects model are labelled "fixed", and results of the random
    effects model are labelled "random" (see Fig. 25).  Methods for
    estimating the parameters of a random-effects model were described by
    DerSimonian & Laird (1986) and Eddy et al. (1992).  The results of the
    analyses are provided in Table 65 (US EPA, 1993).

    Table 64.  Summary of odds ratios from studies on the effects of
               NO2 increased by 0.015 ppm (from: US EPA, 1993)
                                                                        

    Authors                  Estimated odds     2.5 and 97.5 Percentiles
                                 ratio            (confidence interval)
                                                                        

    Melia et al. (1977)          1.28                 1.14 to 1.43
    Melia et al. (1979)          1.22                 1.08 to 1.37
    Melia et al. (1980)          1.49                 1.04 to 2.14
    Melia et al. (1982a,b)       1.11                 0.84 to 1.46
    Ware et al. (1984)           1.07                 0.98 to 1.17
    Neas et al. (1991)           1.40                 1.14 to 1.72
    Ekwo et al. (1983)           1.09                 0.82 to 1.45
    Dijkstra et al. (1990)       0.94                 0.70 to 1.27
    Keller et al. (1979b)        1.09                 0.82 to 1.46
                                                                        

         The first line of Table 65 shows the results of combining all
    nine studies using a fixed model.  The estimated odds ratio is 1.17
    and the 95% confidence limits are 1.11 and 1.23.  The analysis was
    made assuming that the parameters were homogeneous, and this can be
    tested.  The chi-square test for homogeneity for the nine studies was
    12.32 for 8 degrees of freedom, which has a p value of 0.1375.  Thus,
    there is some evidence that the parameters from each study are not
    identical. The estimates for the random-effects model are similar to
    the estimates for the fixed-effects model, but the confidence limits
    are slightly broader.  The conclusion from both models is the same,
    namely that the odds ratio is estimated to be about 1.2, with 95%
    confidence intervals ranging from about 1.1 to 1.3 (Hasselblad et al.,
    1992).  Many researchers have suggested that the random-effects model
    is the more appropriate one, because it does not assume that all
    studies estimate the same parameter.


        Table 65.  Combined analyses of studies on respiratory illness effects of nitrogen dioxide increased by 0.015 ppm
               (from: US EPA, 1993)
                                                                                                                        

    Group                    Number of            Fixed-effects model                       Random-effects model
                              studies                                                                                   
                                            Odds ratio     Confidence interval      Odds ratio     Confidence interval
                                                                                                                        

    All                          9            1.17             1.11 to 1.23            1.18           1.09 to 1.27
    USA                          4            1.11             1.03 to 1.20            1.14           1.00 to 1.29
    United Kingdom               4            1.25             1.15 to 1.35            1.25           1.13 to 1.37
    Measured NO2                 4            1.23             1.08 to 1.41            1.22           0.99 to 1.50
    Gas stove surrogate          5            1.15             1.09 to 1.22            1.16           1.08 to 1.24
    SES adjusted                 3            1.27             1.17 to 1.37            1.27           1.15 to 1.41
    SES not adjusted             6            1.08             1.00 to 1.16            1.09           0.98 to 1.21
    Smoking adjusted             2            1.28             1.09 to 1.52            1.25           0.92 to 1.71
    Smoking not adjusted         7            1.15             1.09 to 1.22            1.16           1.06 to 1.27
    Gender adjusted              5            1.26             1.18 to 1.36            1.27           1.16 to 1.39
    Gender not adjusted          4            1.06             0.98 to 1.15            1.06           0.96 to 1.17
                                                                                                                        
             These studies include results from North America and Europe. 
    Meta-analyses of studies from different countries are common.  For
    example, Canner (1987), Littenberg (1988), and Jaeschke et al. (1990)
    all combined some studies in both North America and Europe and did not
    adjust for geographic differences.  The indoor NO2 studies were
    compared by country.

         The studies were compared by similarity of subjects.  Four of
    them were conducted in the United Kingdom (Melia et al., 1977, 1979),
    and four in the USA (Keller et al., 1979a,b; Ware et al., 1984; Neas
    et al., 1990, 1991; Ekwo et al., 1993).  The United Kingdom studies
    provide a higher estimated odds ratio (1.25) than the USA studies
    (1.11).

         Four of the nine studies used measured NO2 values, whereas the
    other five did not.  The use of a surrogate for exposure should tend
    to reduce the estimate of the effect (Samet & Utell, 1990).  The
    measured NO2 studies gave an estimated odds ratio of 1.23, whereas
    the others gave an estimate of 1.15, which is consistent with a
    measurement error effect.  The chi-square tests for homogeneity were
    not significant at the 0.1 level for either group of studies.

         Table 66 lists the important covariates considered in these nine
    studies and shows if the covariate was used in the study and the
    meta-analysis (US EPA, 1993).  Study design and exposure measurement
    source are also presented.  The effect of having adjusted for various
    covariates can be seen in Table 64.  In general, those studies that
    adjusted for a particular covariate found larger odds ratios than
    those that did not.

         Although there may be reasons to weight certain studies or groups
    of studies more heavily than others, the results indicate that there
    is an increase in the odds of respiratory disease of children exposed
    to NO2, especially those of elementary school age.  The estimates are
    generally centered around an odds ratio of 1.2 with 95% confidence
    limits of 1.1 and 1.3 (Hasselblad et al., 1992), although the studies
    using measured NO2 give a slightly higher estimate of the odds ratio. 
    The estimates are not sensitive to the assumption that each study is
    estimating the same parameter as indicated by the random-effects
    model.  In fact, the finding of increased risk across a wide variety
    of study conditions suggests that the effects seen are not an artifact
    of any one particular study.

        Table 66.  Covariate treatment and other factors in selected NO2 epidemiological studies in meta-analysis
               (from: US EPA, 1993)
                                                                                                                        

    Reference                             Covariatesa                               Exposure measurement source
                                                           
                                  SES     Parental    Gender    Design
                                          smoking
                                                                                                                        

    Melia et al. (1977)            A        NM          A       Cross-sectional     Gas stove vs. electric stove.b
    Melia et al. (1979)            A        M           A       Cross-sectional     Gas stove vs. electric stove.b
    Melia et al. (1980)            M        M           A       Cross-sectional     NO2 measured with Palmes tubes.
                                                                                    Gas stove homes only.
    Melia et al. (1982a,b)         M        M           A       Cross-sectional     NO2 measured with Palmes tubes.
                                                                                    Gas stove homes only.
    Ware et al. (1984)             M        M           M       Cross-sectional     Gas stove vs. electric stove.c
    Neas et al. (1991)             A        A           A       Cross-sectional     NO2 measured with Palmes tubes.
                                                                                    Gas and electric stove homes.
    Ekwo et al. (1983)             NM       A           M       Cross-sectional     Gas stove vs. electric stove.c
    Dijkstra et al. (1990)         M        M           M       Cross-sectional     NO2 measured with Palmes tubes.
                                                                                    NO2 emissions sources in homes.
    Keller et al. (1979b)          M        NM          M       Prospective         Gas stove vs. electric stove.c
                                                                                                                        

    a  SES = Socioeconomic status; A = Covariate included in study and meta-analysis; NM = Not measured in study;
       M = Measured in study but data not available for meta-analysis
    b  Estimate of exposure derived from assumption of gas stove versus electric stove levels in bedrooms in
       England from data in Melia et al. (1980, 1982a,b) of approximately 0.0165 ppm.
    c  Estimate of exposure derived from assumption of gas stove with pilot light versus electric stove levels
       averaged in the home in the USA in Neas et al. (1991) of approximately 0.0173 ppm.
             These results are not sensitive to the inclusion or exclusion of
    any one study.  If the analysis had included the hospitalization
    results of Ekwo et al. (1983), the analysis of the Swiss study, or the
    Berwick et al. (1989) study, there would have been little change in
    the estimated odds ratios or their 95% confidence limits.  It is also
    possible to delete any one study from the analysis, and still obtain
    nearly the same results.  In fact, any two studies can be deleted from
    the analysis, and the estimated odds ratio will have a confidence
    interval that does not include 1.0.

         There is always the concern that the studies described in this
    monograph are not the complete list of studies, but contain primarily
    the positive studies because these are the studies most likely to be
    published.  Alternatively, non-significant results may not be reported
    with sufficient quantitative detail to permit their inclusion.  Both
    of these effects can be considered as "publication bias" (see section
    7.6.3).  It is of interest to contemplate an undiscovered study with
    results so negative that, when combined with the other studies,
    produces a confidence interval for the odds ratio that includes the
    value 1.0. If we assume that the hypothetical study would be the size
    of the Ware et al. (1984) study, then its odds ratio for increased
    respiratory symptoms as the result of a 28.3 µg/m3 (0.015 ppm)
    exposure would have to be 0.77.  Subject to assumptions made for the
    combined analysis for school-aged children, the main conclusion from
    the analysis was that an increased risk of about 20% for respiratory
    symptoms and disease corresponded to each increase of 28 µg/m3
    (0.015 ppm) in estimated 2-week average NO2 exposure, where mean
    weekly bedroom concentrations in studies reporting NO2 levels were
    predominantly between 0.008 and 0.065 ppm NO2 (Hasselblad et al.,
    1992).

    7.8  Synthesis of the evidence for young children

         Various researchers have conducted studies of children less than
    2 years of age (see Table 67).  A major difference for this group of
    studies is that the health outcome measures are less uniform than
    the studies of older children.  For purposes of comparability, a
    meta-analysis similar to the one for older children was made.

        Table 67.  Summary of odds ratios of the effects of nitrogen dioxide, health outcome and exposure estimates in epidemiological studies
               on young children (< 2 year)
                                                                                                                                              

    Reference                Estimated    2.5 and 97.5   Health outcome                          NO2 exposure   Age           Location (date
                             odds ratio   Percentiles                                              estimate                      of study)
                                          (confidence                                               (ppm)
                                          interval)
                                                                                                                                              

    Melia et al. (1983)         0.63      0.36-1.10      Respiratory illness incidence              0.0165a     < 1 year      England (1978)
    Ekwo et al. (1983)          2.4       1.06-3.74      Hospitalization for chest illness          0.0173b     < 2 years     Iowa, USA
                                                         before age 2
    Ware et al. (1984)          1.11      0.97-1.27      Respiratory illness before age 2           0.0173b     < 2 years     Six Cities
                                                                                                                              USA (1974-1979)
    Ogston et al. (1985)        1.14      0.86-1.50      Respiratory illness incidence              0.0165a     < 1 year      Scotland (1980)
    Dockery et al. (1989a)      1.15      0.96-1.37      Respiratory illness before age 2           0.015c      < 2 years     Six Cities USA
                                                                                                                              (1983-1986)
    Margolis et al. (1992)      1.12      0.63-2.04      Persistent lower respiratory symptoms      0.0105d     < 1 year      North Carolina,
                                                                                                                              USA (1986-1988)
    Samet et al. (1993)         0.99e     0.94-1.04e     Lower respiratory illness incidence        0.015f      < 18 months   Albuquerque,
                                                                                                                              USA (1988-1990)
                                                                                                                                              

    a  Estimate of exposure derived from assumption of gas stove versus electric    d  Estimate of exposure derived from assumption of gas
       stove levels in bedrooms in England from data in Melia et al. (1980,            stove versus electric stove levels averaged in the
       1982a) of approximately 0.0165 ppm.                                             home in the Albuquerque study (Samet et al., 1993)
    b  Estimate of exposure derived from assumption of gas stove with pilot            of approximately 0.0105 ppm.
       light versus electric stove levels averaged in the home in the USA in        e  Computed from logistic regression coefficient derived
       Neas et al. (1991) of approximately 0.0173 ppm.                                 from Samet et al. (1993).
    c  Estimate of exposure derived from assumption of gas stove versus electric    f  Exposure level used to convert logistic regression to
       stove levels averaged in the home in USA in Neas et al. (1991) of               an odds ratio.
       approximately 0.015 ppm.
             The seven studies of young children shown in Table 67 show mixed
    results.  A test of homogeneity of the odds ratios gives a chi-squared
    value of 22.66 for 6 degrees of freedom, which has a p value of
    0.0009, implying that the studies are not homogenous.  The variation
    in results could be due to several factors, including different health
    outcome measures and other factors.  Dockery et al. (1989a) noted that
    the associations discussed by Ware et al. (1984) and Dockery et al.
    (1989a) must be viewed with caution because they compared recalled
    respiratory events early in the child's life.  Because of the
    heterogeneity, the studies were combined using a random-effects model.
    Subject to the assumptions made for the meta-analysis, the combined
    odds ratio for the increase in respiratory disease per increase of
    0.015 ppm NO2 was 1.09 with a 95% confidence interval of 0.95 to
    1.26, where mean weekly concentrations in bedrooms were predominately
    between 0.005 and 0.05 ppm NO2 in studies reporting levels.  The
    increase in risk was very small and was not reported consistently by
    all studies.  We cannot conclude that the evidence suggests an effect
    in young children comparable to that seen in older children.

    7.9  Summary

         The evidence from individual studies of the effect of NO2 on
    lower respiratory symptoms and disease in school-aged children is
    somewhat mixed.  The consistency of these studies was examined and
    the evidence synthesized in a combined quantitative analysis
    (meta-analysis) of the subject studies.  Most of the indoor studies
    showed increased lower respiratory morbidity in children associated
    with long-term exposure to NO2.  Mean weekly NO2 concentrations in
    bedrooms in studies reporting NO2 levels were predominately between
    15 and 122 µg/m3 (0.008 and 0.065 ppm) (Hasselblad et al., 1992). 
    Combining the indoor studies as if the end-points were similar gives
    an estimated odds ratio of 1.2 (95% confidence limits of 1.1 and 1.3)
    for the effect per 28.3 µg/m3 (0.015 ppm) increase of NO2 on lower
    respiratory morbidity (Hasselblad et al., 1992).  This suggests that,
    subject to assumptions made for the combined analysis, an increase
    of about 20% in the odds of lower respiratory symptoms and disease
    corresponded to each increase of 28.3 µg/m3 (0.015 ppm) in estimated
    2-week average NO2 exposure.  Thus, the combined evidence is
    supportive for the effects of estimated exposure to NO2 on lower
    respiratory symptoms and disease in children aged 5 to 12 years.

         In the individual indoor studies of young children (2 years of
    age or younger), no consistent relationship was found between
    estimates of NO2 exposure and the prevalence of respiratory symptoms
    and disease.  Based on a meta-analysis of these indoor infant studies,
    subject to the assumptions made for the meta-analysis, the combined
    odds ratio for the increase in respiratory disease per increase of
    28.2 µg/m3 (0.015 ppm) NO2 was 1.09 with a 95% confidence interval
    of 0.95 to 1.26, where mean NO2 weekly concentrations in bedrooms
    were predominately between 9.4 and 94 µg/m3 (0.005 and 0.05 ppm) in
    studies reporting levels.  The increase in risk was very small and was

    not reported consistently by all studies.  We cannot conclude that the
    evidence suggests an effect in infants comparable to that seen in
    older children.  The reasons for these age-related differences are not
    clear.

         The measured NO2 studies gave a higher estimated odds ratio than
    the surrogate estimates, which is consistent with a measurement error
    effect.  The effect of having adjusted for covariates such as
    socioeconomic status, smoking and sex was that those studies that
    adjusted for a particular covariate found larger odds ratios than
    those that did not.

         Although many of the epidemiological studies that involved
    measured NO2 levels used measurements over only 1 or 2 weeks, these
    levels were used to characterize children's exposures over a much
    longer period. The standard respiratory symptom questionnaire used by
    most of these studies summarizes information on health status over an
    entire year.  The 28.2 µg/m3 (0.015 ppm) difference in NO2 levels
    used in the meta-analyses relates to a difference in the household
    annual average exposure between gas and electric cooking stoves.
    Some studies measured NO2 levels only in the winter and may have
    overestimated annual average exposures.  This would tend to have
    underestimated the health effect of a 28.2 µg/m3 (0.015 ppm)
    difference in the annual NO2 exposure.  The study of Neas et al.
    (1991), which was based on household annual average exposure measured
    in both the winter and summer, found a stronger health effect than
    many of the other studies.  The true biologically relevant exposure
    period is unknown, but these exposures extended over a lengthy period
    up to the entire lifetime of the child.

         The association between outdoor NO2 and respiratory health is
    not clear from current research.  There is some evidence that the
    duration of respiratory illness may be increased at higher ambient
    NO2 levels.  A major difficulty in the analysis of outdoor studies is
    distinguishing possible effects of NO2 from those of other associated
    pollutants.

         Several uncertainties need to be considered in interpreting the
    above studies and results of the meta-analysis.  Error in measuring
    exposure is potentially one of the most important methodological
    problems in epidemiological studies of NO2.  Although there is
    evidence that symptoms are associated with indicators of NO2
    exposure, the quality of these exposure estimates may be inadequate to
    determine a quantitative relationship between exposure and symptoms. 
    Most of the studies that measured NO2 exposure did so only for
    periods of 1 to 2 weeks and reported the values as averages.  Few of
    the studies attempted to relate the effects seen to the pattern of
    exposure, such as transient peaks.  Furthermore, measured NO2
    concentration may not be the biologically relevant dose  per se;
    estimating actual exposure requires knowledge of both pollutant levels

    and related human activity patterns.  However, only very limited
    activity and aerometric data are available that examine such factors,
    and the extrapolation to possible patterns of ambient exposure is
    difficult.  In addition, although the level of similarity and common
    elements between the outcome measures in the NO2 studies provide some
    confidence in their use in the quantitative analysis, the symptoms and
    illnesses combined are to some extent different and could indeed
    reflect different underlying processes.  Thus, caution is necessary in
    interpreting the meta-analysis results.

         Other epidemiological studies have attempted to relate some
    measure of indoor and/or outdoor NO2 exposure to changes in pulmonary
    function.  These changes were marginally significant.  Most studies
    did not find any effects, which is consistent with controlled human
    exposure study data (see Chapter 6).  However, there is insufficient
    epidemiological evidence to draw any conclusions about the long- or
    short-term effects of NO2 on pulmonary function.

         On the basis of a background level of 15 µg/m3 (0.008 ppm) and
    the fact that significant health effects occur with an additional
    level of 28.2 µg/m3 (0.015 ppm) or more, an annual guideline value of
    40 µg/m3 (0.023 ppm) is proposed. This value will avoid the most
    severe exposures.  The fact that a no-effect level for subchronic or
    chronic NO2 exposure concentrations has not yet been determined
    should be emphasized.

    8.  EVALUATION OF HEALTH AND ENVIRONMENT RISKS ASSOCIATED WITH
        NITROGEN OXIDES

    8.1  Sources and exposure

         Combustion provides the major source of nitrogen oxides in both
    indoor and outdoor air, producing mostly NO, typically about 90%, with
    some NO2 and small quantities of other species.  Some domestic
    combustion appliances can produce more than 10% of NOx as NO2.  The
    sum of NO and NO2 is generally referred to as NOx.  In the air, NO
    is oxidized to NO2.  This happens rapidly by reaction with ozone, and
    also by a slower photochemical process requiring the presence of
    reactive organic compounds and sunlight.  Nitrogen oxides together
    with reactive organic compounds are precursors for ozone and
    photochemical smog formation.  NO and NO2 may also undergo reactions
    to form a range of other nitrogenous species, including HNO2, HNO3,
    NO3, N2O5, PAN and other organic nitrates.  The complete range of
    gas phase nitrogen oxides is often referred to as NOy.  The
    partitioning of nitrogen among these different compounds is strongly
    dependent on the concentrations of other oxidants and on the
    meteorological history of the air.

         Nitrogenous species have lifetimes in the air ranging from
    minutes to several days.  In general, emissions of NO and NO2 are
    progressively oxidized to HNO3 and nitrate.  Air contaminated by NOx
    emissions and their reaction products can be advected large distances. 
    Exposure tends to be predominantly to NO close to a source, NO2 in
    the local region and HNO3 and nitrates at distances of up to several
    hundred kilometres or more.

         Human and environmental exposure to nitrogen oxides varies
    greatly from indoors to outdoors, from the city to the countryside,
    and with the time of day and season.  The concentrations of NO and
    NO2 typically present outdoors in a range of urban situations is
    relatively well established.  The concentrations encountered indoors
    depend on the specific details of the nature of combustion appliances,
    chimneys and ventilation.  When unvented combustion appliances are
    used for cooking or heating, indoor concentrations of nitrogen oxides
    usually greatly exceed those existing outside.

         Nitrogen oxides are ultimately removed from the atmosphere mostly
    as nitrate by processes of dry and wet deposition.

         Nitrous oxide (N2O) is emitted to the atmosphere from biological
    and some combustion processes.  It is inert in the troposphere but in
    the stratosphere plays a role in the chemistry of stratospheric ozone. 
    N2O is also a greenhouse active gas.

         In indoor air, the concentration and composition of nitrogen
    oxide species is largely the result of indoor combustion sources.
    NO is in greater concentration than NO2, usually by a factor of
    up to ten-fold.  In some cases indoors, HNO2 has been reported at
    concentrations that are more than 10% of those of NO2.  HNO2 may be
    produced from surface reactions of NO or NO2 with water.

         There are several difficulties with measurements of nitrogen
    oxides.  A straightforward interference-free method exists for
    measuring NO (the chemiluminescent reaction with ozone), but this is
    the exception for nitrogenous species.  By firstly converting NO2 to
    NO, this chemiluminescence technique is also commonly used to measure
    NO2.  Unfortunately, the catalysts usually employed for this
    conversion are not specific and, especially for emissions that have
    undergone substantial photochemical reaction, other oxidized nitrogen
    compounds present in the sampled air are also converted to NO and are
    measured as NO2.  For this reason, great care must be taken in
    interpreting the results of the common chemiluminescence analyser in
    terms of NO2, as the signal may represent the sums of NO2 and
    several other nitrogen compounds.  Additional measurement difficulties
    arise because oxidized nitrogen in the atmosphere can also be present
    in both the gas phase and as particulate matter.  For indoor air
    measurements, the Palmes tube technique of NO2 measurement is
    frequently used.  This technique is not suited for measurement of
    short-term peak concentrations.

    8.2  Evaluation of the effects of atmospheric nitrogen species on the
         environment

         Guidance values have been estimated for both critical levels of
    NOx (the air concentration threshold for effects on plants) and
    critical loads of total nitrogen (the deposited nitrogen load to
    ecosystems above which adverse effects can occur).  Since deposited
    nitrogen acts on ecosystems by increasing the nutrient status of
    soils, there is no definitive threshold for effects; all additional
    nitrogen will result in some response.

         The individual nitrogen species present in the polluted
    atmosphere cannot be completely separated with respect to their
    effects on the environment. The relative contribution of NO and NO2
    to the NOx effect on plants is unclear. The vast majority of
    information is on effects of NO2 but available information on NO
    suggests that NO and NO2 have comparable phytotoxic effects. Total
    nitrogen deposition has been used to assess effects on ecosystems
    since it is not possible to identify the relative contribution of
    nitrogen species to nutrient nitrogen elevation.

         Concerning organisms in the environment, information is almost
    exclusively restricted to plants, with minimum data on soil fauna. 
    The evaluation and guidance values are, therefore, expressed in terms
    of nitrogen species effects on vegetation. However, it is expected
    that plants will form the most sensitive component of natural systems
    and that the effect on biodiversity of plant communities is a
    sensitive indicator of biotic effects on the whole ecosystem.

         Gaseous nitrogen species reduce photosynthesis and biomass
    production and increase sensitivity to other stresses (such as frost,
    drought and insect damage) of individual plants. At the level of
    plant communities and ecosystems, eutrophication is more important
    than toxicity, nitrogen causing reduction in biodiversity in
    nutrient-limited habitats.

         Deposited nitrogen will change the chemistry of soils; these
    changes are reflected biologically in total ecosystem effects.
    However, there is one feature that needs separate consideration;
    deposited nitrogen contributes to the leaching of nitrate through soil
    profiles and into surface and groundwater.

         The atmospheric chemistry of nitrogen oxides includes the
    capacity for ozone generation in the troposphere, ozone depletion in
    the stratosphere, and direct and indirect contribution to global
    warming as greenhouse gas.  Nitrogen oxides and ammonia contribute to
    soil acidification (along with sulfur oxides) and thereby to increased
    bioavailability of aluminium.

         The phytotoxic effects of gaseous nitrogen oxides on plants have
    little direct relevance to crop plants when concentrations are close
    to the critical level (see section 8.2.1). However, the role of NOx
    in the generation of ozone and other phytotoxic substances leads to
    crop loss.  Nitrogen deposited on growing crops will represent a very
    small increase in total available nitrogen compared to that added as
    fertilizer.

    8.2.1  Guidance values - critical levels for air concentrations of
           nitrogen oxides

         Critical levels are mostly derived from fumigation experiments.
    In the majority of studies with NO or NO2, no significant effects on
    plants were found at air concentrations below 100 µg/m3.  Most of the
    experiments were designed to evaluate mechanisms of action of nitrogen
    oxides rather than to quantify adverse effect thresholds, and few
    exposure-response relationships have been established.

         Interactions between NOx and other atmospheric pollutants have
    been reported.  Generally SO2 acts synergistically with NO2. In
    mixtures of NO2 with other gases, such as NO, O3 and CO2,
    interactive effects are the exception rather than the rule.

         In order to include the impact of NO, a critical level for NOx
    is proposed instead of one for NO2; for this purpose it has been
    assumed that NO and NO2 act in an additive manner.

         A strong case can be made for the provision of critical levels
    for short-term exposure. However, there are currently insufficient
    data to provide these with confidence.  Current evidence suggests a
    critical level of about 75 µg/m3 for NOx as a 24-h mean.

         The critical level for NOx (NO and NO2 added in ppb and
    expressed as NO2 in µg/m3) is considered to be 30 µg/m3 as an
    annual mean (see section 4.1.8 for detailed reasoning).

         At concentrations slightly above this critical level, growth
    stimulation is the dominant effect (possibly combined with some
    increase in sensitivity to biotic and abiotic stresses) if the ambient
    conditions allow growth and NOx is the only pollutant.  Where biomass
    production is beneficial, for example in agriculture or plantation
    forest, the critical level is probably higher.

         The critical level can be converted to deposition quantities. The
    annual deposition that corresponds to a NOx level of 30 µg/m3 is
    5 to 10 kg/ha. This indicates that at a concentration near to its
    critical level, the contribution of NOx to nitrogen demand is
    negligible for fertilized crops but not for natural vegetation.

    8.2.2  Environment-based guidance values - critical loads for total
           nitrogen deposition

         Critical loads are derived from empirical data and steady-state
    soil models. Estimated critical loads for total nitrogen deposition in
    a variety of natural aquatic and terrestrial ecosystems are given in
    Table 68 (details can be found in section 4.2).  Possible differential
    effects of deposited nitrogen species (NOy and NHx) are
    insufficiently known to differentiate between nitrogen species for
    critical load estimation.

         Atmospheric nitrogen deposition can significantly contribute to
    the leaching of nitrates to surface water and groundwater. There is
    not enough information to provide a guidance value with broad
    applicability for this effect, but the few studies on this subject
    indicate that critical loads are relatively low (see section 4.2.7).


        Table 68.  Summary of guidelines for total nitrogen deposition
               (kg nitrogen per ha per year) in natural and semi-natural
               freshwater and terrestrial ecosystems
                                                                                             

    Ecosystem                         Critical load    Indication
                                                                                             

    Shallow soft-water lakes             5-10a         Decline in isoetid species

    Mesotrophic fens                     20-35b        Increase in tall graminoids;
                                                       decline in diversity

    Ombrotrophic (raised) bogs           5-10b         Decreased Sphagnum and subordinate
                                                       species; increase in tall graminoids

    Calcareous species-rich              14-19a        Increase in tall grasses; decline
      grassland                                        in diversity

    Neutral/acid species-rich            20-30b        Increase in tall grasses; decline
      grassland                                        in diversity

    Montane-subalpine grassland          10-15c        Increase in tall graminoids; decline
                                                       in diversity

    Lowland dry heathland                15-20a        Transition of heather to grass

    Lowland wet heathland                17-22a        Transition of heather to grass

    Species-rich heaths/acid             7-20b         Decline in sensitive species
     grassland

    Arctic and alpine heaths             5-15c         Decline in lichens, mosses and
                                                       evergreen dwarf shrubs; increase
                                                       in grasses and herbs

    Coniferous tree health               11 - > 50b    Nutrient imbalance

    Deciduous tree health                15-20b        Nutrient imbalance; shoot-root ratio

    Acidic (managed) coniferous forest   15-20b        Changes in ground flora

    Acidic (managed) deciduous forest    15-20b        Changes in ground flora

    Calcareous forests                   15-20c        Changes in ground flora
                                                                                             

    a  reliable estimate
    b  reasonably reliable estimate
    c  best guess
             The majority of ecosystems for which there is sufficient
    information to estimate critical loads have temperate climates. The
    few arctic and montane ecosystems included, which might be expected to
    be representative of higher latitudes, have the least reliable basis.
    There is no information on tropical ecosystems. Nutrient-poor tropical
    ecosystems such as rain forests and mangrove swamps are likely to be
    adversely affected by nitrogen deposition. The lack of both deposition
    data and effect thresholds make it impossible to make risk assessments
    for these climatic regions.

         The most sensitive ecosystems for which effects thresholds can be
    estimated show critical loads of 5-10 kg nitrogen per ha per year
    based on decreased biological diversity in plant communities.  A more
    average value for the limited range of ecosystems studied is 15-20 kg
    nitrogen per ha per year.

    8.3  Evaluation of health risks associated with nitrogen oxides

    8.3.1  Concentration-response relationships

         Table 69 summarizes key health effects observed in controlled
    human exposure (clinical) studies with NO2 exposure durations of
    0.5 to 4 h.  At higher exposure levels, i.e. more than 2800 µg/m3
    (1.5 ppma), NO2 exposure results in increased airway responsiveness
    and increased airway resistance in healthy adults.  However, some
    researchers have not observed any NO2-induced changes in airway
    resistance at NO2 levels between 3800 and 7500 µg/m3 (2 to 4 ppm).

         The physiological end-point that, to date, appears to be the most
    sensitive indicator of response is a change in airway responsiveness
    to bronchoconstrictors in asthmatics.  This increase in airway
    responsiveness has been observed in some, but not all, studies. 
    Several individual studies found significant responses at NO2
    concentrations within the range of 380 to 560 µg/m3, with exposure
    periods varying from 30 to 180 min.  A meta-analysis of 20 studies in
    asthmatics suggests that increased airway responsiveness may occur at
    concentrations as low as 200 µg/m3.  However, no individual studies
    showed clearly significant effects on airway responsiveness at
    190 µg/m3 (0.1 ppm) for 60 min.  Additionally, small decreases in
    forced expiratory volume in 1 second (FEV1) or forced vital capacity
    (FVC) in adult or adolescent asthmatics have been observed in response
    to 560 µg/m3 (0.3 ppm) NO2 for 30 min.  However, clear NO2

              

    a  For the purposes of risk evaluation, NO2 concentration presented
       in µg/m3 that were originally derived from concentrations
       expressed as ppm have been rounded off.  The conversion factor
       is 1880 µg/m3 = 1 ppm.


        Table 69.  Key effects of exposure to nitrogen dioxide on human health - clinical studies
                                                                                                                                          

    NO2 concentration   Exposure       Observed effectsa
                        duration
                                                                                                                                          

    376-564 µg/m3       0.5-2.0 h      Trend toward increased airway responsiveness to challenges in asthmatics. However, no significant
    (0.2-0.3 ppm)                      effects observed by same or other investigators at NO2 levels up to 4 ppm. Small (4-6%)
                                       decreases in FEV1 or FVC in adult or adolescent asthmatics, in response to NO2 alone.

    564 µg/m3           3.7 h          Small decreases (5-9%) in FVC and FEV1 in COPD patients with mild exercise. No effects seen
    (0.3 ppm)                          by other investigators for COPD patients at 0.5-2 ppm NO2.

    2820-3760 µg/m3     2-3 h          Increased airway responsiveness to bronchoconstrictors in healthy adults. However, effects not
    (1.5-2.0 ppm)                      detected by other investigators at 2-4 ppm.

    „ 3760 µg/m3        1-3 h          Lung function changes (e.g., increased airway resistance) in healthy subjects. Effects not found
    („ 2.00 ppm)                       by others at 2-4 ppm.
                                                                                                                                          

    a  FEV1 = Forced expiratory volume in 1 second; FVC = Forced vital capacity; COPD = Chronic obstructive pulmonary disease
    
    concentration-response relationships are not evident for either airway
    responsiveness or pulmonary function changes.  Other studies of
    asthmatics exposed to 7500 µg/m3 (4 ppm NO2) for 75 min did not show
    effects on pulmonary function or airway responsiveness.

         A second category of sensitive subjects comprises patients with
    chronic obstructive pulmonary disease (COPD).  Although small
    decreases have been observed in FVC and FEV1 in COPD patients exposed
    to 560 µg/m3 (0.3 ppm) in one study, effects were not seen in other
    studies at higher exposure levels. The collective evidence from
    epidemiological studies examining relationships between estimates of
    exposure to NO2 and lower respiratory symptoms and disease is
    summarized in Table 70.  Lower respiratory symptoms in children are
    generally an indicator of the incidence and severity of respiratory
    illnesses that are often related to viral infections.

         Indoor NO2 exposures are associated with increased
    lower-respiratory illness in children aged 5-12 years, but there is
    no consistent evidence of such an association among younger children
    (0-2 years).  Among primary school children, the risk of lower
    respiratory illness increases by about 20% for an increase of
    28 µg/m3 (0.015 ppm) NO2 indoors (averaged over 1 year).  There
    is no evidence from the epidemiological literature that the
    concentration-response relationship for NO2 and respiratory illness
    differs from linearity.  This is consistent with the concentration-
    response relationship reported from studies of outdoor NO2.

         The epidemiological studies mostly used estimates of NO2
    exposures that were averaged over a period of several weeks to 1 year. 
    Moreover, the health outcomes assessed in these studies were generally
    accumulated over 6-12 months.  This means that the risk estimates
    produced by most of the epidemiological studies (both indoor and
    outdoor) refer to long-term average exposures.

         There is no evidence in the epidemiological literature to
    determine whether there is, or is not, a level of NO2 below which
    health effects are not observed.  However, a quantitative review of
    several large well-conducted studies has shown a statistically
    significant excess of lower respiratory symptoms in homes with gas
    stoves (NO2 levels approximately 38-56 µg/m3 (0.02-0.03 ppm)
    averaged over 12 months) compared with homes with electric stoves
    (average NO2 levels: 9-13 µg/m3 (0.005-0.007 ppm)).

        Table 70.  Key effects of exposure to nitrogen dioxide on human health - epidemiological studies
                                                                                                                                              

                   NO2 exposure                                                              Observed effects
                                                                                                                                              

    0.015-ppm increase, where mean weekly                    A meta-analysis shows an increased risk of lower respiratory symptoms/disease in
    concentrations in bedrooms in studies reporting          children 5 to 12 years old associated with exposure estimates of NO2 levels. The
    levels were mainly between 0.008 and 0.065 ppm NO2       95% confidence interval of the odds ratio was 1.1 to 1.3. Main source of exposure
    (1- and 2-week integrated average NO2 concentration)     contrast is homes with gas and electric stoves.

    0.015-ppm increase in annual average of 2-week           In individual indoor studies of infants < 2 years of age, no consistent relationship
    NO2 levels, where mean weekly concentrations in          was found between estimates of NO2 exposure and prevalence of respiratory symptoms
    bedrooms were mainly between 0.005 and                   and disease. Based on a meta-analyses of these infant studies, the combined odds
    0.050 ppm                                                ratio for the increase in respiratory disease per increase of 0.015 ppm NO2 was
                                                             1.09 with a 95% confidence interval of 0.95 to 1.26. Thus, although the overall
                                                             combined estimate is positive, it contains the no-effect value of 1.0, (i.e., is not
                                                             statistically significant); and so cannot conclude that the evidence suggests an
                                                             effect in infants comparable to that seen in older children.

    > 0.3 ppm (average exposure during work shift)           Elevated prevalence of acute respiratory symptoms

    Episodic exposure during ice hockey game to              Occurrence of acute respiratory symptoms (cough, chest pain, dyspnoea)
    NO2 levels of 1.5 ppm or more

    25 to 100 ppm (episodic occupational exposure)           Bronchitis, bronchiolitis and bronchial pneumonia induced by very high NO2
                                                             exposure.

    > 200 ppm (extreme episodic exposures)                   Extreme exposure health outcomes range from hypoxaemia/transient airway
                                                             obstruction to death
                                                                                                                                              
             Higher levels (> 560 µg/m3, > 0.3 ppm during a shift at work)
    in an occupational setting were related to an elevated prevalence of
    acute respiratory symptoms in adults.  Episodic exposures occurring
    over a period of 1 h or longer at levels possibly as high as
    2800 µg/m3 (1.5 ppm) or more have resulted in the occurrence of acute
    respiratory symptoms.  Exceptionally high acute occupational exposures
    of 47-188 mg/m3 (25 to 100 ppm) NO2 have resulted in broncho-
    pneumonia, bronchitis or bronchiolitis; and very extreme occupational
    NO2 exposures (> 200 ppm) have been associated with effects that
    range from hypoxaemia and transient obstruction of the airways to
    death from adult respiratory distress syndrome.

         Numerous concentration-response studies have been conducted with
    animals using a wide range of exposure durations and end-points.
    The major classes of effects observed at concentrations less than
    1880 µg/m3 (1.0 ppm) include decreases in host defences, alterations
    in lung metabolism (e.g., increased lipid peroxidation and antioxidant
    metabolism), epithelial remodelling of the lower respiratory tract,
    thickening of the centriacinar interstitium, and a variety of
    extrapulmonary changes.  Such findings can be qualitatively
    extrapolated to humans, but major uncertainties in respiratory tract
    dosimetry and species sensitivity currently preclude a quantitative
    extrapolation.  Structural changes in the lung become more severe as
    exposures proceed from weeks to months at a given NO2 concentration. 
    Only substantially higher NO2 concentrations exceeding 22 000 µg/m3
    (12 ppm) have caused emphysema, as defined by criteria developed by
    the US National Institutes of Health. 

         In order to examine the relative importance of concentration (C)
    of NO2 and duration of exposure (T) in the development of increased
    susceptibility to respiratory infection, the effects of different
    C × T products have been evaluated.  Results from infectivity studies
    examining patterns of exposure indicate that, at the same C × T
    product, concentration exerts more influence than duration of exposure
    in increasing susceptibility to respiratory bacterial infection in
    mice.

         Table 71 lists quantitative findings from a few key animal
    studies showing the lowest concentrations that caused several types
    of effects.  Of most importance are findings showing increased
    susceptibility to infection with long-term exposure to NO2 levels as
    low as 940 µg/m3 (0.5 ppm) and of other impacts on host defences with
    exposure to NO2 levels as low as 560 to 940 µg/m3 (0.3 to 0.5 ppm). 
    These findings provide evidence supporting the biological plausibility
    of the association of increased respiratory illness in older children
    in relation to indoor NO2 exposures.

        Table 71.  Key animal toxicological effects of exposure to nitrogen dioxide
                                                                                                                                              

    NO2 concentration (ppm)                     Species   Observed effects
    (exposure duration)
                                                                                                                                              

    0.04 ppm (continuous, 9 months)             Rat       Increased lipid peroxidation (ethane in exhaled breath)

    0.2 ppm (continuous base for 1 year) plus   Mouse     Increased susceptibility to respiratory infection and decreased vital capacity
    0.8 ppm (1-h peak, 2x/day, 5 days/week)               and respiratory system compliance, compared to control or baseline only

    0.25 ppm (7 h/day, 5 days/week, 7 weeks)    Mouse     Systemic effect on cell-mediated immunity

    0.3 ppm (2 h/day, 2 days)                   Rabbit    Decreased phagocytosis of alveolar macrophages

    0.4 ppm (continuous, 4 weeks)               Mouse     Decreased systemic humoral immunity

    0.4 ppm (continuous, 9 months)              Rat       Increased antioxidants and antioxidant metabolism

    0.4 ppm (continuous, up to 27 months)       Rat       Slight increase in thickness of air-blood barrier at 18 months, becoming
                                                          significant by 27 months; also alterations in bronchiolar and alveolar
                                                          epithelium by 27 months

    0.5 ppm (continuous, 3 months)              Mouse     Increased susceptibility to respiratory infection

    0.5-28 ppm (6 min to 1 year)                Mouse     Linear increase in susceptibility to respiratory infection with time, increased slope
                                                          of curve with increased concentration, concentration more important than time

    0.5 ppm (continuous base, 6 weeks) plus     Rat       Alterations in Type 2 cells and increased interstitial matrix of proximal
    1.5 ppm (1-h peak, 2x/day, 5 days/week)               alveolar region, no changes in terminal bronchiolar region of adults
                                                                                                                                              
        8.3.2  Subpopulations potentially at risk

         Certain groups within the population may be more susceptible to
    the effects of NO2 exposure, including people with pre-existing
    respiratory disease, children and the elderly.  The reasons for paying
    special attention to these groups is that: (i) they may be affected by
    lower levels of NO2 than other subpopulations; or (ii) the severity
    of health effects at a given exposure level may be greater.  Some
    causes of heightened susceptibility are better understood than others.
    Subpopulations that already have reduced ventilatory reserves
    (e.g., the elderly and persons with asthma, emphysema and chronic
    bronchitis) are likely to be affected more than other groups by
    similar decreases in pulmonary function.  For example, a healthy young
    person may not notice a small change in pulmonary function, but a
    person whose activities are already limited by reduced lung function
    may not have the reserve to compensate for such a change.

         Several hundred million people suffer from asthma worldwide. 
    Asthmatic individuals appear to be the most susceptible members of
    the population with regard to respiratory responses to NO2.  On
    average, asthmatic persons are much more sensitive to inhaled
    bronchoconstrictors such as histamine, methacholine or carbachol.  The
    airways of asthmatics are also hyperresponsive to a variety of other
    inhaled materials, including pollen, cold dry air, allergens and air
    pollutants.  The potential addition of a further NO2-induced increase
    in airway response to the already heightened responsiveness to other
    substances raises the possibility of exacerbation of this pulmonary
    disease by NO2 in asthmatic individuals.

         Other potentially susceptible groups include patients with COPD,
    such as emphysema and chronic bronchitis.  Several hundred million
    adults worldwide suffer from COPD.  Some of these patients have airway
    hyperresponsiveness to physical and chemical stimuli.  A major concern
    with COPD patients is the absence of an adequate ventilatory reserve,
    a susceptibility factor described above.  In addition, the poor
    distribution of respiratory tract ventilation in COPD may lead to a
    greater delivery of NO2 to the segment of the lung that is well
    ventilated, thus resulting in a greater regional tissue dose. 
    Furthermore, NO2 exposure may alter already impaired defence
    mechanisms, making patients with COPD more susceptible to respiratory
    infection.

         On the basis of epidemiological studies, children aged 5 to
    12 years constitute a subpopulation potentially susceptible to an
    increase in respiratory morbidity associated with NO2 exposure. 
    Worldwide, nearly a billion (109) children fall into the age groups
    at increased potential risk for increased respiratory illnesses
    associated with NO2 exposures.  However, the fraction of the number
    of potentially at-risk children in various age groups that are
    actually exposed to NO2 concentrations/patterns sufficient to induce
    respiratory morbidity has not been determined.

         Another potentially susceptible subpopulation group is
    immunocompromised individuals, who would have an increased
    susceptibility for infectious pulmonary disease as well as other
    health effects.  Such people could be potentially more susceptible to
    agents, such as NO2, that further compromise pulmonary host defences. 
    It is clear that NO2 can affect alveolar macrophages, humoral
    immunity and cell-mediated immunity in otherwise normal animals. 
    However, the animal-to-human extrapolation cannot yet be made
    quantitatively.  Although these immunocompromised groups represent
    potentially susceptible populations for NO2 effects, no human
    research has directly examined the effects of NO2 exposure in these
    groups.

    8.3.3  Derivation of health-based guidance values

         Increased airway responsiveness observed in asthmatic subjects
    with 0.5 to 2.0 h exposures to 380-560 µg/m3 (0.2-0.3 ppm) NO2
    represents an adverse health effect of concern induced by acute,
    short-term human exposures to NO2.  However, some laboratories have
    not observed similar effects with comparable duration NO2 exposures
    at levels above the 380-560 µg/m3 (0.2-0.3 ppm) range.  A possible
    reason might be the difference in severity of asthma of the subjects
    exposed.  Nevertheless, increased airway responsiveness may pose a
    risk for asthmatic individuals (i.e. increased responsiveness to other
    commonly occurring stimuli such as cold air, allergens and other air
    pollutants).

         On the basis of an effect level at 400 µg/m3 and the possibility
    of effects at lower levels, based on a meta-analysis, a 1-h average
    daily maximum NO2 concentration not exceeding 200 µg/m3
    (approx. 0.11 ppm) is recommended as a short-term guideline. 
    This should be adequate to protect most asthmatic subjects from
    experiencing NO2-induced increased airway responsiveness to stimuli
    that might otherwise disrupt their typical daily activities and reduce
    their work productivity.  Similarly, adherence to such a guideline
    should also provide adequate protection against the occurrence of
    pulmonary function decreases in COPD patients or other individuals
    with already compromised lung function.

         Epidemiological observations of associations between increased
    respiratory illness in school children and indoor and outdoor
    exposures to NO2 are suggestive of human health effects associated
    with long-term NO2 exposures.  This is supported by animal
    toxicological findings showing increased susceptibility to respiratory
    infections and impairment of host defences as a result of subchronic
    or chronic exposures to NO2 concentrations near to the ambient
    concentrations.  However, no confident quantitative extrapolation can
    yet be made of these animal toxicological findings to determine
    comparable human exposures, nor can one confidently interpret the
    epidemiological findings as to whether the reported increased

    respiratory illness risk is associated with: (a) chronic low-level
    indoor NO2 exposures; or (b) repeated higher short-term NO2
    excursions that also occur indoors during gas stove use (cooking,
    heating).  However, a quantitative review of several large,
    well-conducted epidemiological studies has shown an excess of lower
    respiratory illness among children aged 5-12 years exposed to annual
    average indoor NO2 concentrations of 38-56 µg/m3 (0.02 to 0.03 ppm).

         On the basis of a background level of 15 µg/m3 (0.008 ppm) and
    the fact that significant adverse health effects occur with an 
    additional level of 28.2 µg/m3 (0.015 ppm) or more, an annual
    guideline value of 40 µg/m3 (0.023 ppm) is proposed.  This value will
    avoid the most severe exposures.  The fact that a no-effect level for
    subchronic or chronic NO2 exposure concentrations has not yet been
    determined should be emphasized.

         There is uncertainty surrounding the lifetime effect because
    studies have not extended beyond individuals older than 12 years. 
    There is no evidence for non-linearity in the concentration-response
    relationship below these levels.  Long-term exposures of experimental
    animals to levels as low as 940 µg/m3 (0.5 ppm) with 1880 µg/m3
    peaks for 5 days increased the mortality for infectious agents,
    indicating an impairment of the immune system.  These animal data
    support the observations of increased respiratory infections seen in
    epidemiological studies.  Chronic and subchronic exposures of
    experimental animals demonstrate biochemical, morphological and
    physiological changes at higher NO2 levels.  Continuing damage occurs
    as the exposure time increases suggesting cumulative effects from
    long-term NO2 exposures.  Although the long-term guidance value does
    not provide a margin of safety, this level will avoid the most severe
    concentrations to which children are commonly exposed.

         With regard to possible health-based guidelines for other
    nitrogen oxides, insufficient information exists on which to base
    guidelines at this time.  Before guidelines can be established for NO,
    HNO2 and other oxidized nitrogen species, which may have important
    health impacts, much more information needs to be gathered in human
    clinical, epidemiological and experimental animal studies.

    9.  CONCLUSIONS AND RECOMMENDATIONS FOR PROTECTION OF HUMAN HEALTH AND
        THE ENVIRONMENT

         Nitrogen oxides can reach concentrations in ambient and indoor
    air that may affect human health.  Short-term NO2 exposure causes
    decreases in lung function and increased airway responsiveness.  Other
    effects include decreases in host defences and alterations in lung
    cells and their activity. Long-term exposure to NO2 is associated
    with respiratory illness.  Individuals with asthma and chronic
    obstructive pulmonary disease are more susceptible than healthy
    individuals.  Children aged 5 to 12 years constitute a subpopulation
    potentially susceptible to an increase in respiratory morbidity
    associated with NO2 exposure.

         On the basis of human controlled exposure studies on asthmatics
    and other high-risk groups, the recommended short-term guidance value
    is for a one-hour average daily maximum NO2 concentration not
    exceeding 200 µg/m3 (0.11 ppm).  The recommended long-term guidance
    value, based on epidemiological studies with increased risk of
    respiratory illness in children is 40 µg/m3 (0.023 ppm) as an annual
    average.

         Limited information exists regarding the health effects of the
    other oxidized nitrogen species (e.g., NO, HNO2, HNO3) as well as of
    mixtures of nitrogenous air pollutants.  Owing to the limited
    database, it is not possible to evaluate potential health risks of
    exposure to these compounds, even though they may be of significance.

         Current total nitrogen deposition in some areas of the world is
    causing reduced biodiversity in ecosystems.  To halt and/or reverse
    these trends, emissions of nitrogen must be reduced.

         Gaseous nitrogen species can reduce photosynthesis and biomass
    and increase the sensitivity of individual plants to other stresses. 
    The critical level for NO2 is considered to be 30 µg/m3 as an annual
    average.

         At the level of plant communities and ecosystems, eutrophication
    dominates over toxicity, with deposited total nitrogen acting as a
    nutrient and causing reduction in biodiversity in nutrient-limited
    habitats. Critical loads for the most sensitive ecosystems are
    estimated at 5-10 kg nitrogen per ha per year; a more average value
    for ecosystems is 15-20 kg nitrogen per ha per year.

         Reversibility of ecosystem effects of deposited nitrogen can
    partly be achieved by managed techniques. In unmanaged systems,
    reversibility, where possible, can be a long-term or a very long-term
    process.  In some cases where erosion and acidification are extreme,
    effects may be irreversible.

         Nitrogen oxides act as greenhouse gases and thus contribute to
    global warming, which may have far-reaching effects on human health
    and the environment.

    10.  FURTHER RESEARCH

    1.   Further epidemiological research is required to resolve the
         issues of:

         a)   the apparent age- and gender-related differences in
              NO2-related health effects;
         b)   the relative importance of chronic or subchronic low-level
              exposure and episodic high-level exposure to NO2;
         c)   the relative importance of NO2 and fine particles to health
              effects of ambient air pollution;
         d)   synergism between exposure to NO2 and other airborne
              contaminants such as ozone, fine particles and bioaerosols;
         e)   modification of the effect of NO2 on the respiratory system
              owing to other environmental factors such as temperature,
              humidity and exposure to viral and other infectious
              pathogens;
         f)   the significance of the observed health effects due to
              low-level NO2 exposure for long-term health outcomes.

    2.   Further human controlled experimental studies are needed on:

         a)   multi-hour, repeated exposure to NO2 to simulate the
              episodic human exposures encountered both outdoors and
              indoors using bronchoalveolar lavage (BAL) and molecular
              biology analysis;
         b)   respiratory responses to HNO2 using BAL and molecular
              biology analysis to correlate respiratory changes with other
              biochemical end-points;
         c)   respiratory and other physiological responses to high
              (approx. 5 ppm) levels of NO to evaluate the effects of
              NO detected indoors;
         d)   research to investigate the relative importance of
              concentration, exposure duration and minute ventilation to
              the health outcome.

    3.   Further animal studies are needed on:

         a)   short- and long-term effects of NO at concentrations ranging
              about those found indoors, with emphasis on mechanisms of
              action so that the animal studies may be related to human
              end-points in epidemiology and controlled human exposure;
         b)   short- and long-term effects of HNO2, with an emphasis on
              mechanism(s) of action and effects on the immune system,
              preferably using head or nose-only studies to reduce
              complications of characterization of the test atmosphere
              resulting from interaction of HNO2 with surfaces;

         c)   identification of mechanism(s) of action NO2 on those
              limited end-points now identified with human health effects
              (e.g., immune defences, airway activity, disease outcome,
              and lung growth);
         d)   effects of well-defined mixtures of nitrogenous air
              pollutants that will simulate those encountered indoors and
              in polluted outdoor air;
         e)   newer animal models of allergic disease and better
              diagnostic procedures for allergic disease in experimental
              animals should be applied to the study of nitrogenous air
              pollutants.

    4.   Further research on atmospheric chemistry regarding:

         a)   exposure to potentially toxic nitrated organic compounds,
              including aromatic/organic nitrates and peroxyacyl nitrates;
         b)   the formation, removal and human exposure pathways of HNO2
              and other potentially toxic compounds produced by the
              interaction of HNO2 with other pollutants.

    5.   Further research on ecosystems is needed concerning:

         a)   the relative roles of different deposited nitrogen species
              (NOx and NHy);
         b)   quantitative data on the effects of NO on plants to
              establish the relative roles of the components of NOx;
         c)   effects of nitrogen deposition on fauna;
         d)   the study of ecosystems representative of tropical climates
              to develop estimates of critical loads relevant to a global
              assessment of the effects of nitrogen;
         e)   effects of nitrogen deposition on montane and arctic
              ecosystems;
         f)   effects of nitrogen deposition on aquatic ecosystems for
              both freshwater and estuarine/marine areas;
         g)   effects of management regimes on grassland, heathland and
              plantation forest in relation to effects of nitrogen
              deposition.

    REFERENCES

    Aaby B (1990) [Nature monitoring report: Monitoring of raised bogs
    1989.] Forest and Nature Administration (in Danish).

    Aaby B (1994) Monitoring Danish raised bogs. In: Grünig A ed. Mires
    and man. Mire conservation in a densely populated country - the Swiss
    experience. Birmensdorf, Kosmos, pp 284-300.

    Abd Aziz SA & Nedwell DB (1986a) The nitrogen cycle of an east coast,
    UK, saltmarsh: I. Nitrogen assimilation during primary production;
    detrital mineralization. Estuar Coast Shelf Sci, 22: 559-575.

    Abd Aziz SA & Nedwell DB (1986b) The nitrogen cycle of an east coast,
    UK, saltmarsh: II. nitrogen fixation, nitrification, denitrification,
    tidal exchange. Estuar Coast Shelf Sci, 22: 689-704.

    Abe M (1967) Effects of mixed NO2-SO2 gas on human pulmonary
    functions: effects of air pollution on the human body. Bull Tokyo Med
    Dent Univ, 14: 415-433.

    Abraham WM, Kim CS, King MM, Oliver W Jr, & Yerger L (1982) Effects of
    nitric acid on carbachol reactivity of the airways in normal and
    allergic sheep. Arch Environ Health, 37: 36-40.

    Abrahamsen G & Thompson WN (1979) A long-term study of the enchytraid
     (Oligochaeta) fauna of a mixed coniferous forest and the effects of
    urea fertilization. Oikos, 1979: 318-327.

    Acton JD & Myrvik QN (1972) Nitrogen dioxide effects on alveolar
    macrophages. Arch Environ Health, 24: 48-52.

    Adams KM, Japar SM, & Pierson WR (1986) Development of a MnO2-coated,
    cylindrical denuder for removing NO2 from atmospheric samples. Atmos
    Environ, 20: 1211-1215.

    Adams WC, Brookes KA, & Schelegle ES (1987) Effects of NO2 alone and
    in combination with O3 on young men and women. J Appl Physiol,
    62: 1698-1704.

    Adaros G, Weigel HJ, & Jäger HJ (1991a) Concurrent exposure to SO2
    alters the growth and yield responses of wheat and barley to low
    concentrations of CO2. New Phytol, 118: 581-591.

    Adaros G, Weigel JH, & Jäger HJ (1991b) Single and interactive effects
    of low levels of O3, SO2, and NO2 on the growth and yield of spring
    rape. Environ Pollut, 72: 269-286.

    Adgate JL, Reid HF, Morris R, Helms RW, Berg RA, Hu PC, Cheng PW, Wang
    OL, Muelenaer PA, Collier AM, & Henderson FW (1992) Nitrogen dioxide
    exposure and urinary excretion of hydroxyproline and desmosine. Arch
    Environ Health, 47(5): 376-384.

    Adkins B Jr, Van Stee EW, Simmons JE, & Eustis SL (1986) Oncogenic
    response of strain A/J mice to inhaled chemicals. J Toxicol Environ
    Health, 17: 311-322.

    Adnot S, Kovyoumdjian C, Defouilloy C, Andrivet P, Sediame S,
    Herigault R, & Fratacci MD (1993) Hemodynamic and gas exchange
    responses to infusion of acetylcholine and inhalation of nitric oxide
    in patients with chronic obstructive lung disease and pulmonary
    hypertension. Am Rev Respir Dis, 148: 310-316.

    Aerts R & Berendse F (1988) The effects of increased nutrient
    availability on vegetation dynamics in wet heathlands. Vegetatio,
    76: 63-69.

    Aerts R, Berendse F, De Caluwe H, & Schmitz M (1990) Competition in
    heathland along an experimental gradient of nutrient availability.
    Oikos, 57: 310-318.

    Aerts R, Wallén B, & Malmer N (1992) Growth-limiting nutrients in
     Sphagnum-dominated bogs subject to low and high atmospheric nitrogen
    supply. J Ecol, 80: 131-140.

    Agren GI (1983) Nitrogen productivity of some conifers. Can J For Res,
    13: 494-500.

    Agren GI & Bosatta E (1988) Nitrogen saturation of terrestrial
    ecosystems. Environ Pollut, 54: 185-197.

    Ahmed T, Dougherty R, & Sackner MA (1983a) Effect of NO2 exposure on
    specific bronchial reactivity in subjects with allergic bronchial
    asthma (Final report). Warren, Michigan, General Motors Research
    Laboratories (Report No. CR-83/07/BI).

    Ahmed T, Dougherty R, & Sackner MA (1983b) Effect of 0.1 ppm NO2 on
    pulmonary functions and non-specific bronchial reactivity of normals
    and asthmatics (Final report). Warren, Michigan, General Motors
    Research Laboratories (Report No. CR-83/11/BI).

    Albritton DL, Liu SC, & Kley D (1984) Global nitrate deposition
    from lightning. In: Aneja VP ed. Environmental impact of natural
    emissions: Proceedings of an Air Pollution Control Association
    specialty conference, Research Triangle Park, March 1984. Pittsburgh,
    Pennsylvania, Air Pollution Control Association.

    Alexander IJ & Fairly RI (1983) Effects of N fertilization on
    populations of fine roots and mycorrhizas in spruce humus. Plant Soil,
    71: 49-53.

    Alexander V & Schnell DM (1973) Seasonal and spatial variation in
    nitrogen fixation in the Barrow, Alaska, tundra. Arct Alp Res,
    5: 77-88.

    Alheid U, Frolich JC, & Förstermann U (1987) Endothelium-derived
    relaxing factor from cultured human endothelial cells inhibits
    aggregation of human platelets. Thromb Res, 47: 561-571.

    Altshuller AP (1986) The role of nitrogen oxides in nonurban ozone
    formation in the planetary boundary layer over N America, W Europe and
    adjacent areas of ocean. Atmos Environ, 20(2): 245-268.

    American Thoracic Society Committee on Standards for Epidemiologic
    Surveys in Chronic Respiratory Disease (1969) Standards for
    epidemiologic surveys in chronic respiratory disease. New York,
    National Tuberculosis and Respiratory Disease Association.

    Amoruso MA, Witz G, & Goldstein BD (1981) Decreased superoxide anion
    radical production by rat alveolar macrophages following inhalation of
    ozone or nitrogen dioxide. Life Sci, 28: 2215-2221.

    Anderson IC & Levine JS (1987) Simultaneous field measurements of
    biogenic emissions of nitric oxide and nitrous oxide. J Geophys Res
    (Atmos), 92: 965-976.

    Anderson LS & Mansfield TA (1979) The effects of nitric oxide
    pollution on the growth of tomato.  Environ Pollut, 20: 113-121.

    Andreae M, Delany AC, Liu S, Logan S, Steele LP, Westberg H, & Zika R
    (1989) Key aspects of species related to global biogeochemical cycles.
    In: Lenschow DH & Hicks BB ed. Global tropospheric chemistry: chemical
    fluxes in the global atmosphere. Boulder, Colorado, National Center
    for Atmospheric Research.

    Angell JK (1988) An update through 1985 of the variations in global
    total ozone and north temperate layer-mean ozone. J Appl Meteorol,
    27: 91-97.

    Anlauf KG, Fellin P, Wiebe HA, Schiff HI, Mackay GI, Braman RS, &
    Gilbert R (1985) A comparison of three methods for measurement of
    atmospheric nitric acid and aerosol nitrate and ammonium. Atmos
    Environ, 19: 325-333.

    Anonymous (1991) [Monitoring of long-range air pollution precipitation
    - Annual report 1989.] Oslo, Norway, State Pollution Control Authority
    (Report No. 437/91) (in Norwegian).

    Anonymous (1993) Air pollution and tree health in the United Kingdom
    (Report of the Department of the Environment). London, Her Majesty's
    Stationary Office (HMSO), 88 pp.

    Aranyi C, Fenters J, Erhlich R, & Gardner D (1976) Scanning electron
    microscopy of alveolar macrophages after exposure to oxygen, nitrogen
    dioxide, and ozone. Environ Health Perspect, 16: 180.

    Aris R, Christian D, & Balmes LR (1991a) The effects of nitric acid
    vapour alone, and in combination with ozone, in exercising, healthy
    subjects as assessed by bronchoalveolar and proximal airway lavage. Am
    Rev Respir Dis, 143(suppl): A97.

    Aris R, Christian D, Sheppard D, & Baumes JR (1991b) The effects of
    sequential exposure to acidic jog and ozone on pulmonary function in
    exercising subjects. Am Rev Respir Dis, 143: 85-91.

    Arner EC & Rhoades RA (1973) Long-term nitrogen dioxide exposure:
    effects on lung lipids and mechanical properties. Arch Environ Health,
    26: 156-160.

    Arnolds E (1988) The changing macromycete flora in The Netherlands.
    Trans Br Mycol Soc, 90: 391-406.

    Arnolds E (1991) Decline of ectomycorrhizal fungi in Europe. Agric
    Ecosys Environ, 35: 209-244.

    Aronsson A (1980) Frost hardiness in Scots pine  (Pinus sylvestris).
    II. Hardiness during winter and spring in young trees of different
    nutritional status. Studia For Suec, 155: 14-50.

    Arts GHP (1990) Deterioration of atlantic soft-water systems and their
    flora, a historical account. The Netherlands, University of Nijmegen
    (Ph.D. thesis).

    Arts GHP, Van Der Velde G, Roelofs JGM, & Vany Swaay CAM (1990)
    Successional changes in the soft-water macrophyte vegetation of
    (sub)atlantic, sandy, lowland regions during this century. Freshwater
    Biol, 24: 287-294.

    Ashenden TW (1979) Effects of SO2 and NO2 pollution on transpiration
    in  Phaseolus vulagris L. Environ Pollut, 18: 45-50.

    Ashenden TW, Bell SA, & Rafarel CR (1990) Effects of nitrogen dioxide
    pollution on the growth of three fern species. Environ Pollut,
    66: 301-308.

    Ashenden TW, Bell SA, Edge CP, Rafarl CR, & Willian TG (1993) Critical
    loads of N & S deposition to semi-natural vegetation. Bangor, United
    Kingdom, Institute for Terrestrial Ecology, 75 pp (Project report
    No. T07064L5).

    Asman WAH (1987) Atmospheric behaviour of ammonia and ammonium. The
    Netherlands, Agricultural University of Wageningen (Ph.D. thesis).

    Atkinson R (1990) Gas-phase tropospheric chemistry of organic
    compounds: a review. Atmos Environ, A24: 1-41.

    Atkinson R & Lloyd AC (1984) Evaluation of kinetic and mechanistic
    data for modelling of photochemical smog. J Phys Chem Ref (Data),
    13: 315-344.

    Atkinson R, Winer AM, & Pitts JN Jr (1986) Estimation of night-time
    N2O5 concentrations from ambient NO2 and NO3 radical
    concentrations and the role of N2O5 in night-time chemistry. Atmos
    Environ, 20: 331-339.

    Atkinson CJ, Wookey P, & Mansfield TA (1991) Atmospheric pollution and
    the sensitivity of stomata on barley leaves to abscisic acid and
    carbon dioxide. New Phytol, 117: 159-166.

    Atlas E (1988) Evidence for > C3 alkyl nitrates in rural and
    remote atmosphere. Nature (Lond), 331: 426-428.

    Avol EL, Linn WS, & Venet TG (1983) A comparison of ambient oxidant
    effects to ozone dose-response relationships. Downey, California,
    Rancho Los Amigos Hospital, Environmental Health Service (Research and
    Development Series No. 83-RD-29).

    Avol EL, Linn WS, Shamoo DA, Valencia LM, Anzar UT, Venet TG, &
    Hackney JD (1985a) Respiratory effects of photochemical oxidant air
    pollution in exercising adolescents. Am Rev Respir Dis, 132: 619-622.

    Avol EL, Linn WS, Venet TG, & Hackney JD (1985b) Short-term health
    effects of ambient air pollution in adolescents: year 2 - health
    effects assessment in children (Final report). Downey, California,
    Rancho Los Amigos Medical Centre, Environmental Health Service
    (Research and Development Series No. 85-RD-43).

    Avol EL, Linn WS, Shamoo DA, Spier CE, Valencia LM, Venet TG, Trim SC,
    & Hackney JD (1987) Short-term respiratory effects of photochemical
    oxidant exposure in exercising children. J Am Pollut Control Assoc,
    37: 158-162.

    Avol EL, Linn WS, Peng RC, Valencia G, Little D, & Hackney JD (1988)
    Laboratory study of asthmatic volunteers exposed to nitrogen dioxide
    and to ambient air pollution. Am Ind Hyg Assoc J, 49: 143-149.

    Avol EL, Linn WS, Peng RC, Whynot JD, Shamoo DA, Little DE, Smith MN,
    & Hackney JD (1989) Experimental exposures of young asthmatic
    volunteers to 0.3 ppm nitrogen dioxide and to ambient air pollution.
    Toxicol Ind Health, 5: 1025-1034.

    Ayaz KL & Csallany AS (1978) Long-term NO2 exposure of mice in the
    presence and absence of vitamin E. II. Effect of glutathione
    peroxidase. Arch Environ Health, 33: 292-296.

    Azoulay E, Soler P, Blayo MC, & Basset F (1977) Nitric oxide effects
    on lung structure and blood oxygen affinity in rats. Bull Eur
    Physiopathol Respir, 13: 629-644.

    Azoulay E, Soler P, & Blayo MC (1978) The absence of lung damage in
    rats after chronic exposure to 2 ppm nitrogen dioxide. Bull Eur
    Physiopathol Respir, 14: 311-325.

    Azoulay E, Soler P, Moreau J, & Blayo MC (1980) Effects of
    low-concentration NOxSO2 gas mixtures on lung structure and
    blood-oxygen affinity in rats. J Environ Pathol Toxicol, 4: 399-409.

    Azoulay E, Bouley G, & Blayo MC (1981) Effects of nitric oxide on
    resistance to bacterial infection in mice. J Toxicol Environ Health,
    7: 873-882.

    Azoulay-Dupuis E, Torres M, Soler P, & Moreau J (1983) Pulmonary NO2
    toxicity in neonate and adult guinea pigs and rats. Environ Res,
    30: 322-339.

    Bakelaar RG & Odum EP (1978) Community and population level responses
    to fertilization in an old-field ecosystem. Ecology, 59: 660-665.

    Balabaeva L & Tabacova S (1985) [Lipid peroxidation in two generations
    of female albino rats exposed to nitrogen dioxide.] Hig Zdraveopaz,
    2: 41-46 (in Bulgarian).

    Balchum OJ, Buckley RD, Sherwin R, & Gardner M (1965) Nitrogen dioxide
    inhalation and lung antibodies. Arch Environ Health, 10: 274-277.

    Baldocchi D (1988) A multi-layer model for estimating sulfur dioxide
    deposition to a deciduous oak forest canopy. Atmos Environ,
    22: 869-884.

    Baldocchi DD, Hicks BB, & Camara P (1987) A canopy stomatal resistance
    model for gaseous deposition to vegetated surfaces. Atmos Environ,
    21: 91-101.

    Balsberg Pählsson AM (1989a) Mineral nutrition, carbohydrates and
    phenolic compounds in leaves of beech  (Fagus sylvatica) in southern
    Sweden as related to environmental factors. Tree Physiol, 5: 485-495.

    Balsberg Pählsson AM (1989b) Influence of nitrogen fertilization on
    minerals, carbohydrates, aminoacids and phenolic compounds in beech
     (Fagus sylvatica L.) leaves. Tree Physiol, 10: 93-100.

    Balsberg Pählsson AM (1992) Influence of nitrogen fertilization on
    minerals, carbohydrates, aminoacids and phenolic compounds in beech
    ( Fagus sylvatica L.) leaves. Tree Physiol, 10: 93-100.

    Bandow H, Okuda M, & Akimoto H (1980) Mechanism of the gas-phase
    reactions of C3H6 and NO3 radicals. J Phys Chem, 84: 3604-3608.

    Barinaga M (1991) Is nitric oxide the "retrograde messenger"? Science,
    254(5036): 1296-1297.

    Barnes PJ (1993) Nitric oxide and airways. Eur Respir J, 6: 163-165.

    Barrie LA & Sirois A (1986) Wet and dry deposition of sulphates and
    nitrates in eastern Canada: 1979-1982. Water Air Soil Pollut,
    30: 303-310.

    Barsdate RJ & Alexander V (1975) The nitrogen balance of arctic
    tundra: pathways, rates, and environmental implications. J Environ
    Qual, 4: 111-117.

    Bauer MA, Utell MJ, Morrow PE, Speers DM, & Gibb FR (1986) Inhalation
    of 0.30 ppm nitrogen dioxide potentiates exercise-induced bronchospasm
    in asthmatics. Am Rev Respir Dis, 134: 1203-1208.

    Baulch DL, Cox RA, Crutzen PJ, Hampson RF, Kerr JA, Troe J, & Watson
    RT (1982) Evaluated kinetic and photochemical data for atmospheric
    chemistry. J Phys Chem Ref (Data), 11(suppl 1): 327-496.

    Becker S, Roger LJ, Devlin RB, & Koren HS (1991) Lymphocyte
    infiltration and increased macrophage phagocytosis in the lungs of
    HNO3-exposed humans. FASEB J, 5: A890.

    Becker S, Roger LJ, Devlin RB, & Koren HS (1992) Increased
    phagocytosis and antiviral activity of alveolar macrophages from
    humans exposed to nitric acid. Am Rev Respir Dis, 145: A429.

    Becker S, Deolen R, Horstman D, Gerrity T, Madden M, Biscardi F, &
    Koren H (1993) Evidence for mild inflammation and change in alveolar
    macrophage function in humans exposed to 2PPM NO2 In: Jaakkola JJK,
    Ilmarinen R, & Seppänen O ed. Indoor air '93 -Proceedings of the 6th
    International Conference on Indoor Air Quality and Climate, Helsinki,
    July 1993. Volume 1: Health Effects, pp 471-476.

    Begon M, Harper JL, & Townsend CR (1990) Ecology. Oxford, Boston,
    London, Blackwell Scientific Publications.

    Beil M & Ulmer WT (1976) [Effect of NO2 in workroom concentrations on
    respiratory mechanics and bronchial susceptibility to acetylcholine in
    normal persons.] Int Arch Occup Environ Health, 38: 31-44 (in German).

    Bell SA, Ashenden TW & Rafael CR (1992) Effects of rural roadside
    levels of nitrogen dioxide on  Poytrichum formosum.  Environ Pollut,
    76: 11-14.

    Bender J, Weigel HJ, & Jäger HJ (1991) Response of nitrogen metabolism
    in bean  (Phaseolus vulgaris) after exposure to ozone and nitrogen
    dioxide, alone and in sequence. New Phytol, 119: 261-267.

    Benedict HM & Breen WH (1955) The use of weeds as a means of
    evaluating vegetation damage caused by air pollution. In: Proceedings
    of the 3rd National Air Pollution Symposium, Los Angeles, pp 177-190.

    Benemansky VV, Prusakov VM, & Leshenko ME (1981) [Blastogenic effect
    of treatment with low concentrations of nitrosodimethylamine,
    dimethylamine and nitrogen dioxide.] Vopr Onkol, 27: 56-62
    (in Russian).

    Benner CL, Eatough DJ, Eatough NL, & Bhardwaja P (1987) Evaluation of
    an annular denuder method for the collection of atmospheric
    nitrogenous species in the southwest desert. Presented at the 80th
    Annual Meeting of the Air Pollution Control Association, New York.
    Pittsburgh, Pennsylvania, Air Pollution Control Association (Paper
    No. 87-63.6).

    Bennet JH, Lee EH, & Heggeststad HE (1990) Inhibition of
    photosynthesis and leaf conductance interactions induced by SO2,
    NO2 and SO2 + NO2. Atmos Environ, 24A: 557-562.

    Benoit FM (1983) Detection of nitrogen and sulfur dioxides in the
    atmosphere by atmospheric pressure ionization mass spectrometry. Anal
    Chem, 55: 2097-2099.

    Berdowski JJM (1987) The catastrophic death of  Calluna vulgaris in
    Dutch heathlands. Utrecht, The Netherlands, University of Utrecht
    (Ph.D. Thesis).

    Berdowski JJM (1993) The effect of external stress and disturbance
    factors on  Calluna-dominated heathland vegetation. In: Aerts R &
    Heil GW ed. Heathlands: Patterns and processes in a changing
    environment. Dordrecht, The Netherlands, Kluwer Academic Publishers,
    pp 85-124.

    Berendse F (1985) The effect of grazing on the outcome of competition
    between plant species with different nutrient requirements. Oikos,
    44: 35-39.

    Berendse F (1988) [The nutrient balance of dry sand terrestrial
    vegetation in connection with eutrophication through air: I. A
    stimulation model as an aid in the control of wet heathlands.]
    Wageningen, The Netherlands, Research Institute for Agrobiology and
    Soil Fertility (in Dutch).

    Berendse F (1990) Organic matter accumulation and nitrogen
    mineralization during secondary succession in heathland ecosystems.
    J Ecol, 78: 413-427.

    Berendse F & Aerts R (1984) Competition between  Erica tetralix L.
    and  Molinia caerulea (L.) Moench.  as affected by the availability
    of nutrients. Acta Oecol/Oecol Plant, 5: 3-14.

    Berendse F, Beltman B, Bobbink R, Kwant, R & Schmitz MB (1987) Primary
    production and nutrient availability in wet heathland ecosystems. Acta
    Oecol/Oecol Plant, 8: 265-276.

    Berglund M, Boström CE, Byhn G, Ewetz L, Gustaffson L, Moldens P,
    Pershagen G, & Victorin K (1993) Health risk evaluation of nitrogen
    oxides. Scan J Work Environ Health, 19(suppl 2): 1-72.

    Berglund M, Bräbäck L, Bylin G, Jonson JO, & Vahter, M (1994) Personal
    NO2 exposure monitoring shows high exposure among ice-skating
    schoolchildren. Arch Environ Health, 49(1): 17.

    Berkey CS, Ware JH, Dockery DW, Ferris BG Jr, & Speizer FE (1986) Air
    pollution and pulmonary function growth in preadolescent children. Am
    J Epidemiol, 123: 250-260.

    Berwick M (1987) Lower respiratory symptoms in children associated
    with nitrogen dioxide exposure [dissertation]. Ann Arbor, Michigan,
    University Microfilms (Publication No. 87-29.173).

    Berwick M, Zagraniski RT, Leaderer BP, & Stolwijk JAJ (1984)
    Respiratory illness in children exposed to unvented combustion
    sources. In: Berglund B, Lindvall T, & Sundell J ed. Indoor air
    '84 - Proceedings of the 3rd International Conference on Indoor Air
    Quality and Climate. Stockholm, Swedish Council for Building Research,
    vol 2, pp 255-260.

    Berwick M, Leaderer BP, Stolwijk JAJ, & Zagraniski RT (1987)
    Association between nitrogen dioxide levels and lower respiratory
    symptoms in children exposed to unvented combustion sources. In:
    Seifert B, Esdorn H, Fischer M, Rueden H, & Wegner J ed. Indoor air
    '87 - Proceedings of the 4th International Conference on Indoor Air
    Quality and Climate. Berlin, Institute for Water, Soil and Air
    Hygiene, vol 2, pp 298-303.

    Berwick M, Leaderer BP, Stolwijk JA, & Zagraniski RT (1989) Lower
    respiratory symptoms in children exposed to nitrogen dioxide from
    unvented combustion sources. Environ Int, 15: 369-373.

    Biermann HW, Tuazon EC, Winer AM, Wallington TJ, & Pitts JN Jr (1988)
    Simultaneous absolute measurements of gaseous nitrogen species in
    urban ambient air by long path length infrared and ultraviolet-visible
    spectroscopy. Atmos Environ, 22: 1545-1554.

    Billick I, Johnson D, Moschandreas D, & Relwani S (1984) An
    investigation of operational factors that influence emission rates
    from gas appliances. In: Berglund B, Lindvall T, & Sundell J ed.
    Indoor air '84 - Proceedings of the 3rd International Conference on
    Indoor Air Quality and Climate. Stockholm, Swedish Council for
    Building Research, vol 4, pp 181-187.

    Billick IH, Ozkaynak H, Butler DA, & Spengler JD (1991) Predicting
    personal exposures to NO2 for population-based exposure and risk
    evaluations. Presented at the 84th Annual Meeting of the Air and Waste
    Management Association. Pittsburgh, Pennsylvania, Air and Waste
    Management Association (Paper No. 91-172.9).

    Billings WD (1978) Plants and the ecosystem, 3rd ed. Belmont,
    California, Wadsworth Publishing Company, Inc., pp 1-62, 83-108.

    Bils RF (1976) The connective tissues and alveolar walls in the lungs
    of normal and oxidant-exposed squirrel monkeys. J Cell Biol, 70: 318.

    Birge RT (1932) The calculation of errors by the method of least
    squares. Phys Rev, 30: 207-227.

    Bjorkman E (1942) [On the conditions of mycorrhiza formation in pine
    and spruce.] Symb Bot Ups, 6: 1-190 (in German).

    Black FM (1989) Motor vehicles as sources of compounds important to
    tropospheric and stratospheric ozone. In: Schneider T, Lee SD, Wolters
    GJR, & Grant LD ed. Atmospheric ozone research and its policy
    implications: Proceedings of the 3rd US-Dutch International Symposium,
    Nijmegen, The Netherlands, 9-13 May 1988. Amsterdam, Oxford, New York,
    Elsevier Science Publishers, pp 85-109.

    Blair WH, Henry MC, & Ehrlich R (1969) Chronic toxicity of nitrogen
    dioxide: II. Effect on histopathology of lung tissue. Arch Environ
    Health, 18: 186-192.

    Blank ML, Dalbey W, Nettesheim P, Price J, Creasia D, & Snyder F
    (1978) Sequential changes in phospholipid composition and synthesis in
    lungs exposed to nitrogen dioxide. Am Rev Respir Dis, 117: 273-280.

    Blankwaardt HFH (1977) [The occurrence of the heath beetle ( Lochmaea
    suturalis Thomson) in the Netherlands since 1915.] Entomol Ber,
    37: 34-40 (in Dutch).

    Bobbink R (1991) Effects of nutrient enrichment in Dutch chalk
    grassland. J Appl Ecol, 28: 28-41.

    Bobbink R & Heil GW (1993) Atmospheric deposition of sulphur and
    nitrogen in heathland ecosystems. In:  Aerts R & Heil GW ed.
    Heathland: Patterns and processes in a changing environment.
    Dordrecht, The Netherlands, Kluwer Academic Publishers, pp 25-50.

    Bobbink R & Willems JH (1987) Increasing dominance of  Brachypodium
    pinnatum (L.) Beauv. in chalk grasslands: a threat to a species-rich
    ecosystem. Biol Conserv, 40: 301-314.

    Bobbink R & Willems JH (1991) Impact of different cutting regimes on
    the performance of  Brachypodium pinnatum in Dutch chalk grassland.
    Biol Conserv, 56: 1-21.

    Bobbink R, Bik L, & Willems JH (1988) Effects of nitrogen
    fertilization on vegetation structure and dominance of  Brachypodium
    pinnatum (L). Beauv. in chalk grassland. Acta Bot Neerl, 37: 231-242.
    Bobbink R, den Dubbelden KC, & Willems JH (1989) Seasonal dynamics
    of phytomass and nutrients in chalk grassland. Oikos, 55: 216-224.

    Bobbink R, Boxman D, Fremstad E, Heil G, Houdijk A, & Roelofs J
    (1992a) Critical load for nitrogen eutrophication of terrestrial and
    wetland ecosystems based upon changes in vegetation and fauna. In:
    Grennfelt P & Thörnelöf E ed. Critical loads for nitrogen: Proceedings
    of a Workshop, Lökeberg, Sweden, April 1992. Copenhagen, Denmark,
    Nordic Council of Ministers, pp 111-159 (Report No. 41).

    Bobbink R, Heil GW, & Raessen MBAG (1992b) Atmospheric deposition and
    canopy exchange processes in heathland ecosystems. Environ Pollut,
    75: 29-37.

    Boettger A, Ehhalt DH, & Gravenhorst G (1978) [Atmospheric cycles of
    nitrogen oxides and ammonia.] Jülich, Germany, Nuclear Research
    Facility Jülich Ltd, Institute of Chemistry, Institute 3: Atmospheric
    Chemistry (Report No. JUEL-1558) (in German).

    Bollinger MJ, Sievers RE, Fahey DW, & Fehsenfeld FC (1983) Conversion
    of nitrogen dioxide, nitric acid, and  n-propyl nitrate to nitric
    oxide by gold-catalyzed reduction with carbon monoxide. Anal Chem,
    55: 1980-1986.

    Borrazzo JE, Osborn JF, Fortmann RC, Keefer RL, & Davidson CI (1987a)
    Modelling and monitoring of CO, NO and NO2 in a modern townhouse.
    Atmos Environ, 21: 299-311.

    Borrazzo JE, Peters C, Peck S, & Davidson CI (1987b) Determination of
    NO2 loss rates from concentration measurements in an occupied urban
    residence. In: Seifert B, Esdorn H, Fischer M, Rueden H, & Wegner J
    ed. Indoor air '87 - Proceedings of the 4th International Conference
    on Indoor Air Quality and Climate. Berlin, Institute for Water, Soil
    and Air Hygiene, vol 1, pp 321-325.

    Borucki WJ & Chameides WL (1984) Lightning: estimates of the rates of
    energy dissipation and nitrogen fixation. Rev Geophys Space Phys,
    22: 363-372.

    Boström C (1993) Nitrogen oxides in ambient air: properties, sources
    and concentrations. Scand J Work Environ Health, 19(suppl 2): 9-13.

    Boumans LJM (1994) [Nitrate in the upper groundwater from regions of
    sandy soil in The Netherlands.] Bilthoven, The Netherlands, National
    Institute of Public Health and Environmental Protection (Report No.
    712300002) (in Dutch).

    Boumans LJM & Beltman W (1991) [Quality of the upper phreatic
    groundwater from forest and heathland in sandy regions of The
    Netherlands.] Bilthoven, The Netherlands, National Institute of Public
    Health and Environmental Protection (Report No. 724901001) (in Dutch).

    Boushey HA Jr, Rubinstein I, Bigby BG, Stites DP, & Locksley RM (1988)
    Studies on air pollution: effects of nitrogen dioxide on airway
    caliber and reactivity in asthmatic subjects; effects of nitrogen
    dioxide on lung lymphocytes and macrophage products in healthy
    subjects; nasal and bronchial effects of sulfur dioxide in asthmatic
    subjects. Sacramento, California, California Air Resources Board
    (Report No. ARB/R-89/384).

    Bowden WB (1986) Gaseous nitrogen emissions from undisturbed
    terrestrial ecosystems: an assessment of their impacts on local and
    global nitrogen budgets. Biogeochemistry, 2: 249-279.

    Bowden WB (1987) The biogeochemistry of nitrogen in freshwater
    wetlands. Biogeochemistry, 4: 313-348.

    Boxman D, Van Dijk H, Houdijk A, & Roelofs J (1988) Critical loads for
    nitrogen, with special emphasis on ammonium. In: Nilsson J & Grennfelt
    P ed. Critical loads for sulphur and nitrogen. Report from a workshop
    held at Skokloster, Sweden, 19-24 March 1988. Copenhagen, Denmark,
    Nordic Council of Ministers, pp 295-323.

    Boxman AW, Krabbendam H, Bellemarkers MJS, & Roelofs JGM (1991)
    Effects of ammonium and aluminium on the development and nutrition of
     Pinus nigra in hydroculture. Environ Pollut, 73: 119-136.

    Boxman AW, Van Dijk HFG, & Roelofs JGM (1994) Soil and vegetation
    responses to decreased atmospheric nitrogen and sulphur inputs into
    Scots pine stand in The Netherlands. For Ecol Manage, 68: 39-45.

    Boxman AW, Van Dam D, Van Dijk HFG, Hogervorst RF, & Koopmans CJ
    (1995) Ecosystem responses to reduced nitrogen and sulphur inputs into
    two coniferous stands in The Netherlands. For Ecol Manage, 71: 7-29.

    Bradshaw JD, Rodgers MO, & Davis DD (1982) Single photon laser-induced
    fluorescence detection of NO and SO2 for atmospheric conditions of
    composition and pressure. Appl Opt, 21: 2493-2500.

    Bradshaw JD, Rodgers MO, Sandholm ST, KeSheng S, & Davis DD (1985)
    A two-photon laser-induced fluorescence field instrument for
    ground-based and airborne measurements of atmospheric NO. J Geophys
    Res (Atmos), 90: 12861-12873.

    Brakenhielm S (1991) [Vegetation monitoring in the PMK reference
    areas. Activity report of 1990.] Stockholm, Swedish Environmental
    Protection Agency (Report No. 3954) (in Swedish, with English
    summary).

    Braman RS, de la Cantera MA, & Han QX (1986) Sequential, selective
    hollow tube preconcentration and chemiluminescence analysis system for
    nitrogen oxide compounds in air. Anal Chem, 58: 1537-1541.

    Branderud TE (1995) The effects of experimental nitrogen addition on
    the mycorrhizal fungus flora in an oligotrophic spruce forest in
    Gardsjön, Sweden. For Ecol Manage, 71: 111-122.

    Braun-Fahrlaender Ch, Ackermann-Liebrich U, Wanner H-U, Rutishauser M,
    Gnehm HE, & Minder ChE (1989) [Effects of air pollutants on the
    respiratory tract in young children.] Schweiz Med Wochenschr,
    119: 1424-1433 (in German).

    Braun-Fahrlaender C, Ackermann-Liebrich U, Schwartz J, Gnehm HP,
    Rutishauser M, & Wanner HU (1992) Air pollution and respiratory
    symptoms in preschool children. Am Rev Respir Dis, 145: 42-47.

    Brauer M & Spengler JD (1994) Nitrogen dioxide exposures inside ice
    skating rinks.  Am J Public Health, 84: 429-433.

    Brauer M, Rasmussen TR, Kjaergaard SK, & Spengler JD (1993) Nitrous
    acid formation in an experimental exposure chamber. Indoor Air,
    3(2): 94-105.

    Brezonik PL (1972) Nitrogen: sources and transformations in natural
    waters. In: Allen HE & Kramer JR ed. Nutrients in natural waters. New
    York, John Wiley & Sons Inc., pp 1-50.

    Brown DJA (1988) Effect of atmospheric N deposition on surface water
    chemistry and the implications for fisheries. Environ Pollut,
    54: 275-284.

    Brown KA, Freer-Smith PH, Howells GD, Skeffington RA, & Wilson RB
    (1988) Rapporteurs' report on discussions at the workshop on excess
    nitrogen deposition, Leatherhead, Surrey, September 1987. Environ
    Pollut, 54: 285-295.

    Bruggink M (1993) Seed bank, germination and establishment of
    ericaceous and gramineous species in heathlands. In: Aerts R & Heil GW
    ed. Heathland: Patterns and processes in a changing environment.
    Dordrecht, The Netherlands, Kluwer Academic Publishers, pp 153-180.

    Bruggink GT, Wolting HG, Dassen JHA, & Bus VM (1988) The effect of
    nitric oxide fumigation at two CO2 concentrations on net
    photosynthesis and stomatal resistance of tomato ( Lycopersicon
     lycopersicum L. cv. Abunda). New Phytol, 110: 185-191.

    Brunekreef B, Fischer P, Houthuijs D, Remijn B, & Boleij J (1987)
    Health effects of indoor NO2 pollution. In: Seifert B, Esdorn H,
    Fischer M, Rueden H, & Wegner J ed. Indoor air '87 - Proceedings of
    the 4th International Conference on Indoor Air Quality and Climate.
    Berlin, Institute for Water, Soil and Air Hygiene, vol 1, pp 304-308.

    Brunsting AMH & Heil GW (1985) The role of nutrients in the
    interaction between a herbivorous beetle and some competing plant
    species in heathlands. Oikos, 44: 23-26.

    Buckley RD & Loosli CG (1969) Effects of nitrogen dioxide inhalation
    on germfree mouse lung. Arch Environ Health, 18: 588-595.

    Buijsman E (1987) Ammonia emission calculation: fiction and reality.
    In: Asman WAH & Diederen HSMA ed. Ammonia and acidification:
    Proceedings of a Symposium of the European Association for the Science
    of Air Pollution (EURASAP). Bilthoven, The Netherlands, European
    Association for the Science of Air Pollution, pp 13-27.

    Buijsman E, Maas JM, & Asman WAH (1987) Anthropogenic NH3 emissions
    in Europe. Atmos Environ, 21: 1009-1022.

    Burns RC & Hardy RWF (1975) Nitrogen fixation in bacteria and higher
    plants. New York, Springer-Verlag. (Molecular Biology, Biochemistry
    and Biophysics, Volume 21).

    Busch RH, Buschbom RL, Cannon WC, Lauhala KE, Miller FJ, Graham JA, &
    Smith LG (1986) Effects of ammonium nitrate aerosol exposure on lung
    structure of normal and elastase-impaired rats and guinea pigs.
    Environ Res, 39: 237-252.

    Bush AF, Glater RA, Dyer J, & Richards G (1962) The effects of engine
    exhaust on the atmosphere when automobiles are equipped with
    afterburners. Berkeley, California, University of California,
    Department of Engineering, pp 1-33 (Report No. 62/63).

    Busey WM, Coate WB, & Badger DW (1974) Histopathologic effects of
    nitrogen dioxide exposure and heat stress in cynomolgus monkeys.
    Toxicol Appl Pharmacol, 29: 130.

    Buttini P, DiPalo V, & Possanzini M (1987) Coupling of denuder and ion
    chromatographic techniques for NO2 trace level determination in air.
    Sci Total Environ, 61: 59-72.

    Bylin G, Lindvall T, Rehn T, & Sundin B (1985) Effects of short-term
    exposure to ambient nitrogen dioxide concentrations on human bronchial
    reactivity and lung function. Eur J Respir Dis, 66: 205-217.

    Bylin G, Hedenstierna G, Lindvall T, & Sundin B (1988) Ambient
    nitrogen dioxide concentrations increase bronchial responsiveness in
    subjects with mild asthma. Eur Respir J, 1: 606-612.

    Cabral-Anderson LJ, Evans MJ, & Freeman G (1977) Effects of NO2 on
    the lungs of aging rats: I. morphology Exp Mol Pathol, 27: 353-365.

    Caceres T, Soto H, Lissi E, & Cisternas R (1983) Indoor house
    pollution: appliance emissions and indoor ambient concentrations.
    Atmos Environ, 17: 1009-1013.

    Calvert JG (1976) Hydrocarbon involvement in photochemical smog
    formation in Los Angeles atmosphere. Environ Sci Technol, 10: 256-262.

    Canner PL (1987) An overview of six clinical trials of aspirin in
    coronary heart disease. Stat Med, 6: 255-263.

    Cape JN (1994) Direct effects of acid rain and cloud on vegetation.
    In: Ashmore MR & Wilson RB, ed. Critical levels of air pollutants for
    Europe: UNECE Workshop on Critical Levels, Egham, United Kingdom,
    23-26 March 1992. New York, Geneva, United Nations, Economic
    Commission for Europe, pp 64-83.

    Cape JN, Fowler D, Eamus D, Murray MB, Sheppard LJ, & Leith ID (1990)
    Effects of acid mist and ozone on frost hardiness of Norway spruce
    seedlings. In: Payer HP, Pffirman T, & Mathi P ed. Environmental
    research with plants in closed chambers. Brussels, Commission of the
    European Communities (Air Pollution Report No. 26).

    Cape JN, Leith ID, Fowler D, Murray MB, Sheppard LJ, Eamus D, & Wilson
    RHF (1991) Sulphate and ammonium in mist impair the frost hardiness of
    red spruce seedlings. New Phytol, 118: 119-126.

    Capron SJM (1989) The effect of oxides of nitrogen and CO2 enrichment
    on photosynthesis and growth of lettuce ( Lactuca sativa L.). New
    Phytol, 111: 473-481.

    Capron TM & Mansfield TA (1976) Inhibition of net photosynthesis in
    tomato in air polluted with NO and NO2.  J Exp Bot, 27: 111-118.

    Capron SJM, Mansfield TA, & Hand DW (1991) Low temperature-enhanced
    inhibition of photosynthesis by oxides of nitrogen in lettuce
     (Lacusa sativa L).  New Phytol, 118: 309-313.

    Capron TM, Hand DW, Mansfield TA, & Wellburn AR (1994) Canopy
    photosynthesis of CO2-enriched lettuce  (Lactuca sativa L). Response
    to short term changes in CO2, temperature and oxides of nitrogen. New
    Phytol, 126: 45-52.

    Capron SJM, Risager M, & Lee JA (1994) Effects of nitrogen supply on
    frost hardiness in  Calluna vulgaris (L). Hull. New Phytol,
    128: 461-468.

    Carroll MA, McFarland M, Ridley BA, & Albritton DL (1985) Ground-based
    nitric oxide measurements at Wallops Island, Virginia. J Geophys Res
    (Atmos), 90: 12853-12860.

    Case GD, Dixon JS, & Schooley JC (1979) Interactions of blood
    metalloproteins with nitrogen oxides and oxidant air pollutants.
    Environ Res, 20: 43-65.

    Cassidy DT & Reid J (1982) Atmospheric pressure monitoring of trace
    gases using tunable diode lasers. Appl Opt, 21: 1185-1190.

    Cavanagh DG & Morris JB (1987) Mucus protection and airway
    peroxidation following nitrogen dioxide exposure in the rat.
    J Toxicol Environ Health, 22: 313-328.

    Central Pollution Control Board (1990) Ambient air quality status of
    some cities/towns in India. New Delhi, Central Pollution Control
    Board, 196 pp (National Ambient Air Quality Monitoring Series,
    Volume 2).

    Chameides WL, Stedman DH, Dickerson RR, Rusch DW, & Cicerone RJ (1977)
    NOx production in lightning. J Atmos Sci, 34: 143-149.

    Chaney S, Blomquist W, DeWitt P, & Muller K (1981) Biochemical changes
    in humans upon exposure to nitrogen dioxide while at rest. Arch
    Environ Health, 36: 53-58.

    Chang TY, Norbeck JM, & Weinstock B (1979) An estimate of the NOx
    removal rate in an urban atmosphere. Environ Sci Technol,
    13: 1534-1537.

    Chang L-Y, Graham JA, Miller FJ, Ospital JJ, & Crapo JD (1986) Effects
    of subchronic inhalation of low concentrations of nitrogen dioxide. I.
    The proximal alveolar region of juvenile and adult rats. Toxicol Appl
    Pharmacol, 83: 46-61.

    Chang L-Y, Mercer RR, Stockstill BL, Miller FJ, Graham JA, Ospital JJ,
    & Crapo JD (1988) Effects of low levels of NO2 on terminal
    bronchiolar cells and its relative toxicity compared to O3. Toxicol
    Appl Pharmacol, 96: 451-464.

    Chapin FS Jr (1974) Human activity patterns in the city. New York,
    Wiley-Interscience Publishers.

    Chapin FS (1980) The mineral nutrition of wild plants. Annu Rev Ecol
    Syst, 11: 233-260.

    Chapin FS III, Bloom AJ, Field CB, & Waring RH (1987) Plant responses
    to multiple environmental factors. Bioscience, 37: 49-57.

    Chapman SB, Hibble J, & Rafael CR (1975) Net aerial production by
     Calluna vulgaris on lowland heath in Britain. J Ecol, 63: 233-258.

    Chen BH, Hong CJ, Pandey MR, & Smith KR (1990) Indoor air pollution in
    developing countries. World Health Stat Q, 43: 127-128.

    Chiodi H & Mohler JG (1985) Effects of exposure of blood hemoglobin to
    nitric oxide. Environ Res, 37: 355-363.

    Chow CK, Dillard CJ, & Tappel AL (1974) Glutathione peroxidase system
    and lysozyme in rats exposed to ozone or nitrogen dioxide. Environ
    Res, 7: 311-319.

    Cicerone RJ, Shetter JD, Stedman DH, Kelly TJ, & Liu SC (1978)
    Atmospheric N2O: measurements to determine its sources, sinks, and
    variations. J Geophys Res (Oceans Atmos), C83: 3042-3050.

    Clark RR (1982) The error-in-variables problem in the logistic
    regression model. Chapel Hill, North Carolina, University of North
    Carolina (Dissertation).

    Clausing P, Mak JK, Spengler JD, & Letz R (1984) Personal NO2
    exposures of high school students. In: Berglund B, Lindvall T, &
    Sundell J ed. Indoor air '84 -  Proceedings of the 3rd International
    Conference on Indoor Air Quality and Climate. Stockholm, Swedish
    Council for Building Research, vol 4, pp 135-139.

    Clausing P, Mak JK, Spengler JD, & Letz R (1986) Personal NO2
    exposures of high school students. Environ Int, 12: 413-417.

    Cochran WG (1937) Problems arising in the analysis of a series of
    similar experiments. J R Stat Soc, 4(suppl): 102-118.

    Coffin DL & Gardner DE (1972) Interaction of biological agents and
    chemical air pollutants. Ann Occup Hyg, 15: 219-234.

    Coffin DL, Gardner DE, Sidorenko GI, & Pinigin MA (1977) Role of time
    as a factor in the toxicity of chemical compounds in intermittent and
    continuous exposures. Part II: Effects of intermittent exposure. J
    Toxicol Environ Health, 3: 821-828.

    Cole JT & Zawacki TS (1985) Emissions from residential gas-fired
    appliances. Final Report (IGT Project No. 30570). Chicago, Illinois,
    Institute of Gas Technology.

    Cole JT, Zawacki TS, Macriss RA, & Moschandreas DJ (1983) Constituent
    source emission rate characterization of three-gas fired domestic
    ranges. Presented at the 76th Annual Meeting of the Air Pollution
    Control Association. Pittsburgh, Pennsylvania, Air Pollution Control
    Association (Paper No. 83-64.3).

    Cote WA, Wade WA III, & Yocom JE (1974) A study of indoor air quality.
    Washington, DC, US Environmental Protection Agency, Office of Research
    and Development (EPA-650/4-74-042).

    Council for Agricultural Science and Technology (1976) Effect of
    increased nitrogen fixation on stratospheric ozone. Ames, Iowa, Iowa
    State University, Department of Agronomy (Report No. 53).

    Cowling DW & Lockyer DR (1981) Increased growth of ryegrass exposed to
    ammonia. Nature (Lond), 292: 337-338.

    Cox RA & Roffey MJ (1977) Thermal decomposition of peroxyacetylnitrate
    in the presence of nitric oxide. Environ Sci Technol, 11: 900-906.

    Crapo JD, Barry BE, Chang L-Y, & Mercer RR (1984) Alterations in lung
    structure caused by inhalation of oxidants. J Toxicol Environ Health,
    13: 301-321.

    Crawley MJ (1983) Herbivory the dynamics of animal/plant interactions.
    Oxford, Boston, Blackwell Scientific Publications.

    Crutzen PJ (1970) The influence of nitrogen oxides on the atmospheric
    ozone content. Q J R Meteorol Soc, 96: 320-325.

    Crutzen PJ (1976) Upper limits on atmospheric ozone reductions
    following increased application of fixed nitrogen to the soil. Geophys
    Res Lett, 3: 169-172.

    Crutzen PJ (1983) Atmospheric interactions - homogeneous gas reactions
    of C, N, and S containing compounds. In: Bolin B & Cook RB ed. The
    major biogeochemical cycles and their interactions. New York, John
    Wiley & Sons, pp 67-114.

    Crutzen PJ (1988) Tropospheric ozone: an overview. In: Isaksen
    ISA ed. Tropospheric ozone - regional and global scale interactions:
    Proceedings of the NATO Advanced Workshop on Regional and Global Ozone
    Interaction and its Environmental Consequences, Lillehammer, Norway,
    June 1987. Dordrecht, The Netherlands, D. Reidel Publishing Company,
    pp 3-32 (NATO Advanced Science Institutes Studies -Series C:
    Mathematical and Physical Sciences, Volume 227).

    Crutzen PJ, Heidt LE, Krasnec JP, Pollock WH, & Seiler W (1979)
    Biomass burning as a source of atmospheric gases CO, H2, N2O, NO,
    CH3Cl and COS. Nature (Lond), 282: 253-256.

    Csallany AS (1975) The effect of nitrogen dioxide on the growth of
    vitamin E deficient, vitamin E supplemented and DPPD supplemented
    mice. Fed Proc Fed Am Soc Exp Biol, 34: 913.

    Curran RD, Ferrari FK, Kispert PH, Stadler J, Stuehr DJ, Simmons RL, &
    Billiar TR (1991) Nitric oxide and nitric oxide-generating compounds
    inhibit hepatocyte protein synthesis. FASEB J, 5: 2085-2092.

    Darley EF, Kettner KA, & Stephens ER (1963) Analysis of peroxyacyl
    nitrates by gas chromatography with electron capture detection. Anal
    Chem, 35: 589-591.

    Dasch JM, Cadle SH, Kennedy KG, & Mulawa PA (1989) Comparison of
    annular denuders and filter packs for atmospheric sampling. Atmos
    Environ, 23: 2775-2782.

    Dassen W, Brunekreef B, Hoek G, Hofschreuder P, Staatsen B, De Groot
    H, Schouten E, & Biersteker K (1986) Decline in children's pulmonary
    function during an air pollution episode. J Air Pollut Control Assoc,
    36: 1223-1227.

    Davis DD (1988) Atmospheric nitrogen oxides, their detection and
    chemistry. In: Third year report to Coordinating Research Council.
    Atlanta, Georgia, Georgia Institute of Technology, pp 1-13.

    Davis DD, Smith G, & Klauber G (1974) Trace gas analysis of power
    plant plumes via aircraft measurement: O3, NOx, and SO2 chemistry.
    Science, 186: 733-736.

    Davis DD, Bradshaw JD, Rodgers MO, Sandholm ST, & KeSheng S (1987)
    Free tropospheric and boundary layer measurements of NO over the
    central and eastern North Pacific Ocean. J Geophys Res (Atmos),
    92: 2049-2070.

    Davis JK, Davidson M, & Schoeb TR (1991) Murine respiratory
    mycoplasmosis: a model to study effects of oxidants. Cambridge,
    Massachusetts, Institute of Health Effects (Research Report No. 47).

    Davison AW, Barnes JD, & Renner CJ (1987) Interactions between air
    pollutants and cold stress. In: Proceedings of the 2nd International
    symposium on Air Pollution and Plant Metabolism, pp 307-328.

    Dawson GA (1977) Atmospheric ammonia from undisturbed land. J Geophys
    Res, 82: 3125-3133.

    De Boer W (1989) Nitrification in Dutch heathland soils. Wageningen,
    The Netherlands, Agricultural University of Wageningen (PhD thesis).

    De Graaf MCC (1994) Ammonium and nitrate in heathland and heathland
    related vegetations:  preferences for nitrogen source and ammonium
    toxicity. Acta Bot Neerl, 43: 393.

    De Hayes DH, Ingle MA, & Waite CE (1989) Nitrogen fertilization
    enhances cold tolerance of red spruce seedings. Can J For Res,
    19: 1037-1043.

    De Kam M, Versteegen CM, Van Den Burg J, & Van der Werf DC (1991)
    Effects of fertilization with ammonium sulphate and potassium sulphate
    on the development of  Sphaeropsis sapinea in Corsican pine. Neth J
    Plant Pathol, 97: 265-274.

    Dekker C, Dales R, Bartlett S, Brunekreef B, & Zwanenburg H (1991)
    Childhood asthma and the indoor environment. Chest, 100: 922-926.

    D'Elia CF, Taft J, Smullen JT, & Macknis J (1982) Nutrient enrichment.
    In: Chesapeake Bay Program technical studies: a synthesis. Annapolis,
    Maryland, US Environmental Protection Agency, pp 36-102 (Publication
    PB84-111202).

    Delwiche CC (1970) The nitrogen cycle. Sci Am, 223: 137-147.

    Den Hartog C (1986) The effects of acid and ammonium deposition on
    aquatic vegetation in the Netherlands. In: Proceedings of the 1st
    International Symposium on Water Milfoil  (Myriophyllum spicatum) and
    Related Haloragaceae Species, Vancouver, Canada, pp 51-58.

    DerSimonian R & Laird N (1986) Meta-analysis in clinical trials.
    Control Clin Trials, 7: 177-188.

    De Santis F, Febo A, Perrino C, Possanzini M, & Liberti A (1985)
    Simultaneous measurements of nitric acid, nitrous acid, hydrogen
    chloride and sulfur dioxide in air by means of high-efficiency annular
    denuders. In: Proceedings of the ECE Workshop on Advancements in Air
    Pollution Monitoring and Procedures. Bonn, Federal Ministry of the
    Interior, pp 68-75.

    De Smidt JT (1979) Origin and destruction of Northwest European heath
    vegetation. In: Wilmanns O & Tüxen R ed. [Appearance and passing of
    plant societies.] Vaduz, J. Cramer, pp 411-435 (in German).

    Detels R, Rokaw SN, Coulson AH, Tashkin DP, Sayre JW, & Massey FJ Jr
    (1979) The UCLA population studies of chronic obstructive respiratory
    disease. I. Methodology and comparison of lung function in areas of
    high and low pollution. Am J Epidemiol, 109: 33-58.

    Detels R, Sayre JW, Coulson AH, Rokaw SN, Massey FJ Jr, Tashkin DP, &
    Wu M-M (1981a) Respiratory effect of long term exposure to two mixes
    of air pollutants in Los Angeles County. Chest, 80(suppl): 27S-29S.

    Detels R, Sayre JW, Coulson AH, Rokaw SN, Massey FJ Jr, Tashkin DP, &
    Wu M-M, (1981b) The UCLA population studies of chronic obstructive
    respiratory disease: IV. respiratory effect of long-term exposure to
    photochemical oxidants, nitrogen dioxide, and sulfates on current and
    never smokers. Am Rev Respir Dis, 124: 673-680.

    Devlin R, Horstman D, Becker S, Gerrity T, Madden M, & Koren H (1992)
    Inflammatory response in humans exposed to 2.0 ppm NO2. Am Rev Respir
    Dis, 145: A456.

    De Vries W (1993) Average critical loads for nitrogen and sulfur and
    its use in acidification abatement policy in the Netherlands. Water
    Air Soil Pollut, 68: 399-434.

    De Vries W (1994) Soil response to acid deposition at different
    regional scales. Wageningen, The Netherlands, Agricultural University
    of Wageningen (PhD Thesis).

    Dewailly E, Allaire S, & Nantel A (1988) Nitrogen dioxide poisoning at
    a skating rink - Quebec. Can Dis Wkly Rep, 14(15): 61-62.

    Dickerson RR (1984) Measurements of reactive nitrogen compounds in the
    free troposphere. Atmos Environ, 18: 2585-2593.

    Dierschke H (1985) [Experimental studies on the composition dynamic of
    calcareous grasslands (Mesobromion) in southern Lower Saxony: I.
    Development of vegetation on permanent lands 1972-1984.] Münst Geogr
    Arb, 20: 9-24 (in German).

    Dignon J (1992) NOx and SOx emissions from fossil fuels: a global
    distribution. Atmos Environ, 26A: 1157-1163.

    Dijkstra L, Houthuijs D, Brunekreef B, Akkerman I, & Boleij JSM (1990)
    Respiratory health effects of the indoor environment in a population
    of Dutch children. Am Rev Respir Dis, 142: 1172-1178.

    Dimmeler S, Lottspeich F, & Brune B (1992) Nitric oxide causes
    ADP-ribosylation and inhibition of glyceraldehyde-3-phosphate
    dehydrogenase. J Biol Chem, 267: 16771-16774.

    Dirkse GM & Martakis GFP (1992)  Effects of fertilizer on Bryophytes
    in Swedish experiments on forest fertilization. Biol Conserv,
    59: 155-161.

    Dirkse GM & Van Dobben HF (1989) [The effect of fertilizer use on the
    composition of ground vegetation of pine woods.] Natura, 9: 208-212
    (in Dutch).

    Dirkse GM, van Dobben HF, & Tamm CO (1991) Effects of fertilization on
    herb and moss layers of a Scots pine stand in Lisselbo (Sweden): a
    multivariate analysis. Leersum, The Netherlands, Research Institute
    for Nature Management, pp 1-40 (Report No. 91/7).

    Dirkse GM (1993) [Forest communities in the Netherlands.] Utrecht, The
    Netherlands, Royal Dutch National Historical Society (Report
    No. WM-208) (in Dutch).

    Dockery DW, Spengler JD, Reed MP, & Ware J (1981) Relationships among
    personal, indoor and outdoor NO2 measurements. Environ Int,
    5: 101-107.

    Dockery DW, Ware JH, Ferris BG Jr, Speizer FE, Cook NR, & Herman SM
    (1982) Change in pulmonary function in children associated with air
    pollution episodes. J Air Pollut Control Assoc, 32: 937-942.

    Dockery DW, Spengler JD, Neas LM, Speizer FE, Ferris BG Jr, Ware JH, &
    Brunekreef B (1989a) An epidemiologic study of respiratory health
    status and indicators of indoor air pollution from combustion sources.
    In: Harper JP ed. Combustion processes and the quality of the indoor
    environment: transactions of an international specialty conference.
    Pittsburgh, Pennsylvania, Air and Waste Management Association,
    pp 262-271 (A&WMA Transactions Series: TR-15).

    Dockery DW, Speizer FE, Stram DO, Ware JH, Spengler JD, & Ferris BG Jr
    (1989b) Effects of inhalable particles on respiratory health of
    children. Am Rev Respir Dis, 139: 587-594.

    Dodge R (1982) The effects of indoor pollution on Arizona children.
    Arch Environ Health, 37(3): 151-155.

    Dohmen GP, McNeill S, & Ell JN (1984) Air pollution increases  Aphis
     fabae pest potential. Nature (Lond), 307: 52-53.

    Dosemeci M, Wacholder S, & Lubin JH (1990) Does nondifferential
    misclassification of exposure always bias a true effect toward the
    null value? Am J Epidemiol, 132: 746-748.

    Douglas WW, Hepper NGG, & Colby TV (1989) Silo-filler's disease. Mayo
    Clin Proc, 64: 291-304.

    Dowell AR, Kilburn KH, & Pratt PC (1971) Short-term exposure to
    nitrogen dioxide: effects on pulmonary ultrastructure, compliance, and
    the surfactant system. Arch Intern Med, 128: 74-80.

    Downing RJ, Hettelingh JP, & De Smet PAM (1993)  Calculation and
    mapping of critical loads in Europe:  status report 1993. Bilthoven,
    The Netherlands, National Institute of Public Health and Environmental
    Protection (RIVM Report No. 259101003).

    Drechsler-Parks DM (1987) Effect of nitrogen dioxide, ozone, and
    peroxyacetyl nitrate on metabolic and pulmonary function. Cambridge,
    Massachusetts, Institute of Health Effects.

    Drechsler-Parks DM, Bedi JF, & Horvath SM (1987) Pulmonary function
    responses of older men and women to NO2. Environ Res, 44: 206-212.

    Driscoll CT, Yatsko CP, & Unangst FJ (1987) Longitudinal and temporal
    trends in the water chemistry of the north branch of the Moose River.
    Biogeochemistry, 3: 37-61.

    Driscoll CT, Schaefer DA, Molot LA, & Dillon PJ (1989) Summary of
    North American data. In: Malanchuk JL & Nilsson J ed. The role of
    nitrogen in the acidification of soils and surface waters. Gotab,
    Sweden, Nordic Council of Ministers, pp 6/1-6/45.

    Driscoll CT, Newton RM, Gubala CP, Baker JP, & Christensen SW (1991)
    Adirondack mountains. In: Charles DF ed. Acidic deposition and aquatic
    ecosystems: regional case studies. New York, Springer-Verlag,
    pp 133-202.

    Drozdz M, Kucharz E, Ludyga K, & Molska-Drozdz T (1976) Studies on the
    effect of long-term exposure to nitrogen dioxide on serum and liver
    proteins level and enzyme activity in guinea pigs. Eur J Toxicol,
    9: 287-293.

    Du YG, Li JQ, & Huang JG (1992) [Indoor NO2 pollution - application
    of a simple NO2 diffusion director.] Environ Monit China, 8: 55-87
    (in Chinese).

    Duan N (1982) Models for human exposure to air pollution. Environ Int,
    8: 305-309.

    Duan N (1991) Stochastic microenvironment models for air pollution
    exposure. J Expo Anal Environ Epidemiol, 1: 235-257.

    Duan XQ, Zhu YZ, & Hou XF (1992) [A study on the characteristics of
    indoor air pollution in Lanzhou city.]  Environ Health, 5: 4-8 (in
    Chinese).

    Dueck TA (1990) Effects of ammonia and sulphur dioxide on the survival
    and growth of Calluna vulgaris (L) Hull seedings. Funct Ecol,
    1990: 109-116.

    Dueck TA & Elderson J (1992)  Influence of NH3 and SO2 on the growth
    and competitive ability of  Arnica montana L. and  Viola canina L.
    New Phytol, 122: 507-514.

    Dueck TA, Dorel FG, Ter Horst R, & Van der Eerden LJM (1990) Effects
    of ammonia, ammonium sulphate, and sulfur dioxide on the frost
    sensitivity of Scots pine ( Pinus sylvestris L.) Water Air Soil
    Pollut, 54: 35-49.

    Dueck TA, Dorel F, Ter Horst R, & Van der Eerden LJ (1991) Effects of
    ammonia and sulphur dioxide on the forts sensitivity of  Pinus
     sylvestri.  Water Air Soil Pollut, 54: 35-49.

    During HJ & Willems JH (1986) The impoverishment of the bryophyte and
    lichen flora of the Dutch chalk grasslands in the thirty years
    1953-1983. Biol Conserv, 36: 143-158.

    Durzan DJ & Steward FC (1983) Nitrogen metabolism. In: Steward FC ed.
    Plant physiology: a treatise. Orlando, Florida, Academic Press, Inc.,
    pp 55-265. (Steward FC, Bidwell RGS ed. Nitrogen metabolism: v. VIII).

    Easter RC, Busness KM, Hales JM, Lee RN, Arbuthnot DA, Miller DF,
    Sverdrup GM, Spicer CW, & Howes JE Jr (1980) Plume conversion rates in
    the SURE region: Volumes 1 and 2. Richland, Washington, Battelle
    Pacific Northwest Laboratories (Report No. EPRI EA-1498).

    Easter RC, Hales JM, Sverdrup GM, & Spicer CW (1983) Plume conversion
    rates in the SURE region: v. 3. Richland, WA, Battelle Pacific
    Northwest Laboratories (Report No. EPRI EA-1498).

    Eddy DM (1989) The confidence profile method: a Bayesian method for
    assessing health technologies. Oper Res, 37: 210-228.

    Eddy DM, Hasselblad V, & Shachter R (1990a) An introduction to a
    Bayesian method for meta-analysis: the confidence profile method. Med
    Decis Making, 10: 15-23.

    Eddy DM, Hasselblad V, & Shachter R (1990b) A Bayesian method for
    synthesizing evidence: the confidence profile method. Int J Technol
    Assess Health Care, 6: 31-55.

    Eddy DM, Hasselblad V, & Shachter R (1992) Meta-analysis by the
    confidence profile method: the statistical synthesis of evidence.
    Boston, Massachusetts, Academic Press, Inc.

    Effler SW, Brooks CM, Auer MT, & Doerr SM (1990) Free ammonia and
    toxicity criteria in a polluted urban lake. Res J Water Pollut Control
    Fed, 62: 771-779.

    Egloff Th (1987) [Does nitrogen (from air) really endanger the last
    (spread) meadows?]  Natur Landsch, 62: 476-478 (in German).

    Ehhalt DH & Drummond JW (1982) The tropospheric cycle of NOx. In:
    Georgii HW & Jaeschke W ed. Chemistry  of the unpolluted and polluted
    troposphere: Proceedings of the NATO Advanced Study Institute, Corfu,
    Greece, September-October 1981. Boston, Massachusetts, D. Reidel
    Publishing Company, vol 96, pp 219-251.

    Ehrlich R (1966) Effect of nitrogen dioxide on resistance to
    respiratory infection. Bacteriol Rev, 30: 604-614.

    Ehrlich R (1975) Interaction between NO2 exposure and respiratory
    infection. In: Scientific seminar on automotive pollutants.
    Washington, DC, US Environmental Protection Agency, Office of
    Research  and Development (EPA-600/9-75-003).

    Ehrlich R & Fenters JD (1973) Influence of nitrogen dioxide on
    experimental influenza in squirrel monkeys. In: Proceedings of the 3rd
    international clean air congress. Düsseldorf, Federal Republic of
    Germany, Society of German Engineers, pp A11-A13.

    Ehrlich R & Henry MC (1968) Chronic toxicity of nitrogen dioxide: I.
    Effect on resistance to bacterial pneumonia. Arch Environ Health,
    17: 860-865.

    Ehrlich R, Silverstein E, Maigetter R, Fenters JD, & Gardner D (1975)
    Immunologic response in vaccinated mice during long-term exposure to
    nitrogen dioxide. Environ Res, 10: 217-223.

    Ehrlich R, Findlay JC, Fenters JD, & Gardner DE (1977) Health effects
    of short-term inhalation of nitrogen dioxide and ozone mixtures.
    Environ Res, 14: 223-231.

    Ehrlich R, Findlay JC, & Gardner DE (1979) Effects of repeated
    exposures to peak concentrations of nitrogen dioxide and ozone on
    resistance to streptococcal pneumonia. J Toxicol Environ Health,
    5: 631-642.

    Ekwo EE, Weinberger MM, Lachenbruch PA, & Huntley WH (1983)
    Relationship of parental smoking and gas cooking to respiratory
    disease in children. Chest, 84: 662-668.

    Ellenberg H (1979) [Indicator values of potted plants of Central
    Europe.] Scripta Geobot, 9: 1-122 (in German).

    Ellenberg H (1985) [Changes in the flora of Central Europe under the
    influence of fertilizer use and emissions.] Schweiz Z Forstwes,
    136: 19-39 (in German).

    Ellenberg H (1987) Floristic changes due to eutrophication. In: Asman
    WAH & Diederen SMA ed. Ammonia and acidification: Proceedings of a
    Symposium of the European Association for the Science of Air Pollution
    (EURASAP). Bilthoven, The Netherlands, European Association for the
    Science of Air Pollution, pp 301-308.

    Ellenberg H (1988a) Vegetation ecology of Central Europe. Cambridge,
    United Kingdom, Cambridge University Press.

    Ellenberg H Jr (1988b) Floristic changes due to nitrogen deposition in
    central Europe. In: Nilsson J & Grennfelt P ed. Critical loads for
    sulphur and nitrogen: Report from a workshop held at Skokloster,
    Sweden, 19-24 March 1988. Copenhagen, Denmark, Nordic Council of
    Ministers, pp 375-383 (Report No. 1988:15).

    Elsayed NM & Mustafa MG (1982) Dietary antioxidants and the
    biochemical response to oxidant inhalation: I. Influence of dietary
    vitamin E on the biochemical effects of nitrogen dioxide exposure in
    rat lung. Toxicol Appl Pharmacol, 66: 319-328.

    Elvebakk A (1985) Higher phytosociological syntax on Svalbard and
    their use in subdivision of the Arctic.  Nord J Bot, 5: 273-284.

    Elwood JW, Sale MJ, Kaufmann PR, & Cada GF (1991) The Southern Blue
    Ridge province. In: Charles DF ed. Acidic deposition and aquatic
    ecosystems: regional case studies. New York, Springer-Verlag.

    Emmett BA, Reynolds B, Stevens PA, Norris DA, Hughes S, Görres J, &
    Lubrecht I (1993) Nitrate leaching from afforested Welsh catchments -
    interactions between stand age and nitrogen deposition. Ambio,
    23: 366-394.

    Enoksson V, Sorensson F, & Graneli W (1990) Nitrogen transformations
    in the Kattegat. Ambio, 19: 159-166.

    Epler GR (1989) Silo-filler's disease: a new perspective. Mayo Clin
    Proc, 64: 368-370.

    Ericsson A, Nordén LG, Näsholm T, & Walheim M (1993)  Mineral nutrient
    imbalances and arginine  concentrations in needles of  Picea abies
     (L.) Karst. from two areas with different levels of airborne
    deposition. Trees, 8: 67-74.

    Eriksson E (1952) Composition of atmospheric precipitation. II.
    Sulfur, chloride, iodine compounds. Bibliography. Tellus, 4: 280-303.

    Evans MJ, Stephens RJ, Cabral LJ, & Freeman G (1972) Cell renewal in
    the lungs of rats exposed to low levels of NO2. Arch Environ Health,
    24: 180-188.

    Evans MJ, Cabral LJ, Stephens RJ, & Freeman G (1973a) Cell division of
    alveolar macrophages in rat lung following exposure to NO2. Am J
    Pathol, 70: 199-208.

    Evans MJ, Cabral LJ, Stephens RJ, & Freeman G (1973b) Renewal of
    alveolar epithelium in the rat following exposure to NO2. Am J
    Pathol, 70: 175-190.

    Evans MJ, Cabral LC, Stephens RJ, & Freeman G (1974) Acute kinetic
    response and renewal of the alveolar epithelium following injury by
    nitrogen dioxide. Chest, 65(suppl): 62S-65S.

    Evans MJ, Cabral LJ, Stephens RJ, & Freeman G (1975) Transformation of
    alveolar Type 2 cells to Type 1 cells following exposure to NO2. Exp
    Mol Pathol, 22: 142-150.

    Evans MJ, Johnson LV, Stephens RJ, & Freeman G (1976) Renewal of the
    terminal bronchiolar epithelium in the rat following exposure to NO2
    or O3. Lab Invest, 35: 246-257.

    Evans MJ, Cabral-Anderson LJ, & Freeman G (1977) Effects of NO2 on
    the lungs of aging rats: II. cell proliferation. Exp Mol Pathol,
    27: 366-376.

    Ewert E (1978) [Damage to vegetation in the surroundings of
    agricultural animal production facilities.] Luft Kältetechn,
    4: 218-420 (in German).

    Ewert E (1979) [Phytotoxicity of ammonia.] Hercynia, 16: 75-80 (in
    German).

    Fahey DW, Hubler G, Parrish DD, Williams EJ, Norton RB, Ridley BA,
    Singh HB, Liu SC, & Fehsenfeld FC (1986) Reactive nitrogen species in
    the troposphere: measurements of NO, NO2, HNO3, particulate nitrate,
    peroxyacetyl nitrate (PAN), O3, and total reactive odd nitrogen
    (NOy) at Niwot Ridge, Colorado.  J Geophys Res (Atmos),
    91:  9781-9793.

    Fahey DW, Murphy DM, Kelly KK, Ko MKW, Proffitt MH, Eubank CS, Ferry
    GV, Loewenstein M, & Chan KR (1989) Measurements of nitric oxide and
    total reactive nitrogen in the Antarctic stratosphere: observations
    and chemical implications. J Geophys Res (Atmos), 94: 16665-16681.

    Falkengren-Grerup U (1986) Soil acidification and vegetation changes
    in deciduous forest in southern Sweden.  Oecologia, 70: 339-347.

    Falkengren-Grerup U & Eriksson H (1990) Changes in soil, vegetation
    and forest yield between 1947 and 1988 in beech and oak sites southern
    Sweden. For Ecol Manage, 38: 37-53.

    Fangmeijer A, Hadwiger-Fangmeier A, Van der Eerden L, & Jäger HJ
    (1994) Effects of atmospheric ammonia on vegetation: A Review. Environ
    Pollut, 86: 43-82.

    Fehsenfeld FC, Dickerson RR, Hubler G, Luke WT, Nunnermacker LJ,
    Williams EJ, Roberts JM, Calvert JG, Curran CM, Delany AC, Eubank CS,
    Fahey DW, Fried A, Gandrud BW, Langford AO, Murphy PC, Norton RB,
    Pickering KE, & Ridley BA (1987) A ground-based intercomparison of NO,
    NOx, and NOy measurement techniques. J Geophys Res (Atmos),
    92: 14710-14722.

    Fehsenfeld FC, Drummond JW, Roychowdhury UK, Galvin PJ, Williams EJ,
    Buhr MP, Parrish DD, Hubler G, Langford AO, Calvert JG, Ridley BA,
    Grahek F, Heikes BG, Kok GL, Shetter JD, Walega JG, Elsworth CM,
    Norton RB, Fahey DW, Murphy PC, Hovermale C, Mohnen VA, Demerjian KL,
    Mackay GI, & Schiff HI (1990) Intercomparison of NO2 measurement
    techniques. J Geophys Res (Atmos), 95:  3579-3597.

    Fennema F (1990) Effects of exposure to atmospheric SO2, NH3 and
    (NH4)2SO4 on survival and extinction of  Arnica montana L. and
     Viola canina L. Arnhem, The Netherlands, Research Institute for
    Forestry and Nature, pp 1-61 (Report No. 90/14).

    Fennema F (1992) SO2 and NH3 deposition as possible causes for the
    extinction of  Arnica montana L. Water Air Soil Pollut, 62: 325-336.

    Fenters JD, Ehrlich R, Findlay J, Spangler J, & Tolkacz V (1971)
    Serologic response in squirrel monkeys exposed to nitrogen dioxide and
    influenza virus. Am Rev Respir Dis, 104: 448-451.

    Fenters JD, Findlay JC, Port CD, Ehrlich R, & Coffin DL (1973) Chronic
    exposure to nitrogen dioxide: immunologic, physiologic, and pathologic
    effects in virus-challenged squirrel monkeys. Arch Environ Health,
    27: 85-89.

    Ferguson P & Lee JA (1980) Some effects of bisulphite and sulphate on
    the growth of  Sphagnum species in the field. Environ. Pollut.,
    A21: 59-71.

    Ferguson P, Robinson RN, Press MC, & Lee JA (1984) Element
    concentrations in five  Sphagnum species in relation to atmospheric
    pollution. J Bryol, 13: 107-114.

    Ferm M (1986) A Na2CO3-coated denuder and filter for determination
    of gaseous HNO3 and particulate NO3- in the atmosphere. Atmos
    Environ, 20: 1193-1201.

    Ferm M & Sjodin A (1985) A sodium carbonate coated denuder for
    determination of nitrous acid in the atmosphere. Atmos Environ,
    19: 979-983.

    Ferrari L, Mc Phail S, & Johnson D (1988) Indoor pollution in
    Australian homes - Results of two winter campaigns. Clean Air,
    22(2): 68-74.

    Ferris BG (1978) Epidemiology standardization project. Am Rev Respir
    Dis, 118: 1-120.

    Ferris BG Jr, Speizer FE, Bishop YMM, & Spengler JD (1979) Effects of
    indoor environment on pulmonary function of children 6-9 years old.
    In: Annual meeting of the American Lung Association and American
    Thoracic Society abstracts. Am Rev Respir Dis, 119: 214.

    Ferris BG Jr, Dockery DW, Ware JH, Speizer FE, & Spiro R III (1983)
    The six-city study: examples of problems in analysis of the data.
    Environ Health Perspect, 52: 115-123.

    Finlayson-Pitts BJ & Pitts JN Jr (1986) Atmospheric chemistry:
    fundamentals and experimental techniques. New York, Wiley
    Interscience, pp 961-1007.

    Finlayson-Pitts BJ, Ezell MJ, & Pitts JN Jr (1989) Formation of
    chemically active chlorine compounds by reactions of atmospheric NaCl
    particles with gaseous N2O5 and ClONO2. Nature (Lond), 337: 241-244.

    Fisher D, Ceraso J, Mathew T, & Oppenheimer M (1988) Polluted coastal
    waters: the role of acid rain. New York, Environmental Defense Fund.

    Fishman J (1985) Ozone in the troposphere. In: Whitten RC & Prasad S
    ed. Ozone in the free atmosphere. New York, Van Nostrand Reinhold
    Company, pp 161-194.

    Fleischer S & Stibe L (1989) Agriculture kills marine fish in the
    1980s. Who is responsible for fish kills in the year 2000? Ambio,
    18: 347-350.

    Fletcher BL & Tappel AL (1973) Protective effects of dietary alpha-
    tocopherol in rats exposed to toxic levels of ozone and nitrogen
    dioxide. Environ Res, 6: 165-175.

    Florey C du V, Melia RJW, Chinn S, Goldstein BD, Brooks AGF, John HH,
    Craighead IB, & Webster X (1979) The relation between respiratory
    illness in primary schoolchildren and the use of gas for cooking: III.
    Nitrogen dioxide, respiratory illness and lung infection. Int J
    Epidemiol, 8: 347-353.

    Florey C du V, Melia R, Goldstein B, Morris R, John HH, & Clark D
    (1982) [The epidemiology of indoor nitrogen dioxide in the U. K.] In:
    Aurand K, Seifert B, & Wegner J ed. [Indoor air quality.] Stuttgart,
    Gustav Fischer Verlag, pp 209-218 (in German).

    Flückiger W & Braun S (1986) Effects of air pollutants on insects
    and host/insect relationship.  In: Proceedings of the CEC Workshop
    organized as part of the Concerted Action: Effects of Air Pollution on
    Terrestrial Ecosystems, Risö, Denmark, March 1986.

    Flückiger W & Braun S (1994) [Forest damage: Report. Studies in
    observation areas of leeches 1984-1993.] Schönenbuch, Switzerland,
    Institute for Applied Plant Biology (in German).

    Focht DD (1974) The effect of temperature, pH, and aeration on the
    production of nitrous oxide and gaseous nitrogenœzero-order kinetic
    model. Soil Sci, 118: 173-179.

    Focht DD & Verstraete W (1977) Biochemical ecology of nitrification
    and denitrification. Adv Microbiol Ecol, 1: 135-214.

    Folinsbee LJ (1988) Human clinical inhalation exposures: experimental
    design, methodology, and physiological responses. In: Gardner DE,
    Crapo JD, & Massaro EJ ed. Toxicology of the lung. New York, Raven
    Press, pp 175-199 (Target Organ Toxicology Series).

    Folinsbee LJ (1992) Does nitrogen dioxide exposure increase airways
    responsiveness? Toxicol Ind Health, 8: 1-11.

    Folinsbee LJ, Horvath SM, Bedi JF, & Delehunt JC (1978) Effect of
    0.62 ppm NO2 on cardiopulmonary function in young male nonsmokers.
    Environ Res, 15: 199-205.

    Folinsbee LJ, Bedi JF, & Horvath SM (1981) Combined effects of ozone
    and nitrogen dioxide on respiratory  function in man. Am Ind Hyg Assoc
    J, 42: 534-541.

    Forrest J, Spandau DJ, Tanner RL, & Newman L (1982) Determination of
    atmospheric nitrate and nitric acid employing a diffusion denuder with
    a filter pack. Atmos Environ, 16: 1473-1485.

    Fortmann RC, Borrazzo JE, & Davidson CI (1984) Characterization of
    parameters influencing indoor pollutant concentrations. In: Berglund
    B, Lindvall T, & Sundell J ed. Indoor air '84 - Proceedings of the 3rd
    International Conference on Indoor Air Quality and Climate. Stockholm,
    Swedish Council for Building Research, vol 4, pp 259-264.

    Fortmann RC, Nagda NL, & Harper JP (1987) Radon mitigation through
    residential pressurization control strategy. In: Seifert B, Esdorn H,
    Fischer M, Rueden H, & Wegner J ed. Indoor air '87 - Proceedings of
    the 4th International Conference on Indoor Air Quality and Climate.
    Berlin, Institute for Water, Soil and Air Hygiene, vol 2, pp 300-304.

    Fowler D, Cape JN, Nicholson IA, Kinnaird JW, & Pateson IS (1980) The
    influence of a polluted atmosphere on outside degradation in Scots
    pine  (Pinus Sylvestris).  In:  Drablos D & Tollan A ed.  Proceedings
    of the International Conference on the Ecological Impact of Acid
    Precipitations. As, Norway, S.N.S.F. Project, pp 156-157.

    Frampton MW, Smeglin AM, Roberts NJ Jr, Finkelstein JN, Morrow PE, &
    Utell MJ (1989a) Nitrogen dioxide exposure  in vivo  and human
    alveolar macrophage inactivation of influenza virus  in vitro.
    Environ Res, 48: 179-192.

    Frampton MW, Finkelstein JN, Roberts NJ Jr, Smeglin AM, Morrow
    PE, & Utell MJ (1989b) Effects of nitrogen dioxide exposure on
    bronchoalveolar lavage proteins in humans. Am J Respir Cell Mol Biol,
    1: 499-505.

    Frampton MW, Morrow PE, Cox C, Gibb FR, Speers DM, & Utell MJ (1991)
    Effects of nitrogen dioxide exposure on pulmonary function and airway
    reactivity in normal humans. Am Rev Respir Dis, 143: 522-527.

    Frampton MW, Voter KZ, Morrow PE, Roberts NJ Jr, Gavras JB, & Utell MJ
    (1992) Effects of NO2 exposure on human host defense. Am Rev Respir
    Dis, 145: A455.

    Frangmeijer AA, Hadwiger-Fangmeier L, Van der Eerden LJ, & Jäger HJ
    (1994)  Effects of atmospheric ammonia on vegetation: A review.
    Environ Pollut, 86: 43-82.

    Fratacci MD, Frostell CG, Chen TY, Wain JC Jr, Robinson DR, & Zapol WM
    (1991) Inhaled nitric oxide: a selective pulmonary vasodilator of
    heparin-protamine vasoconstriction in sheep. Anaesthesiology,
    75: 990-999.

    Freeman G & Haydon GB (1964) Emphysema after low-level exposure to
    NO2. Arch Environ Health, 8: 125-128.

    Freeman BA & Mudd JB (1981) Reaction of ozone with sulfhydryls of
    human erythrocytes.  Arch Biochem Biophys, 208: 212-220.

    Freeman G, Furiosi NJ, & Haydon GB (1966) Effects of continuous
    exposure of 0.8 ppm NO2 on respiration of rats. Arch Environ Health,
    13: 454-456.

    Freeman G, Crane SC, Stephens RJ, & Furiosi NJ (1968a) Environmental
    factors in emphysema and a model system with NO2. Yale J Biol Med,
    40: 566-575.

    Freeman G, Crane SC, Stephens RJ, & Furiosi NJ (1968b) Pathogenesis of
    the nitrogen dioxide-induced lesion in the rat lung: a review and
    presentation of new observations. Am Rev Respir Dis, 98: 429-443.

    Freeman G, Stephens RJ, Crane SC, & Furiosi NJ (1968c) Lesion of the
    lung in rats continuously exposed to two parts per million of nitrogen
    dioxide. Arch Environ Health, 17: 181-192.

    Freeman G, Crane SC, Furiosi NJ, Stephens RJ, Evans MJ, & Moore WD
    (1972) Covert reduction in ventilatory surface in rats during
    prolonged exposure to subacute nitrogen dioxide. Am Rev Respir Dis,
    106: 563-579.

    Freeman G, Juhos LT, Furiosi NJ, Mussenden R, Stephens RJ, & Evans MJ
    (1974) Pathology of pulmonary disease from exposure to interdependent
    ambient gases (nitrogen dioxide and ozone). Arch Environ Health,
    29: 203-210.

    Freer-Smith PH (1984)  The responses of six broadleaved trees during
    long term exposure to SO2 and NO2. New Phytol, 97: 49-61.

    Fremstad E, Aarrestad PA, & Skogen A (1991) [Coastal heathland in West
    Norway and Trondelag. Natural environments and vegetation in danger.]
    NINA Utredning, 29: 1-172 (in Norwegian).

    Frostell C, Fratacci M-D, Wain JC, Jones R, & Zapol WM (1991) Inhaled
    nitric oxide: A selective pulmonary vasodilator reversing hypoxic
    pulmonary vasoconstriction. Circulation, 83(6): 2038-2047.

    Frostell CG, Blomqvist H, Hedenstierna G, Lundberg J, & Zapol WM
    (1993) Inhaled nitric oxide selectively reverses human hyporic
    pulmonary vasoconstriction without causing systemic vasolidation.
    Anaesthesiology, 78: 427-435.

    Fu Y & Blankenhorn EP (1992) Nitric oxide-induced anti-mitogenic
    effects in high and low responder rat strains. J Immunol,
    148: 2217-2222.

    Fugas M (1975) Assessment of total exposure to an air pollutant. In:
    Proceedings of the International Conference on Environmental Sensing
    and Assessment, Las Vegas, Nevada. New York, Institute of Electrical
    and Electronic Engineers, Inc., vol 2, pp 38-45.

    Fujimaki H & Smimizu F (1981) Effects of acute exposure to nitrogen
    dioxide on primary antibody response. Arch Environ Health,
    36: 114-119.

    Fujimaki H, Shimizu F, & Kubota K (1981) Suppression of antibody
    response in mice by acute exposure to nitrogen:  in vitro study.
    Environ Res, 26: 490-496.

    Fujimaki H, Shimizu F, & Kubota K (1982) Effects of subacute exposure
    to NO2 on lymphocytes required for antibody responses. Environ Res,
    29: 280-286.

    Fuller WA (1987) A single explanatory variable. In: Measurement error
    models. New York, John Wiley & Sons, pp 1-9, 13-20.

    Furiosi NJ, Crane SC, & Freeman G (1973) Mixed sodium chloride aerosol
    and nitrogen dioxide in air: Biological effects on monkeys and rats.
    Arch Environ Health, 27: 405-408.

    Gail MH (1985) Adjusting for covariates that have the same
    distribution in exposed and unexposed cohorts. In: Moolgavkar SH &
    Prentice RL ed. Modern statistical methods in chronic disease
    epidemiology, pp 3-18.

    Galbally IE & Roy CR (1978) Loss of fixed nitrogen from soils by
    nitric oxide exhalation. Nature (Lond), 275: 734-735.

    Galbally IE, Roy CR, Elsworth CM, & Rabich HAH (1985) The measurement
    of nitrogen oxide (NO, NO2) exchange over plant/soil surfaces. East
    Melbourne, Australia, Commonwealth Science and Industry Research
    Organization, Division of Atmospheric Research (CSIRO Technical Paper
    No. 8).

    Gallagher CC, Forsberg CA, Pieri RV, Faucher GA, & Calo JM (1985)
    Nitric oxide and nitrogen dioxide content of whole air samples
    obtained at altitudes from 12 to 30 km. J Geophys Res (Atmos),
    90: 7899-7912.

    Galloway JN & Whelpdale DM (1987) WATOX-86 overview and western North
    Atlantic Ocean S and N atmospheric budgets. Global Biogeochem Cycles,
    1: 261-281.

    Galloway JN, Schofield CL, Hendrey GR, Peters NE, & Johannes AH (1980)
    Source of acidity in three lakes acidified during snowmelt. In:
    Drablos D & Tollan A ed. Ecological effects of acid precipitation:
    Proceedings of an International Conference. Mysen, Norway, Johs.
    Grefslie Trykkeri A/S, pp 264-265.

    Galloway JN, Likens GE, Keene WC, & Miller JM (1982) The composition
    of precipitation in remote areas of the world. J Geophys Res (Oceans
    Atmos), 87: 8771-8786.

    Gamble J, Jones W, & Hudak J (1983) An epidemiological study of salt
    miners in diesel and nondiesel mines. Am J Ind Med, 4: 435-458.

    Gamble J, Jones W, & Minshall S (1987) Epidemiological-environmental
    study of diesel bus garage workers: Acute effects of NO2 and
    respirable particulate on the respiratory system. Environ Res,
    42: 201-214.

    Gambrell RP & Patrick WH Jr (1978) Chemical and microbiological
    properties of anaerobic soils and sediments. In: Hook DD & Crawford
    RMM ed. Plant life in anaerobic environments. Ann Arbor, Michigan, Ann
    Arbor Science Publishers Inc., pp 375-423.

    Garber K (1935) [On the physiology of the effects of ammonia gas on
    plants.] Landwirtsch Vers Wes, 122: 227-343 (in German).

    Gardenfors U (1987) Impacts of airborne pollution on terrestrial
    invertebrates with particular reference to molluscs. Solna, Sweden,
    National Environmental Protection Board (Report No. 3362).

    Gardiner TH & Schanker LS (1976) Effect of oxygen toxicity and nitric
    acid-induced lung damage on drug absorption from the rat lung. Res
    Commun Chem Pathol Pharmacol, 15: 107-120.

    Gardner DE (1980) Influence of exposure patterns of nitrogen dioxide
    on susceptibility to infectious respiratory disease. In: Lee SD ed.
    Nitrogen oxides and their effects on health. Ann Arbor, Michigan, Ann
    Arbor Science Publishers Inc., pp 267-288.

    Gardner DE (1982) Toxic response: The significance of local vs.
    systemic effects. Proceedings of the 1982 Summer Toxicology Forum,
    Washington, pp 82-87.

    Gardner DE, Coffin DL, Pinigin MA, & Sidorenko GI (1977a) Role of time
    as a factor in the toxicity of chemical compounds in intermittent and
    continuous exposures. Part I. Effects of continuous exposure. J
    Toxicol Environ Health, 3: 811-820.

    Gardner DE, Miller FJ, Blommer EJ, & Coffin DL (1977b) Relationships
    between nitrogen dioxide concentration, time, and level of effect
    using an animal infectivity model. In: Dimitriades B ed. International
    Conference on Photochemical Oxidant Pollution and its Control:
    Proceedings. Research Triangle Park, North Carolina, US Environmental
    Protection Agency, Environmental Sciences Research Laboratory, vol 2,
    pp 513-525 (EPA-600/3-77-001a).

    Gardner DE, Miller FJ, Blommer EJ, & Coffin DL (1979) Influence of
    exposure mode on the toxicity of NO2. Environ Health Perspect,
    30: 23-29.

    Gardner DE, Miller FJ, Illing JW, Graham JA (1982) Non-respiratory
    function of the lungs: host defenses against infection. In: Schneider
    T & Grant L ed. Air pollution by nitrogen oxides: Proceedings of the
    US-Dutch International Symposium. Amsterdam, Oxford, New York,
    Elsevier Science Publishers, pp 401-415.

    Garg UC & Hassid A (1989) Nitric oxide-generating vasodilators and
    8-bromo-cyclic guanosine monophosphate inhibit mitogenesis and
    proliferation of cultured rat vascular smooth muscle cells. J Clin
    Invest, 83: 1774-1777.

    Garg UC & Hassid A (1991) Nitrogen oxide decreases cytosolic free
    calcium in Balb/c 3T3 fibroblasts by a cyclic GMP-independent
    mechanism. J Biol Chem, 266: 9-12.

    Garland JA & Penkett SA (1976) Absorption of peroxy acetyl nitrate and
    ozone by natural surfaces. Atmos Environ, 10: 1127-1131.

    Garner JHB (1992) Nitrogen oxides, plant metabolism, and forest
    ecosystem responses. In: Proceedings of the Third International
    Symposium on Gaseous Pollutants and Plant Metabolism.

    Garten CT & Hanson (1989) Foliar retention of 15-N nitrate and
    15N-ammonium in Red Maple (Acer rubrum) and white oak (Quercus alba)
    leaves from simulated rain. Environ Exp Bot, 3: 333-342.

    Garthwaite J (1991) Glutamate, nitric oxide and cell-cell signalling
    in the nervous system. Trends Neurosci, 14: 60-67.

    General Environmental Monitoring Station of China (1991) [Environmental
    quality of a number of cities in China in the period of 1986-1990.]
    Beijing, General Environmental Monitoring Station of China (Internal
    Report) (in Chinese).

    Gessel SP, Cole DW, & Steinbrenner EC (1973) Nitrogen balances in
    forest ecosystems of the Pacific Northwest. Soil  Biol Biochem,
    5: 19-34.

    Ghashghaie J & Saugier B (1989) Effects of nitrogen deficiency on leaf
    photosynthetic response of tall fescue to water deficit. Plant Cell
    Environ, 12: 261-271.

    Gimingham CH (1972) Ecology of heathlands. London, Chapman & Hall.

    Gimingham CH & De Smidt JT (1983) Healths as natural and semi-natural
    vegetation. In: Holzner H, Werger MJA, & Ikusima I ed. Man's impact
    upon vegetation. The Hague, Junk, pp 185-189.

    Gimingham CH, Chapman SB, & Webb NR (1979) European heathlands. In:
    Specht RL ed. Ecosystems of the world. Amsterdam, Oxford, New York,
    Elsevier Science Publishers, vol 9A, pp 365-386.

    Giordano AM Jr & Morrow PE (1972) Chronic low-level nitrogen dioxide
    exposure and mucociliary clearance. Arch Environ Health, 25: 443-449.

    Gladen B & Rogan WJ (1979) Misclassification and the design of
    environmental studies. Am J Epidemiol, 109: 607-616.

    Glasgow JE, Pietra GG, Abrams WR, Blank J, Oppenheim DM, & Weinbaum G
    (1987) Neutrophil recruitment and degranulation during induction of
    emphysema in the rat by nitrogen dioxide. Am Rev Respir Dis,
    135: 1129-1136.

    Glass ADM, Siddiqi MY, Ruth TJ, & Rufti TW Jr (1990) Studies of the
    uptake of nitrate in barley. II. Energetics. Plant Physiol,
    93: 1585-1589.

    Glezen WP (1989) Antecedents of chronic and recurrent lung disease:
    Childhood respiratory trouble. Am Rev Respir Dis, 140: 873-874.

    Glime JM (1992) Effects of pollutants on aquatic species. In: Bates JW
    & Farmer AM ed. Bryophytes and lichens in a changing environment.
    Oxford, United Kingdom, Clarendon Press, pp 333-361.

    Goh KM & Haynes RJ (1986) Nitrogen and agronomic practice. In: Haynes
    RJ ed. Mineral nitrogen in the plant-soil system. Orlando, Florida,
    Academic Press, Inc., pp 379-468.

    Goings SAJ, Kulle TJ, Bascom R, Sauder LR, Green DJ, Hebel JR, &
    Clements ML (1989) Effect of nitrogen dioxide exposure on
    susceptibility to influenza A virus infection in healthy adults. Am
    Rev Respir Dis, 139: 1075-1081.

    Gold DR, Tager IB, Weiss ST, Rosteson TD, & Speizer FE (1989) Acute
    lower respiratory illness in childhood as a predictor of lung function
    and chronic respiratory symptoms. Am Rev Respir Dis, 140: 877-884.

    Goldstein E, Eagle MC, & Hoeprich PD (1973) Effect of nitrogen dioxide
    on pulmonary bacterial defense mechanisms. Arch Environ Health,
    26: 202-204.

    Goldstein E, Warshauer D, Lippert W, & Tarkington B (1974) Ozone and
    nitrogen dioxide exposure: Murine pulmonary defense mechanisms. Arch
    Environ Health, 28: 85-90.

    Goldstein BD, Hamburger SJ, Falk GW, & Amoruso MA (1977a) Effect of
    ozone and nitrogen dioxide on the agglutination of rat alveolar
    macrophages by concanavalin A. Life Sci, 21: 1637-1644.

    Goldstein E, Peek NF, Parks NJ, Hines HH, Steffey EP, & Tarkington B
    (1977b) Fate and distribution of inhaled nitrogen dioxide in rhesus
    monkeys. Am Rev Respir Dis, 115: 403-412.

    Goldstein BD, Melia RJW, Chinn S, Florey C du V, Clark D, & John HH
    (1979) The relation between respiratory illness in primary 
    schoolchildren and the use of gas for cooking. II. Factors affecting
    nitrogen dioxide levels in the home. Int J Epidemiol, 8: 339-345.

    Goldstein BD, Melia RJW, & Florey C du V (1981) Indoor nitrogen
    oxides. Bull NY Acad Med, 57: 873-882.

    Goldstein IF, Hartel D, & Andrews LR (1985) Monitoring personal
    exposure to nitrogen dioxide. Presented at the 78th Annual Meeting of
    the Air Pollution Control Association. Pittsburgh, Pennsylvania, Air
    Pollution Control Association (Paper No. 85-85-7).

    Goldstein IF, Andrews LR, & Hartel D (1988) Assessment of human
    exposure to nitrogen dioxide, carbon monoxide and respirable
    particulates in New York inner-city residences. Atmos Environ,
    22: 2127-2139.

    Gooch PC, Luippold HE, Creasia DA, & Brewen JG (1977) Observations on
    mouse chromosomes following nitrogen dioxide inhalation. Mutat Res,
    48: 117-179.

    Goodyear SN & Ormrod DP (1988) Tomato response to concurrent and
    sequential NO2 and O2 exposures. Environ Pollut, 51: 315-326.

    Goren AI, Hellmann S, & Brenner S (1993) Prevalence of respiratory
    symptoms and disease in schoolchildren as related to ETS and other
    combustion products. In: Jaakkola JJK, Ilmarinen R, & Seppänen O ed.
    Indoor air '93 - Proceedings of the 6th International Conference on
    Indoor Air Quality and Climate, Helsinki, July 1993. Volume 1: Health
    Effects, pp 459-464.

    Görsdorf S, Appel KE, Engeholm C, & Obe G (1990) Nitrogen dioxide
    induces DNA single-strand breaks in cultures Chinese hamster cells.
    Carcinogensis, 11: 37-41.

    Goyal SS, Huffaker RC, & Lorenz OA (1982) Inhibitory effects of
    ammoniacal nitrogen on growth of radishplants. II. Investigation on
    the possible causes of ammonium toxicity to radish plants and its
    reversal by nitrate. J Am Soc Hortic Soc, 107: 130-135.

    Grabherr G (1979) Variability and ecology of the alpine dwarf shrub
    community Loiseleurietea- Cetrarietum. Vegetatio, 41: 111-120.

    Graedel TE, Hawkins DT, & Clexton LD (1986) Atmospherical chemical
    compounds: Sources, occurrence, and bioassay. Orlando, Florida,
    Academic Press, Inc.

    Graham JA, Miller FJ, Gardner DE, Ward R, & Menzel DB (1982) Influence
    of ozone and nitrogen dioxide on hepatic microsomal enzymes in mice. J
    Toxicol Environ Health, 9: 849-856.

    Graham JA, Gadner DE, Blommer EJ, House DE, Menache MG, & Miller FJ
    (1987) Influence of exposure patterns of nitrogen dioxide and
    modifications by ozone on susceptibility to bacterial infectious
    diseases in mice. J Toxicol Environ Health, 21: 113-125.

    Graneli E, Wallstrom K, Larsson U, Graneli W, & Elmgren R (1990)
    Nutrient limitation of primary production in the Baltic Sea area.
    Ambio, 19: 142-151.

    Gravenhorst G, Hoefken KD, & Georgii HW (1983a) Acidic input to a
    beech an spruce forest. In: Beilke S & Elshout AJ ed. Acid deposition:
    Proceedings of the CEC Workshop  organized as part of the Concerted
    Action: Physico-chemical behaviour of atmospheric pollutants.
    Dordrecht, The Netherlands, D. Reidel Publishing company, pp 155-171.

    Grayson RR (1956) Silage gas poisoning: Nitrogen dioxide pneumonia, a
    new disease in agricultural workers. Ann Intern Med, 45: 393-408.

    Greenberg SD, Gyorkey F, Jenkins DE, & Gyiokey P (1971) Alveolar
    epithelial cells following exposure to nitric acid: Electron
    microscopic study in rats. Arch Environ Health, 22: 655-662.

    Greene ND & Schneider SL (1978) Effects of NO2 on the response of
    baboon alveolar macrophages to migration inhibitory factor. J Toxicol
    Environ Health, 4: 869-880.

    Greenfelt P & Hultberg H (1986) Effects of nitrogen deposition on the
    acidification of terrestrial and aquatic ecosystems. Water Air Soil
    Pollut, 30: 945-963.

    Greenfelt P & Thörnelöf E (1992) Critical loads for nitrogen: Report
    from a Workshop held at Lökeberg, Sweden, April 1992. Copenhagen,
    Denmark, Nordic Council of Ministers.

    Greenfelt P, Bengston C, & Skärby L (1983a) Deposition and uptake of
    atmospheric nitrogen oxides in a forest  ecosystem. Aquilo Ser Bot,
    19: 208-221.

    Greenfelt P, Bengston C, & Skarby L (1983b) Dry deposition of nitrogen
    dioxide to Scots pine needles. In: Pruppacher HR, Semonin RG, & Slinn
    WGN ed. Precipitation scavenging, dry deposition, and resuspension.
    Volume 2 - Dry deposition and resuspension: Proceedings of the Fourth
    International Conference, Santa Monica, California, November-December
    1982. Amsterdam, Oxford, New York, Elsevier Science Publishers,
    pp 753-762.

    Greenfelt P, Hoem K, Saltbones J, & Schjoldager J (1989) Oxidant data
    collection in OECD-Europe 1985-87 (oxidate): Report on ozone, nitrogen
    dioxide and peroxyacetic nitrate. Lillestrom, Norway, Norwegian
    Institute for Air Research (Report No. NILU-OR-63/89).

    Gregory KL, Malinoski VF, & Sharp CR (1969) Cleveland clinic fire
    survivorship study, 1929-1965. Arch Environ Health, 18: 508-515.

    Gregory RE, Pickrell JA, Hahn FF, & Hobbs CH (1983) Pulmonary effects
    of intermittent subacute exposure to low-level nitrogen dioxide. J
    Toxicol Environ Health, 11: 405-414.

    Gregory GL, Hoell JM Jr, Torres AL, Carroll MA, Ridley BA, Rodgers MO,
    Bradshaw J, Sandholm S, & Davis DD (1990a) An intercomparison of
    airborne nitric oxide measurements: A second opportunity. J Geophys
    Res (Atmos), 95: 10129-10138.

    Gregory GL, Hoell JM Jr, Carroll MA, Ridley BA, Davis DD, Bradshaw J,
    Rodgers MO, Sandholm S, Shiff HI, Hastie DR, Karecki DR, Mackay GI,
    Torres AL, & Fried A (1990b) An intercomparison of airborne nitrogen
    dioxide instruments. J Geophys Res (Atmos), 95: 10103-10127.

    Gregory GL, Hoell JM Jr, Huebert BJ, van Bramer SE, LeBel PJ, Vay SA,
    Marinaro RM, Schiff HI, Hastie DR, Mackay GI, & Karecki DR (1990c) An
    intercomparison of airborne nitric acid measurements. J Geophys Res
    (Atmos), 95: 10089-10102.

    Greven HC (1992) Changes in the moss flora of The Netherlands. Biol
    Conserv, 59: 133-137.

    Griffith DWT & Schuster G (1987) Atmospheric trace gas analysis using
    matrix isolation-Fourier transform infrared spectrometry. J Atmos
    Chem, 5: 59-81.

    Grosjean D & Harrison J (1985) Response of chemiluminescence NOx
    analyzers and ultraviolet ozone analyzers to organic air pollutants.
    Environ Sci Technol, 19: 862-865.

    Gross P, deTreville RTP, Babyak MA, Kaschak M, & Tolker EB (1968)
    Experimental emphysema: Effect of chronic nitrogen dioxide exposure
    and papain on normal and pneumoconiotic lungs. Arch Environ Health,
    16: 51-58.

    Grünhage L, Dämmgen U, Heanel HD, & Jäger HJ (1992) [Vertical flows of
    trace gases in soil-near atmosphere.] In: [Effects of airborne
    substances on grassland ecosystems - Result of a seven-year ecosystem
    research, Part I.] Landbauforsch Völkenrode, 128: 201-245 (in German).

    Guderian R (1988) Critical levels for effects of NOx: Working paper
    for ECE Critical Levels Workshop, Bad Harzburg, 14-18 March 1988. New
    York, Geneva, United Nations, Economic Commission for Europe.

    Gustafsson LE, Leone AM, Persson MG, Widlund NP, & Moncada S (1991)
    Endogenous nitric oxide is present in the exhaled air of rabbits,
    guinea pigs and humans. Biochem Biophys Res Commun, 181: 852-857.

    Haag RW (1974) Nutrient limitations to plan production in two tundra
    communities. Can J Bot, 52: 103-116.

    Hackney JD, Linn WS, Law DC, Karuza SK, Greenberg H, Buckley RD, &
    Pedersen EE (1975a) Experimental studies on human health effects of
    air pollutants: III. Two hour exposure to ozone alone and in
    combination with other pollutant gases. Arch Environ Health,
    30: 385-390.

    Hackney JD, Linn WS, Mohler JG, Pedersen EE, Breisacher P, &
    Russo A (1975b) Experimental studies on human health effects of air
    pollutants: II. Four-hour exposure to ozone alone and in combination
    with other pollutant gases. Arch Environ Health, 30: 379-384.

    Hackney JD, Thiede FC, Linn WS, Pedersen EE, Spier CE, Law DC, &
    Fischer DS (1978) Experimental studies on human health effects of air
    pollutants: IV. Short-term physiological and clinical effects of
    nitrogen dioxide exposure. Arch Environ Health, 33: 176-181.

    Hackney JD, Linn WS, Avol EL, Shamoo DA, Andersn KR, Solmon JC, Little
    DE, & Peng RC (1992) Exposure of older adults with chronic respiratory
    illness to nitrogen dioxide. A combined laboratory and field study. Am
    Rev Respir Dis, 146: 1480-1486.

    Hahn J (1981) Nitrous oxide in the oceans. In: Delwiche CC ed.
    Denitrification, nitrification, and atmospheric nitrous oxide. New
    York, Wiley-Interscience, pp 191-240.

    Hald A (1962) Statistical theory with engineering applications. New
    York, John Wiley & Sons, Inc., pp 243-244.

    Hällgren JE & Häsholm T (1988) Critical load for nitrogen: Effects on
    forest canopies. In: Nilsson J & Greenfelt P ed. Critical loads for
    sulphur and nitrogen. Report from a Workshop held a Skokloster,
    Sweden, 19-24 March 1988. Copenhagen, Denmark, Nordic Council of
    Ministers, pp 323-342 (Report No. 1988:15).

    Hallingbäck T (1991) [Blue-green algae and cyanophilic lichens
    threatened by air pollution and fertilization.] Svensk Bot Tidskr,
    85: 87-104 (in Swedish).

    Halliwell B, Hu ML, Louie S, Duvall TR, Tarkington BK, Motchnik P,
    Cross CE (1992) Interaction of nitrogen dioxide with human plasma:
    Antioxidant depletion and oxidative damage. FEBS Lett, 313(1): 62-66.

    Hansen DA (1989) Measuring trace gases with FM spectroscopy. EPRI J,
    June: 42-43.

    Hansen KF (1991) [Socio-ecological differentiation: Structure and
    dynamics of the Vestna heathlands today.] University of Bergen
    (Thesis) (in Norwegian).

    Hansen J, Lacis A, & Prather M (1989) Greenhouse effect of
    chlorofluorocarbons and other trace gases. J Geophys Res (Atmos),
    94: 16417-16421.

    Hanson PJ, Rott K, Taylor GE Jr, Gunderson CA, Lindberg SE, &
    Ross-Todd BM (1989) NO2 deposition to elements representative of a
    forest landscape. Atmos Environ, 23: 1783-1794.

    Hao WM, Wofsy SC, McElroy MB, Beer JM, & Toqan MA (1987) Sources of
    atmospheric nitrous oxide from combustion. J Geophys Res (Atmos),
    92: 3098-3104.

    Harlos DP, Marbury M, Samet J, & Spengler JD (1987a) Relating indoor
    NO2 levels to infant personal exposures. Atmos Environ, 21: 369-376.

    Harmos DP, Spengler JD, & Billick I (1987b) Continuous nitrogen
    dioxide monitoring during cooking and commuting: Personal and
    stationary exposures. In: Seifert B, Esdorn H, Fischer M, Rueden H, &
    Wegner J ed. Indoor air '87 - Proceedings of the 4th International
    Conference on Indoor Air Quality and Climate. Berlin, Institute for
    Water, Soil and Air Hygiene, vol 2, pp 278-282.

    Harris GW, Carter WPL, Winner AM, Pitts JN Jr, Platt U, & Perner D
    (1982) Observations of nitrous acid in the Los Angeles atmosphere and
    implications for predictions of ozone-precursor relationships. Environ
    Sci Technol, 16: 414-419.

    Hasselblad V, Stead AG, & Creason JP (1980) Multiple probit analysis
    with a nonzero background. Biometrics, 36: 659-663.

    Hasselblad V, Humble CG, Graham MG, & Anderson HS (1981) Indoor
    environmental determinants of lung function in children. Am Rev Respir
    Dis, 123: 479-485.

    Hasselblad V, Eddy DM, & Kotchmar DJ (1992) Synthesis of environmental
    evidence: nitrogen dioxide epidemiology studies. J Air Waste Manage
    Assoc, 42: 662-671.

    Hatch GE, Slade R, Selgrade MK, & Stead AG (1986) Nitrogen oxide
    exposure and lung antioxidants in ascorbic acid-deficient guinea pigs.
    Toxicol Appl Pharmacol, 82: 351-359.

    Hauhs M (1989) Lange-Bramke: An ecosystem study of a forested
    watershed. In: Adiano DC & Salmons W ed. Acidic precipitation. New
    York, Berlin, Springer-Verlag, pp 275-305.

    Hauhs M, Rost-Sieberg K, Raben G, Paces T, & Vigerust B (1989) Summary
    of European data. In: Malanchuk JL & Nilsson J ed. The role of
    nitrogen in the acidification of soils and surface waters. Copenhagen,
    Denmark, Nordic Council of Ministers, pp 5/1-5/37.

    Hayashi Y, Kohno T, & Ohwada H (1987) Morphological effects of
    nitrogen dioxide on the rat lung. Environ Health Perspect,
    73: 135-145.

    Haydon GB, Freeman G, & Furiosi NJ (1965) Covert pathogenesis of NO2
    induced emphysema in the rat. Arch Environ Health, 11: 776-783.

    Haydon GB, Davidson JT, Lillington GA, & Wasserman K (1967) Nitrogen
    dioxide-induced emphysema in rabbits. Am Rev Respir Dis, 95: 797-805.

    Haynes RJ (1986) Uptake and assimilation of mineral nitrogen by
    plants. In: Haynes RJ ed. Mineral nitrogen in the plant-soil system.
    Orlando, Florida, Academic Press, Inc., pp 303-378.

    Hazucha MJ, Ginsberg JF, McDonnell WF, Haak ED Jr, Pimmel RL, House
    DE, & Bromberg PA (1982) Changes in bronchial reactivity of asthmatics
    and normals following exposure to 0.1 ppm NO2. In: Schneider T & Gant
    L ed. Air pollution by nitrogen oxides: Proceedings of the US-Dutch
    International Symposium. Amsterdam, Oxford, New York, Elsevier Science
    Publishers, pp 387-400.

    Hazucha MJ, Ginsberg JF, McDonnell WF, Haak ED Jr, Pimmel RL, Salaam
    SA, House DE, & Bromberg PA (1983) Effects of 0.1 ppm nitrogen dioxide
    on airways of normal and asthmatic subjects. J Appl Pyshio Respir
    Environ Exercise Physiol, 54: 730-739.

    Hazucha MJ, Folinsbee LJ, Seal E, & Bromberg PA (1994) Lung function
    response of healthy women after sequential exposures to NO2 and O3.
    Am J Respir Crit Care Med, 150: 642-647.

    Health Effects Institute, Health Review Committee (1993) Commentary.
    In: Nitrogen dioxide and respiratory illness in children. Cambridge,
    Massachusetts, Institute of Health Effects, pp 51-80 (Research Report
    No. 58)

    Heath RL & Leech Rm (1978( The stimulation of CO2-supported O2
    evolution in intact spinach chloroplasts by ammonium ion. Arch Biochem
    Biophys, 190: 221-226.

    Hecky RE & Kilham P (1988) Nutrient limitation of phytoplankton in
    freshwater and marine environments: A review of recent evidence on the
    effects of enrichment. Limnol Oceanogr, 33: 796 882.

    Hedberg K, Hedberg CW, Iber C, White KE, Osterolm MT, Jones DBW, Flink
    JR, & MacDonald KL (1989) An outbreak of nitrogen dioxide-induced
    respiratory illness among ice hockey players. J Am Med Assoc,
    262: 3014-3017.

    Hedberg K, MacDonald KL, Osterholm M, Hedberg C, & White K (1990)
    Nitrogen dioxide-induced respiratory illness in ice hockey players
    (reply to a letter to the editor). J Am Med Assoc, 263: 3024-3025.

    Hegg DA & Hobbs PV (1979) Some observations of particulate nitrate
    concentrations in coal-fired power plant plumes. Atmos Environ,
    13: 1715-1716.

    Hegg DA, Hobbs PV, Radke LF, & Harrison H (1977) Ozone and nitrogen
    oxides in power plant plumes. In: Dimitriades B ed. International
    Conference on Photochemical Oxidant Pollution and its Control:
    Proceedings. Research Triangle Park, North Carolina, US Environmental
    Protection Agency, Environmental Sciences Research Laboratory, vol 1,
    pp 173-183 (EPA-600/3-77-001a).

    Heij GJ, De Vries W, Posthumus AC, & Mohrem GMJ (1991) Effects of air
    pollution and acid deposition of forest and forest soil. In: Heij GJ &
    Schneider R ed. Acidification research in The Netherlands. Final
    report of the Dutch Priority Programme on Acidification. Amsterdam,
    Oxford, New York, Elsevier Science Publishers, pp 97-137.

    Heikes BG & Thompson AM (1983) Effects of heterogeneous processes on
    NO3, HONO, and HNO3 chemistry in the troposphere. J Geophys Res
    (Oceans Atmos), 88: 10/883-10/895.

    Heil GW & Aerts R (1993) General introduction. In: Aerts R & Heil GW
    ed. Heathland: Patterns and processes in a changing environment.
    Dordrecht, The Netherlands, Kluwer Academic Publishers, pp 1-24.

    Heil GW & Bobbink R (1993a) "Calluna" a simulation model for
    evaluation of impacts of atmospheric nitrogen deposition on dry
    heathlands. Ecol Model, 68: 161-182.

    Heil GW & Bobbink R (1993b) Impact of atmospheric nitrogen deposition
    on dry heathlands: A Stochastic model simulating competition between
     Calluna vulgaris and two grass species. In: Aerts R & Heil GW ed.
    Heathland: Patterns and processes in a changing environment.
    Dordrecht, The Netherlands, Kluwer Academic Publishers, pp 181-200.

    Heil GW & Bruggink M (1987) Competition for nutrients between
     Calluna vulgaris (L.) Hull and  Molinia caerulea (L.) Moench.
    Oecologia, 73: 105-108.

    Heil GW & Diemont WH (1983) Raised nutrient levels change heathland
    into grassland. Vegetatio, 53: 113-120.

    Heil GW, Van Dam D, & Heijne B (1987) Catch of atmospheric deposition
    in relation to vegetation structures of healthlands. In: Asman WAH &
    Diederen HSMA ed. Ammonia and acidification: Proceedings of the
    EURASAP Symposium. Bilthoven, The Netherlands, National Institute of
    Public Health and Environmental Protection, pp 107-123.

    Heil GW, Werger MJA, de Mol W, Van Dam D, & Heijne B (1988) Capture of
    atmospheric ammonium by grassland canopies. Science, 239: 764-765.

    Helas G, Flanz M, & Warneck P (1981) Improved NOx monitor for
    measurements in tropospheric clean air regions. Int J Environ Anal
    Chem, 10: 155-166.

    Helas G, Broll A, Rumpel K-J, & Warneck P (1987) On the origins of
    night-time NO at a rural measurement site. Atmos Environ,
    21: 2285-2295.

    Hemond HP (1983) The nitrogen budget of Thoreau's Bog. Ecology,
    64: 99-109.

    Hemphill CP, Ryan PB, Billick IH, Nagda NL, Koontz MD, & Fortmann RC
    (1987) Estimation of nitrogen dioxide concentrations in homes equipped
    with unvented gas space heaters. In: Seifert B, Esdorn H, Fischer M,
    Rueden H, & Wegner J ed. Indoor air '87: Proceedings of the 4th
    International Conference on Indoor Air Quality and Climate - Volume 1:
    Volatile organic compounds, combustion gases, particles and fibres,
    microbiological agents. Berlin, Germany, Institute for Water, Soil and
    Air Hygiene, pp 420-424.

    Henriksen A (1988) Critical loads of nitrogen to surface water. In:
    Nilsson J & Grennfelt P ed. Critical loads for sulphur and nitrogen.
    Report from a Workshop held at Skokloster, Sweden, 19-24 March 1988.
    Copenhagen, Denmark, Nordic Council of Ministers, pp 385-412 (Report
    No. 1988:15).

    Henriksen A & Brakke DF (1988) Increasing contributions of nitrogen to
    the acidity of surface waters in Norway. Water Air Soil Pollut,
    42: 183-201.

    Henriksen A, Lien L, Traaen TS, Sevaldrud IS, & Brakke DF (1988) Lake
    acidification in Norway: Present and predicted chemical status. Ambio,
    17: 259-266.

    Henry MC, Ehrlich R, & Blair WH (1969) Effect of nitrogen dioxide on
    resistance of squirrel monkeys to  Klebsiella pneumoniae infection.
    Arch Environ Health, 18: 580-587.

    Henry MC, Findlay J, Spangler J, & Ehrlich R (1970) Chronic toxicity
    of NO2 in squirrel monkeys: III. Effect on resistance to bacterial
    and viral infection. Arch Environ Health, 20: 566-570.

    Henry GHR, Freedman B, & Svobota J (1986) Effects of fertilization on
    three tundra plant communities of a polar desert oasis. Can J Bot,
    64: 2502-2507.

    Hering SV, Lawson DR, Allegrini I, Febo A, Perrino C, Possanzini M,
    Sickles JE II, Anlauf KG, Wiebe A, Appel BR, John W, Ondo J, Wall S,
    Braman RS, Sutton R, Cass GR, Solomon PA, Eatough DJ, Eatough NL,
    Ellis EC, Grosjean D, Hicks BB, Womack JD, Horrocks J, Knapp KT,
    Ellestad TG, Paur RJ, Mitchell WJ, Pleasant M, Peake E, MacLean A,
    Pierson WR, Brachaczok W, Schiff HI, Mackay GI, Spicer CW, Stetman DH,
    Winer AM, Biermann HW, & Tuazon EC (1988) The nitric acid shootout:
    Field comparison of measurement methods. Atmos Environ, 22: 1519-1539.

    Hibbs JB, Taintor RR, Vavrin Z, & Rachlin EM (1988) Nitric oxide: A
    cytotoxic activated macrophage effector molecule. Biochem Biophys Res
    Commun, 157: 87-94.

    Hicks BB, Baldocchi DD, Hosker RP Jr, Hutchison BA, Matt DR, McMillen
    RT, & Satterfield LC (1985) On the use of monitored air concentrations
    to infer dry deposition (1985). Silver Spring, Maryland, National
    Oceanic and Atmospheric Administration, Air Resources Laboratory
    (ERL-ARL-141).

    Higashi T, Imasaka T, & Ishibashi N (1983) Thermal lens spectro-
    photometry with argon laser excitation source for nitrogen dioxide
    determination. Anal Chem, 55: 1907-1910.

    Hill AB (1965) The environment and disease: Association or causation?
    Proc R Soc Med, 58: 295-300.

    Hill AC (1971) Vegetation: A sink for atmospheric pollutants. J Air
    Pollut Control Assoc, 21: 341-346.

    Hill AC & Bennet JH (1970) Inhibition of apparent photosynthesis by
    nitrogen oxides. Atmos Environ, 4: 341348.

    Hillam RP, Bice DE, Hahn FF, & Schnizlein CT (1983) Effects of acute
    nitrogen dioxide exposure on cellular immunity after lung
    immunization. Environ Res, 31: 201-211.

    Himmel RL & DeWerth DW (1974) Evaluation of the pollutant emissions
    from gas-fired ranges. Cleveland, Ohio, American Gas Association
    Laboratories (Report No. 1392).

    Hoefken KD & Gravenhorst G (1982) Deposition of atmospheric aerosol
    particles to beech- and spruce forest. In: Georgii H-W & Parinth J ed.
    Deposition of atmospheric pollutants: Proceedings of a Colloquium.
    Dordrecht, The Netherlands, D. Reidel Publishing Company, pp 191-194.

    Hoek G, Brunekreef B, Meijer R, Scholten A, & Boleij J (1984) Indoor
    nitrogen dioxide pollution and respiratory symptoms of schoolchildren.
    Int Arch Occup Environ Health, 55 79-86.

    Hoell JM Jr, Gregory GL, McDougal DS, Caroll MA, McFarland M,
    Ridley BA, Davis DD, Bradshaw J, Rodgers MO, & Torres AL (1985) An
    intercomparison of nitric oxide measurement techniques. J Geophys Res
    (Atmos), 90: 12843-12851.

    Hoell JM Jr, Gregory GL, McDougal DS, Torres AL, Davis DD, Bradshaw J,
    Rodgers MO, Ridley BA, & Carroll MA (1987) Airborne intercomparison
    of nitric oxide measurement techniques. J Geophys Res (Atmos),
    92: 1995-2008.

    Hofmann G, Heinsdorf D, & Kraus HH (1990) [Effects of respirogenic
    nitrogen loads on the productivity and stability of pinewood
    ecosystems.] Beitr Forstwirtschaft, 24: 59-73 (in German).

    Hogg EH, Malmer N, & Wallen B (1994) Microsite and regional variation
    in the potential decay of  Sphagnum magelanicum  in south Swedish
    raised bogs. Ecography, 17: 50-59.

    Högman M, Frostell C, Arnberg H, & Hedenstierna G (1993) Inhalation of
    nitric oxide modulates metacholine-induced bronchoconstriction in the
    rabbit. Eur Respir J, 6: 177-80.

    Holland DM & McElroy FF (1986) Analytical method comparisons by
    estimates of precision and lower detection limit. Environ Sci Technol,
    20: 1157-1161.

    Hollowell CD, Budnitz RJ, & Traynor GW (1977) Combustion-generated
    indoor air pollution. In: Kasuga S, Suzuki N, Yamada T, Kimura G,
    Inagski K, & Onoe K ed. Proceedings of the Fourth International Clean
    Air Congress. Tokyo, Japan, Japanese Union of Air Pollution Prevention
    Associations, pp 684-687.

    Holt PG, Finlay-Jones LM, Keast D, & Papadimitrou JM (1979)
    Immunological function in mice chronically exposed to nitrogen oxides
    (NOx). Environ Res, 19: 154-162.

    Holtan-Hartwig L & Bockman OC (1994) Ammonia exchange between crop and
    air. Norw J Agric Soc, 14(suppl): 1-41.

    Holzworth GC (1967) Mixing depths, wind speeds and air pollution
    potential for selected locations in the United States. J Appl
    Meteorol, 6: 1039-1044.

    Hong CJ (1991) Health aspects of domestic use of biomass fuels and
    coal in China. In: Indoor air pollution from biomass fuels. Geneva,
    World Health Organization, pp 41-77 (WHO/PEP/92.3B).

    Hooftman RN, Kuper CF, & Appelman LM (1988) Comparative sensitivity of
    histo-pathology and specific lung parameters in the detection of lung
    injury. J Appl Toxicol, 8: 59-65.

    Hora FB (1959) Quantitative experiments on toadstool production in
    woods. Trans Br Mycol Soc, 42: 1-14.

    Houdijk ALFM (1993) Atmospheric ammonium deposition and the
    nutritional balance of terrestrial ecosystems. The Netherlands,
    University of Nijmegen (Ph.D. Thesis).

    Houdijk ALFM, Verbeek PJM, Van Dijk HFG, & Roelofs JGM (1993)
    Distribution and decline of endangered herbaceous species in relation
    to the chemical composition of the soil. Plant Soil, 148: 137-143.

    Houlden GS, Mc Neill M, Aminu-Kano M, & Bell JNB (1990) Air pollution
    and agricultural aphid pests: I. Fumigation experiments with SO2 and
    NO2. Environ Pollut, 67: 305-314.

    Houthuijs D, Remijn B, Brunekreef B, & De Koning R (1987) Exposure to
    nitrogen dioxide and tobacco smoke and respiratory health of children.
    In: Seifert B, Esdorn H, Fischer M, Rueden H, & Wegner J ed. Indoor
    air '87: Proceedings of the 4th International Conference on Indoor Air
    Quality and Climate - Volume 1: Volatile organic compounds, combustion
    gases, particles and fibres, microbiological agents. Berlin, Germany,
    Institute for Water, Soil and Air Hygiene, pp 463-467.

    Huebert BJ & Robert CH (1985) The dry deposition of nitric acid to
    grass. J Geophys Res (Atmos), 90: 2085-2090.

    Huebert BJ, Luke WT, Delany AC, & Brost RA (1988) Measurements of
    concentrations and dry surface fluxes of atmospheric nitrates in the
    presence of ammonia. J Geophys Res (Atmos), 93: 7127-7136.

    Hugod C (1979) Effect of exposure to 43 ppm nitric oxide and 3.6 ppm
    nitrogen dioxide on rabbit lung: A light and electron microscopic
    study. Int Arch Occup Environ Health, 42: 159-167.

    Huhta V, Hyvonen R, Koskenniemi A, & Vilkamaa P (1993) Role of
    pH in the effect of fertilization on Nematoda, Oligochaeta and
    microarthropods. In: New trends in soil biology. Ottignies-
    Louvain-la-Neuve, DieuBrichart, pp 61-73.

    Hultberg H (1988) Critical loads for sulphur to lakes and streams. In:
    Nilsson J & Grennfelt P ed. Critical loads for sulphur and nitrogen.
    Report from a Workshop held at Skokloster, Sweden, 19-24 March, 1988.
    Copenhagen, Denmark, Nordic Council of Ministers, pp 185-200 (Report
    No. 1988:15).

    Hutchinson GL, Millington RJ, & Peters DB (1972) Atmospheric ammonia:
    Absorption by plant leaves. Science, 175: 771-772.

    Hyde D, Orthoefer J, Dungworth D, Tyler W, Carter R, & Lum H (1978)
    Morphometric and morphologic evaluation of pulmonary lesions in beagle
    dogs chronically exposed to high ambient levels of air pollutants. Lab
    Invest, 38: 455-469.

    Ichinose T & Sagai M (1982) Studies on biochemical effects of nitrogen
    dioxide: III. Changes of the antioxidative protective systems in rat
    lungs and of lipid peroxidation by chronic exposure. Toxicol Appl
    Pharmacol, 66: 1-8.

    Ichinose T & Sagai M (1989) Biochemical effects of combined gases
    of nitrogen dioxide and ozone: III. Synergistic effects on lipid
    peroxidation and antioxidative protective systems in the lungs of rats
    and guinea pigs. Toxicology, 59: 259-270.

    Ichinose T, Sagai M, & Kubota K (1983) [Changes of lipid peroxidation
    and antioxidative protective systems in lungs of rats exposed acutely,
    subacutely and chronically to nitrogen dioxide.] Taiki Osen Gakkaishi,
    18: 132-146 (in Japanese).

    Ichinose T, Fujii K, & Sagai M (1991) Experimental studies on tumour
    promotion by nitrogen dioxide. Toxicology, 67: 211-25.

    Ide G & Otsu H (1973) [Studies on carcinogenic actions of air
    pollutants.] In: [1972 investigation of air and water pollution on
    human health.] Chiba, Japan, Chiba Prefectural Government, Department
    of Hygiene, pp 99-100 (in Japanese).

    Ignarro LJ (1989) Heme-dependent activation of soluble guanylate
    cyclase by nitric oxide: Regulation of enzyme activity by porphyrins
    and metalloporphyrins. Semin Hematol, 26: 63-76.

    Illing JW, Miller FJ, & Gardner DE (1980) Decreased resistance to
    infection in exercised mice exposed to NO2 and 03. J Toxicol Environ
    Health, 6: 843-851.

    Infante-Rivard C (1993) Childhood asthma and indoor environmental risk
    factors. Am J Epidemiol, 137: 834-844.

    Inoue H, Fukunaga A, & Okubo S (1981) Mutagenic effects of nitrogen
    dioxide combined with methylurea and ethylurea in  Drosophila
     melanogaster . Mutat Res, 88: 281-290.

    Insarova ID, Insarov GE, Brakenhielm S, Hultengren S, Martinsson PO, &
    Semenov SM (1992) Lichen sensitivity and air pollution: A review of
    literature data. Stockholm, Swedish Environmental Protection Agency.

    Iqbal ZM (1984)  In vivo nitrosation of amines in mice by inhaled
    nitrogen dioxide and inhibition of biosynthesis of N-nitrosamines.
    In: O'Neill IK, Von Borstel RC, Miller CT, Long J, & Bartsch H ed.
    N-nitroso compounds: Occurrence, biological effects and relevance to
    human cancer - Proceedings of the VIIIth International Symposium on
    N-Nitroso Compounds. Lyon, International Agency for Research on
    Cancer, pp 291-300 (IARC Scientific Publications No. 57).

    Iqbal ZM, Dahl K, & Epstein SS (1980) Role of nitrogen dioxide in the
    biosynthesis of nitrosamines in mice. Science, 207: 1475-1477.

    Iqbal ZM, Dahl K, & Epstein SS (1981) Biosynthesis of dimethyl-
    nitrosamine in dimethylamine-treated mice after exposure to nitrogen
    dioxide. J Natl Cancer Inst, 67: 137-141.

    Islam MS & Ulmer WT (1979a) [The influence of acute exposure against a
    combination of 5.0 ppm SO2, 5.0 ppm NO2 and 0.1 ppm O3 on the lung
    function in the MAK (lower toxic limit) area (short-time test).] Wiss
    Umwelt, 3: 131-137 (in German).

    Islam MS & Ulmer WT (1979b) [The effects of long-time exposures
    (8 h per day on 4 successive days) to a gas mixture of SO2 + NO2 +
    O3 in the threefold MIC range (maximum emission concentration) on
    lung functions and reactivity of the bronchial system of healthy
    persons.] Wiss Umwelt, 4: 186-190 (in German).

    Isomura K, Chikahira M, Terannishi K, & Hamada K (1984) Induction of
    mutations and chromosome aberrations in lung cells following  in vivo
    exposure of rats to nitrogen oxides. Mutat Res, 136: 119-125.

    Ito K (1971) [Effect of nitrogen dioxide inhalation on influenza virus
    infection in mice]. Nippon Eiseigaku Zasshi, 26: 304-314 (in
    Japanese).

    Ito 0, Okano K, Kuroiwa M, & Totsuka T (1984a) Effects of NO2 and 03
    done or in combination or kidney bean plants: I. Growth, partitioning
    of assimilates and root activities. Tokyo, Japan, National Institute
    of Environmental Studies, pp 1-13 (Research Report No. 6).

    Ito 0, Ohno R, & Totsuka T (1984b) Effects of NO2 and 03 alone or in
    combination on kidney bean plants: II. Amino acid pool size and
    composition. Tokyo, Japan, National Institute of Environmental
    Studies, pp 15-24 (Research Report No. 66).

    Jaakkola JJK, Paunio M, Virtanen M, & Heinonen OP (1991) Low-level air
    pollution and upper respiratory infections in children. Am J Public
    Health, 81: 1060-1063.

    Jackman CH, Frederick JE, & Stolarski RS (1980) Production of odd
    nitrogen in the stratosphere and mesosphere: An intercomparison of
    source strengths. J Geophys Res (Oceans Atmos), 85: 7495-7505.

    Jacob DJ & Wofsy SC (1988) Photochemistry of biogenic emissions over
    the Amazon forest. J Geophys Res (Atmos), 93: 1477-1486.

    Jacobsen M, Smith TA, Hurley JP, Robertson A, & Roscrow R (1988)
    Respiratory infections in coal miners exposed to nitrogen oxides.
    Cambridge, Massachusetts, Institute of Health Effects (Research Report
    No. 18).

    Jacobson JS (1991) The effects of acid precipitation on crops. n:
    Chadwick MJ & Hutton M ed. Acid deposition in Europe. Stockholm,
    Sweden, Environmental Institute (York), pp 81-98.

    Jacobson JS & McManus JM (1985) Pattern of atmospheric sulphur dioxide
    occurrence: An important criterion in vegetation effects assessment.
    Atmos Environ, 19: 501-506.

    Jaeschke R, Oxman AD, & Guyatt GH (1990) To what extent do congestive
    heart failure patients in sinus rhythm benefit from digoxin therapy? A
    systematic overview and meta-analysis. Am J Med, 88: 279-286.

    Jakab GJ (1987) Modulation of pulmonary defense mechanisms by acute
    exposures to nitrogen dioxide. Environ Res, 42: 215-228.

    Jakab GJ (1988) Modulation of pulmonary defense mechanisms against
    viral and bacterial infections by acute exposures to nitrogen dioxide.
    Cambridge, Massachusetts, Institute of Health Effects (Research Report
    No. 20).

    James BR & Riha SJ (1989) Aluminum leaching by mineral acids in forest
    soils: I. Nitric-sulfuric acid differences. Soil Sci Soc Am J,
    53: 259-264.

    Jansen AE & de Vries FW (1988) Qualitative and quantitative research
    on the relation between ectomycorrhiza of  Pseudotsuga menziessi,
    vitality of host and acid rain - Report 1985-1988. Wageningen, The
    Netherlands, Agricultural University, pp 1-73.

    Jeffrey DW & Pigott CD (1973) The response of grasslands on sugar-
    limestone in Teesdale to application of phosphorus and nitrogen.
    J Ecol, 61: 85-92.

    Jeffries DS (1990) Snowpack storage of pollutants, release during
    melting, and impact on receiving waters. In: Norton SA, Lindberg SE, &
    Page AL ed. Acidic precipitation - Volume 4: Soils, aquatic processes,
    and lake acidification. New York, Berlin, Springer-Verlag, pp 107-132.

    Jensen A (1985) The effect of cattle and sheep grazing on salt-marsh
    vegetation at Skallingen, Denmark. Vegetation, 60: 37-48.

    Jentshcke G, Godbold DL, & Huttermann A (1991) Culture of mycorrhizal
    tree seedlings under controlled conditions: Effects of nitrogen and
    aluminium. Physiol Plant, 81: 408-416.

    Joel DD, Chandra P, & Chanana AD (1982) Effects of NO2 on immune
    responses in pulmonary lymph of sheep. J Toxicol Environ Health,
    10: 341-348.

    Johannessen M & Henriksen A (1978) Chemistry of snow meltwater:
    Changes in concentration during melting. Water Resour Res,
    14: 615-619.

    Johansson C (1987) Pine forest: a negligible sink for atmospheric NOx
    in rural Sweden. Tellus Ser, B39: 426-438.

    Johnson DW (1992) Nitrogen retention in forest soils. J Environ Qual,
    21: 1-12.

    Johnson DA, Frampton MW, Winters RS, Morrow PE, & Utell MJ (1990)
    Inhalation of nitrogen dioxide fails to reduce the activity of human
    lung alpha-1-proteinase inhibitor. Am Rev Respir Dis, 142: 758-762.

    Johnson DW, Van Miegroet H, Lindberg SE, Todd DE, & Harrison RB (1991)
    Nutrient cycling in red spruce forests of the Great Smoky Mountains.
    Can J For Res, 21: 769-787.

    Johnston HS (1966) Experimental chemical kinetics. In: Gas phase
    reaction rate theory. New York, The Ronald Press Company, pp 14-34.

    Johnston HS (1971) Reduction of stratospheric ozone by nitrogen oxide
    catalysts from supersonic transport exhaust. Science, 173: 517-522.

    Johnston HS (1982) Odd nitrogen processes. In: Bower FA & Ward RB ed.
    Stratospheric ozone and man. Boca Raton, Florida, CRC Press, Inc., vol
    1, pp 87-140.

    Johnston HS & Selwyn GS (1975) New cross sections for the absorption
    of near ultraviolet radiation by nitrous oxide (N2O). Geophys Res
    Lett, 2: 549-551.

    Jones CL & Seinfeld JH (1983) The oxidation of NO2 to nitrate - day
    and night. Atmos Environ, 17: 2370-2373.

    Jörres R & Magnussen H (1990) Airways response of asthmatics after a
    30 min exposure, at resting ventilation, to 0.25 ppm NO2 or 0.5 ppm
    SO2. Eur Respir J, 3: 132-137.

    Jörres R & Magnussen H (1991) Effect of 0.25 ppm nitrogen dioxide on
    the airway response to methacholine in asymptomatic asthmatic
    patients. Lung, 169: 77-85.

    Jörres R, Nowak D, Grimminger F, Seeger W, Fasske E, Oldigs M, &
    Magnussen H (1992) The effect of 1 ppm nitrogen dioxide on
    bronchoalveolar lavage cells and bronchial biopsy specimens in normal
    and asthmatic subjects. Am Rev Respir Dis, 145(4): A456.

    Joseph DW & Spicer CW (1978) Chemiluminescence method for atmospheric
    monitoring of nitric acid and nitrogen oxides. Anal Chem,
    50: 1400-1403.

    Joshi SB & Bufalini JJ (1978) Halocarbon interferences in
    chemiluminescent measurements of NOX. Environ Sci Technol,
    12: 597-599.

    Kagawa J (1982) Respiratory effects of 2-hr exposure to 1.0 ppm nitric
    oxide in normal subjects. Environ Res, 27: 485-490.

    Kagawa J (1983a) Respiratory effects of two-hour exposure with
    intermittent exercise to ozone, sulfur dioxide and nitrogen dioxide
    alone and in combination in normal subjects. Am Ind Hyg Assoc J,
    44: 14-20.

    Kagawa J (1983b) Effects of ozone and other pollutants on pulmonary
    function in man. In: Lee SD, Mustafa MG, & Mehlman MA ed.
    International Symposium on the Biomedical Effects of Ozone and related
    Photochemical Oxidants. Princeton, New Jersey, Princeton Scientific
    Publishers, Inc., pp 411-422.

    Kagawa J (1986) Experimental studies on human health effects of
    aerosol and gaseous pollutants. In: Lee SD, Schneider T, Grant LD, &
    Verkerk PJ ed. Aerosols: Research, risk assessment and control
    strategies. Proceedings of the Second US-Dutch International
    Symposium, Williamsburg, May 1985. Chelsea, Michigan, Lewis
    Publishers, Inc., pp 683-697.

    Kagawa J (1990) Health effects of exposure to mixtures of nitric oxide
    and nitrogen dioxide in healthy young women. In: Indoor air '90:
    Proceedings of the 5th International Conference on Indoor Air Quality
    and Climate -Volume 1: Human health, comfort and performance. Ottawa,
    Canada, International Conference on Indoor Air Quality and Climate,
    Inc., pp 307-312.

    Kagawa J & Tsuru K (1979) [Respiratory effects of 2-hour exposure to
    ozone and nitrogen dioxide alone and in combination in normal subjects
    performing intermittent exercise.] Nihon Kyobu Shikkan Gakkai Zasshi,
    17: 765-774 (in Japanese).

    Kämäri J, Jeffries DS, Hessen DO, Henriksen A, Posch M & Forsius M
    (1992) Nitrogen critical loads and exceedance for surface waters. In:
    Grennfelt P & Thörnelöf E ed. Critical loads for nitrogen. Copenhagen,
    Denmark, Nordic Council of Ministers, pp 161-200 (Report 1992:41).

    Kamat SR, Godkhindi KD, Shah BW, Mehta AK, Shah VN, Gregrat J, Papewar
    VN, Tyagi NK, Rashid SS, Bhiwankar NT, & Natu RB (1980) Correlation of
    health morbidity to air pollutant levels in Bombay City: Results of
    prospective 3 year survey at one year. J Postgrad Med, 26(1): 45-62.

    Kanner J, Harel S, & Granit R (1992a) Nitric oxide as an antioxidant.
    Arch Biochem Biophys, 289: 130-6.

    Kanner J, Harel S, & Granit R (1992b) Nitric oxide, an inhibitor of
    lipid oxidation by lipoxygenase, cyclooxygenase and hemoglobin. J Natl
    Cancer Inst, 84: 827-831.

    Karlsson PS (1987) Micro-site performance of evergreen and deciduous
    dwarf shrubs in a subarctic heath in relation to nitrogen status.
    Holarctic Ecol, 10: 114-119.

    Katsuki S, Arnold W, Mittal C, & Murad F (1977) Stimulation of
    guanylate cyclase by sodium nitroprusside, nitroglycerin and nitric
    oxide in various tissue preparations and comparison to the effects of
    sodium azide and hydroxylamine. J Cyclic Nucleotide Res, 3: 23-35.

    Kauppi PE, Mielikaeinen K, & Kuusela K (1992) Biomass and carbon
    budget of European forests, 1971 to 1990. Science, 256: 70-74.

    Keeney DR (1973) The nitrogen cycle in sediment-water systems.
    J Environ Qual, 2: 15-29.

    Keller MD, Lanese RR, Mitchell RI, & Cote RW (1979a) Respiratory
    illness in households using gas and electricity for cooking: I. Survey
    of incidence. Environ Res, 19: 495-503.

    Keller MD, Lanese RR, Mitchell Rl, & Cote RW (1979b) Respiratory
    illness in households using gas and electricity for cooking: II.
    Symptoms and objective findings. Environ Res, 19: 504-515.

    Kelly TJ (1986) Modifications of commercial oxides of nitrogen
    detectors for improved response. Upton, New York, US Department of
    Energy, Brookhaven National Laboratory (Report No. BNL-38000).

    Kelly NA (1987) The photochemical formation and fate of nitric acid in
    the metropolitan Detroit area: Ambient, captive-air irradiation and
    modelling results. Atmos Environ, 21: 2163-2177.

    Kelly NA, Wolff GT, & Ferman MA (1984) Sources and sinks of ozone in
    rural areas. Atmos Environ, 18: 1251-1266.

    Kerr HD, Kulle TJ, McIlhany ML, & Swidersky P (1979) Effects of
    nitrogen dioxide on pulmonary function in human subjects: An
    environmental chamber study. Environ Res, 19: 392-404.

    Kershaw KA (1985) Physiological ecology of lichens. Cambridge, United
    Kingdom, Cambridge University Press.

    Kim SU, Koenig JQ, Pierson WE, & Hanley QS (1991) Acute pulmonary
    effects of nitrogen dioxide exposure during exercise in competitive
    athletes. Chest, 99: 815-819.

    Kinnison DE & Wuebbles DJ (1989) Preventing depletion of stratospheric
    ozoneœimplications on future aircraft emissions. Livermore,
    California, US Department of Energy, Lawrence Livermore National
    Laboratory, Atmospheric and Geophysical Sciences Division (Report No.
    UCRL-99926-Rev 1).

    Kinnison DE, Wuebbles DJ, & Johnston HS (1988) A study of the
    sensitivity of stratospheric ozone to hypersonic aircraft emissions.
    Presented at the First International Conference on Hypersonics Flight
    in the 21st Century. Livermore, California, US Department of Energy,
    Lawrence Livermore National Laboratory, Atmospheric and Geophysical
    Sciences Division (Report No. UCRL-98314).

    Kita H & Omichi S (1974) [Effects of air pollutants on cilial movement
    in airway.] Nippon Eiseigaku Zasshi, 29: 100 (in Japanese).

    Kjaergaard S, Ramussen T, Pedersen O, & Braver M (1993) Objective
    effects of nitrous acid gas on eye epithelium in healthy subjects.
    In: Jaakkola JJK, Ilmarinen R, & Seppänen O ed. Indoor air '93 -
    Proceedings of the 6th International Conference on Indoor Air Quality
    and Climate, Helsinki, July 1993. Volume 1: Health Effects,
    pp 483-488.

    Klein RM, Perkinss TD, & Myers HL (1989) Nutrient status and winter
    hardiness of red spruce foliage. Can J For Res, 19: 754-758.

    Kleinerman J & Cowdrey CR (1968) The effects of continuous high level
    nitrogen dioxide on hamsters. Yale J Biol Med, 40: 579-585.

    Kleinerman J, Ip MPC, & Gordon RE (1985) The reaction of the
    respiratory tract to chronic NO2 exposure. In: Scarpelli DG,
    Craighead JE, & Kaufman N ed. The pathologist and the environment.
    Baltimore, Maryland, Williams and Wilkins, pp 200-210 (International
    Academy of Pathology Monograph No. 26).

    Kleinman MT & Mautz WJ (1987) The effects of exercise on respiratory
    tract dosimetry for inhaled gaseous pollutants. Presented at the 80th
    Annual Meeting of the Air Pollution Control Association. Pittsburgh,
    Pennsylvania, Air Pollution Control Association (Paper No. 87-33.5).

    Kleinman MT, Linn WS, Bailey RM, Jones MP, & Hackney JD (1980) Effect
    of ammonium nitrate aerosol on human respiratory function and
    symptoms. Environ Res, 21: 317-326.

    Kleinman MT, Bailey RM, Linn WS, Anderson KR, Whynot JD, Shamoo DA, &
    Hackney JD (1983) Effects of 0.2 ppm nitrogen dioxide on pulmonary
    function and response to bronchoprovocation in asthmatics. J Toxicol
    Environ Health, 12: 815-826.

    Kleinman MT, Bailey RM, Whynot JD, Anderson KR, Linn WS, & Hackney JD
    (1985) Controlled exposure to a mixture of SO2, NO2, and particulate
    air pollutants: Effects on human pulmonary function and respiratory
    symptoms. Arch Environ Health, 40: 197-201.

    Ko MKW, McElroy MB, Weisenstein DK, & Sze ND (1986) Lightning: A
    possible source of stratospheric odd nitrogen. J Geophys Res (Atmos),
    91: 5395-5404.

    Koenig JQ, Covert DS, Morgan MS, Horike M, Horike N, Marshall SG, &
    Pierson WE (1985) Acute effects of 0.12 ppm ozone or 0.12 ppm nitrogen
    dioxide on pulmonary functions in healthy and asthmatic adolescents.
    Am Rev Respir Dis, 132: 648-651.

    Koenig JQ, Pierson WE, Marshall SG, Covert DS, Morgan MS, & Van Belle
    G (1987a) The effects of ozone and nitrogen dioxide on lung function
    in healthy and asthmatic adolescents. Cambridge, Massachusetts,
    Institute of Health Effects (Research Report No. 14).

    Koenig JQ, Covert DS, Marshall SG, Van Belle G, & Pierson WE (1987b)
    The effects of ozone and nitrogen dioxide on pulmonary function in
    healthy and in asthmatic adolescents. Am Rev Respir Dis, 136:1152-1157.

    Koenig JQ, Covert DS, & Pierson WE (1989a) Effects of inhalation of
    acidic compounds on pulmonary function in allergic adolescent
    subjects. In: Symposium on the Health Effects of Acid Aerosols.
    Environ Health Perspect, 79: 173-178.

    Koenig JQ, Hanley QS, Anderson TL, Rebolledo V, & Pierson WE (1989b)
    An assessment of pulmonary function changes and oral ammonia levels
    after exposure of adolescent asthmatic subjects to sulfuric or nitric
    acid. Presented at the 82nd Annual Meeting and Exhibition of the Air
    and Waste Management Association. Pittsburgh, Pennsylvania, Air and
    Waste Management Association (Paper No. 89-92).4.

    Koerselman W, Bakker SA, & Blom M (1990) Nitrogen, phosphorus and
    potassium mass balances for two small fens surrounded by heavily
    fertilized pastures. J Ecol, 78: 428-442.

    Koerselman W & Verhoeven JTA (1992) Nutrient dynamics in mires of
    various trophic status: nutrient inputs and outputs and the internal
    nutrient cycle. In: Verhoeven JTA ed. Fends and bogs in The
    Netherlands: Vegetation, history, nutrient dynamics and conservation.
    Dordrecht, The Netherlands, Kluwer, pp 397-432.

    Kon K, Maeda N, & Shiga T (1977) Effect of nitric oxide on the oxygen
    transport of human erythrocytes. J Toxicol Environ Health,
    2: 1109-1113.

    Koo TC, Ho JH-C, Ho C-Y, Matsuki H, Shimizu H, Mori T, & Tominaga S
    (1990) Personal exposure to nitrogen dioxide and its association with
    respiratory illness in Hong Kong. Am Rev Respir Dis, 141: 1119-1126.

    Kooijman SALM (1987) A safety factor for LC50 values allowing for
    differences in sensitivity among species. Water Res, 22: 269-276.

    Koontz MD, Fortmann RC, Nagda NL, & Billick TH (1986) Protocol for an
    indoor air quality monitoring survey conducted in Texas. Presented at
    the 79th Annual Meeting of the Air Pollution Control Association.
    Pittsburgh, Pennsylvania, Air Pollution Control Association (Paper
    No. 86-6.3).

    Kosmider ST, Misiewicz A, Felus E, Drozdz M, & Ludyga K (1973)
    [Experimental and clinical studies on the effects of nitrogen oxides
    on immunity.] Int Arch Arbeitsmed, 31: 9-23 (in German).

    Kosta-Rick R & Manning WJ (1993) Radish ( raphanus sativus L.) model
    for studying plant responses to air pollutants and other environmental
    stresses. Environ Pollut, 82: 107-138.

    Kowalczyk ML & Bauer E (1981) Lightning as a source of NOx in the
    troposphere. Washington, DC, US Department of Transportation, Federal
    Aviation Administration (Report Nos. IDA-P-1590 and FAA/EE-82-4).

    Kripke BJ & Sherwin RP (1984) Nitrogen dioxide exposure - influences
    on rat testes. Anesth Analg (NY), 63: 526-528.

    Kroeze C, Pegtel DM, & Blom CJC (1989) An Experimental comparison of
    aluminium and manganese susceptibility in  Antennaria dioica, Viola
     canina, Filago minima and Deschampsia lexusa. Acta Bot Neerl,
    38: 165-172.

    Kubota K, Murakami M, Takenaka S, Kawai K, & Kyouo H (1987) Effects of
    long-term nitrogen dioxide exposure on rat lung: morphological
    observations. Environ Health Perspect, 73: 157-169.

    Kuehr J, Hendel-Kramer A, Karmaus W, Moseler M, Weiss K, Stephan V, &
    Urbanek R (1991) [Air pollution and asthma among school children.] Soz
    Praeventivmed, 36: 67-73 (in German).

    Kulle TJ (1982) Effects of nitrogen dioxide on pulmonary function in
    normal health humans and subjects with asthma and chronic bronchitis.
    In: Schneider T & Grant L ed. Air pollution by nitrogen oxides:
    Proceedings of the US-Dutch International Symposium. Amsterdam,
    Oxford, New York, Elsevier Science Publishers, pp 477-486.

    Kulle TJ & Clements ML (1988) Susceptibility to virus infection with
    exposure to nitrogen dioxide. Cambridge, Massachusetts, Institute of
    Health Effects (Research Report No. 15).

    Kupper LL (1984) Effects of the use of unreliable surrogate variables
    on the validity of epidemiologic research studies. Am J Epidemiol,
    120: 643-648.

    Kuppers K & Klump G (1988) Effects of ozone, sulphur dioxide, and
    nitrogen dioxide on gas exchange and starch economy in Norway Spruce
    ( Picea abies [L.]Karsten. GeoJournal, 17(2): 271-275.

    Kwon NS, Stuehr DJ, & Nathan CF (1991) Inhibition of tumour cell
    ribonucleotide reductase by macrophage-derived nitric oxide. J Exp
    Med, 174: 761.

    Kyono H & Kawai K (1982) Morphometric study on age-dependent pulmonary
    lesions in rats exposed to nitrogen dioxide. Ind Health, 20: 73-99.

    Lacis AA, Wuebbles DJ, & Logan JA (1990) Radiative forcing of climate
    by changes in the vertical distribution of ozone. J Geophys Res
    (Atmos), 95: 9971-9981.

    Lafuma C, Harf A, Lange F, Bozzi L, Poncy JL, & Bignon J (1987) Effect
    of low-level NO2 chronic exposure on elastase-induced emphysema.
    Environ Res, 43: 75-84.

    Laiho O (1970) Paxillus involutus as mycorrhizal symbiont of forest
    trees. Acta For Fenn, 79: 1-35.

    Laird NM & Mosteller F (1990) Some statistical methods for combining
    experimental results. Int J Technol Assess Health Care, 6: 5-30.

    Lam C, Kattan M, Collins A, & Kleinerman J (1983) Long-term sequelae
    of bronchiolitis induced by nitrogen dioxide in hamsters. Am Rev
    Respir Dis, 128: 1020-1023.

    Lambert WE, Samet JM, Hunt WC, Skipper BJ, Schwab M, & Spengler JD
    (1993) Assessment of exposure to nitrogen dioxide. In: Nitrogen
    dioxide and respiratory illness in children, part II. Cambridge,
    Massachusetts, Institute of Health Effects, pp 33-50 (Research Report
    No. 58).

    Lane PI & Bell JNB (1984) The effects of simulated urban air pollution
    on grass yield. Part 2: Performance of  Lolium perenne, Phleum
    pratense and  Dactylis glomerata fumigated with SO2, NO2 and/or NO.
    Environ Pollut, A35: 97-124.

    Larsson U, Elmgren R, & Wulff F (1985) Eutrophication and the Baltic
    Sea: causes and consequences. Ambio, 14: 9-14.

    Last JA & Warren DL (1987) Synergistic interaction between nitrogen
    dioxide and respirable aerosols of sulfuric acid or sodium chloride on
    rat lungs. Toxicol Appl Pharmacol, 90: 34-42.

    Last JA, Gerriets JE, & Hyde DM (1983) Synergistic effects on rat
    lungs of mixtures of oxidant air pollutants (ozone or nitrogen
    dioxide) and respirable aerosols. Am Rev Respir Dis, 128: 539-544.

    Leaderer BP (1982) Air pollutant emissions from kerosene space
    heaters. Science, 218: 1113-1115.

    Leaderer BP, Zagraniski RT, Berwick M, & Stolwijk JAJ (1986)
    Assessment of exposure to indoor air contaminants from combustion
    sources: Methodology and application. Am J Epidemiol, 124: 275-289.

    Lebowitz MD, Holberg CJ, Boyer B, & Hayes C (1985) Respiratory
    symptoms and peak flow associated with indoor and outdoor air
    pollutants in the southwest. J Air Pollut Control Assoc,
    35: 1154-1158.

    Lebret E (1987) Errors in exposure variables and their role in
    obscuring exposure-response relations: The case of indoor exposure to
    nitrogen dioxide and fine particles. In: Seifert B, Esdorn H, Fischer
    M, Rueden H, & Wegner J ed. Indoor air '87: Proceedings of the 4th
    International Conference on Indoor Air Quality and Climate -Volume 1:
    Volatile organic compounds, combustion gases, particles and fibres,
    microbiological agents. Berlin, Germany, Institute for Water, Soil and
    Air Hygiene, pp 88-292.

    Lebret E (1990) Errors in exposure measures. Toxicol Ind Health,
    6: 147-156.

    Lechowicz MJ & Shaver GR (1982) A multivariate approach to the
    analysis of factorial fertilization experiments Alaskan arctic tundra.
    Ecology, 63: 1029-1038.

    Lee JA & Studholme CJ (1992) Responses of  Sphagnum species to
    polluted environments. In: Bates JW & Farmer AM ed. Bryophytes and
    lichens in a changing environment. Oxford, United Kingdom, Clarendon
    Press, pp 314-332.

    Lee JA, Press MC, & Woodin SJ (1986) Effects of NO2 on aquatic
    ecosystems. In: Environment and qualiq of life: Study on the need for
    an NO2 long-term limit value for the protection of terrestrial and
    aquatic ecosystems. Luxembourg, Commission of the European
    Communities, pp 99-119.

    Lee JA, Press MC, Woodin S, & Ferguson P (1987) Responses to acidific
    deposition in ombrotrophic mires in the U.K. In: Hutchinson TC & Meema
    KM ed. Effects of atmospheric pollutants on forests, wetlands and
    agricultural ecosystems. New York, Berlin, Springer-Verlag,
    pp 549-560.

    Lee JA, Baddeley JA, & Woodin SR (1989) Effects of acidic deposition
    on semi-natural vegetation. In: Acidification in Scotland. Edinburgh,
    Scottish Development Department.

    Lefkowitz SS, McGrath JJ, & Lefkowitz DL (1983) Effects of NO2 on
    interferon production in mice. In: Lee SD, Mustafa MG, & Mehlman MA
    ed. International Symposium on the Biomedical Effects of Ozone and
    related Photochemical Oxidants, March 1982. Princeton, New Jersey,
    Princeton Scientific Publishers, Inc., vol 5, pp 469-474.

    Lefkowitz SS, McGrath JJ, Lefkowitz DL, & Spicer JB (1984) Interferon
    production following NO2 exposure. Int J Immunopharmacol, 6: 275-278.

    Lefkowitz SS, McGrath JJ, & Lefkowitz DL (1986) Effects of NO2 on
    immune responses. J Toxicol Environ Health, 17: 241-248.

    Lefohn AS & Tingey DT (1984) The co-occurrence of potentially
    phytotoxic concentrations of various gaseous air pollutants. Atmos
    Environ, 18: 2521-2526.

    Lefohn AS, Davis CE, Jones CK, Tingey DT, & Hogsett WE (1987)
    Co-occurrence patterns of gaseous air pollutant pairs at different
    minimum concentrations in the United States. Atmos Environ,
    21: 2435-2444.

    Lefohn AS, Benkovitz CM, Tanner RL, Smith LA, & Shadwick DS (1991) Air
    quality measurements and characterizations for terrestrial effects
    research. In: Irving PM ed. Acidic deposition: State of science and
    technology - volume 1: Emissions, atmospheric processes, and
    deposition. Washington, DC, The US National Acid Precipitation
    Assessment Program (State of Science and Technology Report No. 7).

    Lepoivre M, Fieschi F, Coves J, Thelander L, & Fontecave M (1991)
    Inactivation of ribonucleotide reductase by nitric oxide. Biochem
    Biophys Res Commun, 179: 442-448.

    Leu M-T (1988) Heterogeneous reactions of N2O5 with H2O and HCI on
    ice surfaces: Implications for Antarctic ozone depletion. Geophys Res
    Lett, 15: 851-854.

    Leuven RSEW, Den Hartog C, Christiaans MMC, & Heyligers WHC (1986)
    Effects of water acidification on the distribution pattern and the
    reproductive success of amphibians. Experientia (Basel), 42: 495-503.

    Levander T (1990) The relative contributions to the greenhouse effect
    from the use of different fuels. Atmos Environ, A24: 2707-2714.

    Levine JS, Rogowski RS, Gregory GL, Howell WE, & Fishman J (1981)
    Simultaneous measurements of NOX, NO, and O3 production in a
    laboratory discharge: Atmospheric implications. Geophys Res Lett,
    8: 357-360.

    Levine JS, Augustsson TR, Anderson IC, Hoell JM Jr, & Brewer DA (1984)
    Tropospheric sources of NOx: lightning and biology. Atmos Environ
    18: 1797-1804.

    Lilievald J & Crutzen PJ (1994) Role of deep cloud convection in the
    Ozone budget of the troposphere. Science, 264: 1759-1761.

    Liljelund LE & Torstensson P (1988)  Critical load of nitrogen with
    regards to effects on plant composition.  In: Nilsson J & Grennfelt P
    ed. Critical loads for sulphur and nitrogen: report from a workshop.
    Copenhagen, Denmark, Nordic Council of Ministers, pp 363-373.

    Lindberg SE, Lovett GM, & Meiwes KJ (1987) Deposition and forest
    canopy interactions of airborne nitrate. In: Hutchinson TC & Meema KM
    ed. Effects of atmospheric pollutants on forests, wetlands and
    agricultural ecosystems. Berlin, Springer-Verlag, pp 117-130.

    Linn WS & Hackney JD (1983) Short-term human respiratory effects of
    nitrogen dioxide: determination of quantitative dose-response profile.
    Phase I - Exposure of healthy volunteers to 4 ppm NO2. Atlanta,
    Georgia, Coordinating Research Council Inc. (Report No. CRC-APRAC-
    CAPM-48-83).

    Linn WS & Hackney JD (1984) Short-term human respiratory effects of
    nitrogen dioxide: determination of quantitative dose-response
    profiles. Phase II - Exposure of asthmatic volunteers to 4 ppm NO2.
    Atlanta,Georgia, Coordinating Research Council Inc. (Report No.
    CRC-CAPM-48-83-02).

    Linn WS, Jones MP, Bailey RM, Kleinman MT, Spier CE, Fischer DA, &
    Hackney JD (1980a) Respiratory effects of mixed nitrogen dioxide and
    sulfur dioxide in human volunteers under simulated ambient exposure
    conditions. Environ Res, 22: 431-438.

    Linn WS, Jones MP, Bachmayer EA, Spier CE, Mazur SF, Avol EL, &
    Hackney JD (1980b) Short-term respiratory effects of polluted ambient
    air: a laboratory study of volunteers in a high-oxidant community. Am
    Rev Respir Dis, 121: 243-252.

    Linn WS, Shamoo DA, Spier CE, Valencia LM, Anzar UT, Venet TG, Avol
    EL, & Hackney JD (1985a) Controlled exposure of volunteers with
    chronic obstructive pulmonary disease to nitrogen dioxide. Arch
    Environ Health, 40: 313-317.

    Linn WS, Solomon JC, Trim SC, Spier CE, Shamoo DA, Venet TG, Avol EL,
    & Hackney JD (1985b) Effects of exposure to 4 ppm nitrogen dioxide in
    healthy and asthmatic volunteers. Arch Environ Health, 40: 234-239.

    Linn WS, Shamoo DA, Avol EL, Whynot JD, Anderson KR, Venet TG, &
    Hackney JD (1986) Dose-response study of asthmatic volunteers exposed
    to nitrogen dioxide during intermittent exercise. Arch Environ Health,
    41: 292-296.

    Lipari F (1984) New solid-sorbent method for ambient nitrogen dioxide
    monitoring. Anal Chem, 56: 1820-1826.

    Lipschultz F, Zafiriou OC, Wofsy SC, McElroy MB, Valois FW, & Watson
    SW (1981) Production of NO and N2 by soil nitrifying bacteria. Nature
    (Lond), 294: 641-643.

    Littenberg B (1988) Aminophylline treatment in severe, acute asthma: a
    meta-analysis. J Am Med Assoc, 259: 1678-1684.

    Liu SC, Kley D, McFarland M, Mahlman JD, & Levy H II. (1980) On the
    origin of tropospheric ozone. J Geophys Res (Oceans Atmos),
    85: 7546-7552.

    Liu SC, Trainer M, Fehsenfeld FC, Parrish DD, Williams EJ, Fahey DW,
    Huebler G, & Murphy PC (1987) Ozone production in the rural
    troposphere and the implications for regional and global ozone
    distributions. J Geophys Res (Atmos), 92: 4191-4207.

    Logan JA (1983) Nitrogen oxides in the troposphere: global and
    regional budgets. J Geophys Res (Oceans Atmos), 88: 10785-10807.

    Logan JA (1985) Tropospheric ozone: seasonal behavior, trends, and
    anthropogenic influence. J Geophys Res  (Atmos), 90: 10463-10482.

    Lovett G (1992) Atmospheric deposition and canopy interactions of
    nitrogen. In: Johnson DW & Lindberg SE ed. Atmospheric deposition and
    forest nutrient cycling: a synthesis of the integrated forest study.
    New York, Springer-Verlag.

    Lovett GM, & Lindberg SE (1984) Dry deposition and canopy exchange in
    a mixed oak forest as determined by analysis of throughfall. J Appl
    Ecol, 21: 1013-1027.

    Lovett GM, & Lindberg SE (1986) Dry deposition of nitrate to a
    deciduous forest. Biogeochemistry, 2: 137-148.

    Lowry T & Schuman LM (1956) "Silo-filler's disease" - a syndrome
    caused by nitrogen dioxide. J Am Med Assoc, 162: 153-160.

    Loye-Pilot MD, Martin JM, & Morelli J (1990) Atmospheric input of
    inorganic nitrogen to the Western Mediterranean. Biogeochemistry,
    9: 117-134.

    Lucas PW (1990) The effects of prior exposure to sulphur dioxide and
    nitrogen dioxide on the water relations of Timothy grass  (Phleum
     pratense) under drought conditions. Environ Pollut, 66: 117-138.

    Luke WT & Dickerson RR (1987) The flux of reactive nitrogen compounds
    from eastern North America to the western Atlantic Ocean. Global
    Biogeochem Cycles, 1: 329-343.

    Lyons WA & Cole HS (1976) Photochemical oxidant transport: mesoscale
    lake breeze and synoptic-scale aspects. J Appl Meteorol, 15: 733-743.

    McCall T & Vallance P (1992)  Nitric oxide tabes centre stage with
    newly defined roles.  Trends Pharmacol Sci, 13: 1-6.

    McConnell JC (1973) Atmospheric ammonia. J Geophys Res, 78: 7812-7821.

    Machta L  (1983) Effects of non-CO2 greenhouse gases. In: Changing
    climate: report of the Carbon Dioxide Assessment Committee.
    Washington, DC, National Academy Press, pp 285-291.

    McElroy MB (1980) Sources and sinks for nitrous oxide. Washington, DC,
    US Department of Transportation (Report No. FAA-EE-80-20).

    McElroy MB, Salawitch RJ, Wofsy SC, & Logan JA (1986) Reductions of
    Antarctic ozone due to synergistic interactions of chlorine and
    bromine. Nature (Lond), 321: 759-762.

    McElroy MB & Salawitch RJ (1989) Stratospheric ozone: impact of human
    activity. Planet Space Sci, 37: 1653-1672.

    McGrath JJ & Oyervides J (1985)  Effects of nitrogen dioxide on
    resistance to Klebsiella pneumoniae in mice.  J Am Coll Toxicol,
    4: 227-231.

    Macriss RA & Elkins RH (1977) Control of the level of NOx in the
    indoor environment. In: Kasuga S, Suzuki N, Yamada T, Kimura G,
    Inagaki K, & Onoe K ed. Proceedings of the Fourth International Clean
    Air Congress. Tokyo, Japan, Japanese Union of Air Pollution Prevention
    Associations, pp 510-514.

    Maeda Y, Aoki K, & Munemori M (1980) Chemiluminescence method for the
    determination of nitrogen dioxide. Anal Chem, 52: 307-311.

    Maeda N, Imaizumi K, Kon K, & Shiga T (1987) A kinetic study on
    functional impairment of nitric oxide-exposed rat erythrocytes. 
    Environ Health Perspect, 73: 171-177.

    Magee PN (1971) Toxicity of nitrosamines: their possible human health
    hazards. Food Cosmet Toxicol, 9: 207-218.

    Magee PN, Montesano R, & Preussmann R (1976) N-nitroso compounds
    and related carcinogens. In: Searle CE ed. Chemical carcinogens.
    Washington, DC, American Chemical Society, pp 491-625 (ACS Monograph
    No. 173).

    Maigetter RZ, Fenters JD, Findlay JC, Ehrlich R0, & Gardner DE (1978) 
    Effect of exposure to nitrogen dioxide on T and B cells in mouse
    spleens. Toxicol Lett, 2:157-161.

    Malko MW & Troe J (1982) Analysis of the unimolecular reaction N2O5
    + M = NO2 + NO3 + M. Int J Chem Kinet, 14: 399-416.

    Mann C (1990) Meta-analysis in the breech. Science, 249: 76-480.

    Mansfield TA & McCune DC (1988) Problems of crop loss assessment when
    there is exposure to two or more gaseous pollutants. In: Heck WW,
    Taylor OC, & Tingey DT ed. Assessment of crop loss from air
    pollutants. London, Elsevier Applied Science, pp 317-344.

    Mantel N & Haenszel W (1959) Statistical aspects of the analysis of
    data from retrospective studies of disease. J Natl Cancer Inst,
    22: 719-748.

    Maples KR, Sandström T, Su Y-F, & Henderson RF (1991) The nitric
    oxide/heme protein complex as a biological marker of exposure to
    nitrogen dioxide in humans, rats and  in vitro models. Am J Respir
    Cell Mol Biol, 4: 538-543.

    Marchetto A, Mosello R, Psenner R, Barbieri A, Bendetta G, Tait D, &
    Tartari GA (1994)  Evaluation of the level of acidification and the
    critical loads for Alpine lakes. Ambio, 23: 150-154.

    Marbury MC, Harlos DP, Samet JM, & Spengler JD (1988) Indoor
    residential NO2 concentrations in Albuquerque, New Mexico. J Air
    Pollut Control Assoc, 38: 392-398.

    Margolis PA, Greenberg RA, Keyes LL, Lavange LM, Chapman RS, Denny FW,
    Bauman KE, & Boat BW (1992) Lower respiratory illness in infants and
    low socioeconomic status. Am J Public Health, 82: 1119-1126.

    Marrs RH (1993)  An assessment of change in  Calluna heathland. Biol
    Conserv, 65: 133-139.

    Marshall VG (1977) Effects of manure and fertilizers on soil fauna: a
    review. Harpenden, United Kingdom, Rothamsted Experimental Station,
    Commonwealth Bureau of Soils, pp 1-79 (Special Publication No. 3).

    Marshall JD & Cadle SH (1989) Evidence for trans-cuticular uptake of
    HNO3 vapor by foliage of eastern white pine ( Pinus strobus L.).
    Environ Pollut, 60: 15-28.

    Massachusetts Institute of Technology (1976) Experimental evaluation
    of range-top burner modification to reduce NOx formation. American
    Gas Association (Report No. M40677).

    Matheson Company (1966) Matheson gas data book, 4th ed. East
    Rutherford, New Jersey, The Matheson Company, Inc.

    Matthews RD, Sawyer RF, & Schefer RW (1977) Interferences in
    chemiluminescent measurement of NO and NO2 emissions from combustion
    systems. Environ Sci Technol, 11:1092-1096.

    Mehlhorn H & Wellburn AR (1987)  Stress ethylene formation determines
    plant sensitivity to ozone. Nature (Lond), 327: 417-418.

    Melia RJW, Florey C du V, Altman DG, & Swan AV (1977) Association
    between gas cooking and respiratory disease in children. Br Med J,
    2: 149-152.

    Melia RJW, Florey C du V, Darby SC, Palmes ED, & Goldstein BD (1978)
    Differences in NO2 levels in kitchens with gas or electric cookers.
    Atmos Environ, 12: 1379-1381.

    Melia RJW, Florey C du V, & Chinn S (1979) The relation between
    respiratory illness in primary schoolchildren and the use of gas for
    cooking: I. Results from a national survey. Int J Epidemiol,
    8: 333-338.

    Melia RJW, Florey C du V, Chinn S, Goldstein BD, Brooks AGF, John HH,
    Clark D, Craighead IB, & Webster X (1980) The relation between indoor
    air pollution from nitrogen dioxide and respiratory illness in primary
    schoolchildren. Clin Respir Physiol, 16: 7P-8P.

    Melia RJW, Florey C du V, Morris RW, Goldstein BD, Clark D, & John HH
    (1982a) Childhood respiratory illness and the home environment. I.
    Relations between nitrogen dioxide, temperature and relative humidity.
    Int J Epidemiol, 11: 155-163.

    Melia RJW, Florey C du V, Morris RW, Goldstein BD, John HH, Clark D,
    Craighead IB, & Mackinlay JC (1982b) Childhood respiratory illness and
    the home environment: II. Association between respiratory illness and
    nitrogen dioxide, temperature and relative humidity. Int J Epidemiol,
    11: 164-169.

    Melia RJW, Florey C, Sittampalam Y, & Watkins C (1983) The relation
    between respiratory illness in infants and gas cooking in the UK: a
    preliminary report. In: Proceedings of the VIth World Congress on Air
    Quality Paris, Air Pollution Prevention Association, pp 263-269.

    Melia RJW, Florey C du V, Chinn S, Morris RW, Goldstein BD, John HH, &
    Clark D (1985) Investigations into the relations between respiratory
    illness in children, gas cooking and nitrogen dioxide in the UK Tokai.
    J Exp Clin Med, 10: 375-378.

    Melia RJW, Chinn S, & Rona RJ (1988) Respiratory illness and home
    environment of ethnic groups. Br Med J, 296: 1438-1441.

    Melia RJW, Chinn S, & Rona RJ (1990) Indoor levels of NO2 associated
    with gas cookers and kerosene heaters in inner city areas of England.
    Atmos Environ (Urban Atmos), 24: 177-180.

    Menge JA & Grand LF (1978) Effects of fertilization on the production
    of epigeous basidiocarps by mycorrhizal fungi in lobloly pine
    plantations.  Can J Bot, 56: 2357-2362.

    Menzel DB (1970) Toxicity of ozone, oxygen and radiation. Annu Rev
    Pharmacol,  10: 379-394.

    Menzel DB (1976) The role of free radicals in the toxicity of air
    pollutants (nitrogen oxides and ozone).  In: Pryor WA ed. Free
    radicals in molecular biology and pathology. New York, London,
    Academic Press, Inc., vol 2, pp 181-202.

    Menzel DB, Roehm JN, & Lee SD (1972) Vitamin E: The biological and
    environmental antioxidant. J Agric Food Chem, 20(3): 481-486.

    Menzel DB, Abou-Donia NB, Roe CR, Ehrlich R, Gardner DE, & Coffin DL
    (1977) Biochemical indices of nitrogen dioxide intoxication of guinea
    pigs following low level-long term exposure. In: Proceedings of
    International Conference on Photochemical Oxidant Pollution and its
    Control.  Research Triangle Park, North Carolina, US Environmental
    Protection Agency,  Environmental Sciences Research Laboratory, vol
    II, pp 577-587.

    Mercer RR, Costa DL, & Crapo JD (1995) Effects of prolonged exposure
    to low doses of nitric oxide or nitrogen dioxide on the alveolar septa
    of the adult rat lung. Lab Invest, 73(1): 20-38.

    Meulenbelt J & Sangster B (1990) Acute nitrous oxide intoxication:
    clinical symptoms, pathophysiology and treatment. Neth J Med,
    37: 132-138.

    Meyer FH (1988) Ectomycorrhiza and decline of trees. In: AE Jansen,
    Dighton J, & Bresser AHM ed. Ectomycorrhiza and acid rain. 
    Proceedings of the Workshop on Ectomycorrhiza. Luxembourg, Commission
    of the European Communities, pp 9-31 (CEC Air Pollution Research
    Report No. 12).

    Meyrahn H, Helas G, & Warneck P (1987) Gas chromatographic
    determination of peroxyacetyl nitrate: two convenient calibration
    techniques. J Atmos Chem, 5: 405-415.

    Michie RM Jr, Sokash JA, Fritschel BP, McElroy FF, & Thompson VL
    (1983) Performance test results and comparative data for designated
    reference methods for nitrogen dioxide. Research Triangle Park, North
    Carolina, US Environmental Protection Agency, Environmental Monitoring
    Systems Laboratory (EPA-600/4-83-019).

    Miflin BJ (1980) Amino acids and derivatives. Volume 5 - The
    biochemistry of plants: A comprehensive treatise. New York, London,
    Academic Press, Inc.

    Miller DP (1984) Ion chromatographic analysis of Palmes tubes for
    nitrite.  Atmos Environ, 18: 891-892.

    Miller DF, Alkezweeny AJ, Hales JM, & Lee RN (1978) Ozone formation
    related to power plant emissions. Science, 202: 1186-1188.

    Miller FJ, Graham JA, Illing JN, & Gardner DE (1980)  Extrapulmonary
    effects of NO2 as reflected by pento-barbital-induced sleeping time
    in mice. Toxicol Lett, 6: 267-274.

    Miller FJ, Overton JH, Myers ET, & Graham JA (1982) Pulmonary
    dosimetry of nitrogen dioxide in animals and man. In: Schneider R &
    Grant L ed. Air pollution by nitrogen oxides. Amsterdam, Oxford, New
    York, Elsevier Science Publishers, pp  377-386.

    Miller FJ, Overton JH Jr., Jaskot RH, & Menzel DB (1985) A model of
    the regional uptake of gaseous pollutants in the lung: I. The
    sensitivity of the uptake of ozone in the human lung to lower
    respiratory tract secretions and exercise. Toxicol Appl Pharmacol,
    79: 11-27.

    Miller FJ, Graham JA, Raub JA, Illing JW, Ménache MG, House DE, &
    Gardner DE (1987) Evaluating the toxicity of urban patterns of oxidant
    gases: II. Effects in mice from chronic exposure to nitrogen dioxide.
    J Toxicol Environ Health,  21: 99-112.

    Mink SN, Coalson JJ, Whitley L, Greville H, & Jadue C (1984) Pulmonary
    function tests in the detection of small airway obstruction in a
    canine model of bronchiolitis obliterans. Am Rev Respir Dis,
    130: 1125-1133.

    Mirvish SS (1970) Kinetics of dimethylamine nitrosation in relation to
    nitrosamine carcinogenesis. J Natl Cancer Inst, 44: 633-639.

    Mirvish SS, Issenberg P, & Sams JP (1981) N-nitroso-morpholine
    synthesis in rodents exposed to nitrogen dioxide and morpholine. In:
    N-nitroso compounds. Washington, DC, American Chemical Society,
    pp 181-191 (ACS Symposium Series No. 174).

    Mirvish SS, Sams JP, & Issenberg P (1983) The nitrosating agent in
    mice exposed to nitrogen dioxide: improved extraction method and
    localization in the skin. Cancer Res, 43: 2550-2554.

    Mirvish SS, Babcook DM, Deshpande AD, & Nagel DL (1986) Identification
    of cholesterol as a mouse skin lipid that reacts with nitrogen dioxide
    to yield a nitrosating agent, and of cholesteryl nitrite as the
    nitrosating agent produced in a chemical system from cholesterol.
    Cancer Lett, 31: 97-104.

    Mirvish SS, Ramm MD, Sams JP, & Babcook DM (1988) Nitrosamine
    formation from amines applied to the skin of mice after and before
    exposure to nitrogen dioxide. Cancer Res, 48: 1095-1099.

    Mitchell RI, Williams R, Cote RW, Lanese RR, & Keller MD (1975)
    Household survey of the incidence of respiratory disease in relation
    to environmental pollutants. In: Recent advances in the assessment of
    the health effects of environmental pollution: Proceedings of an
    International Symposium. Luxembourg, Commission of the European
    Communities, vol 2, pp 47-61 (Publication No. 5360).

    Mitsch WJ & Gosselink JG (1986) Wetlands. New York, Van Nostrand
    Reinhold Company.

    Miyazaki T (1984) Adsorption characteristics of NOx by several kinds
    of interior materials. In: Berglund B, Lindvall T, & Sundell J ed.
    Indoor air '84 - Proceedings of the 3rd International Conference on
    Indoor Air Quality and Climate. Stockholm, Swedish Council for
    Building Research, vol 4, pp 103-110.

    Mochitate K, Kava K, Miura T, & Kubota K (1984)  In vivo effects of
    nitrogen dioxide on membrane constituents in lung and liver of rats.
    Environ Res, 33: 17-28.

    Mochitate K, Miura T, & Kubota K (1985) An increase in the activities
    of glycolytic enzymes in rat lungs produced by nitrogen dioxide. J
    Toxicol Environ Health, 15: 323-331.

    Mochitate K, Takahashi Y, Ohsumi T, & Miura T (1986) Activation and
    increment of alveolar macrophages induced by nitrogen dioxide. J
    Toxicol Environ Health, 16: 229-239.

    Mohsenin V (1987a) Airway responses to nitrogen dioxide in asthmatic
    subjects. J Toxicol Environ Health, 22: 371-380.

    Mohsenin V (1987b) Effect of vitamin C on NO2-induced airway
    hyperresponsiveness in normal subjects: a randomized double-blind
    experiment. Am Rev Respir Dis, 136: 1408-1411.

    Mohsenin V (1988) Airway responses to 2.0 ppm nitrogen dioxide in
    normal subjects. Arch Environ Health, 43: 242-246.

    Mohsenin V (1991) Lipid peroxidation and antielastase activity in the
    lung under oxidant stress: role of antioxidant defenses. J Appl
    Physiol, 70: 1456-1462.

    Mohsenin V & Gee JBL (1987) Acute effect of nitrogen dioxide exposure
    on the functional activity of alpha-1-protease inhibitor in
    bronchoalveolar lavage fluid of normal subjects. Am Rev Respir Dis,
    136: 646-650.

    Moinard J, Manier G, Pillet O, & Castaing Y (1994) Effect of inhaled
    nitric oxide on haemodynamics and Va/Q inequalities in patients with
    chronic obstructive pulmonary disease. Am J Respir Crit Care Med,
    149: 1482-1487.

    Molina LT & Molina MJ (1987) Production of chlorine oxide (Cl2O2)
    from the self-reaction of the ClO radical. J Phys Chem, 91: 433-436.

    Molina MJ, Tso T-L, Molina LT, & Wang FCY (1987) Antarctic
    stratospheric chemistry of chlorine nitrate, hydrogen chloride, and
    ice: release of active chlorine. Science, 238: 1253-1257.

    Moncada S (1992) The 1991 Ulf v Euler lecture. The L-arginine: nitric
    oxide pathway.  Acta Physiol Scand, 145: 201-227.

    Moncada S, Radomski MW, & Palmer RMJ (1988) Endothelium-derived
    relaxing factor: Identification as nitric oxide and role in the
    control of vascular tone and platelet function. Biochem Pharmacol,
    37: 2495-2501.

    Moncada S, Palmer RMJ, & Higgs EA (1991) Nitric oxide: Physiology,
    pathophysiology, and pharmacology. Pharmacol Rev, 43(2): 109-142.

    Mooi J (1984) [Effects of SO2, NO2, O3 and their mixtures on
    poplars and other plant species.] Forst-Holzwirt, 39: 438-444 (in
    German).

    Moore DRJ & Keddy PA (1989) The relationship between species richness
    and standing crop in wetlands: the importance of scale. Vegetatio,
    79: 99-106.

    Morgan SA, Lee JA, & Ashenden TW (1992)  Effects of nitrogen oxides on
    nitrate assimilation in bryophytes.  New Phytol, 120: 89-97.

    Morris JT (1991) Effects of nitrogen loading on wetland ecosystems,
    with particular references to atmospheric deposition. Annu Rev Ecol
    Syst, 22: 257-279.

    Morris JT, Houghton RA, & Botkin DB (1984) Theoretical limits of below
    ground production by  Spartina alterniflora:  an analysis through
    modelling. Ecol Model, 26: 155-175.

    Morrow PE & Utell MJ (1989) Responses of susceptible subpopulations to
    nitrogen dioxide. Cambridge, Massachusetts, Institute of Health
    Effects (Research Report No. 23).

    Morrow PE, Utell MJ, Bauer MA, Smeglin AM, Frampton MW, Cox C, Speer
    DM, & Gibb FR (1992) Pulmonary performance of elderly normal subjects
    and subjects with chronic obstructive pulmonary disease exposed to
    0.30 ppm nitrogen dioxide. Am Rev Respir Dis, 145: 291-300.

    Mortensen LM (1985) Nitrogen oxides produced during CO2 enrichment:
    II: Effects on different tomato and lettuce cultivars. New Phytol,
    101: 411-415.

    Mortensen LM (1986) nitrogen oxides produced CO2 enrichment: III
    Effects on tomato at different photon flux densities. New Phytol,
    104: 653-660.

    Moschandreas DJ, Relwani SM, Macriss RA, & Cole JT (1984) Differences
    and similarities of two techniques used to measure emission rates from
    unvented gas appliances. In: Berglund B, Lindvall T, & Sundell J ed.
    Indoor air '84 - Proceedings of the 3rd International Conference on
    Indoor Air Quality and Climate. Stockholm, Swedish Council for
    Building Research, vol 4, pp 375-379.

    Moschandreas DJ, Relwani SM, O'Neill HJ, Cole JT, Elkins RH, & Macriss
    RA (1985) Characterization of emission rates from indoor combustion
    sources. Chicago, Illinois, Gas Research Institute (Report No. GRI
    85/0075).

    Mosier AR, Stillwel M, Paton WJ, & Woodmansee RG (1981) Nitrous oxide
    emissions from a native shortgrass prairie. Soil Sci Soc Am J,
    45: 617-619.

    Moss B (1988) Ecology of fresh waters: man and medium. Oxford, Boston,
    Blackwell Scientific Publications.

    Motomiya K, Ito K, Yoshida A, Idewara S, Otsu Y, & Nakajima Y (1973)
    [The effects of exposure to NO2 gas on the infection of influenza
    virus of mouse - long term experiment in low concentration.]  Kankyo
    Kagaku Kenkyu Hokoku (Chiba Daigak), 1: 27-33 (in Japanese).

    Mountford MO, Lakhani KH, & Holland (1994) The effects of nitrogen on
    species diversity and agricultural production on the Somerset Moors.
    Phase II: (a) after seven years of fertilizers application; (b) after
    cessation of fertilizer input for three years. Huntingdon, United
    Kingdom, Institute of Terrestrial Ecology.

    Muelenaer P, Reid H, Morris R, Saltzman L, Horstman D, Collier A, &
    Henderson F (1987) Urinary hydroxyproline excretion in young males
    exposed experimentally to nitrogen dioxide. In: Seifert B, Esdorn H,
    Fischer M, Rueden H, & Wegner J ed. Indoor air '87 - Proceedings of
    the 4th International Conference on Indoor Air Quality and Climate.
    Berlin, Institute for Water, Soil and Air Hygiene, vol 2, pp 97-103.

    Mulik JD & Williams D (1986) Passive sampling devices for NO2.
    In: Proceedings of the 1986 EPA/APCA Symposium on Measurement of
    Toxic Air Pollutants, Raleigh, North Carolina, April 1986. Pittsburgh,
    Pennsylvania, Air Pollution Control Association, pp 61-70 (EPA-600/
    9-86-013).

    Mulik JD & Williams DE (1987) Passive sampling device measurements of
    NO2 in ambient air. In: Proceedings of the 1987 EPA/APCA Symposium on
    Measurement of Toxic and Related Air Pollutants, Research Triangle
    Park, North Carolina. Pittsburgh, Pennsylvania, Air Pollution Control
    Association, pp 387-397 (EPA-600/9-87-010).

    Muniz IP (1991) Freshwater acidification:  its effects on species and
    communities of freshwater microbes, plants and animals. Proc Royal Soc
    Edinb, 97b: 227-254.

    Murphy SD, Ulrich CE, Frankowitz SH, & Xintaras C (1964) Altered
    function in animals inhaling low concentrations of ozone and nitrogen
    dioxide. Am Ind Hyg Assoc J, 25: 246-253.

    Murray AJS & Wellburn AR (1980) Differences in nitrogen metabolism
    between contuvars of tomato and pepper during exposure to glasshouse
    atmosphere containing oxides of nitrogen. Environ Pollut, 39: 303-316.

    Murray F, Clarke K, & Wilson S (1992) Effects of NO2 on hoop pine can
    be counteracted by SO2. Eur J For Pathol, 22: 403-409.

    Mustafa MG, Elsayed N, Lim JST, Postlethwait E, & Lee SD (1979)
    Effects of nitrogen dioxide on lung metabolism. In: Grosjean D ed.
    Nitrogenous air pollutants: Chemical and biological implications. Ann
    Arbor, Michigan, Ann Arbor Science Publishers Inc., pp 165-178.

    Mustafa MG, Elsayed NM, Von Dohlen FM, Hassett CM, Postlethwait EM,
    Quinn CI, Graham JA, & Gardner DE (1984) A comparison of biochemical
    effects of nitrogen dioxide, ozone, and their combination in mouse
    lung.  I. Intermittent exposures. Toxicol Appl Pharmacol, 72: 82-90.

    Muzio LJ & Kramlich JC (1988) An artifact in the measurement of N2O
    from combustion sources. Geophys Res Lett, 15: 1369-1372.

    Nadziejko CE, Nansen L, Mannix RC, Kleinman MT, & Phalen RF (1992) The
    effect of nitric acid vapor on the response to inhaled ozone. Inhal
    Toxicol, 4: 343-358.

    Nakajima T & Kusumoto S (1968) [Effect of nitrogen dioxide exposure on
    the contents of reduced glutathione in mouse lung.] Osaka-furitsu
    Koshu Eisei Kenkyusho Kenkyu Hokoku Rodo Eisei Hen, 6: 17-21 (in
    Japanese).

    Nakajima T, Hattori S, Tateisni R, & Horai T (1972) [Morphological
    changes in the bronchial alveolar system of mice following continuous
    exposure to low concentrations of nitrogen dioxide and carbon
    monoxide.] Nihon Kyobu Shikkan Gakkai Zasshi, 10: 16-22 (in Japanese).

    Nakajima T, Oda H, Kusumoto S, & Nogami H (1980) Biological effects of
    nitrogen dioxide and nitric oxide. In: Lee SD ed. Nitrogen oxides and
    their effects on health. Ann Arbor, Michigan, Ann Arbor Science
    Publishers Inc., pp 121-141.

    Nakaki T, Nakayama M, & Kato R (1990) Inhibition by nitric oxide and
    nitric oxide-producing vasodilators of DNA synthesis in vascular
    smooth muscle cells. Eur J Pharmacol, 189: 347-53.

    Namiesnik J, Gorecki T, Kozlowski E, Torres L, & Mathieu J (1984)
    Passive dosimeters - an approach to atmospheric pollutants analysis.
    Sci Total Environ, 38: 225-258.

    NASA (1983) Assessment of techniques for measuring tropospheric
    NxOy: Proceedings of a Workshop. Hampton, Virginia, National
    Aeronautics and Space Administration, Langley Research Center (NASA
    Conference Publication NASA-CP-2292).

    Näsholm T, Högberg P, & Edvast AB (1991) Uptake of NOx by mycorrhizal
    and non-mycorrhizal Scots pine seedlings: quantities and effects on
    amino acid and protein concentrations. New Phytol, 119: 83-92.

    Nathan C (1992) Nitric oxide as a secretory product of mammalian
    cells. FASEB J, 6: 3051-3064.

    National Acid Precipitation Assessment Program (1990) Atmospheric
    processes research and process model development. In: Acidic
    deposition: State of science and technology - Volume I: Emissions,
    atmospheric processes and deposition. Washington, DC, National Acid
    Precipitation Assessment Program (NAPAP Report No. 2).

    National Institutes of Health (1991) Guidelines for the diagnosis and
    management of asthma. Bethesda, Maryland, National Institute of
    Health, National Heart, Lung, and Blood Institute, National Asthma
    Education Program (Publication No. 91-3/042).

    National Research Council (1971) Guides for short-term exposures of
    the public to air pollutants. I. Guide for oxides of nitrogen.
    Washington, DC, National Academy of Sciences.

    National Research Council (1977) Nitrogen oxides. Washington, DC,
    National Academy of Sciences.

    National Research Council (1978) Nitrates: an environmental
    assessment. Washington, DC, National Academy of Sciences.

    National Research Council (1983) Acid deposition: atmospheric
    processes in eastern North America, a review of current scientific
    understanding. Washington, DC, National Academy Press.

    National Research Council (1986) Environmental tobacco smoke:
    measuring exposures and assessing health effects. Washington, DC,
    National Academy Press, pp 9-11, 130, 208, 223-249, 284-288.

    Neas LM, Ware JH, Dockery DW, Spengler JD, Ferris BG Jr, & Speizer FE
    (1990) The association of indoor nitrogen dioxide levels with
    respiratory symptoms and pulmonary function in children. In: Indoor
    air '90 - Proceedings of the 5th International Conference on Indoor
    Air Quality and Climate. Ottawa, Canada, International Conference on
    Indoor Air Quality and Climate Inc., vol 1, pp 381-386.

    Neas LM, Dockery DW, Ware JH, Spengler JD, Speizer FE, & Ferris BG Jr
    (1991) Association of indoor nitrogen dioxide with respiratory
    symptoms and pulmonary function in children. Am J Epidemiol,
    134: 204-219.

    Neas LM, Dockery DW, Spengler JD, Speizer FE, & Ferris BG Jr (1992)
    Variations in the association between indoor nitrogen dioxide and
    childhood respiratory symptoms by sampling location, season and
    source. Am Rev Respir Dis, 145: A93.

    Neighbour EA, Cottam DA, & Mansfield TA (1988) Effects of sulphur
    dioxide and nitrogen dioxide on the control of water loss by birch
    (Betula spp.) New Phytol, 108: 149-157.

    Nguya T, Brunson D, Crespi CL, Penman BW, Wishnok JS, & Tannenbaum SR
    (1992)  DNA damage and mutation in human cells exposed to nitric oxide
     in vitro.  Proc Natl Acad Sci (USA), 89: 3030-3034.

    Nilsson J (1978) [Heathlands and their management. Communication from
    the Research Group for the Management of Nature Reserve 3.] Lund,
    Sweden, Lund University Department of Ecological Botany (In Swedish).

    Nilsson J ed. (1986) Critical loads for nitrogen and sulfur.
    Copenhagen, Denmark, Nordic Council of Ministers.

    Nilsson J & Grennfelt P ed. (1988) Critical loads for sulphur and
    nitrogen: report from a workshop held at Skokloster, Sweden, 19-24
    March 1988. Copenhagen, Denmark, Nordic Council of Ministers, pp 1-418
    Report No. 1988:15).

    Nitta H & Maeda K (1982) Personal exposure monitoring to nitrogen
    dioxide. Environ Int, 8: 243-248.

    Nixon SW & Pilson MEQ (1983) Nitrogen in estuarine and coastal marine
    ecosystems. In: Carpenter EJ & Capone DG ed. Nitrogen in the marine
    environment. New York, London, Academic Press, Inc., pp 565-648.

    Nohrstedt HO, Wedin M, & Gerhardt K (1988) [Effects of forest
    fertilization on nitrogen-fixing lichens.] Uppsala, Sweden, Institute
    for Forest Improvement (Report No. 4( (in Swedish with English
    summary).

    Norby RJ, Weerasuriya Y, & Hanson PJ (1989) Induction of nitrate
    reductase activity in red spruce needles by NO2 and HNO3 vapor. Can
    J For Res, 19: 889-896.

    Norkus EP, Boyle S, Kuenzig W, & Mergens WJ (1984) Formation of
    N-nitrosomorpholine in mice treated with morpholine and exposed to
    nitrogen dioxide. Carcinogenesis, 5: 549-554.

    Novichkova-Ivanova LN (1971) Soil and aerial algae of polar desert and
    arctic tundra.  In: Wielgolaski FE & Rosswall TH ed. Proceedings of
    the 4th International Meeting on the Biological Productivity of
    Tundra, pp 261-265.

    Noxon JF (1976) Atmospheric nitrogen fixation by lightning. Geophys
    Res Lett, 3: 463-465.

    Noxon JF (1978) Tropospheric NO2. J Geophys Res (Oceans Atmos),
    83: 3051-3057.

    Noxon JF, Norton RB, & Marovich E (1980) NO3 in the troposphere.
    Geophys Res Lett, 7: 125-128.

    Noy D, Lebret E, Boleij J, & Brunekreef B (1984) Integrated NO2
    exposure estimates. In: Berglund B, Lindvall T, & Sundell J ed. Indoor
    air '84 - Proceedings of the 3rd International Conference on Indoor
    Air Quality and Climate. Stockholm, Swedish Council for Building
    Research, vol 4, pp 37-42.

    O'Connor G, Sparrow D, Taylor D, Segal M, & Weiss S (1987) Analysis of
    dose-response curves to methacholine: an approach suitable for
    population studies. Am Rev Respir Dis, 136: 1412-1417.

    Oda H, Kusumoto S, & Nakajima T (1975) Nitrosyl-hemoglobin formation
    in the blood of animals exposed to nitric oxide. Arch Environ Health,
    30: 453-456.

    Oda H, Nogami H, Kusumoto S, Nakajima T, Kurata A, & Imai K (1976)
    [Long-term exposure to nitric oxide in mice.] Taiki Osen Kenkyu,
    11: 150-160 (in Japanese).

    Oda H, Nogami H, Kusumoto S, Nakajima T, & Kurata A (1980a) Lifetime
    exposure to 2.4 ppm nitric oxide in mice. Environ Res, 22: 254-263.

    Oda H, Nogami H, & Nakajima T (1980b) Reaction of hemoglobin with
    nitric oxide and nitrogen dioxide in mice. J Toxicol Environ Health,
    6: 673-678.

    Oda H, Tsubone H, Suzuki A, Ichinose T, & Kubota K (1981) Alterations
    of nitrite and nitrate concentrations in the blood of mice exposed to
    nitrogen dioxide. Environ Res, 25: 294-301.

    Oda H, Nogami H, & Nakajima T (1979) Alteration of hemoglobin reacted
    with nitrogen oxides  in vitro. Proceedings of the Sixth Meeting on
    Study of Toxic Effects, Osaka.  J Toxicol Sci, 4: 299-300.

    Oezkaynak H, Ryan PB, Allen GA, & Turner WA (1982) Indoor air quality
    modeling: compartmental approach with reactive chemistry. Environ Int,
    8: 461-471.

    Oezkaynak H, Ryan PB, Spengler JD, & Laird NM (1986) Bias due to
    misclassification of personal exposures in epidemiologic studies of
    indoor and outdoor air pollution. In: Berglund B, Berglund U, Lindvall
    T, Spengler J, & Sundell J ed. Indoor air quality: Papers from the 3rd
    International Conference on Indoor Air Quality and Climate, August
    1984, Stockholm, Sweden. Environ Int, 12: 389-393.

    Ogren JA, Blumenthal DL, & Vanderpol AH (1977) Oxidant measurements in
    western power plant plumes. Volume I: Technical analysis and Volume
    II: Data. Palo Alto, California, Electric Power Research Institute 
    (Report No. EPRI EA-421).

    Ogston SA, Florey C du V, & Walker CHM (1985) The Tayside infant
    morbidity and mortality study: effect on health of using gas for
    cooking. Br Med J, 290: 957-960.

    Ohenoja E (1988)  Behaviour of mycorrhizal fungi in fertilized
    forests.  Karstenia, 28: 27-30.

    Okano K & Totsuka T (1986) Absorption of NO2 by sunflower plants
    grown at various levels of nitrate. New Phytol, 102: 551-562.

    Okano K, Totsuka T, Fukuzawa T, & Tazaki T (1985b) Growth responses of
    plants of various concentrations of nitrogen dioxide. Environ Pollut,
    38: 361-373.

    Okita T, Morimoto S, Izawa M, & Konno S (1976) Measurement of gaseous
    and particulate nitrates in the atmosphere. Atmos Environ,
    10: 1085-1089.

    Oltmans SJ & Komhyr WD (1986) Surface ozone distributions and
    variations from 1973-1984 measurements at the NOAA geophysical
    monitoring for climatic change baseline observatories. J Geophys Res
    (Atmos), 91: 5229-5236.

    Orehek J, Massari JP, Gayrard P, Grimaud C, & Charpin J (1976) Effect
    of short-term, low-level nitrogen dioxide exposure on bronchial
    sensitivity of asthmatic patients. J Clin Invest, 57: 301-307.

    Orehek J, Grimaldi F, Muls E, Durand JP, Viala A, & Charpin J (1981)
    Reponse bronchique aux allergenes apres exposition controlee au
    dioxide d'azote [Bronchial response to allergens after controlled NO2
    exposure]. Bull Eur Physiopathol Respir, 17: 911-915.

    Oshima H, Friesen M, Brouet I, & Bartsch H (1990)  Nitrotyrosine as a
    new marker for endogenous nitrosation and nitration of proteins.  Food
    Chem Toxicol, 28(9): 647-52.

    Otsu H & Ide G (1975) [Effect of nitrogen dioxide on tumorigenesis
    induced by injection of 4-nitroquinoline-1-oxide.] Taiki Osen Kenkyu,
    9: 702-707 (in Japanese).

    Ott W (1989) Human activity patterns: a review of the literature for
    estimating time spent indoors, outdoors, and in-transit. In: Starks TH
    ed. Proceedings of the Research Planning Conference on Human Activity
    Patterns. Las Vegas, Nevada, US Environmental Protection Agency 
    (EPA-600/4-89/004).

    Overton JH Jr (1984) Physicochemical processes and the formulation of
    dosimetry models. In: Miller FJ & Menzel DB ed. Fundamentals of
    extrapolation modeling of inhaled toxicants: ozone and nitrogen
    dioxide. Washington, DC, Hemisphere Publishing Corporation, pp 93-114.

    Overton JH Jr, Graham RC, & Miller FJ (1987a)  Mathematical
    modeling of ozone absorption in the lower respiratory tract.  In:
    Pharmacokinetics in risk assessment - Volume 8: Drinking water and
    health. Washington, DC, National Academy Press, pp 302-311.

    Overton JH, Graham RC, & Miller FJ (1987b) A model of the regional
    uptake of gaseous pollutants in the lung:  II . The sensitivity of
    ozone uptake in laboratory animal lungs to anatomical and ventilatory
    parameters. Toxicol Appl Pharmacol,  88: 418-432.

    Palmer MS, Exley RW, & Coffin DL (1972) Influence of pollutant gases
    on benzpyrene hydroxylase activity.  Arch Environ Health, 25: 439-442.

    Palmes ED, Gunnison AF, DiMattio J, & Tomczyk C (1976) Personal
    sampler for nitrogen dioxide. Am Ind Hyg Assoc J, 37: 570-577.

    Parker RF, Davis JK, Cassell GH, White H, Dziedzic D, Blalock DK,
    Thorp RB, & Simecka JW (1989) Short-term exposure to nitrogen dioxide
    enhances susceptibility to murine respiratory mycoplasmosis and
    decreases intrapulmonary killing of  Mycoplasma pulmonis.  Am Rev
    Respir Dis, 140: 502-512.

    Parrish DD, Fahey DW, Williams EJ, Liu SC, Trainer M, Murphy PC,
    Albritton DL, & Fehsenfeld FC (1986) Background ozone and
    anthropogenic ozone enhancement at Niwot Ridge, Colorado. J Atmos
    Chem, 4: 63-80.

    Patel JM, Edwards DA, Block ER, & Raizada MK (1988) Effect of nitrogen
    dioxide on surface membrane fluidity and insulin receptor binding of
    pulmonary endothelial cells. Biochem Pharmacol, 37: 1497-1507.

    Pattemore PK, Asher MI, Harrison AC, Mitchell EA, Rea HH, & Stewart AW
    (1990) The interrelationship among bronchial hyperresponsiveness, the
    diagnosis of asthma, and asthma symptoms. Am Rev Respir Dis,
    142: 549-554.

    Pearson K (1904) Report on certain enteric fever inoculation
    statistics. Br Med J, 2: 1243-1246.

    Pearson J & Stewart GR (1993) The deposition of atmospheric ammonia
    and its effects on plants. New Phytol, 125: 283-305.

    Peat JK, Wachinger SL, Toelle BG, & Woolcock AJ (1990) A preliminary
    investigation of the relation of domestic gas appliances and indoor
    nitrogen dioxide levels to bronchial hyperresponsiveness and
    respiratory symptoms in a sample of Australian children. Aust NZ J
    Med, 20(Suppl 1): 516.

    Pegtel DM (1987) Effects of ionic Al in culture solutions on the
    growth of  Arnica montana L. and  Deschampsia flexuosa (L.) Trin.
    Plant Soil, 102: 85-92.

    Pelzer A-M & Thomson ML (1966) Effect of age, sex, stature, and
    smoking habits on human airway conductance. J Appl Physiol,
    21: 469-476.

    Penkett SA (1991) Changing ozone: evidence for a perturbed atmosphere.
    Environ Sci Technol, 25: 631-635.

    Penning de Vries FWT (1982) Crop production in relation to
    availability of nitrogen. In: Penning de Vries FWT & van Laar HH ed.
    Simulation of plant growth and crop production. Wageningen, The
    Netherlands, Centre for Agricultural Publishing and Documentation,
    pp 213-221.

    Pérez-Soba M & Van der Eerden LJ (1993) Nitrogen deposition in needles
    of Scots pine in relation to a gaseous ammonia exposure and a
    15N-labelled ammonium supply to the soil. Plant Soil, 153: 231-242.

    Pérez-Soba M, Stulen I, & Van der Eerden LJM (1990) Effects of NH3
    on the N metabolism of  Pinus Sylvestri.  In:Proceedings of the
    International Congress on Forest Decline Research, Friedrichshafen,
    Germany, October 1989. Karlsruhe, Germany, Centre for Nuclear Research
    (Poster Summary).

    Pérez-Soba M, Van der Eerden LJ, & Stulen I (1994) Combined effects of
    gaseous ammonia and sulphur dioxide on the nitrogen metabolism and
    needles of Scots pine trees. Plant Physiol Biochem, 32(4): 539-546.

    Persson H & Ahlstrom K (1991) The effect of forest liming and
    fertilization on fine-root growth. Water Air Soil Pollut, 54: 365-375.

    Persson MG, Gustafsson LE, Wiklund NP, Moncada S, & Hedqvist P
    (1990) Endogenous nitric oxide as a probable modulator of pulmonary
    circulation and hypoxic pressor response  in vivo.  Acta Physiol
    Scand, 140: 449-457.

    Petreas M, Liu K-S, Chang B-H, Hayward SB, & Sexton K (1988) A survey
    of nitrogen dioxide levels measured inside mobile homes. J Air Pollut
    Control Assoc, 38: 647-651.

    Pietila M, Lahdesmaki P, Pietilainen P, Ferm A, Hytonen J, & Patila A
    (1991) High nitrogen deposition causes changes in amino acid
    concentrations and protein spectra in needles of the Scots pine
     (Pinus sylvestris). Environ Pollut, 72: 103-115.

    Pierotti D & Rasmussen RA (1977) The atmospheric distribution of
    nitrous oxide. J Geophys Res, 82: 5823-5832.

    Pilotto LS & Douglas RM (1993) Indoor low-level nitrogen dioxide
    exposure and respiratory illness in 6 to 11 year olds - Final report.
    Canberra, National Health and Medical Research Council.

    Pilotto LSJ (1994) Indoor nitrogen dioxide exposure and respiratory
    illness in children. Canberra, Australian National University (Thesis
    submitted for the degree of Doctor of Philosophy).

    Pitcairn CER, Fowler D, & Grace J (1991) Changes in species
    composition of semi-natural vegetation associated with the increase in
    atmospheric inputs of nitrogen. Edinburgh, United Kingdom, Nature
    Conservancy Council, Institute of Terrestrial Ecology.

    Pitelka LF & Raynal DJ (1989) Forest decline and acidic deposition.
    Ecology, 70: 2-10.

    Pitts JN Jr (1987) Nitration of gaseous polycyclic aromatic
    hydrocarbons in simulated and ambient urban atmospheres: a source of
    mutagenic nitroarenes. Atmos Environ, 21: 2531-2547.

    Pitts JN Jr, Winer AM, Aschmann SM, Carter WPL, & Atkinson R (1985)
    Experimental protocol for determining hydroxyl radical reaction rate
    constants: estimation of atmospheric reactivity. Research Triangle
    Park, North Carolina, US Environmental Protection Agency
    (EPA-600/3-85-000).

    Placet M, Battye RE, Fehsenfeld FC, & Bassett GW (1991) Emissions
    involved in acidic deposition processes. In: Irving PM ed. Acidic
    deposition: state of science and technology. Volume I: Emissions,
    atmospheric processes, and deposition. Washington, DC, National Acid
    Precipitation Assessment Program (State of Science and Technology
    Report No. 1).

    Platt UF & Perner D (1983) Measurements of atmospheric trace gases
    by long path differential UV/visible absorption spectroscopy. In:
    Killinger DK & Mooradian A ed. Optical and laser remote sensing. New
    York, Berlin, Springer-Verlag, pp 97-105.

    Platt U, Perner D, Schroeder J, Kessler C, & Toennissen A (1981) The
    diurnal variation of NO3. J Geophys Res (Oceans Atmos),
    86: 11965-11970.

    Poizat O & Atkinson GH (1982) Determination of nitrogen dioxide by
    visible photoacoustic spectroscopy. Anal Chem, 54: 1485-1489.

    Ponka A (1991) Asthma and low level air pollution in Helsinki. Arch
    Environ Health, 46: 262-270.

    Port CD, Ketels KV, Coffin DL, & Kane P (1977) A comparative study of
    experimental and spontaneous emphysema. J Toxicol Environ Health,
    2: 589-604.

    Posin C, Clark K, Jones MP, Patterson JV, Buckley RD, & Hackney JD
    (1978) Nitrogen dioxide inhalation and human blood biochemistry. Arch
    Environ Health, 33: 318-324.

    Postlethwait EM & Mustafa MG (1981) Fate of inhaled nitrogen dioxide
    in isolated perfused rat lung. J Toxicol Environ Health, 7: 861-872.

    Press MC & Lee JA (1982) Nitrate reductase activity of  Sphagnum
    species in the South Pennines. New Phytol, 92: 487-494.

    Press MC, Woodin SJ, & Lee JA (1986) The potential importance of an
    increased atmospheric nitrogen supply to the growth of ombrotrophic
     Sphagnum species. New Phytol, 103: 45-55.

    Prinz B (1982) [Damage to forests in the Federal Republic of Germany.]
    Hessen, Germany, Institute for Air Pollution Control for the State of
    Hessen (LIS Report No. 28) (in Germany).

    Purdue LJ & Hauser TR (1980) Review of US Environmental Protection
    Agency NO2 monitoring methodology requirements. In: Lee SD ed
    Nitrogen oxides and their effects on health. Ann Arbor, Michigan, Ann
    Arbor Science Publishers Inc., pp 51-76.

    Purvis MR & Ehrlich R (1963) Effect of atmospheric pollutants on
    susceptibility to respiratory infection: II. effect of nitrogen
    dioxide. J Infect Dis, 113: 72-76.

    Quackenboss JJ, Kanarek MS, Spengler JD, & Letz R (1982) Personal
    monitoring for nitrogen dioxide exposure: methodological
    considerations for a community study. Environ Int, 8: 249-258.

    Quackenboss JJ, Spengler JD, Kanarek MS, Letz R, & Duffy CP (1986)
    Personal exposure to nitrogen dioxide: relationship to indoor/outdoor
    air quality and activity patterns. Environ Sci Technol, 20: 775-783.

    Rao CNR & Bhaskar KR (1969) Spectroscopy of the nitroso group. In:
    Feuer H ed. The chemistry of the nitro and nitroso groups, part 1. New
    York, Interscience Publishers, pp 137-163.

    Rasmusser TR, Kjaergaard SK, Tard V, & Pederson OF (1992) Delayed
    effects of NO2 exposure on alveolar permeability and glutathione
    peroxidase in healthy humans. Am Rev Respir Dis, 146: 654-659.

    Ratcliffe DA (1984) Post-medieval and recent changes in British
    vegetation: the culmination of human influence. New Phytol,
    98: 73-100.

    Raven JA (1988) Acquisition of nitrogen by the shoots of land plants:
    its occurrence and implications for acid-base regulation. New Phytol,
    109: 1-20.

    Rebmann H, Huenges R, Wichmann HE, Malin EM, Huebner HR, Roell A,
    Hoerz G, Hub R, Walter C, Doeller G, & Gerth H-J (1991) [Croup and
    air-pollution: results of a two-year prospective longitudinal study.]
    Zent.bl Hyg Umweltmed, 192: 104-115 (in German).

    Redbo-Torstensson P (1994) The demographic consequence of nitrogen
    fertilization of a population of sundew,  Drosera rotundifolia . Acta
    Bot Neerl, 1994: 175-188.

    Reddy KR & Patrick WH (1984) Nitrogen transformations and loss in
    flooded soils and sediments. Crit Rev Environ Control, 13: 273-309.

    Rehfuess KE (1987) Perceptions on forest diseases in central Europe.
    Forestry, 60: 1-11.

    Rehn T, Svartengren M, Philipson K, & Camner P (1982) [Mucociliary
    transport in the lung and nose after exposure to nitrogen dioxide.]
    Vallingby, Swedish State Power Board (Project KHM Technical Report
    No. 40) (in Swedish).

    Reif DW & Simons RD (1990) Nitric oxide mediates iron release from
    ferritin. Arch Biochem Biophys, 283: 537-541.

    Reiners WA & Olson RK (1984) Effects of canopy components on
    throughfall chemistry: an experimental analysis. Oecologia,
    63: 320-330.

    Research Triangle Institute (1990) An investigation of infiltration
    and indoor air quality. Albany, New York, New York State Energy
    Research and Development Authority (Report No. NYERDA 90-11).

    Richters A & Damji KS (1988) Changes in T-lymphocyte subpopulations
    and natural killer cells following exposure to ambient levels of
    nitrogen dioxide. J Toxicol Environ Health, 25: 247-256.

    Richters A & Damji KS (1990) The relationship between inhalation of
    nitrogen dioxide, the immune system, and progression of a
    spontaneously occurring lymphoma in AKR mice. J Environ Pathol Toxicol
    Oncol, 10: 225-230.

    Richters A & Kuraitis K (1981) Inhalation of NO2 and blood borne
    cancer cell spread to the lungs. Arch Environ Health, 36: 36-39.

    Richters A & Kuraitis K (1983) Air pollutants and the facilitation of
    cancer metastasis. Environ Health Perspect, 52: 165-168.

    Richters A & Richters V (1983) A new relationship between air
    pollutant inhalation and cancer. Arch Environ Health, 38: 69-75.

    Richters A, Richters V, & Alley WP (1985) The mortality rate from lung
    metastases in animals inhaling nitrogen dioxide (NO2). J Surg Oncol,
    28: 63-66.

    Rickman EE Jr & Wright RS (1986) Interference of nitrogenous compounds
    on chemiluminescent measurement of nitrogen dioxide. Research Triangle
    Park, North Carolina, Research Triangle Institute (Report No.
    RTI/3180/24-01F).

    Rickman EE Jr, Green AH, Wright RS, & Sickles JE II (1988) Laboratory
    and field evaluations of extrasensitive sulfur dioxide and nitrogen
    dioxide analyzers for acid deposition monitoring. Research Triangle
    Park, North Carolina, Research Triangle Institute, pp 3-8 (Report No.
    RTI/3999/18-02F).

    Rietjens IMCM, Poelen MCM, Hempenius RA, Gijbels MJJ, & Alink GM
    (1986) Toxicity of ozone and nitrogen dioxide to alveolar macrophages:
    comparative study revealing differences in their mechanism of toxic
    action. J Toxicol Environ Health, 19: 555-568.

    Ritter G (1990) [On the effect of nitrogen deposition on root system
    and mycorrhiza formation in pine stocks.] Beitr Forstwirtschaft,
    24: 100-104 (in German).

    Ritter G & Tölle H (1978) [Nitrogen fertilizer use in pine stocks and
    its effect on mycorrhiza formation and fructification in symbiotic
    fungi.] Beitr Forstwirtsch, 4: 162-166 (in German).

    Roberts JM (1990) The atmospheric chemistry of organic nitrates. Atmos
    Environ, A24: 243-287.

    Roberts JM, Norton RB, Goldan PD, & Fehsenfeld FC (1987) Evaluation
    of the tungsten oxide denuder tube technique as a method for the
    measurement of low concentrations of nitric acid in the troposphere.
    J Atmos Chem, 5: 217-238.

    Robertson A, Dodgson J, Collings P, & Seaton A (1984) Exposure to
    oxides of nitrogen: respiratory symptoms and lung function in British
    coalminers. Br J Ind Med, 41: 214-219.

    Robinson JP (1977) How Americans use their time: a social
    psychological analysis of everyday behavior. New York, Praeger
    Publishers.

    Rodgers MO & Davis DD (1989) A UV-photofragmentation/laser-induced
    fluorescence sensor for the atmospheric detection of HONO. Environ Sci
    Technol, 23: 1106-1112.

    Rodgers MO, Asai K, & Davis DD (1980) Photofragmentation-laser induced
    fluorescence: a new method for detecting atmospheric trace gases. Appl
    Opt, 19: 3597-3605.

    Rodhe H (1990) A comparison of the contribution of various gases to
    the greenhouse effect. Science, 248: 1217-1219.

    Roelofs JGM (1983) Impact of acidification and eutrophication on
    macrophyte communities in soft waters in The Netherlands: I. Field
    observations. Aquat Bot, 17: 139-155.

    Roelofs JGM (1986) The effect of airborne sulphur and nitrogen
    deposition on aquatic and terrestrial heathland vegetation.
    Experientia (Basel), 42: 372-377.

    Roelofs JGM, Schuurkes JAAR, & Smits AJM (1984) Impact of
    acidification and eutrophication on macrophyte communities in soft
    waters in the Netherlands: II. Experimental studies. Aquat Bot,
    18: 389-411.

    Roelofs JGM, Kempers AJ, Houdijk ALFM, & Jansen J (1985) The effect of
    air-borne ammonium sulphate on  Pinus nigra var.  maritime in the
    Netherlands. Plant Soil, 84: 45-56.

    Roelofs JGM, Boxman AW, & Van Dijk HFG (1987) Effects of airborne
    ammonium on natural vegetation and forests. In: Asman WAH & Diederen
    HSMA ed. Ammonia and acidification: Proceedings of a Symposium of the
    European Association for the Science of Air Pollution (EURASAP).
    Bilthoven, The Netherlands, European Association for the Science of
    Air Pollution, pp 266-276.

    Roger LJ, Horstman DH, McDonnell W, Kehrl H, Ives PJ, Seal E, Chapman
    R, & Massaro E (1990) Pulmonary function, airway responsiveness, and
    respiratory symptoms in asthmatics following exercise in NO2. Toxicol
    Ind Health, 6: 155-171.

    Rokaw SN, Detels R, Coulson AH, Sayre JW, Tashkin DP, Allwright SS, &
    Massey FJ Jr (1980) The UCLA population studies of chronic obstructive
    respiratory disease: 3. Comparison of pulmonary function in three
    communities exposed to photochemical oxidants, multiple primary
    pollutants, or minimal pollutants. Chest, 78: 252-262.

    Rombout PJA, Dormans JAMA, Marra M, & Van Esch GJ (1986) Influence of
    exposure regimen on nitrogen dioxide-induced morphological changes in
    the rat lung. Environ Res, 41: 466-480.

    Rose RM, Fuglestad JM, Skornik WA, Hammer SM, Wolfthal SF, Beck BD, &
    Brain JD (1988) The pathophysiology of enhanced susceptibility to
    murine cytomegalovirus respiratory infection during short-term
    exposure to 5 ppm nitrogen dioxide. Am Rev Respir Dis, 137: 912-917.

    Rosen K (1988) Effects of biomass accumulation and forestry on
    nitrogen in forest ecosystems. In: Nilsson J & Grennfelt P ed.
    Critical loads for sulphur and nitrogen: Report from a workshop.
    Copenhagen, Denmark, Nordic Council of Ministers, pp 269-293.

    Rosen K, Gundersen P, Tegnhammar L, & Johansson M (1992) Nitrogen
    enrichment of Nordic forest ecosystems: The concept of critical loads.
    Ambio, 21: 364-368.

    Rosenberg R, Elmgren R, Fleischer S, Jonsson P, Persson G, & Dahlin H
    (1990) Marine eutrophication case studies in Sweden. Ambio,
    19: 102-108.

    Rosenthal R (1979) The `file drawer problem' and tolerance for null
    results. Psychol Bull, 86: 639-641.

    Rossi OVJ, Kinnula VL, Tienari J, & Huhti E (1993) Association of
    severe asthma attacks with weather, pollen and air pollutants. Thorax,
    48: 244-248.

    Rowland AJ, Drew MC, & Wellburn AR (1987) Foliar entry and
    incorporation of atmospheric nitrogen oxide into barley plants of
    different nitrogen status. New Phytol, 107: 357-371.

    Roy-Burman P, Pattengale PK, & Sherwin RP (1982) Effect of low levels
    of nitrogen dioxide inhalation on endogenous retrovirus gene
    expression. Exp Mol Pathol, 36: 144-155.

    Roze F (1988) Nitrogen cycle in Brittany heathland. Acta Oecol Oecol
    Plant, 9: 371-379.

    Rubinstein I, Bigby BG, Reiss TF, & Boushey HA Jr (1990) Short-term
    exposure to 0.3 ppm nitrogen dioxide does not potentiate airway
    responsiveness to sulfur dioxide in asthmatic subjects. Am Rev Respir
    Dis, 141: 381-385.

    Rubinstein I, Reiss TF, Bigby BG, Stites DP, & Bousley HA Jr
    (1991) Effects of 0.60 ppm nitrogen dioxide on circulating and
    bronchoalveolar lavage lymphocyte phenotypes in healthy subjects.
    Environ Res, 55: 18-30.

    Rudd JWM, Kelly CA, Schindler DW, & Turner MA (1988) Disruption of the
    nitrogen cycle in acidified lakes. Science, 240: 1515-1517.

    Ruhling A & Tyler D (1991) Effects of simulated nitrogen deposition
    to the forest floor on macrofungal flora of a beech forest. Ambio,
    20: 261-263.

    Runeckles VC & Palmer K (1987) Pretreatment with nitrogen dioxide
    modifies plant response to ozone -short communication. Atmos Environ,
    21(3): 717-719.

    Russell AG, McRae GJ, & Cass GR (1985) The dynamics of nitric acid
    production and the fate of nitrogen oxides. Atmos Environ,
    19: 893-903.

    Rutishauser M, Ackermann U, Braun Ch, Gnehm HP, & Wanner HU (1990a)
    Significant association between outdoor NO2 and respiratory symptoms
    in preschool children. In: Matthys H ed. Eighth Congress of the
    European Society of Pneumology, European Pediatric Respiratory
    Society; September 1989; University of Freiburg, Federal Republic of
    Germany. Lung, 168(suppl): 347-352.

    Rutishauser M, Ackermann U, Braun Ch, Gnehm HP, & Wanner U (1990b)
    [Association between airway symptoms in young children and NO2
    concentrations in the outside air.] Pneumologie, 44: 245-246.

    Ryan PB, Spengler JD, & Letz R (1983) The effects of kerosene heaters
    on indoor pollutant concentrations: a monitoring and modeling study.
    Atmos Environ, 17: 1339-1345.

    Ryan PB, Soczek ML, Treitman RD, Spengler JD, & Billick IH (1988) The
    Boston residential NO2 characterization study: II. Survey methodology
    and population concentration estimates. Atmos Environ, 22: 2115-2125.

    Rydberg L, Edler L, Floderus S, & Graneli W (1990) Interaction between
    supply of nutrients, primary production, sedimentation and oxygen
    consumption in SE Kattegat. Ambio, 19: 134-141.

    Sabarathnam S, Gupta G, & Mulchi C (1988a) Effects of nitrogen dioxide
    on leaf chlorophyll and nitrogen content of soybean. Environ Pollu,
    51: 113-120.

    Sabarathnam S, Gupta G, & Mulchi C (1988b) Nitrogen dioxide effects on
    photosynthesis in soyabean. J Environ Qual, 17: 143-146.

    Sackner MA, Dougherty RD, Chapman GA, Zarzecki S, Zarzemski L, &
    Schreck R (1979) Effects of sodium nitrate aerosol on cardiopulmonary
    function of dogs, sheep, and man. Environ Res, 18: 421-436.

    Sackner MA, Broudy M, Friden A, & Cohn MA (1980) Effects of breathing
    low levels of nitrogen dioxide for four hours on pulmonary function of
    normal adults. Am Rev Respir Dis, 121: 254S.

    Sagai M, Ichinose T, & Kubota K (1984) Studies on the biochemical
    effects of nitrogen dioxide. IV. Relation between the change of lipid
    peroxidation and the antioxidative protective system in rat lungs upon
    life span exposure to low levels of NO2. Toxicol Appl Pharmacol,
    73: 444-456.

    Sagai M, Arakawa K, Ichinose T, & Shimojo N (1987) Biochemical
    effects on combined gases of nitrogen dioxide and ozone: I. Species
    differences of lipid peroxides and phospholipids in lungs. Toxicology,
    46: 251-265.

    Samet JM (1978) A historical and epidemiologic perspective on
    respiratory symptoms questionnaires. Am J Epidemiol, 108: 435-446.

    Samet JM & Spengler JD (1989) Nitrogen dioxide and respiratory
    infection: pilot investigations. Cambridge, Massachusetts, Institute
    of Health Effects (Research Report No. 28).

    Samet JM & Utell MJ (1990) The risk of nitrogen dioxide: what have we
    learned from epidemiological and clinical studies? Toxicol Ind Health,
    6: 247-262.

    Samet JM, Tager IB, & Speizer FE (1983) The relationship between
    respiratory illness in childhood and chronic air-flow obstruction in
    adulthood. Am Rev Respir Dis, 127: 508-523.

    Samet JM, Marbury MC, & Spengler JD (1987) Health effects and sources
    of indoor air pollution. Part I. Am Rev Respir Dis, 136: 1486-1508.

    Samet JM, Marbury MC, & Spengler JD (1988) Health effects and sources
    of indoor air pollution. Part II. Am Rev Respir Dis, 137: 221-242.

    Samet JM, Lambert WE, Skipper BJ, Cushing AH, McLaren LC, Schwab M, &
    Spengler JD (1992) A study of respiratory illnesses in infants and
    nitrogen dioxide exposure. Arch Environ Health, 47: 57-63.

    Samet JM, Lambert WE, Skipper BJ, Cushing AH, Hunt WC, Young SA,
    McLaren LC, Schwab M, & Spengler JD (1993) Health outcomes. In:
    Nitrogen dioxide and respiratory illness in children, part I.
    Cambridge, Massachusetts, Institute of Health Effects, pp 1-32
    (Research Report No. 58).

    Sandhu R & Gupta G (1989) Effects of nitrogen dioxide on growth and
    yield of black turtle bean  (Phaseolus vulagris L.) cv Domino.
    Environ Pollut, 59: 337-344.

    Sandstroem T, Kolmodin-Hedman B, Stjernberg N, & Andersson MC (1989)
    Inflammatory cell response in bronchoalveolar fluid after nitrogen
    dioxide exposure of healthy subjects. Am Rev Respir Dis, 139(suppl):
    A124.

    Sandstroem T, Andersson MC, Kolmodin-Hedman B, Stjernberg N, &
    Angstrom T (1990a) Bronchoalveolar mastocytosis and lymphocytosis
    after nitrogen dioxide exposure in man: a time-kinetic study. Eur
    Respir J, 3: 138-143.

    Sandstroem T, Bjermer L, Kolmodin-Hedman B, & Stjernberg N (1990b)
    Nitrogen dioxide (NO2) induced inflammation in the lung; attenuated
    response after repeated exposures. Am Rev Respir Dis, 141(suppl): A73.

    Saul RL & Archer MC (1983) Nitrate formation in rats exposed to
    nitrogen dioxide. Toxicol Appl Pharmacol, 67: 284-291.

    Savant NK & De Datta SK (1982) Nitrogen transformations in wetland
    rice soils. Adv Agron, 35: 241-302.

    Saxe H (1986a) Effects of NO2 and CO2 on net photosynthesis, dark
    respiration and transpiration of potplants.  New Phytol, 103: 185-197.

    Saxe H (1986b) Stomatal-dependent and stomatal-independent uptake of
    NOx. New Phytol. 103: 199-205.

    Saxe H (1994) Relative sensitivity of greenhouse pot plants to long
    term exposures of NO and NO2-containing air. Environ Pollut,
    85: 283-290.

    Saxe H & Voight Christensen O (1984) Effects of carbon dioxide with
    and without nitric oxide pollution on growth, morphogenesis and
    production time of potted plants. Acta Hortic, 162: 179-186.

    Schaefer DA, Driscoll CT Jr, Van Dreason R, & Yatsko CP (1990) The
    episodic acidification of Adirondack lakes during snowmelt. Water
    Resour Res, 26: 1639-1647.

    Schafer DW (1987) Covariate measurement error in generalized linear
    models. Biometrika, 74: 385-391.

    Schaminee JHJ, Westhoff V, & Arts GHP (1992) [The "shore plant"
    populations (Littorelletea Br.-Bl. et Tx. 43) of the Netherlands put
    in a European frame.] Phytocoenologia, 20: 529-558 (in German).

    Schenk M & Wehrman J (1979) The influence of ammonia in nutrient
    solution on growth and metabolism of cucumber plants. Plant Soil,
    52: 1287-1297.

    Schenker MB, Samet JM, & Speizer FE (1983) Risk factors for childhood
    respiratory disease: the effect of host factors and home environmental
    exposures. Am Rev Respir Dis, 128: 1038-1043.

    Schiff HI, Hastie DR, Mackay GI, Iguchi T, & Ridley BA (1983) Tunable
    diode laser systems for measuring trace gases in tropospheric air: a
    discussion of their use and the sampling and calibration procedures
    for NO, NO2, and HNO3. Environ Sci Technol, 17: 352A-364A.

    Schiff HI, Mackay GI, Castledine C, Harris GW, & Tran Q (1986) A
    sensitive direct measurement NO2 instrument. In: Proceedings of the
    1986 EPA/APCA Symposium on Measurement of Toxic Air Pollutants.
    Pittsburgh, Pennsylvania, Air Pollution Control Association,
    pp 834-844 (EPA-600/9-86-013).

    Schimel DS, Simkins S, Rosswall T, Mosier AR, & Parton WJ (1988)
    Scale and the measurement of nitrogen-gas fluxes from terrestrial
    ecosystems. In: Rosswall T, Woodmansee RG, & Risser PG ed. Scales and
    global change. New York, John Wiley & Sons, pp 179-193.

    Schlechte G (1986) [Concerning the mycorrhizas in damaged forest
    stocks.] Z Mikol, 52: 225-232 (in German).

    Schlesinger RB (1987a) Effects of intermittent inhalation exposures to
    mixed atmospheres of NO2 and H2SO4 on rabbit alveolar macrophages.
    J Toxicol Environ Health, 22: 301-312.

    Schlesinger RB (1987b) Intermittent inhalation of nitrogen dioxide:
    effects on rabbit alveolar macrophages. J Toxicol Environ Health,
    21: 127-139.

    Schlesinger WH (1991) Biogeochemistry an analysis of global change.
    London, New York, San Diego, Academic Press, Inc.

    Schlesinger RB & Gearhart JM (1987) Intermittent exposures to mixed
    atmospheres of nitrogen dioxide and sulfuric acid: effect on particle
    clearance from the respiratory region of rabbit lungs. Toxicology,
    44: 309-319.

    Schlesinger RB, Driscoll KE, & Vollmuth TA (1987) Effect of repeated
    exposures to nitrogen dioxide and sulfuric acid mist alone or in
    combination on mucociliary clearance from the lungs of rabbits.
    Environ Res, 44: 294-301.

    Schlesinger RB, Driscoll KE, Gunnison AF, & Zelikoff JT (1990)
    Pulmonary arachidonic acid metabolism following acute exposures to
    ozone and nitrogen dioxide. J Toxicol Environ Health, 31: 275-290.

    Schlesinger RB, Weideman PA, & Zelikoff JT (1991) Effects of repeated
    exposure to ozone and nitrogen dioxide on respiratory tract
    prostanoids. Inhal Toxicol, 3: 27-36.

    Schneider T & Bresser AHM (1988) [Evaluation raport: acidification.]
    Bilthoven, The Netherlands, National Institute of Public Health and
    Environmental Protection, pp 1-190 (in Dutch).

    Schoof-van Pelt MM (1973) Littorelletea, a study of the vegetation of
    some amphiphytic communities of western Europe. Nijmegen, The
    Netherlands, Catholic University of Nijmegen (Ph.D. Thesis).

    Schulze ED & Freer-Smith PH (1991) An evaluation on forest decline
    based on field observations focussed on Norway spruce,  Picea abies.
    Proc R Soc Edinb, B97: 155-168.

    Schulze ED, Lange OL, & Oren R (1989a) Forest decline and air
    pollution. In: Volume 77: Ecological studies. New York, Berlin,
    Springer-Verlag.

    Schulze ED, De Vries W, Hauhs M, Rosen K, Rasmussen L, Tamm C-O, &
    Nilsson J (1989b) Critical loads for nitrogen deposition on forest
    ecosystems. Water Air Soil Pollut, 48: 451-456.

    Schulze ED, Oren R, & Lange OL (1989c) Nutrient relations of trees in
    healthy and declining Norway spruce stands. In: Volume 77: Ecological
    studies. New York, Berlin, Springer-Verlag, pp 392-417.

    Schuurkes JAAR, Kok CJ, & Den Hartog C (1986) Ammonium and nitrate
    uptake by aquatic plants from poorly buffered and acidified waters.
    Aquat Bot, 24: 131-146.

    Schuurkes JAAR, Elbers MA, Gudden JJF, & Roelofs JGM (1987) Effects of
    simulated ammonium sulphate and sulphuric acid rain on acidification,
    water quality and flora of small-scale soft water systems. Aquat Bot,
    28: 199-225.

    Schwab M, Colome SD, Spengler JD, Ryan PB, & Billick IH (1990)
    Activity patterns applied to pollutant exposure assessment: data from
    a personal monitoring study in Los Angeles. Toxicol Ind Health,
    6: 517-532.

    Schwartz J (1989) Lung function and chronic exposure to air pollution:
    a cross-sectional analysis of NHANES II. Environ Res, 50: 309-321.

    Schwartz J & Zeger S (1990) Passive smoking, air pollution, and acute
    respiratory symptoms in a diary study of student nurses. Am Rev Respir
    Dis, 141: 62-67.

    Schwartz J, Spix C, Wichmann HE, & Malin E (1991) Air pollution and
    acute respiratory illness in five German communities. Environ Res,
    56: 1-14.

    Sega K & Fugas M (1991) Different approaches to the assessment of
    human exposure to nitrogen dioxide. J Expo Anal Environ Epidemiol,
    1: 227-234.

    Sekharam KM, Patel JM, Block ER (1991) Plasma membrane-specific
    phospholipase A1, activation by nitrogen dioxide in pulmonary artery
    endothelial cells. Toxicol Appl Pharmacol, 107: 545-554.

    Selgrade MK, Mole ML, Miller FJ, Hatch GE, Gardner DE, & Hu PC (1981)
    Effect of NO2 inhalation and vitamin C deficiency on protein and
    lipid accumulation in the lung. Environ Res, 26: 422-437.

    Selgrade MK, Daniels MJ, & Grose EC (1991) Evaluation of
    immunotoxicity of an urban profile of nitrogen dioxide: acute,
    subchronic, and chronic studies. Inhal Toxicol, 3: 389-403.

    Seto K, Kon M, Kawakami M, Yagishita S, Sugita K, & Shishido M (1975)
    [Influence of nitrogen dioxide inhalation on the formation of protein
    in the lung.] Igaku to Seibutsugaku, 90: 103-106 (in Japanese).

    Sevanian A, Hacker AD, & Elsayed N (1982) Influence of vitamin E and
    nitrogen dioxide on lipid peroxidation in rat lung and liver
    microsomes. Lipids, 17: 269-277.

    Shaffer G & Rönner U (1984) Denitrification in the Baltic proper deep
    water. Deep Sea Res, 31: 197-220.

    Shalamberidze OP & Tsereteli NT (1971) Effect of low concentrations of
    sulfur and nitrogen dioxides on the estrual cycle and reproductive
    functions of experimental animals. Hyg Sanit (USSR), 36: 178-182.

    Shaver GR & Chapinn FAS (1980) Response to fertilization by various
    plant growth forms in an Alskan tundra: nutrient accumulation and
    growth. Ecology, 61: 662-675.

    Sheppard LJ (1994) Causal mechanisms by which sulphate, nitrate and
    acidity influence frost hardiness in red spruce: review and
    hypothesis. New Phytol, 127: 69-82.

    Sheppard LJ, Cape JN & Leith ID (1993) Influence of acidic mist on
    frost hardiness and nutrient concentrations in red spruce seedlings.
    Part I: Exposure of the foliage and rooting environment. New Phytol,
    124: 595-605.

    Sherwin RP & Carlson DA (1973) Protein content of lung lavage
    fluid of guinea pigs exposed to 0.4 ppm nitrogen dioxide: disc-gel
    electrophoresis for amount and types. Arch Environ Health, 27: 90-93.

    Sherwin RP & Layfield LJ (1974) Proteinuria in guinea pigs exposed to
    0.5 ppm nitrogen dioxide. Arch Environ Health, 28: 336-341.

    Sherwin RP & Richters V (1982) Hyperplasia of type 2 pneumocytes
    following 0.34 ppm nitrogen dioxide exposure: quantitation by image
    analysis. Arch Environ Health, 37: 306-315.

    Sherwin RP, Margolick JB, & Azen SP (1973) Hypertrophy of alveolar
    wall cells secondary to an air pollutant: a semi-automated
    quantitation. Arch Environ Health, 26: 297-299.


    Sherwin RP, Shih JC, Lee JD, & Ransom R (1986) Serotonin content of
    the lungs, brains, and blood of mice exposed to 0.45 ppm nitrogen
    dioxide. J Am Coll Toxicol, 5: 583-588.

    Sherwood RL, Lippert WE, Goldstein E, & Tarkington B (1981) Effect of
    ferrous sulfate aerosols and nitrogen dioxide on murine pulmonary
    defense. Arch Environ Health, 36: 130-135.

    Shiraishi F & Bandow H (1985) The genetic effects of the photochemical
    reaction products of propylene plus NO2 on cultured Chinese hamster
    cells exposed  in vitro. J Toxicol Environ Health, 15: 531-538.

    Shoaf CR, Wolpert RL, & Menzel DB (1989) Factors controlling
    nitrosamine formation in the lung:  a unique uptake system. Inhal
    Toxicol, 1: 167-179.

    Shy CM, Kleinbaum DG, & Morgenstern H (1978) The effect of
    misclassification of exposure status in epidemiological studies of
    air pollution health effects. Bull NY Acad Med, 54: 1155-1165.

    Sickles JE II (1987) Sampling and analytical methods development for
    dry deposition monitoring. Research Triangle Park, North Carolina,
    Research Triangle Institute (Report No. RTI/2823/00-15F).

    Sickles JE II (1992) Sampling and analysis for ambient oxides of
    nitrogen and related species. In: Nriagu JO ed. Gaseous pollutants:
    Characterization and cycling. New York, John Wiley & Sons Inc.,
    pp 51-128.

    Sickles JE II & Wright RS (1979) Atmospheric chemistry of selected
    sulfur-containing compounds: outdoor smog chamber study - phase 1.
    Research Triangle Park, North Carolina, US Environmental Protection
    Agency, Environmental Sciences Research Laboratory, pp 45-49
    (EPA-600/7-79-227).

    Sickles JE II, Hodson LL, Rickman EE Jr, Saeger ML, Hardison DL,
    Turner AR, Sokol CK, Estes ED, & Paur RJ (1989) Comparison of
    the annular denuder system and the transition flow reactor for
    measurements of selected dry deposition species. J Air Pollut Control
    Assoc, 39: 1218-1244.

    Sickles JE II, Grohse PM, Hodson LL, Salmons CA, Cox KW, Turner AR, &
    Estes ED (1990) Development of a method for the sampling and analysis
    of sulfur dioxide and nitrogen dioxide from ambient air. Anal Chem,
    62: 338-346.

    Siddiqi MY, Glass ADM, Ruth TJ, & Rufty TW Jr (1990) Studies of the
    uptake of nitrate in barley: I. kinetics of 13NO3- influx. Plant
    Physiol, 93: 1426-1432.

    Simpson JC & Olsen AR (1990) Wet deposition temporal and spatial
    patterns in North America, 1987. Research Triangle Park, North
    Carolina, US Environmental Protection Agency, Atmospheric Research and
    Exposure Assessment Laboratory (EPA-600/4-90-019).

    Sims J & Kjellström T (1991) Biomass fuel and indoor air pollution:
    underlying issues from a social perspective. In: Indoor air pollution
    from biomass fuels. Geneva, World Health Organization, pp 142-161
    (WHO/PEP/92.3B).

    Sinclair TR & Van Houtte RF (1982) Simulative analysis of ammonia
    exchange between the atmosphere and plant communities. Agric Environ,
    7: 237-242.

    Singh HB (1987) Reactive nitrogen in the troposphere: chemistry and
    transport of NOx and PAN. Environ Sci Technol, 21: 320-327.

    Singh HB & Salas LJ (1983) Methodology for the analysis of
    peroxyacetyl nitrate (PAN) in the unpolluted atmosphere. Atmos
    Environ, 17: 1507-1516.

    Singh HB, Salas LJ, Ridley BA, Shetter JD, Donahue NM, Fehsenfeld FC,
    Fahey DW, Parrish DD, Williams EJ, Liu SC, Hubler G, & Murphy PC
    (1985) Relationship between peroxyacetyl nitrate and nitrogen oxides
    in the clean troposphere. Nature (Lond), 318: 347-349.

    Sinn JP & Pell EJ (1984) Impact of repeated nitrogen dioxide exposures
    on composition and yield of potato foliage and tubers. J Am Soc Hortic
    Sci, 109: 481-484.

    Skarby L, Bengtson C, Bostrom CA, Grennfelt P, & Troeng E (1981)
    Uptake of NOx in Scots pine. Silva Fenn, 15: 396-398.

    Skeffington RA & Wilson EJ (1988) Excess nitrogen deposition: issues
    for consideration. Environ Pollut, 54: 159-184.

    Skiba U, Hargreaves KJ, Fowler D, & Smith KH (1992) Fluxes of nitric
    and nitrous oxides from agricultural soils in a cool temperate
    climate. Atmos Environ, A26: 2477-2488.

    Skogen A (1979) Conversion of Norwegian coastal heath landscape
    through development of potential natural vegetation. In: Vegetation
    ecology and creation of new environments. Proceedings of an
    International Symposium. Tokyo, Tokai University Press, pp 196-204.

    Skoogh B-E (1973) Normal airways conductance at different lung
    volumes. Scand J Clin Lab Invest, 31: 429-441.

    Slade R, Highfill JW, Hatch GE (1989) Effects of depletion of ascorbic
    acid or nonprotein sulfhydryls on the acute inhalation toxicity of
    nitrogen dioxide, ozone, and phosgene. Inhal Toxicol, 1: 261-271.

    Slemr F & Seiler W (1984) Field measurements of NO and NO2 emissions
    from fertilized and unfertilized soils. J Atmos Chem, 2: 1-24.

    Smith VH (1982) The nitrogen and phosphorus dependence of algal
    biomass in lakes: an empirical and theoretical analysis. Limnol
    Oceanogr, 27: 1101-1112.

    Smith KR (1986) Biomass combustion and indoor air pollution: the
    bright and dark sides of small is beautiful. Environ Manage, 1986(1).

    Smith KR (1987) Biofuels, air pollution, & health: a global review.
    New York, Plenum Publishing Company (Modern Perspectives in Energy
    series).

    Smith CT, Elston J, & Bunting AH (1971) The effects of cutting and
    fertilizer treatment on the yield and botanical composition of chalk
    turfs. J Brit Grassl Soc, 26: 213-219.

    Smith RG, Bryan RJ, Feldstein M, Levadie B, Miller FA, & Stephens ER
    (1972) Tentative method of analysis for peroxyacetyl nitrate (PAN) in
    the atmosphere (gas chromatographic method). In: Methods of air
    sampling and analysis. Washington, DC, American Public Health
    Association, pp 215-219.

    Smith RA, Alexander RB, & Wolman MG (1987) Water-quality trends in the
    nation's rivers. Science, 235: 1607-1615.

    Smith W, Anderson T, Anderson HA, & Remington PL (1992) Nitrogen
    dioxide and carbon monoxide intoxication in an indoor ice arena -
    Wisconsin, 1992. Morbid Mortal Wkly Rep, 41: 383-385.

    Smullen JT, Taft JL, & Macknis J (1982) Nutrient and sediment loads to
    the tidal Chesapeake Bay system. In: Chesapeake Bay Program technical
    studies: a synthesis. Annapolis, Maryland, US Environmental Protection
    Agency, pp 150-251.

    Snyder SH & Bredt DS (1992) Biological roles of nitric oxide. Sci Am,
    266: 68-77.

    Soederlund R & Svensson BH (1976) The global nitrogen cycle. In:
    Svensson BH & Soederlund R ed. Nitrogen, phosphorus and sulphur -
    global cycles. Ecol Bull, 22: 23-73 (SCOPE Report No. 7).

    Solomon PA, Fall T, Salmon L, Lin P, Vasquez F, & Cass GR (1988)
    Acquisition of acid vapor and aerosol concentration data for use in
    dry deposition studies in the South Coast Air Basin: Volume I.
    Pasadena, California, California Institute of Technology,
    Environmental Quality Laboratory (EQL Report No. 25).

    Southern California Gas Company (1986) Residential indoor air quality
    characterization study of nitrogen dioxide. Phase I: Volumes 1, 2 and
    3. Southern California Gas Company.

    Speizer FE, Ferris B Jr, Bishop YMM, & Spengler J (1980) Respiratory
    disease rates and pulmonary function in children associated with NO2
    exposure. Am Rev Respir Dis, 121: 3-10.

    Spengler JD, Ferris BG Jr, Dockery DW, & Speizer FE (1979) Sulfur
    dioxide and nitrogen dioxide levels inside and outside homes and the
    implications on health effects research. Environ Sci Technol,
    13: 1276-1280.

    Spengler JD, Duffy CP, Letz R, Tibbitts TW, & Ferris BG Jr (1983)
    Nitrogen dioxide inside and outside 137 homes and implications for
    ambient air quality standards and health effects research. Environ Sci
    Technol, 17: 164-168.

    Spengler JD, Allen GA, Foster S, Severance P, & Ferris B Jr (1986)
    Sulfuric acid and sulfate aerosol events in two US cities. In: Lee SD,
    Schneider T, Grant LD, & Verkerk PJ ed. Aerosols: Research, risk
    assessment and control strategies:  Proceedings of the Second US-Dutch
    International Symposium. Chelsea, Michigan, Lewis Publishers Inc.,
    pp 107-120.

    Spengler JD, Ryan PB, Schwab M, Colome SD, & Wilson AL (1992) Nitrogen
    dioxide exposure studies. Volume IV: Personal exposure to nitrogen
    dioxide in the Los Angeles basin. Chicago, Illinois, Gas Research
    Institute (Report No. GRI-92/0426).

    Spengler J, Neas L, Nakai S, Dockery D, Speizer F, Ware J, & Raizenne
    M (1993) Respiratory symptoms and housing characteristics. In:
    Jaakkola JJK, Ilmarinen R, & Seppänen O ed. Indoor air '93 -
    Proceedings of the 6th International Conference on Indoor Air Quality
    and Climate, Helsinki, July 1993. Volume 1: Health Effects,
    pp 165-170.

    Spicer CW (1982) Nitrogen oxide reactions in the urban plume of
    Boston. Science, 215: 1095-1097.

    Spicer CW & Sverdrup GM (1981) Trace nitrogen chemistry during the
    Philadelphia oxidant data enhancement study (1979). Research Triangle
    Park, North Carolina, US Environmental Protection Agency, Office of
    Air Quality Planning and Standards.

    Spicer CW, Ward GF, & Gay BW Jr (1978a) A further evaluation of
    microcoulometry for atmospheric nitric acid monitoring. Anal Lett,
    A11: 85-95.

    Spicer CW, Joseph DW, & Ward GF (1978b) Investigations of nitrogen
    oxides within the plume of an isolated city. New York, Coordinating
    Research Council Inc. (Report No. CRC-APRAC-CAPA-9-77).

    Spicer CW, Sverdrup GM, & Kuhlman MR (1981) Smog chamber studies of
    NOx chemistry in power plant plumes. Atmos Environ, 15: 2353-2365.

    Spicer CW, Howes JE Jr, Bishop TA, Arnold LH, & Stevens RK (1982)
    Nitric acid measurement methods: an intercomparison. Atmos Environ,
    16: 1487-1500.

    Spicer CW, Joseph DW, & Ward GF (1983) Studies of NOx reactions
    and O3 transport in southern California - fall, 1976. Research
    Triangle Park, North Carolina, US Environmental Protection Agency,
    Environmental Sciences Research Laboratory (EPA-600/3-83-026).

    Spicer CW, Coutant RW, & Ward GF (1986) Investigation of nitrogen
    dioxide removal from indoor air [final report (December 1984 -
    September 1986)]. Chicago, Illinois, Gas Research Institute (Report
    No. GRI-86/0303).

    Srivastava HS & Ormrod DP (1986) Effects of nitrogen dioxide and
    nitrate nutrition on nodulation, nitrogenase activity, growth, and
    nitrogen content of bean plants. Plant Physiol, 81: 737-741.

    Srivastava HS, Jolliffe PA, & Runeckless VC (1975) The effects of
    environmental conditions on the inhibition of leaf gas exchange by
    NO2. Can J Bot, 53: 475-482.

    Stacy RW, Seal E Jr, House DE, Green J, Roger LJ, & Raggio L (1983) A
    survey of effects of gaseous and aerosol pollutants on pulmonary
    function of normal males. Arch Environ Health, 38: 104-115.

    Stadler J, Billiar TR, Curran RD, Stuehr DJ, Ochoa JB, & Simmons
    RL (1991) Effect of exogenous and endogenous nitric oxide on
    mitochondrial respiration of rat hepatocytes. Am J Physiol,
    260: C910-C916.

    Stara JF, Dungworth DL, Orthoefer JG, & Tyler WS ed. (1980) Long-term
    effects of air pollutants: in canine species. Cincinnati, Ohio, US
    Environmental Protection Agency, Office of Health and Environmental
    Assessment, Environmental Criteria and Assessment Office
    (EPA-600/8-80-014).

    Stavert DM, Archuleta DC, Holland LM, & Lehnert BE (1986) Nitrogen
    dioxide exposure and development of pulmonary emphysema. J Toxicol
    Environ Health, 17: 249-267.

    Steadman BL, Jones RA, Rector DE, & Siegel J (1966) Effects on
    experimental animals of long-term continuous inhalation of nitrogen
    dioxide. Toxicol Appl Pharmacol, 9: 160-170.

    Stedman DH & Shetter RE (1983) The global budget of atmospheric
    nitrogen species. Adv Environ Sci Technol, 12: 411-454.

    Stefanski LA & Carroll RJ (1985) Covariate measurement error in
    logistic regression. Ann Stat, 13: 1335-1351.

    Stephens ER (1969) The formation, reactions, and properties of
    peroxyacyl nitrates (PANs) in photochemical air pollution. In: Pitts
    JN Jr & Metcalf RL ed. Advances in environmental science and
    technology, Volume 1. New York, Wiley-Interscience, pp 119-146.

    Stephens ER & Price MA (1973) Analysis of an important air pollutant:
    peroxyacetyl nitrate. J Chem Educ, 50: 351-354.

    Stephens RJ, Freeman G, Crane SC, & Furiosi NJ (1971a) Ultrastructural
    changes in the terminal bronchiole of the rat during continuous,
    low-level exposure to nitrogen dioxide. Exp Mol Pathol, 14: 1-19.

    Stephens RJ, Freeman G, & Evans MJ (1971b) Ultrastructural changes in
    connective tissue in lungs of rats exposed to NO2. Arch Intern Med,
    127: 873-883.

    Stephens RJ, Freeman G, & Evans MJ (1972) Early response of lungs to
    low levels of nitrogen dioxide: light and electron microscopy. Arch
    Environ Health, 24: 160-179.

    Stephens RJ, Sloan MF, & Groth DG (1976) Effects of long term, low
    level exposure of NO2 or O3 on Rat lungs. Environ Health Perspect,
    16: 178-179.

    Stephens RJ, Sloan MF, Groth DG, Negi DS, & Lunan KD (1978) Cytologic
    response of postnatal rat lungs to O3 or NO2 exposure. Am J Pathol,
    93: 183-200.

    Stevens MA, Menache MG, Crapo JD, Miller FJ, & Graham JA (1988)
    Pulmonary function in juvenile and young adult rats exposed to
    low-level NO2 with diurnal spikes. J Toxicol Environ Health,
    23: 229-240.

    Stockwell WR & Calvert JG (1983) The mechanism of NO3 and HONO
    formation in the nighttime chemistry of the urban atmosphere. J
    Geophys Res (Oceans Atmos), 88: 6673-6682.

    Stoddard JL & Kellogg JH (1993) Trends and patterns in lake
    acidification in the state of Vermont: evidence from the Long-Term
    Monitoring Project. Water Air Soil Pollut, 67(3/4): 301-317.

    Stoddard JL & Murdoch PS (1991) Catskill Mountains: an overview of
    chronic and episodic acidity in dilute Catskill Mountain streams. In:
    Charles DF ed. Acid deposition and aquatic ecosystems: regional case
    studies of acid deposition. Berlin, Springer-Verlag.

    Summers CF (1978) Production of montane dwarf shrub communities. Ecol
    Studies, 27: 263-276.

    Sutton MA, Pitcairn CER, & Fowler D (1993) The exchange of ammonia
    between the atmosphere and plant communities. Adv Ecol Res,
    24: 301-393.

    Swift MJ, Heal OW, & Anderson JM (1979) Decomposition in terrestrial
    ecosystems. Oxford, London, Edinburgh, Blackwell Scientific
    Publishers.

    Sykes MT & Van Der Maarel E (1991) Spatial and temporal patterns in
    species turnover in the limestone grasslands of Oland, Sweden.
    Abstracts of the 34th IAVS Symposium on Mechanisms in Vegetation
    Dynamics, Eger, Hungary, p 36.

    Suzuki T & Ishikawa K (1965) [Research on the effects of smog on the
    human body: report of the specialized study on prevention of air
    pollution No. 2.] Tokyo, Japan, Research Coordination, Bureau of the
    Science and Technology Agency, pp 199-221 (in Japanese).

    Suzuki T, Ikeda S, Kanoh T, & Mizoguchi I (1986) Decreased
    phagocytosis and superoxide anion production in alveolar macrophages
    of rats exposed to nitrogen dioxide. Arch Environ Contam Toxicol,
    15: 733-739.

    Szalai A ed. (1972) The use of time: daily activities of urban and
    suburban populations in 12 countries. The Hague, The Netherlands,
    Mouton and Company.

    Tabacova S & Balabaeva L (1988) Nitrogen dioxide embryotoxicity and
    lipid peroxidation. 16th Conference of the European Teratology
    Society, 19-22 September 1988, Baveno, Italy. Teratology, 38(2): 29A.

    Tabacova S, Balabaeva L, & Vardev F (1984) Nitrogen dioxide: Maternal
    and fetal effects. In: Abstracts of the 25th Congress of the European
    Society of Toxicology, Budapest, Hungary, 11-14 June 1984, p 40.

    Tabacova S, Nikiforov B, & Balabaeva L (1985) Postnatal effects of
    maternal exposure to nitrogen dioxide. Neurobehav Toxicol Teratol,
    7: 785-789.

    Takahashi Y, Mochitate K, & Miura T (1986) Subacute effects of
    nitrogen dioxide on membrane constituents of lung, liver, and kidney
    of rats. Environ Res, 41: 184-194.

    Takano T & Miyazaki Y (1984) Combined effect of nitrogen dioxide and
    cold stress on the activity of the hepatic cytochrome P-450 system in
    rats. Toxicology, 33: 239-244.

    Tallis JH (1964) Studies on southern Pennine peats: III. The behaviour
    of  Sphagnum.  J Ecol, 52: 345-353.

    Tamm CO (1991) Nitrogen in terrestrial ecosystems: Questions of
    productivity, vegetational changes, and ecosystem stability. New York,
    Berlin, Springer-Verlag.

    Tanner RL, Kelly TJ, Dezaro DA, & Forrest J (1989) A comparison of
    filter, denuder, and real-time chemiluminescence techniques for nitric
    acid determination in ambient air. Atmos Environ, 23: 2213-2222.

    Taylor OC & Eaton FM (1966) Suppression of plant growth by NO2. Plant
    Physiol, 41: 132-135.

    Taylor GE Jr & Pitelka LF (1992) Genetic diversity of plant
    populations and the role of air pollution. In: Barker JR & Tingey DT
    ed. Air pollution effects on biodiversity. New York, Van Nostrand
    Reinhold Company, pp 111-130.

    Taylor HJ, Ashmore MR, & Bell JNB (1987) Air pollution injury to
    vegetation. London, IEHO, 68 pp.

    Tepper JS, Costa DL, Winsett DW, Stevens MA, & Doerfler DL (1993)
    Near-lifetime exposure of the rat to a simulated urban profile of
    nitrogen dioxide: pulmonary function evaluation. Toxicol Appl
    Pharmacol, 20(1): 88-96.

    Termorshuizen AJ (1990) Decline of carpophores of mycorrhizal fungi
    in stands of  Pinus sylvestris.  Wageningen, The Netherlands,
    Agricultural University of Wageningen (PhD. Thesis).

    Termorshuizen AJ & Schaffers AP (1987) Occurrence of carpophores of
    mycorrhizal fungi in selected stands of  Pinus sylvestris in the
    Netherlands in relation to stand vitality and air pollution. Plant
    Soil, 104: 209-217.

    Termorshuizen AJ, Schaffers AP, Ket PC, & Ter Stege EA (1988) The
    significance of nitrogen pollution on the mycorrhizas of  Pinus
     sylvestris.  In: Jansen AE, Dighton J, & Bresser AHML ed.
    Ectomycorrhiza and acid rain.  Proceedings of a Workshop, Berg en Dal,
    The Netherlands.

    Thijsse ThR (1978) Gas chromatographic measurement of nitrous oxide
    and carbon dioxide in air using electron capture detection. Atmos
    Environ, 12: 2001-2003.

    Thijse G & Baass P (1990) `Natural' and  NH3-induced variation in
    epicuticular needle wax morphology of  Pseudotsuga menziessi (Mirb.)
    Franco. Trees, 4, 111-119.

    Thimonier A, Dupouey JL, Bost F, & Becker M (1994) Simultaneous
    eutrophication and acidification of a forest in North-East France. New
    Phytol, 126: 533-539.

    Thomas HV, Mueller PK, & Wright R (1967) Response of rat lung mast
    cells to nitrogen dioxide inhalation. J Air Pollut Control Assoc,
    17: 33-35.

    Thomas HV, Mueller PK, & Lyman RL (1968) Lipoperoxidation of lung
    lipids in rats exposed to nitrogen dioxide. Science, 159: 532-534.

    Thompson CR, Kats G, & Hensel E (1970) Effects of ambient levels of
    NO2 on navel oranges. Environ Sci Technol, 5: 1017-1019.

    Thrasher WH & DeWerth DW (1979) Evaluation of the pollutant emissions
    from gas-fired room heaters. Cleveland, Ohio, American Gas Association
    Laboratories (Research Report No. 1515).

    Tikalsky S, Reisdorf K, Flickinger J, Totzke D, Haywood J, Annen L,
    Kanarek M, Kaarakka P, & Prias E, (1987) Gas range/oven emissions
    impact analysis (Final report - July 1985-December 1987). Chicago,
    Illinois, Gas Research Institute (Report No. GRI-87/0119).

    Tobacco Research Council (1976) Statistics of smoking in the United
    Kingdom, 7th ed. London, Tobacco Research Council.

    Tolbert MA, Rossi MJ, Malhotra R, & Golden DM (1987) Reaction of
    chlorine nitrate with hydrogen chloride and water at Antarctic
    stratospheric temperatures. Science, 238: 1258-1260.

    Tolbert MA, Rossi MJ, & Golden DM (1988a) Antarctic ozone depletion
    chemistry: reactions of N2O5 with H2O and HCl on ice surfaces.
    Science, 240: 1018-1021.

    Tolbert MA, Rossi MJ, & Golden DM (1988b) Heterogeneous interactions
    of chlorine nitrate, hydrogen chloride, and nitric acid with sulfuric
    acid surfaces at stratospheric temperatures. Geophys Res Lett,
    15: 847-850.

    Totten RS & Moran TJ (1961) Cortisone and atypical pulmonary
    "epithelial" hyperplasia: effects of pretreatment with cortisone on
    repair of chemically damaged rabbit lungs. Am J Pathol, 38: 575-586.

    Touraine B, Grignon N, & Grignon C (1988) Charge balance in NO3--fed
    soybean: estimation of K+ and carboxylate recirculation. Plant
    Physiol, 88: 605-612.

    Toyama T, Tsunoda T, Nakaza M, Higashi T, & Nakadate T (1981) [Airway
    response to short-term inhalation of NO2, O3 and their mixture in
    healthy men.] Sangyo Igaku, 23: 285-293 (in Japanese).

    Trainer M, Williams EJ, Parrish DD, Buhr MP, Allwine EJ, Westberg HH,
    Fesenfeld FC, & Liu SC (1987) Models and observations of the impact of
    natural hydrocarbons on rural ozone. Nature (Lond), 329: 705-707.

    Traynor GW, Allen JR, Apte MG, Dillworth JF, Girman JR, Hollowell CD,
    & Koonce JF Jr (1982a) Indoor air pollution from portable
    kerosene-fired space heaters, wood-burning stoves, and wood-burning
    furnaces. In: Proceedings of the Air Pollution Control Association
    Specialty Conference on Residential Wood and Coal Combustion.
    Pittsburgh, Pennsylvania, Air Pollution Control Association,
    pp 253-263.

    Traynor GW, Anthon DW, & Hollowell CD (1982b) Technique for
    determining pollutant emissions from a gas-fired range. Atmos Environ,
    16: 2979-2987.

    Traynor GW, Girman JR, Allen JR, Apte MG, Carruthers AR, Dillworth JF,
    & Martin VM (1983a) Indoor air pollution due to emissions from
    unvented gas-fired space heaters. Presented at the 76th Annual Meeting
    of the Air Pollution Control Association. Pittsburgh, Pennsylvania,
    Air Pollution Control Association (Paper No. 83-9.6).

    Traynor GW, Allen JR, Apte MG, Girman JR, & Hollowell CD (1983b)
    Pollutant emissions from portable kerosene-fired space heaters.
    Environ Sci Technol, 17: 369-371.

    Traynor GW, Apte MG, Carruthers AR, Dillworth JF, Grimsrud DT, &
    Gundel LA (1984a) Indoor air pollution to emissions from wood burning
    stoves. Presented at the 77th Annual Meeting of the Air Pollution
    Control Association. Pittsburgh, Pennsylvania, Air Pollution Control
    Association (Paper No. 84-33.4).

    Traynor GW, Apte MG, Carruthers AR, Dillworth JF, & Grimsrud DT
    (1984b) Pollutant emission rates from unvented infrared and convective
    gas-fired space heaters. Berkeley, California, US Department of
    Energy, Lawrence Berkeley Laboratory (Report No. LBL-18258).

    Traynor GW, Aceti JC, Apte MG, Smith BV, Green LL, Smith-Reiser A,
    Novak KM, & Moses DO (1987) Macromodel for assessing indoor exposures
    to combustion-generated pollutants. In: Seifert B, Esdorn H, Fischer
    M, Rueden H, & Wegner J ed. Indoor air '87 -Proceedings of the 4th
    International Conference on Indoor Air Quality and Climate. Berlin,
    Institute for Water, Soil and Air Hygiene, vol 1, pp 273-277.

    Tsuda H, Kushi A, Yoshida D, & Goto F (1981) Chromosomal aberrations
    and sister-chromatid exchanges induced by gaseous nitrogen dioxide in
    cultured Chinese hamster cells. Mutat Res, 89: 303-309.

    Tuazon EC, Graham RA, Winer AM, Easton RR, Pitts JN Jr, & Hanst PL
    (1978) A kilometer pathlength Fourier-transform infrared system for
    the study of trace pollutants in ambient and synthetic atmospheres.
    Atmos Environ, 12: 865-875.

    Tuazon EC, Atkinson R, Plum CN, Winer AM, & Pitts JN Jr (1983) The
    reaction of gas phase N2O5 with water vapor. Geophys Res Lett,
    10: 953-956.

    Tuazon EC, Carter WPL, Atkinson R, Winer AM, & Pitts JN Jr (1984)
    Atmospheric reactions of  N -nitrosodimethylamine and dimethyl-
    nitramine. Environ Sci Technol, 18: 49-54.

    Tyler G (1987) Probable effects of soil acidification and nitrogen
    deposition on the floristic composition of oak  (Quercus robur L.)
    forests. Flora, 179: 165-170.

    Tyler M (1988) Contribution of atmospheric nitrate deposition to
    nitrate loading in the Chesapeake Bay. Annapolis, Maryland, Department
    of Natural Resources, Chesapeake Bay Research and Monitoring Division
    (Report No. AD-88-7).

    Tyler G, Balsberg Pahlsson AM, Bergkvist B, Falkengren-Grerup U,
    Folkeson L, Nihlgard B, Ruhling A, & Stjernquist I (1992) Chemical and
    biological effects of simulated nitrogen deposition to the ground in a
    Swedish beech forest. Scan J For Res, 7: 515-532.

    Ulrich B (1983a) Interaction of forest canopies with atmospheric
    constituents: SO2, alkali and earth alkali cations and chloride. In:
    Ulrich B & Pankrath J ed. Effects of accumulation of air pollutants in
    forest ecosystems. Dordrecht, The Netherlands, D. Reidel Publishing
    Company, pp 33-45.

    Ulrich B (1983b) Soil acidity and its relations to acid deposition.
    In: Ulrich B & Pankrath J ed. Effects of accumulation of air
    pollutants in forest ecosystems: Proceedings of a Workshop, May 1982.
    Dordrecht, The Netherlands, D. Reidel Publishing Company, pp 127-146.

    UNECE (1988) ECE-Critical Levels Workshop, Bad Harzburg, 14-18 March
    1988. New York, Geneva, United Nations, Economic Commission for
    Europe.

    UNECE (1993) Manual on methodologies for mapping critical
    levels/loads. New York, Geneva, United Nations, Economic Commission
    for Europe.

    UNECE (1994) Proceedings of the Workshop on Critical Levels, Egham,
    United Kingdom, 23-26 March 1993. New York, Geneva, United Nations,
    Economic Commission for Europe.

    Urban NR & Eisenreich SJ (1988) Nitrogen cycling in a forested
    Minnesota bog. Can J Bot, 66: 435-449.

    Uren S (1992) The effects of wet and dry deposited ammonia on
     Calluna vulgaris . South Kensington, Imperial College (Ph.D.
    Thesis).

    US Bureau of the Census (1982) 1980 census of population and housing:
    supplementary report: provisional estimates of social, economic, and
    housing characteristics: states and selected standard metropolitan
    statistical areas. Washington, DC, US Department of Commerce, Bureau
    of the Census (Report No. PHC 80-S1-1).

    US EPA (1978) Diagnosing vegetation injury caused by air pollution.
    Research Triangle Park, North Carolina, US Environmental Protection
    Agency, Office of Air and Waste Management.

    US EPA (1982a) Air quality criteria for oxides of nitrogen. Research
    Triangle Park, North Carolina, US Environmental Protection Agency,
    Office of Health and Environmental Assessment, Environmental Criteria
    and Assessment Office (EPA-600/8-82-026).

    US EPA (1982b) Air quality criteria for particulate matter and sulfur
    oxides. Research Triangle Park, North Carolina, US Environmental
    Protection Agency, Environmental Criteria and Assessment Office,
    volumes I, II and III (EPA-600/8-82-029aF-cF).

    US EPA (1982c) Review of the national ambient air quality standards
    for nitrogen oxides:  Assessment of scientific and technical
    information. Research Triangle Park, North Carolina, US Environmental
    Protection Agency, Office of Air Quality Planning and Standards
    (EPA-450/5-82-002).

    US EPA (1986a) Second addendum to air quality criteria for particulate
    matter and sulfur oxides (1982): assessment of newly available health
    effects information. Research Triangle Park, North Carolina, US
    Environmental Protection Agency, Office of Health and Environmental
    Assessment, Environmental Criteria and Assessment Office
    (EPA-600/8-86-020F).

    US EPA (1986b) Part 58, Ambient air quality surveillance: Appendix A -
    Quality assurance requirements for state and local air monitoring
    stations (SLAMS). Fed Reg, 51: 9595.

    US EPA (1987a) Diagnosis vegetation injury caused by air pollution.
    Research Triangle Park, North Carolina, US Environmental Protection
    Agency, Office of Air and Waste Management.

    US EPA (1987b) National primary and secondary ambient air quality
    standards. Code Fed Reg, 40: sect 50.

    US EPA (1987c) Ambient air monitoring reference and equivalent
    methods. Code Fed Reg, 40: sect 53.

    US EPA (1991) National air quality and emissions trends report, 1989.
    Research Triangle Park, North Carolina, US Environmental Protection
    Agency, Office of Air Quality Planning and Standards (EPA/450/
    4-91/003).

    US EPA (1993) Air quality criteria for oxides of nitrogen. Research
    Triangle Park, North Carolina, US Environmental Protection Agency,
    Office of Health and Environmental Assessment, Environmental Criteria
    and Assessment Office (EPA/600/8-91/049F).

    Utell MJ, Swinburne AJ, Hyde RW, Speers DM, Gibb FR, & Morrow PE
    (1979) Airway reactivity to nitrates in normal and mild asthmatic
    subjects. J Appl Physiol Respir Environ Exercise Physiol, 46: 189-196.

    Utell MJ, Aquilina AT, Hall WJ, Speers DM, Douglas RG Jr, Gibb FR,
    Morrow PE, & Hyde RW (1980) Development of airway reactivity to
    nitrates in subjects with influenza. Am Rev Respir Dis, 121: 233-241.

    Utell MJ, Framphon MW, Roberts NJ, Finkelstein JN, Cox C, & Morrow PE
    (1991) Mechanisms of nitrogen dioxide toxicity in humans. Cambridge,
    Massachusetts, Institute of Health Effects (Research Report No. 43).

    Valiela I & Teal JM (1979) The nitrogen budget of a salt marsh
    ecosystem. Nature (Lond), 280: 652-656.

    Valiela I, Wilson J, Buchsbaum R, Rietsma C, Bryant D, Foreman K, &
    Teal J (1984) Importance of chemical composition of salt marsh litter
    on decay rates and feeding by detritivores. Bull Mar Sci, 35: 261-269.

    Van Breemen N & Van Dijk HFG (1988) Ecosystem effects of atmospheric
    deposition of nitrogen in The Netherlands. In: Dempster JP, Manning
    WJ, & Skeffington RA ed. Excess nitrogen deposition: Papers from the
    workshop, September 1987, Leatherhead, Surrey, United Kingdom. Environ
    Pollut, 54: 249-274.

    Van Breemen N, Burrough PA, Velthorst EJ, Dobben HF van, Wit T de,
    Ridder TB, & Reijnders HFR (1982)  Soil acidification from atmospheric
    ammonium sulphate in forest canopy throughfall. Nature (Lond),
    299: 548-550.

    Van Dam D (1990) Atmospheric deposition and nutrient cycling in chalk
    grassland. Utrecht, The Netherlands, University of Utrecht (Ph.D.
    Thesis).

    Van Dam H & Buskens RFM (1993) Ecology and management of moorland
    pools: balancing acidification and eutrophication. Hydrobiologia,
    265: 225-263.

    Van Dam D, Van Dobben HF, Ter Braak CFJ, & De Wit T (1986) Air
    pollution as a possible cause for the decline of some phanerogamic
    species in The Netherlands. Vegetatio, 65: 47-52.

    Van Dam D, Heil GW, Bobbink R, & Heijne B (1992) Impact of atmospheric
    deposition on nutrient cycling in chalk grassland: throughfall, canopy
    exchange, nitrogen turnover and input/output budgets. Oecologia.

    Van de Geijn SC, Goudriaan J, Van der Eerden LJ, & Rozema J (1993)
    Problems and approaches to integrating the concurrent impacts of
    elevated CO2, temperature, UVB radiation and O3 on crop production.
    In: International crop science I. Madison, Wisconsin, Crop Science of
    America, pp 333-338.

    Van Den Bergh JP (1979) Changes in the composition of mixed
    populations of grassland species. In: Werger MJA ed. The study of
    vegetation. The Hague, Junk Publishing Company, pp 59-80.

    Van der Eerden LJM (1982) Toxicity of ammonia to plants. Agric
    Environ, 7: 223-235.

    Van der Eerden LJ & Duym NJ (1988) An evaluation method for combined
    effects of SO2 and NO2 on vegetation. Environ Pollut, 53(1-4):
    468-470.

    Van der Eerden LJM & Pérez-Soba M (1992) Physiological responses of
     Pinus sylvestris atmospheric ammonia.  Trees, 6: 48-53.

    Van der Eerden LJM, Dueck TA, Elderson J, Van Dobben HF, Berdowski
    JJM, Latuhihin M, & Prins AH (1990) Effects on NH3 and (NH4)2 SO4
    deposition on terrestrial semi-natural vegetation on nutrient-poor
    soils. Report IPO/RIN, pp 124-125.

    Van der Eerden LJ, Dueck TA, Berdowski JJM, Greven H, & Van Dobben HF
    (1991) Influence of NH3 and (NH4)2SO4 on heathland vegetation.
    Acta Bot Neerl, 40: 281-296.

    Van der Eerden LJM, Tonneijck AEG, Jarosz W, Bestboer S, & Dueck TA
    (1994) Influence of nitrogenous air pollutants on carbon dioxide and
    ozone effects on vegetation. In: Jackson M & Black CR ed. Interacting
    stress seen plants in changing climate. Heidelberg, Berlin,
    Springer-Verlag, pp 125-137.

    Van Der Maas MP (1990) Hydrochemistry of two douglas fir ecosystems
    and a heather ecosystem in the Veluwe, The Netherlands. Wageningen,
    The Netherlands, Agricultural University of Wageningen.

    Van der Molen J, Bussink DW, Vertregt N, Van Faassen HG, & Den Boer DJ
    (1989) Ammonia volatilization from arable and grassland soils. In:
    Hansen JA & Henriksen K ed. Nitrogen in organic wastes applied to
    soils. New York, London, Academic Press, Inc., pp 185-201.

    Van Dijk G (1992) The status of semi-natural grasslands in Europe.
    Strasburg, Council of Europe.

    Van Dijk HFG & Roelofs JGM (1988) Effects of excessive ammonium
    deposition on the nutritional status and condition of pine needles.
    Physiol Plant, 73: 494-501.

    Van Dijk HFG, Creemers RCM, Rijniers JPLWN, & Roelofs JGM (1989)
    Impact of artificial ammonium-enriched rainwater on soils and young
    coniferous tree in a greenhouse: Part 1 - Effects on the soils.
    Environ Pollut, 62: 317-336.

    Van Dijk HFG, De Louw MHJ, Roelofs JGM, & Verburgh JJ (1990) Impact of
    artificial ammonium-enriched rainwater on soils and young coniferous
    trees in a greenhouse: Part 2 - Effects on the trees. Environ Pollut,
    63: 41-60.

    Van Dijk HFG, Boxman AW, & Roelofs JGM (1992a) Effects of a decrease
    in atmospheric deposition of nitrogen and sulphur on the mineral
    balance and vitality of a Scots pine stand in the Netherlands. For
    Ecol Manage, 51: 207-215.

    Van Dijk HF, Van der Gaag PJM, & Roelofs JGM (1992b) Nutrient
    availability in Corsican pine stands in the Netherlands and the
    occurrence of  Sphaeropsis sapinea (Fr) Dyko & Sutton; a field study.
    Can J Bot, 70: 870-875.

    Van Dobben HF (1991) Integrated effects (low vegetation). In: Heij GJ
    & Schneider T ed. Acidification research in the Netherlands: Final
    report of the Dutch Priority Programme on Acidification. Amsterdam,
    Oxford, New York, Elsevier Science Publishers, pp 464-524.

    Van Dobben HF (1993) Vegetation as a monitor for deposition of
    nitrogen and acidity. Utrecht, The Netherlands, University of Utrecht
    (Ph.D. Thesis).

    Van Haut H & Prinz B (1979) [Evaluation of relative phytotoxicity of
    organic air pollutants in the LIS short-term test.] Staub Reinhalt
    Luft, 39: 408-413 (in German).

    Van Hecke P, Impens I, & Behaeghe TJ (1981) Temporal variation of
    species composition and species diversity in permanent grassland plots
    with different fertilizer treatments. Vegetatio, 47: 221-232.

    Van Hove LWA, Adema EH, Vredenberg WJ, & Pieters GA (1989) A study of
    the adsorption of NH3 and SO2 on leaf surfaces. Atmos Environ,
    23: 1479-1486.

    Van Kootwijk EJ & Van der Voet H (1989) [The mapping of the
    heatherlands in the Netherlands using the Landsat thematic mapper
    satellite pictures.] Arnhem, The Netherlands, Research Institue for
    Forestry and Nature (Report No. RIN 89/2) (in Dutch).

    Van Stee EW, Sloane RA, Simmons JE, & Brunnemann KD (1983)  In vivo
    formation of  N-nitrosomorpholine in CD-1 mice exposed by inhalation
    to nitrogen dioxide and by gavage to morpholine. J Natl Cancer Inst,
    70: 375-379.

    Van Tooren BF, Oden B, During HJ, & Bobbink R (1990) Regeneration of
    species richness in the bryophyte layer of  Dutch chalk grasslands.
    Lindbergia, 16: 153-160.

    Vaughan TR Jr, Jennelle LF, & Lewis TR (1969) Long-term exposure to
    low levels of air pollutants: effects on pulmonary function in the
    beagle. Arch Environ Health, 19: 45-50.

    Vedal S, Schenker MB, Munoz A, Samet JM, Batterman S, & Speizer FE
    (1987) Daily air pollution effects on children's respiratory symptoms
    and peak expiratory flow. Am J Public Health, 77: 694-698.

    Verhoeven JTA & Schmitz MB (1991) Control of plant growth by nitrogen
    and phosphorus in mesotrophic fens. Biogeochemistry, 12: 135-148.

    Verhoeven JTA, Koerselman W, & Beltman B (1988) The vegetation of fens
    in relation to their hydrology and nutrient dynamics: a case study.
    In: Symoens JJ ed. Vegetation of inland waters. Dordrecht, The
    Netherlands, Kluwer Academic Publishers, pp 249-282.

    Vermeer JG (1986) The effects of nutrients on shoot biomass and
    species composition of wetland and hayfield communities. Acta
    Oecol/Oecol Plant, 7: 31-41.

    Vermeer JG & Berendse F (1983) The relationship between nutrient
    availability, shoot biomass and species richness in grassland and
    wetland communities. Vegetatio, 53: 121-126.

    Victorin K (1994) Review of the genotoxicity of nitrogen oxides. Mutat
    Res, 317: 43-55.

    Victorin K & Stahlberg M (1988) A method for studying the mutagenicity
    of some gaseous compounds in  Salmonella typhimurium.  Environ Mol
    Mutagen, 11: 65-77.

    Victorin K, Busk L, Cederberg H, & Magnusson J (1990) Genotoxic
    activity of 1,3-butadiene and nitrogen dioxide and their photochemical
    reaction products in Drosophila and in mouse bone marrow micronucleus
    assay. Mutat Res, 228: 203-209.

    Vierkorn-Rudolph B, Rudolph J, & Diederich S (1985) Determination of
    peroxyacetyl nitrate (PAN) in unpolluted areas. Int J Environ Anal
    Chem, 20: 131-140.

    Vilkamaa P & Huhta V (1986) Effects of fertilization and pH upon
    communities of Collembola in pine forest soil. Ann Zool Fennici,
    23: 167-174.

    Vöge M (1988) [Studies of underwater vegetation in Scandinavian lakes
    taking into consideration isoetidal vegetation.] Limnologica (Berlin),
    19: 89-107 (in German).

    Vollmuth TA, Driscoll KE, & Schlesinger RB (1986) Changes in early
    alveolar particle clearance due to single and repeated nitrogen
    dioxide exposures in the rabbit. J Toxicol Environ Health,
    19: 255-266.

    Von Liebig J (1827) Une note sur la nitrification. Ann Chim Phys,
    35: 329-333.

    Von Nieding G & Wagner HM (1977) Experimental studies on the
    short-term effect of air pollutants on pulmonary function in man:
    two-hour exposure to NO2, O3 and SO2 alone and in combination.
    In: Kasuga S, Suzuki N, Yamada T, Kimura G, Inagaki K, & Onoe K ed.
    Proceedings of the Fourth International Clean Air Congress. Tokyo,
    Japan, Japanese Union of Air Pollution Prevention Associations,
    pp 5-8.

    Von Nieding G & Wagner HM (1979) Effects of NO2 on chronic
    bronchitis. Environ Health Perspect, 29: 137-142.

    Von Nieding G, Wagner HM, Krekeler H, Smidt U, & Muysers K (1970)
    Absorption of NO2 in low concentrations in the respiratory tract and
    its acute effects on lung function and circulation. Presented at the
    Second International Clean Air Congress, Washington, DC (Paper No.
    MB-15G).

    Von Nieding G, Wagner M, Krekeler H, Smidt U, & Muysers K (1971)
    [Minimum concentrations of NO2 causing acute effects on the
    respiratory gas exchange and airway resistance in patients with
    chronic bronchitis.] Int Arch Arbeitsmed, 27: 338-348 (in German).

    Von Nieding G, Krekeler H, Fuchs R, Wagner M, & Koppenhagen K (1973a)
    Studies of the acute effects of NO2 on lung function: influence on
    diffusion, perfusion and ventilation in the lungs. Int Arch
    Arbeitsmed, 31: 61-72.

    Von Nieding G, Wagner HM, & Krekeler H (1973b) Investigation of the
    acute effects of nitrogen monoxide on lung function in man. In:
    Proceedings of the Third International Clean Air Congress, October,
    Dusseldorf, Federal Republic of Germany. Dusseldorf, Society of German
    Engineers, pp A14-A16.

    Von Nieding G, Wagner M, Loellgen H, & Krekeler H (1977) [The acute
    effect of ozone on the pulmonary function of man.] VDI Ber,
    270: 123-129 (in German).

    Von Nieding G, Wagner HM, Krekeler H, Loellgen H, Fries W, & Beuthan A
    (1979) Controlled studies of human exposure to single and combined
    action of NO2, O3, and SO2. Int Arch Occup Environ Health,
    43: 195-210.

    Von Nieding G, Wagner HM, Casper H, Beuthan A, & Smidt U (1980) Effect
    of experimental and occupational exposure to NO2 in sensitive and
    normal subjects. In: Lee SD ed. Nitrogen oxides and their effects on
    health. Ann Arbor, Michigan, Ann Arbor Science Publishers, Inc.,
    pp 315-331.

    Vossler TL, Stevens RK, Paur RJ, Baumgardner RE, Bell JP (1988) 
    Evaluation of improved inlets and annular denuder systems to measure
    inorganic air pollutants. Atmos Environ, 22: 1729-1736.

    Vukovich FM, Bach WD Jr, Crissman BW, & King WJ (1977) On the
    relationship between high ozone in the rural surface layer and high
    pressure systems. Atmos Environ, 11: 967-983.

    Wade WA III, Cote WA, & Yocom JE (1975) A study of indoor air quality.
    J Air Pollut Control Assoc, 25: 933-939.

    Wagner H-M (1970) [Absorption of NO and NO2 in mik- and mak-
    concentrations during inhalation.] Staub Reinhalt Luft, 30: 380-381
    (in German).

    Wagner WD, Duncan BR, Wright PG, & Stokinger HE (1965) Experimental
    study of threshold limit of NO2. Arch Environ Health, 10: 455-466.

    Walega JG, Stedman DH, Shetter RE, Mackay GI, Iguchi T, & Schiff HI
    (1984) Comparison of a chemiluminescent and a tunable diode laser
    absorption technique for the measurement of nitrogen oxide, nitrogen
    dioxide, and nitric acid. Environ Sci Technol, 18: 823-826.

    Walker DA & Crofts AR (1970) Photosynthesis. Annu Rev Biochem,
    39: 389-428.

    Walker AM & Blettner M (1985) Comparing imperfect measures of exposure.
    Am J Epidemiol, 121: 783-790.

    Wallace LA & Ott WR (1982) Personal monitors: A state of the art
    survey. J Air Pollut Control Assoc, 32: 601-610.

    Walles SA, Victorin K, & Lundborg M (1995) DNA damage in lung cells
     in vivo and  in vitro by 1,3-butadiene and nitrogen dioxide and
    their photochemical reaction products. Mutat Res, 328: 11-19.

    Ware JH, Dockery DW, Spiro A III, Speizer FE, & Ferris BG Jr (1984)
    Passive smoking, gas cooking, and respiratory health of children
    living in six cities. Am Rev Respir Dis, 129: 366-374.

    Waring RH (1987) Nitrate pollution: a particular danger to boreal
    and subalpine coniferous forests. In: Fujimori T & Kimura M ed.
    Human impacts and management of mountain forests: Proceedings of a
    Symposium. Ibaraki, Japan, Forestry and Forest Products Research
    Institute, pp 93-105.

    Warneck P (1988) Chemistry of the natural atmosphere. New York,
    London, Academic Press, Inc.

    Warren R (1994) Determination of nitrogen dioxide emission of gas
    appliances. Adelaide, South Australia, South Australian Gas Company
    Ltd (Laboratory report).

    Wasterlund I (1982) [Do pine mycorrhizal fungi disappear following
    fertilizer treatment?] Svensk Bot Tidskr, 76: 411-417 (in Swedish).

    Watanabe H, Fukase O, & Isomura K (1980) Combined effects of nitrogen
    oxides and ozone on mice.  In Lee SD ed. Nitrogen oxides and their
    effects on health. Ann Arbor, Michigan, Ann Arbor Science Publishers
    Inc., pp 181-189.

    Wayne RP, Barnes I, Biggs P, Burrows JP, Canosa-Mas CE, Hjorth J, Le
    Bras G, Moortgat GK, Perner D, Poulet G, Restelli G, & Sidebottom H
    (1991) The nitrate radical: physics, chemistry, and the atmosphere.
    Atmos Environ, A25: 1-203.

    Weast RC, Astle MJ, & Beyer WH ed. (1986) CRC handbook of chemistry
    and physics: a ready-reference book of chemical and physical data,
    67th ed. Boca Raton, Florida, CRC Press Inc., pp B/111-B/-112.

    Weinberg ED (1992) Iron depletion: A defense against intracellular
    infection and neoplasia. Life Sci, 50: 1289-1297.

    Weiss RF & Craig H (1976) Production of atmospheric nitrous oxide by
    combustion. Geophys Res Lett, 3: 751-753.

    Wellburn A (1988) Air pollution and acid rain. White Plains, New York,
    Longman Publishing Company.

    Wellburn AR (1990) Why are atmospheric oxides of nitrogen usually
    phytotoxic and not alternative fertilizers?  Tansley Review 24. New
    Phytol, 115: 395-429.

    Wellburn AR, Majernik O, & Wellburn FAN (1972) Effects of SO2 and
    NO2 polluted air upon the ultrastructure of chloroplasts. Environ
    Pollut, 3: 37049.

    Wellburn AR, Wilson J, & Aldridge PH (1980) Biochemical responses on
    nitric oxide polluted atmospheres. Environ Pollut, A22: 219-228.

    Wellburn AR, Higginson C, Robinson D, & Walmsey C (1981) Biochemical
    explanation of more than additive inhibitory low atmospheric levels of
    SO2 + NO2 upon plants. New Phytol, 88: 223-237.

    Wells TCE (1974) Some concepts of grassland management. In: Duffey E
    ed. Grassland ecology and wildlife management. London, Chapman & Hall,
    pp 163-174.

    Wells TCE, Sparks TH, Cox R, & Frost A (1993) Critical loads for
    nitrogen assessment and effects on southern heathlands and grasslands:
    Report to National Power. Huntingdon, UK, Institute of Terrestrial
    Ecology, Monks Wood Experimental Station.

    Wendel GJ, Stedman DH, Cantrell CA, & Damrauer L (1983) Luminol-based
    nitrogen dioxide detector. Anal Chem, 55: 937-940.

    Westberg H, Sexton K, & Roberts E (1981) Transport of pollutants along
    the western shore of Lake Michigan. J Air Pollut Control Assoc,
    31: 385-388.

    Wetselaar R & Farquhar GD (1980) Nitrogen losses from tops of plants.
    Adv Agron, 33: 263-302.

    Wheeler BD & Giller KE (1982) Species richness of herbaceous fen
    vegetation in Broadland, Norfolk in relation to the quantity of
    above-ground plant material. J Ecol, 70: 179-200.

    White WH (1977) NOx-O3 photochemistry in power plant plumes:
    comparison of theory with observation. Environ Sci Technol,
    11: 995-1000.

    Whitmore M (1985) Relationship between dose of SO2 and NO2 mixtures
    and growth of  Poa pratensis . New Phytol, 99: 545-553.

    Whitmore M & Freer-Smith PH (1982) Growth effects of SO2 and/or NO2
    and/or NO2 on woody plants and grasses during spring and summer.
    Nature (Lond), 300(5887): 55-57.

    Whittemore AS & Keller JB (1988) Approximations for regression with
    covariate measurement error. J Am Stat Assoc, 83: 1057-1066.

    Whitmore ME & Mansfield TA (1983) Effects of long term exposures to
    SO2 and NO2 on  Poa pratensi and other grasses. Environ Pollut,
    A31: 217-235.

    WHO (1977) Environmental health criteria 4: Oxides of nitrogen.
    Geneva, World Health Organization.

    WHO (1984) Biomass fuel combustion and health. Geneva, World Health
    Organization (EFP/84.64).

    WHO (1987) The effects of nitrogen on vegetation. In: Air quality
    guidelines for Europe. Copenhagen, Denmark: World Health Organization,
    Regional Office for Europe, pp 373-385 (WHO Regional Publications,
    European Series No. 23).

    WHO (1988) Assessment of urban air quality. Geneva, World Health
    Organization and United Nations Environment Programme, Global
    Environment Monitoring System (GEMS Report).

    WHO (1992) Indoor air pollution from biomass fuel. Report of a WHO
    Consultation, June 1991. Geneva, World Health Organization
    (WHO/PEP/92.3A).

    WHO/UNEP (World Health Organization/United Nations Environment
    Programme) (1992) Urban air pollution in megacities of the world.
    Oxford, United Kingdom, Blackwell Publishers.

    Wigington PJ, Davies TD, Tranter M, & Eshleman K (1989) Episodic
    acidification of surface waters due to acidic deposition. Washington,
    DC, National Acid Precipitation Assessment Program (State-of-Science
    Technology Report No. 12).

    Willems JH (1980) An experimental approach to the study of species
    diversity and above-ground biomass in chalk grassland. Proc Koninkl
    Nederl Akad Wetensch, C83: 279-306.

    Willems JH (1982) Phytosociological and geographical survey of
    Mesobromion communities in Western Europe.  Vegetatio, 48: 227-240.

    Willems JH, Peek RK, & Bik L (1993) Changes in chalk-grassland
    structure and species richness resulting from selective nutrient
    additions. J Veg Sci, 4: 203-212.

    Willett W (1989) An overview of issues related to the correction of
    non-differential exposure measurement error in epidemiologic studies.
    Stat Med, 8: 1031-1040, 1071-1073.

    Williams ED (1978) Botanical composition of the Park Grass plots at
    Rothamsted 1856-1976. Harpenden, United Kingdom, Rothamsted
    Experimental Station (Internal Report).

    Willis AJ (1963) Braunton Burrows: The effects on the vegetation of
    the addition of mineral nutrients to the dune soils. J Ecol,
    51: 353-374.

    Winer AM, Peters JW, Smith JP, & Pitts JN Jr (1974) Response of
    commercial chemiluminescent NO-NO2 analyzers to other
    nitrogen-containing compounds. Environ Sci Technol, 8: 1118-1121.

    Winer AM, Atkinson R, & Pitts JN Jr (1984) Gaseous nitrate radical:
    possible nighttime atmospheric sink for biogenic organic compounds.
    Science, 224: 156-159.

    Winer AM, Atkinson R, Arey J, Biermann HW, Harger WP, Tuazon EC, &
    Zielinska B (1987) The role of nitrogenous pollutants in the formation
    of atmospheric mutagens and acid deposition. Sacramento, California,
    California Air Resources Board (Report No. ARB-R-87/308).

    Wingsle G, Näsholm T, Lundmark T, & Ericsson A (1987) Induction of
    nitrate reductase in needles of Scots pine seedling by NOx and NO3.
    Physiol Plant, 70: 399-403.

    Wink DA, Kasprzak KS, Maragos CM, Elespuru RK, Misra M, & Dunams TM
    (1991) DNA deaminating ability and genotoxicity of nitric oxide and
    its progenitors. Science, 254: 1001-1003.

    Wisheu IC & Keddy PA (1989) The conservation and management of a
    threatened coastal plain plant community in eastern North America
    (Nova Scotia, Canada). Biol Conserv, 48: 229-238.

    Witschi H (1988) Ozone, nitrogen dioxide and lung cancer: a review of
    some recent issues and problems. Toxicology, 48: 1-20.

    Wittig R (1982) [Spread of littorelletea species in the Bay of
    Westphalia.] Decheniana (Bonn), 135: 14-21 (in German).

    WMO (1988) BAPMoN data for 1983 - Volume II: Precipitation chemistry,
    continuous atmospheric carbon dioxide and suspended particulate
    matter. Geneva, World Meteorological Organization, Environmental
    Pollution Monitoring and Research Programme (Report No. 54).

    WMO (1989) WMO BAPMoN data for 1984 and 1985 - Volume II:
    Precipitation chemistry, continuous atmospheric CO2 and suspended
    particulate matter. Geneva, World Meteorological Organization,
    Environmental Pollution Monitoring and Research Programme (Report
    No. 60).

    WMO (1991) Scientific assessment of ozone depletion: 1991. Geneva,
    World Meteorological Organization, Global Ozone Research Monitoring
    Project (Report No. 25).

    Wolfenden J, Pearson M, & Francis BJ (1991) Effects of over-winter
    fumigation with sulphur and nitrogen dioxides on biochemical
    parameters and spring growth in red spruce  (Picea rubensi Sarg.).
    Plant Cell Environ,  14: 35-45.

    Wolff GT (1984) On the nature of nitrate in coarse continental
    aerosols. Atmos Environ, 18: 977-981.

    Wolff EW, Mulvaney R, & Oates K (1989) Diffusion and location of
    hydrochloric acid in ice: implications for polar stratospheric clouds
    and ozone depletion. Geophys Res Lett, 16: 487-490.

    Wolkinger F & Plank S (1981) Dry grasslands of Europe. Strasbourg,
    Council of Europe.

    Wollenheber B & Raven JA (1993) Implications of N acquisition from
    atmospheric NH3 for acid-base and cation-anion balance in
     Lolium perenne . Physiol Plant, 89: 519-523.

    Woodin SJ & Farmer AM (1993) Impacts of sulphur and nitrogen
    deposition on sites and species of nature conservation importance in
    Great Britain. Biol Conserv, 63: 23-30.

    Woodin SJ & Lee JA (1987) The fate of some components of acidic
    deposition in ombrotrophic mires. Environ Pollut, 45: 61-72.

    Woodin SJ, Press MC, & Lee JA (1985) Nitrate reductase activity in
     Sphagnum fuscum in relation to wet deposition of nitrate from the
    atmosphere. New Phytol, 99: 381-388.

    Wood T & Bormann FH (1975) Increases in foliar leaching caused by
    acidification of an artificial mist. Ambio, 4: 169-171.

    Woods JE (1983) Sources of indoor air contaminants. ASHRAE Trans,
    89: 462-497.

    Woodwell GM (1970) Effects of pollution on the structure and
    physiology of ecosystems: changes in natural ecosystems caused by many
    different types of disturbances are similar and predictable. Science,
    168: 429-433.

    Wuebbles DJ (1989) On the mitigation of non-CO2 greenhouse gases.
    Washington, DC, US Department of Energy (Report No. UCRL-101523).

    Wuebbles DJ, Grant KE, Connell PS, & Penner JE (1989) The role of
    atmospheric chemistry in climate change. J Am Pollut Control Assoc,
    39: 22-28.

    Wulff F, Stigebrandt A, & Rahm L (1990) Nutrient dynamics of the
    Baltic Sea. Ambio, 19: 126-133.

    Yamamoto I & Takahashi M (1984) Ultrastructural observations of rat
    lung exposed to nitrogen dioxide for 7 months. Kitasato Arch Exp Med,
    57: 57-65.

    Yamanaka S (1984) Decay rates of nitrogen oxides in a typical Japanese
    living room. Environ Sci Technol, 18: 566-570.

    Yamanaka S, Hirose H, & Takada S (1979) Nitrogen oxides emissions from
    domestic kerosene-fired and gas-fired appliances. Atmos Environ,
    13: 407-412.

    Yanagisawa Y & Nishimura H (1982) A badge-type personal sampler for
    measurement of personal exposure to NO2 and NO in ambient air.
    Environ Int, 8: 235-242.

    Yanagisawa Y, Matsuki H, Osaka F, Kasuga H, & Nishimura H (1984)
    Annual variation of personal exposure to nitrogen dioxide. In:
    Berglund B, Lindvall T, & Sundell J ed. Indoor air '84 - Proceedings
    of the 3rd International Conference on Indoor Air Quality and Climate.
    Stockholm, Swedish Council for Building Research, vol 4, pp 33-36.

    Yanagisawa Y, Nishimura H, Matsuki H, Osaka F, & Kasuga H (1986)
    Personal exposure and health effect relationship for NO2 with urinary
    hydroxyproline to creatinine ratio as indicator. Arch Environ Health,
    41: 41-48.

    Yang YS, Skelly JM, & Chevone BI (1983) Effects of pollutant
    combinations at low doses on growth of forest trees. Aquilo Ser Bot,
    19: 406-418.

    Yockey CC, Eden BM, & Byrd RB (1980) The McConnell missile accident:
    clinical spectrum of nitrogen dioxide exposure. J Am Med Assoc,
    244: 1221-1223.

    Yokoyama E (1968) Uptake of SO2 and NO2 by the isolated upper
    airways. Bull Inst Public Health (Tokyo), 17: 302-306.

    Yokoyama E, Ichikawa I, & Kawai K (1980) Does nitrogen dioxide modify
    the respiratory effects of ozone? In: Lee SD ed. Nitrogen oxides and
    their effects on health. Ann Arbor, Michigan, Ann Arbor Science
    Publishers Inc., pp 217-229.

    Yoneyama T, Saskawa H, Ishizuka S, & Totsuka T (1979) Absorption of
    atmospheric NO2 by plants and soil.  II.  Nitrite accumulation,
    nitrite reductase activity and diuralchange of NO2 absorption in
    leaves. Soil Sci Plant Nutr, 25: 267-276.

    Yoshida K & Kasama K (1987) Biotransformation of nitric oxide. Environ
    Health Perspect, 73: 201-206.

    Yoshida K, Kasama K, Kitabatake M, Okuda M, & Imai M (1980) Metabolic
    fate of nitric oxide. Int Arch Occup Environ Health, 46: 71-77.

    Yoshida K, Kasama K, Kitabatake M, Wakabayashi K, & Imai M (1981)
    Changing of nitric oxide in the airway: experiment with model-airway
    and perfused lung. Rep Environ Sci, 6: 57-61.

    Yoshimura I (1990) The effect of measurement error on the dose-response
    curve. Environ Health Perspect, 87: 173-178.

    Yusuf S, Peto R, Lewis J, Collins R, & Sleight P (1985) Beta blockade
    during and after myocardial infarction: an overview of the randomized
    trials. Prog Cardiovasc Dis, 27: 335-371.

    Zafiriou OC & McFarland M (1981) Nitric oxide from nitrite photolysis
    in the central equatorial Pacific. J Geophys Res (Oceans Atmos),
    86: 3173-3182.

    Zapol WM, Rimar S, Gillis N, Marletta M, & Bosken C (1994) Nitric
    oxide and the lung. Am J Respir Crit Care Med, 149: 1375-1380.

    Zawacki TS, Cole JT, Huang VMS, Banasiuk H, & Macriss RA (1984)
    Efficiency and emissions improvement of gas-fired space heaters. Task
    2. Unvented space heater emission reduction (Final report). Chicago,
    Illinois, Gas Research Institute (Report No. GRI-84/0021).

    Zawacki TS, Cole JT, Jasionowski WJ, & Macriss RA (1986) Measurement
    of emission rates from gas-fired space heaters (Final report for IGT
    Project No. 30570-13). Chicago, Illinois, Institute of Gas Technology.
    Zeger SL & Liang K-Y (1986) Longitudinal data analysis for discrete
    and continuous outcomes. Biometrics, 42: 121-130.

    Zemba SG, Golomb D, & Fay JA (1988) Wet sulfate and nitrate deposition
    patterns in eastern North America. Atmos Environ, 22: 2751-2761.

    Zierock KH, Mansfield T, Postumus A, Guderian R, Lee J, & De Leeuw F
    (1986)  Studies on the need of a NO2 long term limit value for the
    protection of terrestrial and aquatic ecosystems: CEC Final report.
    Luxembourg, Commission of the European Communities (EUR 10-546-EN).

    Zimmernann FK (1977) Genetic effects of nitrous acid. Mutat Res,
    39: 127-148.

    RESUME

    1. Oxydes d'azote et composés apparentés

         Les oxydes d'azote peuvent être présents en quantités importantes
    dans l'air ambiant et dans l'air intérieur. La nature et la
    concentration des dérivés azotés dépendent largement du lieu, de
    l'heure et de la saison. Les émissions d'oxydes d'azote sont
    principalement imputables aux processus de combustion. Les centrales
    thermiques à combustibles fossiles, les véhicules à moteur et les
    appareils et ustensiles ménagers qui font appel à la combustion sont
    des sources d'oxydes d'azote, émis principalement sous la forme
    d'oxyde nitrique (NO) et, pour une moindre part (en général moins
    de 10%), de dioxyde d'azote (NO2). Des réactions chimiques qui se
    produisent dans l'air conduisent à l'oxydation du NO en NO2 et autres
    composés. Il existe également des processus biologiques qui provoquent
    la libération de dérivés azotés par le sol, notamment de l'oxyde
    nitreux (N2O). Les émissions de N2O peuvent nuire à la couche
    d'ozone stratosphérique.

         Il peut y avoir atteinte à la santé humaine en présence de
    concentrations importantes de NO2 ou d'autres espèces azotées comme
    le nitrate de peroxyacyle (PAN), l'acide nitrique (HNO3), l'acide
    nitreux (HNO2) et certains autres dérivés nitrés. En outre, les
    nitrates et l'acide nitrique peuvent, lorsqu'ils se déposent sur le
    sol, avoir des effets nocifs sur la santé et sur les écosystèmes.

         On désigne généralement par NOx l'ensemble NO + NO2.  Une fois
    libéré dans l'air, NO est oxydé en NO2 par les oxydants présents (en
    particulier l'ozone, O3). Dans certaines conditions, la réaction est
    rapide dans l'air extérieur; à l'intérieur, le processus est en
    général beaucoup plus lent. Les oxydes d'azote sont des précurseurs
    déterminants de la pollution atmosphérique par les oxydants
    photochimiques, qui débouche sur la formation d'ozone et de smog;
    l'interaction entre les oxydes d'azote (sauf N2O) et certaines
    espèces organiques réactives conduisent, sous l'action du rayonnement
    solaire, à la formation d'ozone dans la troposphère et de smog dans
    les zones urbaines.

         NO et NO2 peuvent également subir des réactions conduisant à la
    formation, dans l'air extérieur ou intérieur, d'autres oxydes et
    dérivés oxygénés de l'azote, notamment HNO3, HNO2, NO3 (trioxyde
    d'azote), N2O5 (pentoxyde de diazote), du nitrate de peroxyacyle et
    d'autres nitrates organiques. L'ensemble des oxydes d'azote présents
    dans ce mélange gazeux complexe est désigné par NOy. La proportion
    des oxydes d'azote dans ce mélange dépend fortement de la concentration
    des autres oxydants et des antécédents météorologiques. 

         L'acide nitrique HNO3 se forme par réaction de OH- sur NO2.
    C'est un important piège à azote actif et il entre également dans la
    composition des dépôts acides. Parmi les pièges physiques et chimiques
    potentiels à HNO3, on peut citer les dépôts humides et secs, la
    photolyse, la réaction avec les radicaux OH ainsi que la réaction avec
    l'ammoniac gazeux qui conduit à la formation d'aérosols de nitrate
    d'ammonium. 

         Les nitrates de peroxyacyle se forment par la réaction de
    radicaux peroxy organiques sur NO2. Le nitrate de peroxyacyle est le
    nitrate organique le plus abondant dans la troposphère et il peut
    servir de réservoir temporaire d'azote actif susceptible de
    déplacements régionaux.

         Le radical NO3, une espèce de type NOy à courte vie, qui se
    forme dans la troposphère, principalement par réaction de O3 sur
    NO2, subit une photolyse rapide à la lumière du jour ou réagit sur
    NO. Sa concentration est appréciable pendant la nuit.

         N2O5 est principalement un constituant nocturne de l'air
    ambiant qui se forme par réaction de NO3 sur NO2. Dans l'air
    ambiant, N2O5 réagit en milieu hétérogène avec l'eau pour donner de
    l'acide nitrique qui se dépose à son tour.

         N2O est un composé ubiquitaire car il résulte de processus
    naturels qui se déroulent dans le sol. Toutefois, il n'est pas, autant
    qu'on sache, impliqué dans des réactions au sein de la troposphère.
    Dans la haute atmosphère, N2O participe à des réactions qui
    contribuent à la réduction de la couche d'ozone stratosphérique et
    c'est également un gaz à puissant effet de serre qui intervient dans
    le réchauffement général du climat.

    1.1  Transport atmosphérique

         Le transport et la dispersion des diverses espèces azotées
    dans la basse atmosphère dépend de paramètres chimiques et
    météorologiques. Des processus tels que l'advection, la diffusion et
    les transformations chimiques se combinent pour déterminer la durée de
    leur séjour dans l'atmosphère.  La durée de séjour dans l'atmosphère
    détermine à son tour l'ampleur du déplacement de tel ou tel
    composé.  Les émissions de surface se dispersent verticalement et
    horizontalement sous l'action de processus turbulents qui dépendent
    largement du gradient vertical de température et de la vitesse du
    vent.

         Par suite des processus météorologiques, les NOx émis en ville
    dans les premières heures de la matinée subissent une dispersion
    verticale caractéristique et se déplacent avec le vent au fil de la
    journée. Pendant les journées ensoleillées d'été, la majorité des NOx

    auront été transformés en HNO3 et en PAN lorsqu'arrivera le
    crépuscule, avec formation concomitante d'ozone. L'acide nitrique
    s'évacue en grande partie par dépôt lors du déplacement de la masse
    d'air, mais le HNO3 et le PAN entraînés avec les couches supérieures
    (au-dessus de de la couche d'inversion nocturne mais au-dessous de
    l'inversion de subsidence à altitude plus élevée) peuvent être
    transportés sur de grandes distances par des masses d'air chargées
    d'oxydants.

    1.2  Dosage

         On dispose d'un certain nombre de méthodes pour le dosage des
    dérivés azotés aéroportés. Le présent document donne un aperçu des
    méthodes actuelles généralement utilisées pour la surveillance
     in situ de leur concentration, tant dans le milieu ambiant que dans
    l'air intérieur. Les dérivés envisagés sont NO, NO2, NOx, l'azote
    réactif non usuel total (NOy), le PAN et autres nitrates organiques,
    HNO3, HNO2, N2O5, NO3- et N2O.

         Le dosage des oxydes d'azote n'a rien d'évident.Il existe certes
    une méthode simple, praticable un peu partout, pour le dosage de NO
    (réaction de chimioluminescence avec l'ozone), mais il s'agit là d'une
    exception. La chimioluminescence est également la méthode la plus
    couramment utilisée pour NO2 (que l'on réduit préalablement en NO). 
    Malheureusement, le catalyseur utilisé pour la réduction n'est pas
    spécifique et il est d'une efficacité variable selon l'oxyde d'azote à
    réduire. Dans ces conditions, il faut être très prudent lorsque l'on
    interprète les résultats d'un dosage de NO2 par cette méthode car le
    signal peut correspondre en fait à la superposition des signaux de
    nombreux autres produits. En outre, des difficultés supplémentaires
    peuvent surgir du fait de la répartition des oxydes d'azote entre la
    phase gazeuse et la phase particulaire, tant dans l'atmosphère qu'au
    cours du prélèvement des échantillons.

    1.3  Exposition

         L'exposition humaine et environnementale aux oxydes d'azote varie
    beaucoup selon qu'il s'agit de l'air intérieur ou extérieur, d'une
    zone urbaine ou rurale, ou encore en fonction de l'heure ou de la
    saison. On connaît relativement bien la concentration de NO et de NO2
    dans l'air extérieur qui caractérise certaines situations urbaines. A
    l'intérieur, la concentration de ces composés dépend de la nature
    exacte des appareils domestiques de chauffage ou de cuisson ou encore
    des cheminées et de la ventilation. En cas d'utilisation d'appareils
    de chauffage ou de cuisson à combustion dans des locaux non ventilés,
    la concentration des oxydes d'azote dans l'air intérieur se
    caractérise par des valeurs beaucoup plus élevées qu'à l'extérieur.
    Des travaux récents ont montré que dans ces conditions, la

    concentration de HNO2 peut être élevée. C'est ainsi qu'il a été
    montré que la concentration de HNO2 peut représenter plus de 10% de
    la concentration totale en oxydes d'azote (généralement indiquée en
    NO2).

    2.  Effets des dérivés azotés présents dans l'atmosphère, et notamment
        des oxydes d'azote, sur la végétation

         C'est dans les écosystèmes (semi-)naturels aquatiques et
    terrestres que la biodiversité se manifeste la plupart du temps dans
    sa plénitude. Dans nombre de ces écosystèmes, l'azote est un nutriment
    qui joue le rôle de facteur limitant pour la croissance des végétaux.
    La plupart des espèces végétales qui peuplent ces biotopes sont
    adaptées à un faible apport de nutriments et la compétition avec
    d'autres plantes ne peut leur être favorable que sur des sols pauvres
    en azote.

         L'activité humaine, qu'elle soit agricole ou industrielle, a eu
    pour conséquence d'accroître considérablement la quantité de dérivés
    azotés biodisponibles, perturbant ainsi le cycle naturel de l'azote.
    Les polluants atmosphériques azotés existent sous diverses formes: les
    principales sont NO, NO2 et l'ammoniac (NH3) en dépôt sec; les
    nitrates (NO3-) et les sels d'ammonium (NH4+) en dépôt humide.  Il
    peut également y avoir des dépôts occultes provenant de brouillards ou
    de nébulosités diverses. En fait, les polluants atmosphériques azotés
    sont bien plus nombreux (par exemple, N2O5, le PAN, N2O, des amines
    etc.) Mais nous n'en tiendrons pas compte ici, soit parce qu'ils ne
    contribuent, semble-t-il, que trop peu aux dépôts azotés, soit parce
    que leur concentration est probablement très inférieure au seuil
    d'apparition des effets.

         Les polluants atmosphériques azotés peuvent nuire à la
    végétation, soit indirectement par l'intermédiaire de produits de
    réaction photochimiques, soit directement par dépôt sur les plantes,
    le sol ou l'eau. La voie de contamination indirecte n'est guère
    abordée dans le présent document, encore que les processus qui y sont
    à l'oeuvre soient tout à fait interessants et méritent d'être pris en
    considération lors de l'évaluation de l'impact global des polluants
    atmosphériques azotés: le NO2 est un précurseur de l'ozone
    troposphérique qui agit à la fois comme phytotoxine et comme gaz à
    effet de serre.

         L'impact d'un dépôt accru de dérivés azotés sur les systèmes
    biologiques peut résulter, soit d'une fixation directe de ces produits
    par le feuillage, soit d'un captage au niveau du sol. Pour ce qui
    est de la plante elle-même, les effets les plus significatifs sont
    des lésions tissulaires, une modification de la biomasse et une
    sensibilité accrue aux facteurs secondaires de stress. En ce qui
    concerne la végétation dans son ensemble, l'azote ainsi déposé joue le

    rôle d'un nutriment; il en résulte une modification des conditions de
    compétition entre les différentes espèces et une diminution de la
    biodiversité.  La valeur critique de la charge azotée dépend i) de la
    nature de l'écosystème; ii) de l'exploitation et de l'aménagement
    passés et présents des sols; et iii) des conditions abiotiques du lieu
    (en particulier celles qui influent sur la capacité de nitrification
    et le taux d'immobilisation dans le sol).

         L'adsorption de dérivés azotés à la surface de la feuille peut
    endommager la couche cireuse de la cuticule, mais on n'a pas encore la
    preuve que cela soit quantitativement important sur le terrain.
    L'existence d'un gradient de concentration entre l'atmosphère et le
    mésophylle favorise la fixation des NOx et de l'ammoniac. Cette
    fixation est généralement, mais pas systématiquement, directement liée
    à la conductance des stomates et dépend donc des facteurs qui en
    conditionnent l'ouverture. On est de plus en plus fondé à penser que
    la fixation de l'azote par les feuilles réduit sa fixation par les
    racines. La fixation et l'échange d'ions à la surface de la feuille
    sont des processus relativement lents et qui ne peuvent donc prendre
    de l'ampleur que si la surface foliaire reste humide suffisamment
    longtemps.

         NO n'est que légèrement soluble dans l'eau, mais la présence
    d'autres substances peut en modifier la solubilité. NO2 est davantage
    soluble et NH3 beaucoup plus. NO2- (principal produit de réaction
    des NOx), NH3 et NH4+ sont tous très phytotoxiques et peuvent très
    bien être à l'origine des effets nocifs provoqués par les polluants
    atmosphériques azotés. Le radical libre *N=O peut jouer un rôle dans
    la phytotoxicité de NO.

         Des effets dépassant la simple additivité (synergie) ont été
    observés dans presque toutes les études relatives à SO2 en présence
    de NO2. Dans le cas des autres mélanges contenant NO2, par exemple
    en présence de NO, O3 et CO2, les effets interactifs sont
    l'exception plutôt que la règle.

         Lorsque conditions climatiques et apport d'autres nutriments
    permettent la production de biomasse, les NOx et NHy ont pour effet
    de stimuler la croissance à faible concentration et de la réduire à
    concentration élevée. Toutefois la concentration à partir de laquelle
    la stimulation se change en inhibition est beaucoup plus faible dans
    le cas de NOx que dans celui de NHy.

         On a pu constater que les plantes sont plus sensibles lorsque
    l'intensité lumineuse est faible (par exemple la nuit ou en hiver) et
    la température basse (juste au-dessus de 0°C). NOx et NHy peuvent
    accroître la sensibilité des végétaux au gel, à la sécheresse, au vent
    et aux ravageurs.

         Il existe une corrélation entre la chimie du sol et la
    sensibilité de la végétation aux dépôts de composés azotés; cette
    dernière dépend en effet du pH et de la disponibilité de l'azote.

         La contribution relative de NO et de NO2 aux effets des NOx sur
    les plantes n'est pas connue avec certitude. La très grande majorité
    des données dont on dispose concerne les effets de NO2, mais ce que
    l'on sait de NO incite à penser que NO et NO2 ont une action
    phytotoxique comparable.

         Les valeurs-guides pour la qualité de l'air sont basées sur la
    notion de seuil d'apparition d'effets indésirables. On distingue deux
    types de seuils: les niveaux critiques (CLE) et les charges critiques
    (CLO). Par niveau critique, on entend la concentration d'un polluant
    atmosphérique à partir de laquelle des effets indésirables directs
    peuvent, selon nos connaissances actuelles, se produire sur certains
    recepteurs, qu'il s'agisse de plantes, d'écosystèmes ou de matériaux.
    Par charge critique, on entend la valeur estimative de l'exposition
    (dépôt) à un ou plusieurs polluants au-dessous de laquelle il ne se
    produit pas, autant qu'on sache, d'effets délétères sur les éléments
    sensibles de l'environnement.

         Dans la pratique, on obtient les niveaux critiques en
    déterminant, par une méthode graphique, la concentration la plus
    faible qui provoque un effet indésirable sur les fonctions
    physiologiques ou la croissance des végétaux (en excluant les effets
    biochimiques).

         Pour tenir compte des effets dus à NO, on a proposé un niveau
    critique pour NOx plutôt que pour NO2; à cette fin on a posé que,
    par hypothèse, les effets de NO et ceux de NO2 ne sont pas additifs.
    L'établissement de niveaux critiques pour une exposition de brève
    durée est tout à fait défendable, mais on ne possède pas actuellement
    de données en nombre suffisant pour proposer des valeurs de bonne
    fiabilité.Les résultats dont on dispose conduisent à proposer un
    niveau critique pour NOx d'environ 75 µg/m3 en moyenne sur 24 h. 

         On estime à 30 µg/m3 en moyenne annuelle le niveau critique pour
    NOx (NO et NO2 en parties par milliard exprimés sous forme de NO2
    en µg/m3).

         Les données relatives aux biotes présents dans l'environnement
    concernent presque exclusivement les plantes, avec un minimum
    de renseignements sur la faune terricole. C'est pourquoi les
    valeurs-guides qui sont proposées ici se rapportent aux effets des
    divers dérivés azotés sur la végétation. On pense toutefois que la
    végétation est l'élément le plus fragile des écosystèmes naturels et
    que l'effet sur la biodiversité végétale est un indicateur sensible
    des effets exercés sur l'ensemble des écosystèmes.

         Les charges critiques s'obtiennent à partir de données
    expérimentales et de modèles pédologiques stationnaires. On trouvera
    dans la présente évaluation la valeur estimative de la charge
    critique pour les dépôts azotés sur divers écosystèmes terrrestres et
    aquatiques. On ne connaît pas suffisamment bien les effets imputables
    aux différentes espèces chimiques (NOx et NHy) pour pouvoir
    différencier ces différents composés par rapport à leur charge
    critique.

         Les écosystèmes sur lesquels on possède suffisamment de données
    pour établir des charges critiques sont en grande majorité situés dans
    la zone tempérée.

         Les quelques écosystèmes arctiques ou montagnards qui figurent
    dans ce groupe et dont on pourrait attendre qu'ils soient
    représentatifs de la situation aux latitudes élevées, constituent
    en fait la base de données la moins fiable. On ne sait rien des
    écosystèmes tropicaux et pas grand chose des écosystèmes marins ou
    estuariels, quelle que soit la zone climatique où ils se situent. Les
    écosystèmes tropicaux pauvres en nutriments, comme la forêt ombrophile
    et les mangroves auraient probablement à souffrir des dépôts de
    dérivés azotés. En l'absence de données sur ces dépôts et sur les
    seuils d'apparition des effets, il est impossible de se livrer à une
    évaluation du risque dans ces zones climatiques.

         Dans les écosystèmes les plus fragiles (marais ombrotrophiques,
    lacs aux eaux douces et peu profondes, hauteurs arctiques et alpines)
    où l'on a pu évaluer la charge critique, on a obtenu des valeurs de
    l'ordre de 5-10 kg N. ha-1.année-1. Ces évaluations sont basées sur
    la diminution de la diversité biologique de la végétation. On a obtenu
    la valeur plus moyenne de 15-20 kg N.ha-1 .année-1 pour les quelques
    écosystèmes étudiés, valeur qui s'applique aux arbres des forêts.

         La chimie atmosphérique des oxydes d'azote concerne leur action
    sur la capacité de régénération de l'ozone troposphérique et sur la
    réduction de la couche d'ozone stratosphérique ainsi que leur
    contribution au réchauffement général de la planète par effet de
    serre. Avec l'ammoniac et les oxydes de soufre, ils contribuent
    à l'acidification des sols et augmentent par conséquent la
    biodisponibilité de l'aluminium.

         Lorsque leur concentration ne dépasse que marginalement le niveau
    critique, les oxydes d'azote n'exercent sur les récoltes que des
    effets phytotoxiques négligeables. Il n'empêche que par leur action
    sur la formation d'ozone et d'autres substances phytotoxiques dans la
    troposphère, comme les nitrates organiques par exemple, les NOx
    peuvent causer des dommages aux récoltes. Les dérivés azotés déposés
    sur les plantes en culture ne représentent qu'une partie infime de
    l'azote disponible total, comparativement à l'apport d'azote par les
    engrais.

    3.  Effets sanitaires de l'exposition aux oxydes d'azote

         On a effectué de nombreuses études dans le but d'évaluer les
    effets des NOx sur la santé. L'un de ces composés, NO2, a été
    extrêmement étudié. Dans ce qui suit, on s'attache principalement à
    NO2, NO, HNO2 et HNO3, sans trop s'attarder sur les nitrates.

    3.1  Etudes sur les effets des dérivés azotés chez les animaux de
         laboratoire

         L'extrapolation à l'homme des résultats obtenus chez l'animal
    de laboratoire comporte des aspects qualitatifs et des aspects
    quantitatifs. Comme on l'explique succintement dans ce qui suit, NO2
    exerce toute une gamme d'effets chez plusieurs espèces animales,
    en particulier, il affecte les défenses de l'hôte contre les
    pneumopathies infectieuses et peut modifier la biochimie et le
    métabolisme pulmonaires ainsi que la structure et la fonction de
    l'appareil respiratoire. Du fait de l'existence d'analogies
    structurales et métaboliques chez tous les mammifères, qu'il s'agisse
    de l'homme ou des animaux de laboratoire, le fait de retrouver chez
    plusieurs espèces animales à peu près les mêmes résultats conduit à
    conclure que, selon toute vraisemblance, NO2 produit les mêmes effets
    chez l'homme. Toutefois, en raison des différences qui existent
    malgré tout entre les espèces mammaliennes, on ne peut pas dire avec
    certitude quel effet telle ou telle exposition produirait sur l'homme.
    C'est là le domaine de l'extrapolation quantitative. Les quelques
    recherches qui ont été consacrées à la modélisation dosimétrique de
    l'extrapolation quantitative (c'est-à-dire la détermination de la dose
    au tissu ou à la cellule cibles qui produit effectivement un effet
    toxique), incitent à penser que que le NO2 se répartit de façon
    similaire dans les voies respiratoires de l'homme et des animaux, sans
    toutefois que l'on puisse en tirer des valeurs extrapolables de
    l'animal à l'homme. On ne dispose malheureusement que de très peu de
    données sur un autre aspect fondamental de l'eaxtrapolation, à savoir
    la sensibilité selon l'espèce (c'est-à-dire la réaction des tissus à
    une dose donnée chez les différentes espèces). Ainsi, nous savons,
    grâce à ces études sur l'animal, quels effets NO2 est susceptible de
    produire chez l'homme, mais nous ne sommes pas pour autant en mesure
    de déterminer de manière fiable quels effets telle ou telle dose
    inhalée de NO2 produit  effectivement.

         Compte tenu de ce qui vient d'être dit, on trouvera ci-dessous
    une récapitulation de la base de données toxicologiques relative à
    NO2 par centres d'intérêt et par principaux types d'effets.  Il est
    certain que les effets de NO2 ne sont pas strictement localisés aux
    poumons, mais l'interprétation de ces effets généraux eu égard au
    risque qu'ils représentent pour l'homme, demeure incertaine. Ils ne
    sont donc pas évoqués dans ce qui suit, mais abordés dans les
    chapitres suivants. Les interactions qui peuvent se produire entre
    NO2 et d'autres polluants comme l'ozone ou l'acide sulfurique

    (H2SO4) sont d'une grande importance, en particulier en cas de
    synergie, mais au stade actuel, la base de données ne permet pas de
    tirer des conclusions qui conduiraient à évaluer la possibilité de
    telles interactions en situation réelle.

    3.1.1  Mode d'action des oxydes d'azote au niveau cellulaire et
           biochimique

         NO2 se comporte comme un puissant oxydant. Il oxyde facilement
    les lipides insaturés en donnant principalement naissance à des
    peroxydes. L'acide ascorbique (vitamine C) et l'alpha-tocophérol
    (vitamine E) inhibent tous deux la peroxydation des lipides insaturés.
    Lorsque l'acide ascorbique est emprisonné dans une double couche
    liposomique, il est rapidement oxydé par le NO2.  L'effet protecteur
    de l'alpha-tocophérol et de l'acide ascorbique chez l'homme et
    l'animal est dû à l'inhibition de l'oxydation par le NO2. NO2 oxyde
    également les protéines membranaires. L'oxydation des lipides ou des
    protéines membranaires conduit à la disparition du mécanisme de
    régulation de la perméabilité cellulaire. Dans la lumière pulmonaire
    des sujets humains et des animaux de laboratoire exposés au NO2, on
    constate la présence d'une plus grande quantité de protéines. Ces
    phénomènes sont à l'origine du recrutement de cellules inflammatoires
    et des altérations qui se produisent au niveau pulmonaire.

         Les propriétés oxydantes de NO2 mettent en action différentes
    voies de détoxication: la voie de la glutathion-peroxydase, celle
    de la glutathion-réductase et celle de la glucose-6-phosphate
    déshydrogénase. Après exposition au NO2, la montée de la voie de
    détoxication peroxydique suit une relation de type dose-réponse.

         Le mode d'action du NO n'est pas aussi clair.Il y a d'abord
    oxydation en NO2 avant que n'intervienne la peroxydation. En cas
    d'exposition à NO, il y a toujours une certaine exposition à NO2 qui
    se produit simultanément de sorte qu'il est difficile de démêler les
    effets imputables à chacun des composés. NO se comporte comme un
    second messager intracellulaire qui module toutes sortes d'enzymes
    essentielles et qui, par rétroaction négative, inhibe sa propre
    production. NO active la guanilate-cyclase qui accroît à son tour la
    concentration intracellulaire de cGMP. Quant aux nitrates, il est
    possible qu'ils agissent en libérant l'histamine présente dans les
    granules des mastocytes. Les polluants atmosphériques acides
    constitués de dérivés azotés, en particulier HNO3, pourraient agir
    en modifiant le pH intracelllulaire.

         Le PAN se décompose dans l'eau en donnant de l'eau oxygénée
    (peroxyde d'hydrogène). On sait très peu de chose sur son mode
    d'action, mais il est probable qu'il agit, comme ses congénères, en
    provoquant un stress oxydatif.

         Il se pourrait, comme on l'a d'ailleurs indiqué plus haut, que
    l'action des nitrates inorganiques consiste à modifier le pH
    intracellulaire. L'ion nitrate est transporté dans les cellules
    alvéolaires de type 2 dont il provoque l'acidification. Il mobilise
    également l'histamine des mastocytes. HNO2 pourrait également modfier
    le pH intracelllulaire, mais son mode d'action n'est pas encore
    vraiment élucidé.

         Le mode d'action des autres oxydes d'azote n'est pas connu.

         Une exposition aiguë à NO2 à la concentration de 750 µg/m3,
    soit 0,4 ppm, peut provoquer la peroxydation des lipides. NO2 peut
    oxyder les lipides insaturés qui entrent dans la composition de la
    membrane cellulaire ainsi que les groupes fonctionnels de protéines,
    par exemple, de protéines solubles présentes à l'intérieur de la
    cellule, comme les enzymes, ou encore de protéines de structure,
    comme les protéines membranaires. Ces réactions d'oxydation (qui
    s'effectuent par l'intermédiaire de radicaux libres) sont le mécanisme
    par lequel NO2 exerce son action toxique sur les cellules
    pulmonaires.A l'appui de l'existence de ce mode d'action, on peut
    citer des études sur animaux de laboratoire qui montrent l'importance
    des défenses antioxydantes du poumon, qu'elles soient endogènes (par
    exemple, maintien d'un taux suffisant de glutathion intrapulmonaire)
    ou exogènes (par exemple, apport alimentaire de vitamines C et E),
    dans la protection contre les effets de NO2. Selon de nombreuses
    études, les diverses enzymes pulmonaires, et notamment la
    glutathion-peroxydase, la superoxyde-dismutase et la catalase,
    pourraient également avoir pour rôle de protéger le poumon contre les
    attaques oxydantes.

    3.1.2  Effets sur les défenses de l'hôte

         Bien que la fonction essentielle de l'arbre respiratoire soit
    d'assurer des échanges gazeux efficaces, cet organe constitue
    également la première ligne de défense de l'organisme contre les
    agents aéroportés, viables ou non, qu'inhale le sujet. Une abondante
    base de données montre que l'exposition à NO2 peut entraîner la
    perturbation de ces défenses et, par voie de conséquence, une plus
    grande sensibilité aux affections respiratoires d'origine infectieuse.
    Parmi les éléments de ces défenses qui peuvent être affectés par NO2,
    figurent notamment l'activité biochimique et fonctionnelle de
    certaines cellules pulmonaires, les macrophages alvéolaires,
    l'immunocompétence, la sensibilité aux infections respiratoires
    expérimentales et la vitesse d'élimination par l'ascenseur
    mucociliaire.

         NO2 s'attaque aux macrophages alvéolaires. Ces cellules ont
    pour fonction de maintenir la stérilité de la région pulmonaire en en
    éliminant les particules étrangères et en assurant également des
    fonctions immunologiques. Parmi les altérations fonctionnelles qui ont

    été relevées on peut citer les suivantes: suppression de la capacité
    de phagocytose et stimulation de la clairance pulmonaire à la dose de
    560 µg/m3 (0,3 ppm) 2 h par jour pendant 13 jours; diminution de
    l'activité bactéricide à la dose de 4320 µg/m3 (2,3 ppm) pendant
    17 h; affaiblissement de la réponse au facteur d'inhibition de la
    migration à la dose de 3760 µg/m3 (2,0 ppm) 8 h par jour, 5 jours par
    semaine, pendant 6 mois. L'exposition prolongée des macrophages à NO2
    provoque une modification morphologique de ces cellules.

         L'importance des défenses de l'hôte saute aux yeux lorsqu'on
    observe des animaux de laboratoire porteurs d'infections respiratoires
    expérimentales. La mortalité des animaux exposés à NO2 et qui
    succombent à l'infection bactérienne ou virale, dépend de la dose.  La
    mortalité augmente également à mesure qu'augmente la concentration de
    NO2 ou la durée de l'exposition. En cas d'exposition aiguë, on
    observe des effets dès la dose de 3760 µg/m3 (2 ppm). Sur modèle
    d'infectiosité, on constate des effets dans les 6 mois suivant
    l'exposition à une dose ne dépassant pas 940 µg/m3 (0,5 ppm).

         L'exposition à NO2 affecte les défenses humorales comme les
    défenses à médiation cellulaire. Dans les cas où l'on a étudié le
    comportement du système immunitaire, on a pu observer des effets après
    une exposition de courte durée à des concentrations supérieures ou
    égales à 9400 µg/m3 (5 ppm). Les effets sont complexes car le
    sens de la modification (augmentation ou diminution) dépend de la
    concentration de NO2 et de la durée de l'exposition.

    3.1.3  Effets d'une exposition prolongée sur l'apparition d'une
           pneumopathie chronique

         L'homme est exposé en permanence au NO2. C'est pourquoi ce type
    d'exposition a été assez largement étudié chez l'animal en ayant
    recours à des méthodes morphologiques ou morphométriques. En règle
    générale, ce genre de travaux montre que diverses modifications de
    structure, avec leurs corrélats fonctionnels, se produisent au niveau
    pulmonaire. Certaines de ces modifications peuvent se révéler
    reversibles lorsque cesse l'exposition.

         Chez l'animal de laboratoire, une exposition chronique au dioxyde
    d'azote peut entraîner une altération de la fonction respiratoire.
    Après exposition à du dioxyde d'azote pendant 4 mois à la dose de
    7520 µg/m3, soit 4,0 ppm, on a observé une déterioration des échanges
    gazeux qui se traduisait par une réduction de la pression partielle
    d'oxygène dans le sang artériel, une diminution de la condition
    physique et une augmentation du métabolisme anaérobie.

         Il est certain que le dioxyde d'azote provoque des modifications
    morphologiques au niveau des voies respiratoires, mais il peut arriver
    que la base de données soit un peu trompeuse sur ce point en raison

    des variations qualitatives et quantitatives qui se manifestent dans
    la sensibilité des différentes espèces, voire à l'intérieur d'une
    même espèce. Le rat, qui est l'animal le plus fréquemment utilisé
    pour l'évaluation de l'exposition sur la base des modifications
    morphologiques, se révèle relativement résistant au NO2. Une
    exposition de brève durée à des concentrations de 9400 µg/m3
    (5,0 ppm) ou moins, n'a généralement guère d'effets sur le rat, alors
    que dans les mêmes conditions le cobaye présente des lésions de
    l'épithélium centroacinaire.

         Une exposition de plus longue durée peut, chez certaines espèces,
    provoquer des lésions à des doses ne dépassant pas 560 à 940 µg/m3
    (0,3 à 0,5 ppm). Elles se caractérisent par un remodelage de
    l'épithélium similaire à celui qui a été décrit plus haut, mais avec
    extension aux voies aériennes proximales et épaississement du tissu
    interstitiel. Toutefois, nombre de ces altérations finissent par
    disparaître, même si l'exposition se poursuit, et il faut que celle-ci
    se situe au moins à 3760 µg/m3 (2,0 ppm) pour que des dommages plus
    étendus et plus persistants se produisent au niveau des poumons.
    Certains effets sont relativement persistants, (par exemple, la
    bronchiolite) alors que d'autres manifestent une tendance à la
    réversibilité et sont limités, même si l'exposition se poursuit. De
    toute façon, il semble que la réponse soit davantage liée à la dose
    qu'à la durée - brève ou longue - de l'exposition. On a de bonnes
    raisons de penser qu'une exposition de longue durée à de fortes
    concentrations de NO2 provoque, chez plusieurs espèces animales, des
    lésions affectant la morphologie pulmonaire. La destruction de la
    paroi alvéolaire, qui constitue un critère supplémentaire essentiel
    d'emphysème chez l'homme, a été constatée quelquefois à l'occasion
    d'études tout à fait dignes de foi effectuées sur l'animal. Ces
    résultats ne permettent toutefois pas de déterminer quelle est la
    concentration de NO2 la plus faible à partir de laquelle apparaissent
    des lésions pulmonaires emphysémateuses.

    3.1.4  Effets cancérogènes ou co-cancérogènes potentiels

         On a montré que NO2 était mutagène pour les salmonelles, mais
    une étude indique qu'il ne l'est pas pour des cellules mammaliennes en
    culture. D'autres travaux sur cultures cellulaires ont montré
    l'existence d'échanges entre chromatides soeurs ainsi que des ruptures
    au niveau d'un des brins de l'ADN. Auncun effet génotoxique n'a été
    mis en évidence  in vivo dans les lymphocytes, les spermatocytes
    ou les cellules de la moelle osseuse, mais deux études au cours
    desquelles on a fait inhaler pendant 3 h ou 6 h (aux doses respectives
    de 50 760 et 56 400 µg/m3, soit 27 et 30 ppm) le produit à des
    animaux, ont révélé la présence de tels effets dans les poumons. 

         Les études bibliographiques qui ont été effectuées sur ce sujet
    n'ont pas révélé l'existence de travaux comportant une étude
    toxicologique classique sur l'animal avec exposition de longue durée,

    dans le but d'étudier le pouvoir cancérogène du NO2. Les études
    effectuées sur des souris présentant un taux élevé de tumeurs
    spontanées, n'ont fourni que des résultats équivoques. Dans une étude,
    on a observé qu'à la concentration de 18 800 µg/m3 (10 ppm) le NO2
    augmentait légèrement l'incidence des adénomes pulmonaires chez une
    souche de souris sensibles (A/J). On a bien effectué un certain nombre
    d'études de co-cancérogénicité, mais des problèmes de méthodologie et
    d'interprétation empêchent d'en tirer des conclusions. Quant à savoir
    si l'exposition au NO2 rend les tumeurs pulmonaires plus aptes à
    métastasier, les études qui ont été consacrées à ce problème ne
    permettent guère de conclure. Dans d'autres études, on s'est attaché à
    rechercher si l'exposition au NO2 pouvait entraîner la formation de
    nitrates ou de nitrites susceptibles de donner naissance à des
    nitrosamines par réaction sur les amines présentes dans l'organisme.
    Certains résultats donnent à penser que des nitrosamines se forment
    chez les animaux exposés au NO2 auxquels on administre des amines
    à haute dose, mais d'autres travaux montrent en revanche que la
    formation de nitrosamines est improbable.

    3.1.5  Sensibilité en fonction de l'âge

         Les travaux consacrés à cette question sont insuffisants et les
    résultats obtenus jusqu'ici sont équivoques.

    3.1.6  Influence des modalités de l'exposition

         Un certain nombre d'études toxicologiques ont permis d'expliciter
    les relations entre la concentration C et la durée T de l'exposition.
    Ces relations se révèlent complexes.La plupart des travaux utilisent
    le modèle d'infectiosité. Les premières études consacrées à la
    relation Effet =  f  (C,T), ont montré que la concentration avait
    davantage d'influence sur la mortalité que la durée de l'exposition.
    Les relations Effet =  f (C, T) ne permettent pas d'évaluer la
    toxicité de NO2.

    3.2  Exposition contrôlée aux oxydes d'azote: études sur l'homme

         On a étudié les réactions humaines à divers dérivés oxygénés de
    l'azote. La base de données de loin la plus abondante et la mieux
    adaptée à l'évaluation du risque est celle qui a été établie à partir
    des résultats d'expositions contrôlées au NO2. La base de données sur
    les réactions de l'organisme humain à une exposition à NO, HNO3 et
    HNO2 en phase vapeur et à divers nitrates inorganiques sous forme
    d'aérosols, n'est pas aussi fournie. On a examiné un certain nombre de
    sous-groupes sensibles ou potentiellement sensibles, notamment des
    adolescents et des adultes asthmatiques, ainsi que des adultes
    d'âge mûr atteints d'une pneumopathie obstructive chronique et
    d'hypertension pulmonaire. On a constaté que lorsque l'exposition à
    ces composés s'accompagne d'un exercice physique, il y a accroissement

    de leur absorption et modification de leur répartition à l'intérieur
    du poumon. La proportion relative de NO2 déposé dans les voies
    respiratoires inférieures est également augmentée par l'exercice
    physique.  Chez les personnes qui s'adonnent à une activité physique
    tout en étant exposées à des dérivés oxygénés de l'azote, les effets
    de ces composés peuvent donc se trouver accrus.

         Comme chaque fois que l'organisme humain est exposé par la voie
    respiratoire à des gaz ou à des particules, sa réponse biologique au
    NO2 se caractérise par une certaine variabilité. Les sujets en bonne
    santé ont tendance à moins réagir aux effets du NO2 que les individus
    atteints d'une pneumopathie. Il est certain que les asthmatiques
    constituent le groupe le plus sensible au NO2 qui ait été étudié
    jusqu'ici. Les sujets atteints d'une pneumopathie obstructive
    chronique pourraient être plus sensibles que les sujets sains, mais
    comme leur capacité de réaction au NO2 est limitée, il est difficile
    de procéder à une évaluation quantitative. On ne possède pas
    suffisamment de données pour déterminer si l'âge et le sexe jouent un
    rôle dans la réaction au NO2.

         Un sujet normal peut déceler l'odeur du NO2, quelquefois à une
    concentration inférieure à 188 µg/m3 (0,1 ppm). D'une façon générale,
    l'exposition au NO2 n'a provoqué aucune augmentation des symptômes
    respiratoires chez les sujets étudiés.

         Le NO2 entraîne une réduction de la fonction pulmonaire et en
    particulier, une augmentation de la résistance des voies aériennes
    chez le sujet sain au repos exposé pendant 2 h à une concentration
    de 4700 µg/m3 (approx. 2,5 ppm). Les données disponibles
    sont insuffisantes pour permettre d'expliciter la relation
    concentration-réponse.

         L'exposition pendant 1 h ou plus à une concentration de NO2 ne
    dépassant pas 2800 µg par m3, cest-à-dire approx. 1,5 ppm, rend les
    voies aériennes plus sensibles aux agents bronchoconstricteurs chez
    les sujets sains non fumeurs pratiquant une activité physique.

         Chez les asthmatiques exposés au NO2, on observe, du moins
    chez certains d'entre eux, une augmentation de la sensibilité des
    voies aériennes à divers agents, en particulier des substances
    cholinergiques et des antihistaminiques, ou encore au SO2 ou à l'air
    froid.  Les réactions de ce type semblent dépendre du protocole
    expérimental, et notamment de la présence ou de l'absence d'une
    activité physique pendant l'exposition. Elles peuvent se produire à
    des concentrations ne dépassant pas 380 µg/m3 (0,2 ppm). Lorsqu'on
    soumet ces résultats à une méta-analyse, on est amené à penser que les
    réactions précitées peuvent se produire à des concentrations encore
    plus faibles. On a cependant constaté l'existence d'une relation
    concentration-réponse indiscutable entre 350 et 1150 µg/m3
    (approx. 0,2 à 0,6 ppm).

         On ne voit pas très bien ce que signifie cette tendance générale,
    mais une sensibilité accrue des voies aériennes pourrait entraîner
    une exacerbation des réactions aux allergènes ou l'aggravation
    temporaire d'un asthme, avec pour conséquences une augmentation de
    la consommation de médicaments, voire même des hospitalisations.

         Chez les malades porteurs d'une pneumopathie obstructive
    chronique, on peut observer une augmentation modérée de la résistance
    des voies aériennes après une brève exposition (15-60 min) à des
    concentrations de NO2 ne dépassant pas 2800 µg/m3 (approx. 1,5 ppm)
    et une diminution des valeurs spirométriques peut également s'observer
    dès que la concentration atteint 600 µg/m3 (approx. 0,3 ppm) sur 3 h:
    le volume maximal expiré en une seconde (VEMS) est en baisse de 3 à
    8%.

         L'exposition à des concentrations de NO2 dépassant 2800 µg/m3
    (approx.1,5 ppm) peut modifier le nombre et le type des cellules
    inflammatoires présentes dans la partie distale des voies aériennes et
    des alvéoles. Ce gaz peut également perturber le fonctionnement des
    cellules intrapulmonaires ainsi que la production de médiateurs
    susceptibles de jouer un rôle important dans les défenses pulmonaires.
    Cet ensemble de perturbations interessant les défenses de l'hôte, la
    modification des cellules pulmonaires et l'altération de leurs
    fonctions, de même que les anomalies affectant la production de
    certains médiateurs biochimiques, correspondent bien aux résultats des
    études épidémiologiques, à savoir que l'exposition au NO2 accroît la
    sensibilité des voies respiratoires du sujet.

         D'après des études portant sur des mélanges de polluants
    contenant du NO2, il ne semble pas que la présence de NO2 accroisse
    les réactions aux autres polluants au-delà de ce qui serait observé en
    présence de ces polluants seuls. Il y a toutefois une exception
    notable, à savoir le fait qu'une exposition préalable à ce gaz rend
    les voies aériennes encore plus sensibles à l'ozone, comme on a pu le
    constater chez des sujet sains exerçant une activité physique en
    présence de NO2, puis exposés à de l'ozone. Cette observation incite
    à penser que la réponse au NO2 peut être retardée ou persistante.

         Si l'on considère l'intervalle de concentration pour lequel il
    serait interessant d'évaluer le risque que représente une exposition
    au NO2 (c'est-à-dire 100-600 µg/m3), on constate que les données
    disponibles ne permettent pas d'établir une relation concentration-
    réponse concernant divers symptômes et notamment les effets aigus sur
    la fonction pulmonaire ou sur la sensibilité des voies aériennes aux
    agents bronchoconstricteurs.

         En se basant sur l'effet constaté à 400 µg/m3 et la possibilité
    d'effets à concentration plus faible telle qu'elle ressort d'une
    méta-analyse des données, on recommande de prendre comme valeur-guide
    de la concentration moyenne maximale journalière de NO2 sur 1 h, le
    chiffre de 200 µg/m3 (approx.0,11 ppm).

         Il est admis que le NO joue un rôle important comme deuxième
    messager au sein de divers organes. Lorsqu'il est inhalé à une
    concentration supérieure à 6000 µg/m3 (approx.5 ppm), il peut
    provoquer une vasodilatation des vaisseaux pulmonaires qui ne s'étend
    pas à la circulation générale. On n'a pas établi quelle est la
    concentration minimale capable de produire cet effet. Pour le moment,
    les données dont on dispose sur les effets qu'une exposition au NO
    serait susceptible d'avoir sur la fonction et les défenses pulmonaires
    sont trop limitées pour qu'on puisse en tirer la moindre conclusion.
    Des concentrations relativement élevées ont été utilisées en clinique
    (> 40 000 µg/m3) pendant de courtes périodes (< 1 h) sans que l'on
    n'observe d'effets indésirables.

         Dans l'intervalle de concentration de 250-500 µg/m3
    (97-194 parties par milliard), l'acide nitrique peut avoir des effets
    indésirables sur la fonction pulmonaire chez l'asthmatique adolescent
    mais pas chez l'adulte en bonne santé.

         Les données limitées dont dispose sur HNO2 incitent à penser que
    cet acide peut provoquer une inflammation oculaire à la concentration
    de 760 µg/m3 (0,40 ppm). Rien n'a été publié jusqu'ici sur la manière
    dont le poumon humain réagit à une exposition à HNO2. 

         Les données relatives aux nitrates organiques sont également
    limitées et indiquent que sous la forme d'aérosols, ces composés n'ont
    pas d'effets sur la fonction pulmonaire à des concentrations
    inférieures ou égales à 7000 µg/m3.

    3.3  Etudes épidémiologiques sur le dioxyde d'azote

         Les études épidémiologiques consacrées aux effets des oxydes
    d'azote portent essentiellement sur le NO2. Nombre d'entre elles ont
    été menées en extérieur ou en intérieur afin de déterminer la nature
    des effets de ce composé sur la santé humaine. Deux types d'effets
    sanitaires sont généralement pris en considération pour l'étude de
    l'exposition au NO2, à savoir le retentissement sur la fonction
    pulmonaire et les affections ou symptômes respiratoires.  Les études
    effectuées sur des écoliers au sujet des effets (symptômes et
    maladies) que le NO2 exerce au niveau de voies respiratoires
    inférieures ont donné des résultats quelque peu contrastés. Ces
    travaux ont fait l'objet d'un examen visant à en vérifier la cohérence
    et une synthèse en a été élaborée sous la forme d'une analyse
    quantitative (méta-analyse). La plupart des études effectuées en
    intérieur font ressortir une augmentation de la morbidité affectant

    les voies respiratoires inférieures chez les enfants durablement
    exposés au NO2. Les concentrations hebdomadaires moyennes relevées
    dans les chambres à coucher se situaient essentiellement entre 15 et
    122 µg/m3 (0,008 et 0,065 ppm). Une synthèse des résultats obtenus
    en intérieur en supposant des points d'aboutissement toxicologique
    communs donne, pour les effets sur les voies respiratoires
    inférieures, un  odds ratio de 1,2 (limites de confiance à 95%: 1,1
    et 1,3) par incrément de 28,3 µg/m3 (0,015 ppm) de l'exposition
    moyenne au NO2 calculée sur 2 semaines. On est donc amené à penser,
    compte tenu des hypothèses sur lesquelles repose cette analyse
    globale, que chaque fois que l'exposition moyenne sur deux semaines
    augmente de 28,3 µg/m3 (0,015 ppm) les chances de symptômes ou de
    maladie affectant les voies respiratoires inférieures augmentent de
    20%. Cet ensemble de résultats milite donc en faveur de l'hypothèse
    selon laquelle une exposition au NO2 provoque, chez les enfants de 5
    à 12 ans, des effets au niveau de voies respiratoires inférieures.

         Des études également menées en intérieur, mais cette fois au
    niveau individuel, chez des enfants de 2 ans au plus, n'ont pas permis
    de dégager une relation systématique entre les estimations de
    l'exposition au NO2 et la prévalence des symptômes ou des maladies
    affectant les voies respiratoires inférieures. En se basant sur une
    méta-analyse de ces données et compte tenu des hypothèses formulées
    à cette fin, on a trouvé que  l'odds ratio combiné pour une
    augmentation égale à 28,2 µg/m3 (0,015 ppm) de l'exposition au NO2,
    était de 1,09, avec un intervalle de confiance à 95% de 0,95-1,26,
    lorsque la concentration hebdomadaire moyenne du NO2 dans les
    chambres à coucher se situait entre 9,4 et 94 µg/m3 (0,005 et
    0,050 ppm). L'accroissement du risque était très faible et n'a
    d'ailleurs pas été mentionné systématiquement dans toutes les études.
    Finalement, on ne peut pas conclure que ces résultats indiquent
    l'existence, chez les enfants en bas âge, d'effets analogues à ceux
    qui ont été constatés chez les enfants plus âgés. Les raisons de cette
    différence due à l'âge restent obscures.

         Les études dans lesquelles l'exposition au NO2 avait
    effectivement été mesurée ont donné un  odds ratio systématiquement
    plus élevé que celles dans lesquelles ces estimations avaient été
    obtenues de façon indirecte, ce qui s'explique par les erreurs de
    mesure. Les corrections apportées pour tenir compte de covariables
    aléatoires comme la situation socio-économique, le tabagisme et le
    sexe ont eu pour conséquence que les études dans lesquelles des
    corrections de ce type avaient été faites, ont donné un  odds ratio 
    plus élevé que celles où elles ne l'avaient pas été.

         Bien que nombre des études épidémiologiques basées sur des
    mesures effectives de l'exposition au NO2 n'aient utilisé que des
    données obtenues sur 1 à 2 semaines tout au plus, on en a tout de même
    déduit l'exposition des enfants sur une période beaucoup plus longue.
    Le questionnaire standard utilisé dans la plupart des cas pour

    enregistrer les symptômes respiratoires récapitule des informations
    sur l'état de santé des sujets qui s'étendent sur toute une année. Le
    chiffre de 28,2 µg/m3 (0,015 ppm) utilisé dans les méta-analyses
    correspond à la différence d'exposition annuelle moyenne au NO2,
    selon que le ménage utilisait une cuisinière à gaz ou une cuisinière
    électrique. Dans certaines études, on n'a mesuré la concentration de
    NO2 que pendant l'hiver, d'où une possible surestimation de
    l'exposition annuelle moyenne. Dans ces conditions, il y aurait eu
    sous-estimation de l'effet sanitaire d'une différence de 28,2 µg/m3
    (0,015 ppm) dans l'exposition annuelle au NO2. Dans une étude basée
    sur l'exposition annuelle moyenne dans les ménages, mesurée en hiver
    et en été, l'effet observé a été plus important que dans beaucoup des
    autres études. On ignore quelle est la période qui serait vraiment
    significative sur le plan biologique, mais il est à noter que
    l'exposition prise en considération dans ces travaux s'est poursuivie
    pendant de longues périodes, voire pendant toute la vie.

         Les travaux actuels ne mettent pas en évidence d'association
    claire entre la concentration de NO2 à l'extérieur et l'intégrité de
    la fonction respiratoire. Un certain nombre de résultats indiquent que
    les affections respiratoires pourraient se prolonger lorsque l'air est
    fortement chargé en NO2. L'analyse des études portant sur l'air
    extérieur se heurte à une difficulté majeure: distinguer les effets
    imputables au NO2 de ceux qui sont dus à d'autres polluants.

         L'interprétation des résultats des études précitées et de la
    méta-analyse doit prendre en considération plusieurs incertitudes qui
    subsistent.L'erreur de mesure sur l'exposition pourrait être l'un des
    problèmes méthodologiques les plus importants qui se posent dans les
    études épidémiologiques sur le NO2. Les résultats expérimentaux
    incitent à admettre l'existence d'une association entre certains
    symptômes et les indicateurs de l'exposition au NO2, mais ces
    estimations de l'exposition ne seraient pas suffisamment fiables pour
    permettre d'établir une relation quantitative entre exposition et
    symptômes. Dans la plupart des études au cours desquelles il a été
    procédé à des mesures de l'exposition, ces mesures ne portaient que
    sur une durée de 1 à 2 semaines et ont été rapportées sous la forme de
    valeurs moyennes. On a rarement cherché à établir une relation entre
    les effets observés et les modalités de l'exposition, par exemple
    l'existence de pics transitoires de concentration. En outre, il est
    possible que la concentration de NO2 mesurée n'ait pas été égale à
    la dose biologiquement significative. D'ailleurs, l'estimation de
    l'exposition effective suppose la connaissance de l'espèce chimique en
    cause, de sa concentration et du type d'activité humaine qui lui a
    donné naissance. On ne dispose toutefois que d'un nombre limité de
    données sur l'activité humaine et les conditions météorologiques en
    rapport avec ces facteurs. L'extrapolation à d'autres modalités
    d'exposition reste un exercice difficile. En outre, même si, du fait
    des analogies et des éléments communs qui existent entre les variables
    mesurées dans ces études, on peut avoir une certaine confiance dans

    leur utilisation en vue d'une analyse quantitative, les symptômes
    et les maladies constatés sont quand même différents, jusqu'à un
    certain point, et peuvent parfaitement correspondre à des processus
    sous-jacents d'une autre nature. Dans ces conditions, la prudence
    s'impose dans l'interprétation des résultats de la méta-analyse.

         Dans d'autres études épidémiologiques, on s'est efforcé
    d'établir une relation entre certaines mesures de l'exposition au
    NO2 à l'intérieur ou à l'extérieur et l'altération de la
    fonction pulmonaire. Il s'agissait en fait d'anomalies respiratoires
    d'importance marginale. La plupart des études ne sont pas parvenues
    à déceler le moindre effet, résultat qui cadre avec ceux des
    études contrôlées sur l'homme.  Quoi qu'il en soit, les données
    épidémiologiques sont insuffisantes pour que l'on puisse tirer des
    conclusions sur les effets qu'une exposition de courte ou de longue
    durée au NO2 pourrait avoir au niveau pulmonaire. 

         En se basant sur un niveau de fond de 15 µg/m3 (0,008 ppm) et le
    fait que des effets indésirables significatifs apparaissent lorsque
    l'exposition augmente d'au moins 28,2 µg/m3, c'est-à-dire 0,015 ppm,
    on peut proposer une valeur-guide de 40 µg/m3 (0,023 ppm) en moyenne
    annuelle. Cette valeur permettra d'éviter les expositions les plus
    graves. Il reste cependant à souligner qu'il n'a pas encore été
    possible de déterminer la valeur de la concentration correspondant à
    l'absence d'effet en cas d'exposition chronique ou subchronique au
    NO2.

    3.4  Valeurs-guides à visée sanitaire pour le dioxyde d'azote

         Le résultats des études contrôlées sur l'homme conduisent à
    adopter, en cas d'exposition à court terme, une valeur-guide de
    200 µg/m3 (0,11 ppm) pour la concentration journalière maximale de
    NO2 calculée en moyenne sur 1 h.  Dans le cas d'une exposition à long
    terme, on recommande, en se basant sur les études épidémiologiques
    attestant un risque accru d'affections respiratoires chez l'enfant,
    une valeur-guide de 40 µg/m3 (0,023 ppm) en moyenne annuelle.

    RESUMEN

    1.  Oxidos de nitrógeno y compuestos afines

         Los óxidos de nitrógeno pueden alcanzar concentraciones
    considerables en el aire del medio ambiente y de espacios cerrados.
    Los tipos y concentraciones de los compuestos de nitrógeno presentes
    pueden variar notablemente de unos lugares a otros, con la hora del
    día y con la estación. Las fuentes principales de emisión de óxidos de
    nitrógeno son los procesos de combustión. Las centrales eléctricas que
    funcionan con combustibles fósiles, los vehículos de motor y los
    aparatos de combustión domésticos emiten óxidos de nitrógeno, sobre
    todo óxido nítrico (NO), y en algunos casos (normalmente menos del
    10 por ciento) dióxido de nitrógeno (NO2). En el aire se producen
    reacciones químicas que oxidan el NO a NO2 y otros productos.
    Hay también procesos biológicos que liberan del suelo productos
    nitrogenados, incluso óxido nitroso (N2O). Las emisiones de N2O
    pueden producir alteraciones en la capa de ozono estratosférica.

         La salud humana puede verse afectada por la presencia de
    concentraciones importantes de NO2 u otros productos nitrogenados,
    como por ejemplo el nitrato de peroxiacetilo (NPA), el ácido nítrico
    (NO3H), el ácido nitroso (NO2H) y los compuestos orgánicos
    nitrogenados. Además, cuando los nitratos y el ácido nítrico se
    depositan en la tierra pueden tener efectos en la salud y
    repercusiones considerables sobre los ecosistemas.

         El conjunto de NO y NO2 suele recibir el nombre de NOx. Una vez
    liberado en el aire, el NO se oxida a NO2 por acción de los oxidantes
    presentes (en particular el ozono, O3). Esta reacción, en
    determinadas condiciones, es muy rápida al aire libre; en el aire de
    espacios cerrados suele ser un proceso mucho más lento. Los óxidos de
    nitrógeno son un precursor que controla la contaminación del aire por
    oxidantes fotoquímicos, dando lugar a la formación de ozono y de
    bruma; las interacciones de los óxidos de nitrógeno (excepto el N2O)
    con compuestos orgánicos reactivos y la luz solar producen ozono en la
    troposfera y bruma en las zonas urbanas.

         El NO y el NO2 pueden sufrir asimismo reacciones que producen
    una serie de óxidos de nitrógeno, tanto en espacios abiertos como
    cerrados, entre ellos NO2H, NO3H, trióxido de nitrógeno (NO3),
    pentóxido de nitrógeno (N2O5), NPA y otros nitratos orgánicos. La
    gama compleja de óxidos de nitrógeno gaseosos recibe el nombre de
    NOy. El reparto de los óxidos de nitrógeno entre estos compuestos
    depende fundamentalmente de las concentraciones de otros oxidantes y
    de los antecedentes meteorológicos del aire.

         El NO3H es producto de la reacción entre el OH- y el NO2. Es
    el sumidero principal del nitrógeno activo y contribuye también a la
    deposición acídica. Entre los posibles sumideros físicos y químicos

    del NO3H figuran la deposición húmeda y seca, la fotolisis la
    reacción con radicales OH y la reacción con amoníaco gaseoso para
    formar un aerosol de nitrato de amonio.

         Los NPA se forman mediante la combinación de radicales peroxilo
    orgánicos con NO2. El NPA es el nitrato orgánico más abundante en la
    troposfera y puede servir como reservorio temporal de nitrógeno
    reactivo, que se puede transportar de una zona a otra.

         El radical NO3, compuesto NOy que se forma en la troposfera
    fundamentalmente por reacción del NO2 con el O3, sufre una fotolisis
    rápida a la luz del día o una reacción con el NO. Durante la noche se
    observan concentraciones apreciables.

         El N2O5 es básicamente un componente nocturno del aire
    atmosférico, puesto que se forma a partir de la reacción del NO3 y el
    NO2. En el aire de la atmósfera, el N2O5 sufre una reacción
    heterogénea con el agua y forma NO3H, que a su vez se deposita.

         El N2O está presente en todas partes, debido a que es un
    producto de procesos biológicos naturales del suelo. No se sabe, sin
    embargo, si interviene en alguna reacción en la troposfera. El N2O
    participa en reacciones de la capa superior de la atmósfera,
    contribuyendo a la reducción del ozono (O3) de la estratosfera, y
    también es un gas de efecto de invernadero relativamente potente, que
    contribuye al calentamiento mundial.

    1.1  Transporte en la atmósfera

         El transporte y la dispersión de los diversos compuestos
    nitrogenados en la capa inferior de la troposfera dependen de
    parámetros tanto meteorológicos como químicos. La advección, la
    difusión y las transformaciones químicas combinadas determinan los
    tiempos de permanencia en la atmósfera. Éstos, a su vez, ayudan a
    establecer el alcance geográfico del transporte de un compuesto
    concreto. Las emisiones superficiales se dispersan en sentido vertical
    y horizontal a través de la atmósfera mediante procesos mixtos
    turbulentos que dependen en gran medida de la estructura vertical de
    la temperatura y de la velocidad del viento.

         Como consecuencia de los procesos meteorológicos, los NOx
    emitidos en las primeras horas de la mañana, en una zona urbana, se
    suelen dispersar en sentido vertical y desplazarse en el sentido del
    viento a medida que avanza el día. En los días soleados de verano, la
    mayoría de los NOx se habrán convertido en NO3H y NPA al atardecer,
    con la consiguiente formación de ozono. Una gran parte del NO3H se
    elimina por deposición con el transporte de las masas de aire, pero el
    NO3H y el NPA arrastrados a las capas altas (por encima de la capa de

    inversión nocturna, pero por debajo de una inversión de subsidencia
    superior) se pueden transportar potencialmente a grandes distancias en
    masas de aire ricas en oxidantes.

    1.2  Medición

         Son varios los métodos disponibles para medir los compuestos de
    nitrógeno presentes en el aire. En el presente documento se describen
    brevemente las metodologías utilizables o de uso general en la
    actualidad para la vigilancia in situ de las concentraciones en el
    aire en ambientes tanto externos como internos. Los compuestos
    examinados son el NO, el NO2, el NOx, el nitrógeno complejo reactivo
    total (NOy), el NPA y otros nitratos orgánicos, el NO3H, el NO2H,
    el N2O5, el radical nitrato NO3- y el N2O.

         La medición de las concentraciones de óxidos de nitrógeno no es
    sencilla. Aunque existe un método fácil muy utilizado para la medición
    del NO (reacción quimioluminiscente con el ozono), es una excepción
    para los óxidos de nitrógeno. La quimioluminiscencia es también la
    técnica más utilizada para el NO2; éste se reduce en primer lugar a
    NO. Por desgracia, el catalizador utilizado normalmente para la
    reducción no es específico y tiene diversas eficacias de conversión
    para otros compuestos de nitrógeno oxidados. Por este motivo hay que
    tener mucho cuidado a la hora de interpretar los resultados del
    analizador común de quimioluminiscencia en cuanto al NO2, puesto que
    la señal puede incluir otros muchos compuestos. Se añaden nuevas
    dificultades por el hecho de que los óxidos de nitrógeno se pueden
    dividir entre las fases gaseosa y particulada tanto en la atmósfera
    como en el procedimiento de muestreo.

    1.3  Exposición

         La exposición humana y ambiental a los óxidos de nitrógeno varía
    mucho entre los espacios cerrados y abiertos, entre las ciudades y el
    campo y con la hora del día y la estación. Las concentraciones de NO y
    NO2 que suelen estar presentes en los espacios abiertos de una serie
    de situaciones urbanas están relativamente bien definidas. Las
    concentraciones en los espacios cerrados dependen de los detalles
    específicos del tipo de los aparatos de combustión, las chimeneas y la
    ventilación. Cuando se utilizan aparatos de combustión para cocinar o
    calentar sin ventilación, las concentraciones de óxidos de nitrógeno
    en el interior superan en general con mucho las que hay en el
    exterior. En investigaciones recientes se ha comprobado que en esas
    circunstancias el NO2H puede alcanzar concentraciones considerables.
    En un informe se señalaba que el NO2H puede representar más del
    10 por ciento de las concentraciones que se suelen dar como NO2.

    2.  Efectos en la vegetación de los compuestos de nitrógeno de la
        atmósfera, en particular los óxidos de nitrógeno

         La mayor parte de la biodiversidad del planeta se encuentra en
    ecosistemas (semi)naturales de hábitats tanto acuáticos como
    terrestres. El nitrógeno es el factor nutriente limitante para el
    crecimiento de las plantas en muchos ecosistemas (semi)naturales. La
    mayoría de las especies vegetales de estos hábitats están adaptadas a
    condiciones con escasez de nutrientes y solamente pueden competir con
    éxito en suelos con concentraciones bajas de nitrógeno.

         Las actividades humanas, tanto industriales como agrícolas, han
    aumentado considerablemente la cantidad de compuestos de nitrógeno
    disponibles desde el punto de vista biológico, alterando así el ciclo
    natural del nitrógeno. Hay diversas formas de nitrógeno que contaminan
    el aire, sobre todo el NO, el NO2 y el amoníaco (NH3) como
    deposición sólida y los nitratos (NO3-) y el amonio (NH4+) como
    deposición líquida. El NHy es la suma del NH3 y el NH4+. Otra
    parte corresponde a la deposición oculta (niebla y nubes). Hay muchos
    más contaminantes del aire que contienen nitrógeno (por ejemplo
    N2O5, NPA, N2O, aminas), pero éstos no se tienen en cuenta, debido
    a que se considera que su contribución a la deposición total de
    nitrógeno es pequeña o a que sus concentraciones están probablemente
    muy por debajo de los umbrales con efectos.

         Los contaminantes del aire con nitrógeno pueden afectar a la
    vegetación de manera indirecta, por medio de sus productos de reacción
    fotoquímica, o bien directamente, tras depositarse en la vegetación,
    el suelo, o la superficie del agua. La vía  indirecta apenas se tiene
    en cuenta aquí, aunque comprende procesos muy importantes y se debe
    tener presente al evaluar los efectos totales de los contaminantes del
    aire con nitrógeno: el NO2 es un precursor del O3 de la troposfera,
    que actúa como fitotoxina y como gas del efecto de invernadero.

         Los efectos de la mayor deposición de nitrógeno en los sistemas
    biológicos pueden deberse a la absorción directa del follaje o bien
    a través del suelo. Si se consideran las plantas individuales, los
    efectos más destacados son la lesión de los tejidos, los cambios
    en la producción de biomasa y la mayor susceptibilidad a factores
    secundarios de tensión. En relación con la vegetación, el nitrógeno
    depositado actúa como nutriente; esto produce cambios en las
    relaciones competitivas entre las especies y pérdida de biodiversidad.
    Las cargas críticas del nitrógeno dependen de: i) el tipo de
    ecosistema, ii) la utilización y ordenación de la tierra en el pasado
    y en el presente; y iii) las condiciones abióticas (especialmente las
    que influyen en el potencial de nitrificación y el índice de
    inmovilización en el suelo).

         En la superficie externa de las hojas se produce adsorción que
    puede ocasionar daños en las capas céreas de la cutícula, pero todavía
    no se ha demostrado la importancia cuantitativa para la situación en
    el campo. La absorción de NOx y NH3 depende del gradiente de
    concentración entre la atmósfera y el mesófilo. En general, aunque no
    siempre, está directamente determinada por la conductancia de los
    estomas, por lo que depende de factores que influyen en la apertura
    de éstos. Hay cada vez más pruebas de que la absorción foliar de
    nitrógeno reduce la que se produce por las raíces. La absorción y el
    intercambio de iones a través de la superficie de las hojas es
    un proceso relativamente lento, de manera que únicamente tiene
    importancia si la superficie se mantiene húmeda durante períodos
    prolongados.

         El NO sólo es ligeramente soluble en agua, pero la presencia de
    otras sustancias puede alterar la solubilidad. El NO2 tiene una
    solubilidad mayor, mientras que la del NH3 es mucho más elevada. El
    NO2- (producto primario de la reacción del NOx), el NH3 y el NH4+
    son muy fitotóxicos y podrían ser sin duda la causa de los efectos
    adversos de los contaminantes del aire que contienen nitrógeno. El
    radical libre *N=O  puede desempeñar una función en la fitotoxicidad
    del NO.

         Se han encontrado efectos superiores a los aditivos (sinergia) en
    casi todos los estudios relativos al SO2 más NO2. Con otras mezclas
    del NO2 (NO, O3 y CO2), los efectos interactivos son la excepción
    en lugar de la regla.

         Cuando las condiciones climáticas y el suministro de otros
    nutrientes permiten la producción de biomasa, tanto el NOx como el
    NHy estimulan el crecimiento a concentraciones bajas y lo reducen
    cuando las concentraciones son más elevadas. Sin embargo, el nivel de
    exposición al cual se pasa del estímulo del crecimiento a su
    inhibición es mucho más bajo para el NOx que para el NHy.

         Hay pruebas de que las plantas son más sensibles con una
    intensidad de luz escasa (por ejemplo de noche y en invierno) y a
    temperaturas bajas (ligeramente por encima de 0°C). El NOx y el NHy
    pueden aumentar la sensibilidad de las plantas a las heladas, la
    sequía, el viento y los daños de los insectos.

         Existe interacción entre la química del suelo y la sensibilidad
    de la vegetación a la deposición de nitrógeno; este proceso está
    relacionado con el pH y la disponibilidad de nitrógeno.

         No está clara la contribución relativa del NO y el NO2 al efecto
    del NOx en las plantas. La inmensa mayoría de la información
    disponible se refiere a los efectos del NO2, pero los datos
    existentes sobre el NO parecen indicar que éste y el NO2 tienen
    efectos fitotóxicos comparables.

         Las directrices sobre la calidad del aire se refieren a los
    umbrales para los efectos adversos. Existen dos tipos distintos de
    umbrales para los efectos: los niveles críticos y las cargas críticas.
    El nivel crítico se define como la concentración en la atmósfera por
    encima de la cual, según los conocimientos actuales, pueden producirse
    efectos adversos directos en los receptores, como las plantas, los
    ecosistemas o los materiales. La carga crítica se define como la
    estimación cuantitativa de una exposición (deposición) a uno o más
    contaminantes por debajo de la cual, según los conocimientos actuales,
    no hay efectos nocivos significativos en elementos sensibles
    específicos del medio ambiente.

         De acuerdo con la práctica actual, los niveles críticos se han
    derivado de la evaluación de las concentraciones mínimas de exposición
    que causan efectos adversos en la fisiología o el crecimiento de las
    plantas (se excluyeron los efectos bioquímicos), utilizando un método
    gráfico.

         A fin de incluir los efectos del NO, se propone un nivel crítico
    para el NOx en lugar de para el NO2; con este fin, se ha partido de
    la hipótesis de que el NO y el NO2 actúan de manera aditiva. Se
    pueden aducir razones sólidas a favor del establecimiento de niveles
    críticos para la exposición a corto plazo. Sin embargo, en la
    actualidad no se dispone de datos adecuados para definirlos con
    suficiente confianza. Las pruebas actuales parecen indicar un nivel
    crítico aproximado de 75 µg/m3 para el NOx como media de 24 horas.

         El nivel crítico para el NOx (NO y NO2 añadidos en ppmm y
    expresados como NO2 en µg/m3) se considera que es de 30 µg/m3 como
    media anual.

         La información acerca de los organismos en el medio ambiente se
    limita casi exclusivamente a las plantas, con datos mínimos sobre la
    fauna del suelo. Por consiguiente, los valores de esta evaluación y de
    orientación se expresan en función de los efectos de los compuestos de
    nitrógeno en la vegetación. Sin embargo, cabe prever que las plantas
    formen el componente más sensible de los sistemas naturales y que
    el efecto en la biodiversidad de las comunidades vegetales sea un
    indicador aceptable de los efectos en todo el ecosistema.

         Las cargas críticas se derivan de datos empíricos y de modelos
    estables del suelo. Se dan cargas críticas estimadas para la
    deposición total de nitrógeno en una serie de ecosistemas acuáticos y
    terrestres naturales. Los posibles efectos diferenciales de los
    compuestos de nitrógeno depositados (NOx y NHy) no se conocen
    suficientemente para diferenciar entre los distintos compuestos en la
    estimación de la carga crítica.

         La gran mayoría de los ecosistemas acerca de los cuales se
    dispone de suficiente información para estimar las cargas críticas son
    de climas templados. Los escasos ecosistemas árticos y montañosos
    incluidos, que cabría esperar que fueran representativos de altitudes
    mayores, tienen la base menos fidedigna. No hay información sobre
    ecosistemas tropicales y es muy poca la relativa a ecosistemas de
    estuarios o marinos de cualquier zona climática. Es probable que
    los ecosistemas tropicales con escaso nitrógeno, como las selvas
    tropicales y los manglares pantanosos, se vean afectados negativamente
    por la deposición de nitrógeno. La falta de datos sobre la deposición
    y de umbrales de los efectos hacen que sea imposible efectuar
    evaluaciones del riesgo para esas regiones climáticas.

         Los ecosistemas más sensibles (turberas ombrotróficas, lagos poco
    profundos de agua blanda y brezales árticos y alpinos) para los que
    pueden estimarse umbrales de los efectos muestran cargas críticas de
    5-10 kg de N/ha/año, tomando como base la menor diversidad biológica
    de las comunidades vegetales. Un valor más medio para la gama limitada
    de ecosistemas estudiados es de 15-20 kg de N/ha/año, que es aplicable
    a los árboles de los bosques.

         La química atmosférica de los óxidos de nitrógeno comprende la
    capacidad de generación de ozono en la troposfera, la reducción del
    ozono en la estratosfera y la contribución al calentamiento mundial
    como gases del efecto de invernadero. Los óxidos de nitrógeno y el
    amoníaco contribuyen a la acidificación del suelo (junto con
    los óxidos de azufre) y, por consiguiente, al aumento de la
    biodisponibilidad de aluminio.

         Los efectos fitotóxicos de los óxidos de nitrógeno en las
    plantas tienen escaso interés directo para las cultivadas cuando las
    concentraciones superan marginalmente el nivel crítico. Sin embargo,
    la función del NOx en la generación de ozono y otras sustancias
    fitotóxicas, por ejemplo nitratos orgánicos, da lugar a la pérdida
    de cultivos. El nitrógeno depositado en las plantas en fase de
    crecimiento representa un aumento muy pequeño del nitrógeno total
    disponible en comparación con el que se añade como fertilizante.

    3.  Efectos de la exposición al dióxido de nitrógeno en la salud

         Se han realizado numerosos estudios con objeto de evaluar los
    efectos del NOx para la salud. De los compuestos del NOx, el más
    estudiado ha sido el NO2. El examen de esta sección se concentra en
    el NO2, el NO, el NO2H y el NO3H, mientras que los nitratos se
    mencionan brevemente.

    3.1  Estudios sobre los efectos de los compuestos de nitrógeno en
         animales de experimentación

         La extrapolación a las personas de los datos obtenidos en
    animales tiene componentes tanto cualitativos como cuantitativos. Como
    se señala a continuación de manera resumida, el NO2 produce una serie
    de efectos en varias especies animales, en particular sobre las
    defensas del huésped frente a las enfermedades infecciosas pulmonares,
    en el metabolismo/bioquímica de los pulmones, la función de éstos y
    su estructura. Debido a las analogías fisiológicas, metabólicas y
    estructurales básicas de todos los mamíferos (animales de laboratorio
    y personas), el conjunto de las observaciones realizadas en varias
    especies animales lleva a la conclusión razonable de que el NO2
    podría ocasionar tipos parecidos de efectos en las personas. Sin
    embargo, debido a las diferencias entre las especies de mamíferos, no
    se sabe todavía con exactitud qué exposiciones darían lugar en la
    práctica a esos efectos. Éste es el aspecto de la extrapolación
    cuantitativa. Las limitadas investigaciones sobre la creación de
    modelos relativos al aspecto dosimétrico (es decir, la dosis que
    realmente produce toxicidad en el tejido/célula destinatario) de la
    extrapolación cuantitativa parecen indicar que la distribución de la
    deposición de NO2 en el aparato respiratorio de los animales y las
    personas es análoga, aunque no se dispone todavía de valores adecuados
    que puedan utilizarse para la extrapolación de los animales a las
    personas. Por desgracia, es muy poca la información disponible sobre
    el otro aspecto básico de la extrapolación, la sensibilidad específica
    (es decir, la respuesta de los tejidos de distintas especies a una
    dosis determinada). Así, gracias a los estudios sobre animales
    actualmente disponibles sabemos qué efectos puede tener el NO2 para
    la salud humana. No estamos en condiciones de definir con gran
    precisión los efectos que produce  realmente una dosis determinada de
    NO2 inhalada.

         Teniendo en cuenta lo expuesto, a continuación se resume la base
    de datos sobre la toxicología del NO2 en los animales, de acuerdo con
    las principales clases de efectos y los temas de especial interés.
    Aunque es evidente que los efectos de la exposición al NO2 van más
    allá de los límites de los pulmones, no está clara la interpretación
    de estos efectos sistémicos en relación con el posible riesgo para la
    salud humana. Por consiguiente, no se sigue hablando de ellos aquí,
    sino que se examinan en capítulos posteriores. Aunque las interacciones
    del NO2 y otros contaminantes que lo acompañan, como el O3 y
    el ácido sulfúrico (SO4H2), pueden ser bastante importantes,
    especialmente si se produce sinergia, la base de datos no permite
    todavía llegar a conclusiones a partir de las cuales se puedan evaluar
    las interacciones potenciales en la realidad.

    3.1.1  Mecanismos de acción bioquímicos y celulares de los óxidos de
           nitrógeno

         El NO2 actúa como oxidante fuerte. Los lípidos insaturados se
    oxidan fácilmente, con peróxidos como producto predominante. Tanto el
    ácido ascórbico (vitamina C) como el alpha-tocoferol (vitamina E)
    inhiben la peroxidación de los lípidos insaturados. Cuando el ácido
    ascórbico queda encerrado herméticamente dentro de liposomas de doble
    capa, el NO2 oxida con rapidez el ácido ascórbico englobado. Los
    efectos protectores del alpha-tocoferol y el ácido ascórbico en los
    animales y las personas se deben a la inhibición de la oxidación por
    el NO2. Éste también oxida las proteínas de las membranas. La
    oxidación de los lípidos o las proteínas de las membranas provoca la
    pérdida del control de la permeabilidad celular. Los pulmones de las
    personas y de los animales experimentales expuestos al NO2 tienen
    cantidades mayores de proteínas en el lumen. La aparición de células
    inflamatorias y los cambios en los pulmones se deben a esa acción.

         Las propiedades oxidantes del NO2 también inducen la vía de
    destoxificación de los peróxidos de la glutatión peroxidasa, la
    glutatión reductasa y la glucosa-6-fosfato deshidrogenasa. Tras la
    exposición al NO2, se registra una relación exposición-respuesta en
    el aumento de la vía de destoxificación de los peróxidos en los
    animales.

         El mecanismo de acción del NO es menos claro. Se oxida fácilmente
    a NO2 y luego se produce una peroxidación. Debido a que en las
    exposiciones a NO hay también presente algo de NO2, es difícil
    distinguir los efectos de ambos. El NO actúa como segundo mensajero
    intracelular que modula una gran variedad de enzimas esenciales e
    inhibe su propia producción (por ejemplo, mediante retroinhibición).
    El NO activa la guanilato ciclasa, que a su vez eleva los niveles de
    GMPc intracelular. Un posible mecanismo de acción de los nitratos se
    puede producir por medio de la liberación de histamina de los gránulos
    de los mastocitos. Los contaminantes atmosféricos nitrogenados ácidos,
    en particular el NO3H, pueden actuar alterando el pH intracelular.

         El NPA se descompone en el agua, formando peróxido de hidrógeno.
    Apenas se conoce el mecanismo de acción, pero es probable que haya
    presión oxidativa para el NPA y las sustancias análogas.

         Los nitratos inorgánicos pueden actuar mediante alteraciones del
    pH intracelular. El ión nitrato se transporta a las células alveolares
    de tipo 2 y las acidifica. También moviliza la histamina de los
    mastocitos. El NO2H podría actuar también alterando el pH
    intracelular, pero este mecanismo no está claro.

         No se conocen los mecanismos de acción de los demás óxidos de
    nitrógeno.

         La exposición aguda al NO2 a una concentración de 750 µg/m3
    (0,4 ppm) puede dar lugar a una peroxidación de los lípidos. El
    NO2 puede oxidar los ácidos grasos poliinsaturados de las
    membranas celulares, así como grupos funcionales de proteínas
    (proteínas solubles de la célula, como las enzimas, o bien proteínas
    estructurales, como los componentes de las membranas celulares). Tales
    reacciones de oxidación (con la intervención de radicales libres) son
    un mecanismo mediante el cual el NO2 produce una toxicidad directa
    en células pulmonares. Este mecanismo de acción se ha comprobado en
    estudios con animales, en los que se pone de manifiesto la importancia
    de las defensas antioxidantes de los pulmones, tanto endógenas (por
    ejemplo el mantenimiento de los niveles de glutatión de los pulmones)
    como exógenas (por ejemplo las vitaminas C y E de la alimentación) en
    la protección frente a los efectos del NO2. En numerosos estudios se
    ha observado que diversas enzimas de los pulmones, entre ellas la
    glutatión peroxidasa, la superóxido dismutasa y la catalasa, pueden
    actuar también defendiendo los pulmones del ataque de los oxidantes.

    3.1.2  Efectos en la defensa de los huéspedes

         Aunque la función primaria de las vías respiratorias es asegurar
    un intercambio eficaz de gases, ese sistema orgánico proporciona
    también al cuerpo la primera línea de defensa frente a los agentes
    presentes en la atmósfera, viables y no viables, que se inhalan. En
    una amplia base de datos se pone claramente de manifiesto que la
    exposición al NO2 puede provocar la disfunción de estas defensas del
    huésped, aumentando la susceptibilidad a las enfermedades infecciosas
    de las vías respiratorias. Los parámetros de defensa del huésped
    afectados por el NO2 incluyen la actividad funcional y bioquímica
    de las células de los pulmones, los macrófagos alveolares, la
    competición inmunológica, la susceptibilidad a infecciones de las vías
    respiratorias inducidas experimentalmente y la tasa de eliminación
    mucociliar.

         Los macrófagos alveolares se ven afectados por el NO2. Estas
    células se encargan de mantener la esterilidad de la región pulmonar,
    eliminando las partículas de ella y participando en las funciones
    inmunológicas. Entre los cambios funcionales que se han descrito cabe
    mencionar los siguientes: supresión de la capacidad fagocítica y del
    estímulo de la limpieza de  los pulmones a 560 µg/m3 (0,3 ppm) dos
    horas/día durante 13 días; disminución de la actividad bacteriana a
    4320 µg/m3 (2,3 ppm) durante 17 horas; y una disminución de la
    respuesta al factor de inhibición de la migración a 3760 µg/m3
    (2,0 ppm) ocho horas/día y cinco días/semana durante seis meses. El
    aspecto morfológico de estas células de defensa cambia tras la
    exposición crónica al NO2.

         La importancia de las defensas del huésped se pone de manifiesto
    cuando los animales tienen que hacer frente a infecciones pulmonares
    inducidas en el laboratorio. Los animales expuestos a NO2 sucumben a
    las infecciones bacterianas o víricas de manera dependiente de la
    concentración. También aumenta la mortalidad con la elevación de la
    concentración de NO2 o la duración de la exposición. Tras una
    exposición aguda, se observan efectos a concentraciones de apenas
    3760 µg/m3 (2 ppm). La exposición a concentraciones de sólo
    940 µg/m3 (0,5 ppm) produce efectos en el modelo de infectividad
    después de seis meses.

         La exposición al NO2 modifica tanto el sistema de defensa
    humoral como el celular. En los casos en que se ha investigado el
    sistema inmunitario, se han observado efectos tras una exposición
    breve a concentraciones  9400 µg/m3 (5 ppm). Los efectos son
    complejos, puesto que la dirección del cambio (es decir, el aumento o
    disminución) depende de la concentración de NO2 y de la duración de
    la exposición.

    3.1.3  Efectos de la exposición crónica en la evolución de las
           neumopatías crónicas

         Las personas están crónicamente expuestas al NO2. Por
    consiguiente, dicha exposición se ha estudiado en animales con
    bastante detenimiento, normalmente utilizando métodos morfológicos y/o
    morfométricos. Esta investigación ha demostrado en general que en los
    pulmones se producen diversas alteraciones estructurales, acompañadas
    de otras funcionales. Algunos de estos cambios pueden ser reversibles
    cuando cesa la exposición.

         La función pulmonar de animales experimentales se puede alterar
    tras la exposición crónica al NO2. Después de una exposición a
    7520 µg/m3 (4,0 ppm) de NO2 durante cuatro meses se registró un
    desequilibrio del intercambio de gases, y esto se puso de manifiesto
    en una menor tensión arterial de O2, una disminución del rendimiento
    físico y un aumento del metabolismo anaerobio.

         Aunque el NO2 produce cambios morfológicos en las vías
    respiratorias, la base de datos es a veces confusa, debido a la
    variabilidad cuantitativa y cualitativa de la capacidad de respuesta
    en distintas especies, e incluso en la misma. La rata, que es el
    animal experimental más utilizado en evaluaciones morfológicas de la
    exposición, parece ser relativamente resistente al NO2. La exposición
    de corta duración a concentraciones de 9400 µg/m3 (5,0 ppm) o menores
    tiene en general escasos efectos en la rata, mientras que exposiciones
    similares en el cobaya pueden producir algunos daños en el epitelio
    centriacinar.

         La exposición de más larga duración provoca lesiones en algunas
    especies con concentraciones de sólo 560-940 µg/m3 (0,3-0,5 ppm).
    Éstas se caracterizan por una modificación del epitelio parecida a la
    descrita más arriba, pero con la intervención de vías respiratorias
    más proximales y el engrosamiento del intersticio. Sin embargo, muchos
    de estos cambios desaparecen incluso con una exposición continuada,
    necesitándose una exposición de larga duración a niveles por encima de
    un valor aproximado de 3760 µg/m3 (2,0 ppm) para que aparezcan
    cambios más extensos y permanentes en los pulmones. Algunos efectos
    son relativamente persistentes (por ejemplo la bronquiolitis),
    mientras que otros tienden a ser reversibles y limitados, incluso con
    una exposición continuada. En cualquier caso, parece que tanto en la
    exposición de corta duración como en la larga la respuesta depende más
    de la concentración que del tiempo de exposición.

         Hay pruebas bastante convincentes de que la exposición de
    larga duración de varias especies de animales de laboratorio a
    concentraciones elevadas de NO2 da lugar a lesiones morfológicas en
    los pulmones. En un número limitado de estudios bastante fidedignos se
    ha descrito la destrucción de las paredes alveolares de los pulmones
    de animales como otro criterio esencial para el enfisema humano. A
    partir de estos estudios publicados no se puede determinar la
    concentración más baja de NO2 para la duración más breve de
    exposición que provoca lesiones pulmonares enfisematosas.

    3.1.4  Posibles efectos carcinógenos o cocarcinógenos

         Se ha demostrado que el NO2 tiene una acción mutagénica sobre la
    bacteria  Salmonella, pero dicha acción no se puso de manifiesto en
    un estudio realizado con un cultivo de células de mamífero. En otros
    estudios realizados con cultivos de células se han descubierto
    intercambios entre cromatidios hermanos y roturas de cadenas sencillas
    de ADN. No se han observado efectos genotóxicos  in vivo en relación
    con linfocitos, espermatocitos o células de la médula ósea, aunque en
    dos estudios de inhalación con concentraciones elevadas (50 760 y
    54 400 µg/m3, 27 y 30 ppm) durante 3 y 6 horas, respectivamente, se
    han demostrado dichos efectos en las células pulmonares.

         En la bibliografía no se han encontrado informes publicados de
    estudios sobre el NO2 utilizando bioensayos crónicos clásicos de
    carcinogénesis con animales enteros. Las investigaciones con ratones
    que espontáneamente tenían un índice elevado de tumores eran
    equívocas. En un estudio con una concentración de NO2 de
    18 800 µg/m3 (10 ppm) se detectó un ligero aumento de la frecuencia
    de adenomas pulmonares en una raza de ratones sensible (A/J). Si bien
    se han realizado varias investigaciones de cocarcinogénesis, no se ha
    podido sacar ninguna conclusión debido a problemas de metodología e
    interpretación. Los informes sobre si el NO2 facilita la formación de
    metástasis de tumores en los pulmones son también insuficientes para
    sacar conclusiones. Otras investigaciones se han concentrado en la

    posibilidad de que el NO2 forme nitratos y nitritos que, al
    reaccionar con las aminas del organismo, podían producir nitrosaminas.
    En un pequeño número de estudios parece que se forman nitrosaminas en
    organismos tratados con dosis elevadas de aminas y expuestos a NO2,
    pero en otros estudios se ha señalado que no es probable la formación
    de nitrosaminas.

    3.1.5  Susceptibilidad en función de la edad

         Las investigaciones sobre la dependencia de la edad no son
    suficientes y los resultados hasta ahora son equívocos.

    3.1.6  Influencia de las modalidades de exposición

         En varios estudios toxicológicos realizados con animales se ha
    puesto de manifiesto la relación entre concentración (C) y duración
    (T) de la exposición, indicando que ésta es compleja. En la mayor
    parte de estas investigaciones se ha utilizado el modelo de la
    infectividad.

         En los primeros estudios de C × T se demostró que la concentración
    tenía más efectos en la mortalidad que la duración de la exposición.
    Una evaluación de la toxicidad de la exposición al NO2 no se puede
    definir por la relación C × T.

    3.2  Estudios de exposición humana controlada a óxidos de nitrógeno

         Se han evaluado las respuestas humanas a una serie de compuestos
    de nitrógeno oxidado. Con diferencia, la base de datos más amplia y la
    más adecuada para la evaluación del riesgo es la disponible para
    exposiciones controladas al NO2. La base de datos sobre la respuesta
    humana al NO, NO3H gaseoso, NO2H gaseoso y aerosoles de nitratos
    inorgánicos no es tan amplia. Se han examinado varios subgrupos
    sensibles o potencialmente sensibles, incluidos adolescentes y adultos
    asmáticos, ancianos, y pacientes con neumopatía obstructiva crónica e
    hipertensión pulmonar. El ejercicio durante la exposición aumenta la
    absorción total y altera la distribución del material inhalado dentro
    de los pulmones. La proporción relativa del NO2 depositado en las
    vías respiratorias inferiores aumenta también con el ejercicio. Esto
    puede acentuar los efectos de los compuestos citados anteriormente en
    personas que están en movimiento durante la exposición.

         Como suele ocurrir en la respuesta biológica humana a partículas
    y gases inhalados, la correspondiente al NO2 es variable. Las
    personas sanas tienden a ser menos receptivas a los efectos del NO2
    que las que padecen enfermedades pulmonares. Los asmáticos son
    claramente el grupo más sensible al NO2 entre los estudiados hasta
    ahora. Las personas con neumopatía obstructiva crónica pueden ser más
    sensibles que las sanas, pero tienen una capacidad de respuesta
    limitada al NO2, por lo que son difíciles de evaluar las diferencias

    cuantitativas entre este tipo de pacientes y otras personas. No se
    dispone por el momento de información suficiente para determinar si la
    edad y el sexo desempeñan una función en la respuesta al NO2.

         Las personas sanas pueden detectar el olor del NO2, en algunos
    casos a concentraciones inferiores a 188 µg/m3 (0,1 ppm). En general,
    la exposición a este compuesto no aumentó los síntomas respiratorios
    en ninguno de los grupos sometidos a prueba.

         El NO2 provoca una disminución de la función pulmonar, en
    particular una mayor resistencia al paso del aire en personas sanas en
    reposo sometidas a concentraciones de sólo 4700 µg/m3 (2,5 ppm)
    durante dos horas. Los datos disponibles son insuficientes para
    determinar la naturaleza de la relación concentración-respuesta.

         La exposición de personas no fumadoras sanas en movimiento a
    concentraciones de NO2 de sólo 2800 µg/m3 (1,5 ppm) durante una
    hora o más produce una mayor sensibilización de las vías respiratorias
    a los agentes broncoconstrictores.

         La exposición de asmáticos al NO2 causa, en algunos pacientes,
    una mayor sensibilización de las vías respiratorias a diversos
    mediadores reactivos, incluidos productos químicos colinérgicos e
    histaminérgicos, el SO2 y el aire frío. En presencia de estas
    respuestas parece que influye el procedimiento de exposición, en
    particular el hecho de que ésta tenga lugar con ejercicio o sin él.
    Las respuestas pueden comenzar a concentraciones de apenas 380 µg/m3
    (0,2 ppm). De un metanálisis parece desprenderse que los efectos
    pueden  presentarse a concentraciones incluso más bajas. Sin embargo,
    se ha observado una relación concentración-respuesta inequívoca entre
    350 y 1150 µg/m3 (0,2 a 0,6 ppm).

         Los efectos de esta tendencia general no están claros, pero una
    mayor sensibilización de las vías respiratorias podría producir en
    potencia una respuesta más intensa a los alergenos del aire o un
    recrudecimiento del asma, lo cual posiblemente llevaría a un aumento
    de la medicación o incluso de las hospitalizaciones.

         Puede producirse un incremento moderado de la resistencia de las
    vías respiratorias en pacientes con neumopatía obstructiva crónica
    sometidos a exposiciones breves (15-60 minutos) a concentraciones de
    NO2 de sólo 2800 µg/m3 (1,5 ppm), y también puede observarse una
    disminución en las mediciones espirométricas de la función pulmonar
    (cambio del 3 al 8% en el VEF1 (volumen espiratorio forzado en un
    segundo)) con exposiciones más prolongadas (3 horas) a concentraciones
    de sólo 600 µg/m3 (0,3 ppm).

         La exposición a concentraciones de NO2 superiores a 2800 µg/m3
    (1,5 ppm) puede alterar el número de células inflamatorias y sus
    tipos en las vías distales o los alvéolos. También pueden modificar el

    funcionamiento de las células dentro de los pulmones y la producción
    de mediadores que pueden ser importantes para las defensas pulmonares
    del huésped. El conjunto de cambios en dichas defensas, las
    alteraciones de las células pulmonares y de sus actividades y los
    cambios en los mediadores bioquímicos están en consonancia con los
    hallazgos epidemiológicos de una mayor susceptibilidad del huésped
    relacionada con la exposición al NO2.

         En estudios sobre mezclas de NO2 con otros contaminantes no
    se ha observado que éste aumente la respuesta frente a los demás
    contaminantes presentes por encima del nivel que se detectaría si
    éstos se encontrasen solos. Una excepción importante es la observación
    de que una exposición previa al NO2 de personas sanas en movimiento
    potenciaba los cambios inducidos por el ozono en la sensibilización de
    las vías respiratorias cuando posteriormente se las sometía a una
    exposición al ozono. Esta observación parece poner de manifiesto la
    posibilidad de respuestas retardadas o persistentes al NO2.

         Dentro de la gama de concentraciones de NO2 que puede interesar
    con vistas a la evaluación del riesgo (es decir, 100-600 µg/m3), los
    datos disponibles no permiten determinar las características de la
    relación concentración-respuesta para cambios drásticos de la función
    pulmonar, la capacidad de respuesta de las vías respiratorias a
    agentes broncoconstrictores o los síntomas.

         A partir de un efecto a 400 µg/m3 y de la posibilidad de efectos
    a niveles más bajos, tomando como base un metanálisis, se recomienda a
    título indicativo para un período breve un promedio diario de una hora
    a una concentración máxima de NO2 de 200 µg/m3 (0,11 ppm).

         Se sabe que el NO es un segundo mensajero endógeno importante en
    varios sistemas del organismo. La inhalación de concentraciones de NO
    superiores a 6000 µg/m3 ( 5 ppm) puede producir vasodilatación en la
    circulación pulmonar sin afectar a la sistémica. No se ha determinado
    la concentración eficaz mínima. La información sobre la función
    pulmonar y las defensas de los pulmones del huésped después de la
    exposición al NO es demasiado limitada para que por el momento se
    puedan sacar conclusiones. No se ha informado de reacciones
    secundarias tras la utilización, en aplicaciones clínicas, de
    concentraciones relativamente altas (> 40 000 µg/m3) durante
    períodos breves (<1 hora).

         Concentraciones de ácido nítrico del orden de 250-500 µg/m3
    (97-194 ppmm) pueden producir cierta respuesta en la función pulmonar
    de adolescentes asmáticos, pero no en adultos sanos.

         De la limitada información disponible sobre el NO2H cabe deducir
    que puede causar inflamación ocular a 760 µg/m3 (0,40 ppm). En la
    actualidad no existen datos publicados sobre respuesta pulmonar del
    ser humano al NO2H.

         Los limitados datos sobre nitratos inorgánicos de que se dispone
    indican que los aerosoles de nitratos a una concentración de
    7000 µg/m3 o inferior no tienen efectos en la función pulmonar.

    3.3  Estudios epidemiológicos sobre el dióxido de nitrógeno

         Los estudios epidemiológicos sobre los efectos de los óxidos de
    nitrógeno en la salud se han concentrado principalmente en el NO2. Se
    han realizado numerosos estudios epidemiológicos en espacios cerrados
    y abiertos para determinar los efectos del NO2 en la salud. En
    general se consideran dos mediciones del estado de salud en la
    exposición al NO2: las mediciones de la función pulmonar y los
    síntomas y enfermedades de las vías respiratorias.

         Las pruebas de estudios individuales de los efectos del NO2
    sobre los síntomas y las enfermedades de las vías respiratorias
    inferiores de niños en edad escolar son algo confusas. Se examinó la
    concordancia de los estudios y las pruebas se resumieron en un
    análisis cuantitativo combinado (metanálisis) de ellos. En la mayor
    parte de los estudios realizados en espacios cerrados se observó una
    mayor morbilidad de las vías respiratorias inferiores en niños
    asociada a exposiciones prolongadas al NO2. En la mayoría de los
    estudios en los que se notificaron los niveles de NO2 las
    concentraciones medias semanales en los dormitorios fueron
    predominantemente de 15 a 122 µg/m3 (0,008 y 0,065 ppm). La
    combinación de los estudios en espacios cerrados como si los
    resultados finales fueran semejantes da una razón de posibilidades
    estimada de 1,2 (límites de confianza del 95 por ciento de 1,1 y 1,3)
    para el efecto de un aumento de la concentración de NO2 de
    28,3 µg/m3 (0,015 ppm) en la morbilidad de las vías respiratorias
    inferiores. Esto indica que, teniendo en cuenta la hipótesis hechas
    para el análisis combinado, a cada aumento de 28,3 µg/m3 (0,015 ppm)
    en la exposición media estimada al NO2 durante dos semanas
    corresponde un aumento de alrededor del 20 por ciento en las
    posibilidades de síntomas y enfermedades en las vías respiratorias
    inferiores. Así pues, las pruebas combinadas confirman los efectos de
    la exposición estimada para el NO2 en los síntomas y las enfermedades
    de las vías respiratorias inferiores de niños con edades comprendidas
    entre cinco y 12 años.

         En estudios individuales con niños de dos años o menos en
    espacios cerrados no se encontró una relación uniforme entre las
    estimaciones de la exposición al NO2 y la prevalencia de síntomas
    y enfermedades de las vías respiratorias. Tomando como base un
    metanálisis de estos estudios realizados con niños en espacios
    cerrados, en función de las hipótesis hechas para el metanálisis, la
    razón combinada de posibilidades para el aumento de las enfermedades
    respiratorias con incrementos de 28,2 µg/m3 (0,015 ppm) de NO2 fue
    de 1,9, con un intervalo de confianza del 95 por ciento de 0,95 a
    1,26, siendo las concentraciones semanales medias de NO2 en los

    dormitorios predominantemente de 9,4 a 94 µg/m3 (0,005 y 0,050 ppm).
    El aumento de riesgo fue muy pequeño y no en todos los estudios se
    notificaron resultados uniformes. No se puede concluir que las pruebas
    indiquen un efecto en los niños pequeños comparable al observado en
    otros de más edad. No están claras las razones de estas diferencias
    relacionadas con la edad.

         En los estudios de medición del NO2 se obtuvo una razón de
    posibilidades superior a las estimaciones sustitutivas, lo que
    concuerda con un efecto de error de medición. El efecto de haber
    ajustado covariantes como la situación socioeconómica, la condición de
    fumador y el sexo fue que en los estudios en que se ajustó una
    variante particular se encontraron razones de posibilidades más altas
    que en los otros.

         Si bien en muchos de los estudios epidemiológicos con niveles
    conocidos de NO2 sólo se realizaron mediciones durante una o
    dos semanas, estos niveles se utilizaron para caracterizar las
    exposiciones de los niños durante un período mucho más largo. El
    cuestionario normalizado sobre síntomas respiratorios utilizado en la
    mayor parte de estos estudios resume la información sobre el estado de
    salud durante todo un año. La diferencia de 28,2 µg/m3 (0,015 ppm) en
    los niveles de NO2 utilizados en los metanálisis corresponde a la
    diferencia en la exposición media anual en el hogar entre las
    cocinas de cocinar de gas y eléctricas. En algunos estudios las
    concentraciones de NO2 se midieron sólo durante el invierno y puede
    haber una sobreestimación de la exposición media anual. Esto podría
    inducir a subestimar el efecto sobre la salud de la diferencia de
    28,2 µg/m3 (0,015 ppm) en la exposición anual al NO2. En un estudio
    basado en dicha exposición en el hogar, medida tanto en invierno como
    en verano, se puso de manifiesto un efecto sobre la salud mayor que en
    muchos de los otros estudios. Se desconoce el período de exposición
    verdaderamente importante desde el punto de vista biológico, pero
    estas exposiciones se prolongaban durante un período largo, que podía
    durar incluso toda la vida del niño.

         De las investigaciones actuales no es posible deducir una
    relación clara entre el NO2 de los espacios abiertos y la salud de
    las vías respiratoria. Hay algunas pruebas que demuestran que la
    duración de estas enfermedades puede aumentar en ambientes con
    concentraciones de NO2 más altas. Una dificultad importante para el
    análisis de los estudios realizados en estos espacios radica en la
    distinción entre los posibles efectos debidos al NO2 y los de otros
    contaminantes que lo acompañan.

         Hay que considerar varios aspectos dudosos a la hora de
    interpretar los estudios y metanálisis expuestos más arriba. El error
    en la medición de la exposición es posiblemente uno de los problemas
    metodológicos más importantes en los estudios epidemiológicos del
    NO2. Si bien hay pruebas de que los síntomas están relacionados con

    indicadores de la exposición al NO2, la calidad de estas estimaciones
    de la exposición puede ser insuficiente para determinar una relación
    cuantitativa entre la exposición y los síntomas. En la mayor parte de
    los estudios en que se midió la exposición al NO2 se hizo sólo por
    períodos de una a dos semanas y se dieron los valores como promedios.
    En pocos de los estudios se intentó relacionar los efectos observados
    con las modalidades de exposición (por ejemplo, niveles máximos
    transitorios de NO2). Además, es posible que la concentración de NO2
    medida no sea una dosis biológicamente importante; la estimación de la
    exposición real exige el conocimiento de los tipos de contaminantes,
    sus niveles y las pautas correspondientes de la actividad humana. Sin
    embargo, los datos disponibles sobre actividad y aerométricos en los
    que se examinen dichos factores son muy limitados. La extrapolación a
    posibles pautas de exposición ambiental es difícil. Además, aunque el
    nivel de semejanza y de elementos comunes entre las medidas de los
    resultados en los estudios del NO2 proporcione cierta confianza en su
    uso en el análisis cuantitativo, los síntomas y las enfermedades
    combinados son en cierto sentido diferentes y podrían reflejar de
    hecho procesos básicos distintos. Así pues, hay que ser prudentes a la
    hora de interpretar los resultados del metanálisis.

         En otros estudios epidemiológicos se ha tratado de relacionar
    alguna medida de la exposición al NO2 en espacios cerrados y/o
    abiertos con cambios en la función pulmonar. Estos cambios fueron
    marginalmente significativos. En la mayoría de los estudios no se
    encontró efecto alguno, lo que corrobora los datos obtenidos en los
    estudios de exposición humana controlada. Sin embargo, no hay pruebas
    epidemiológicas suficientes que permitan sacar conclusiones sobre los
    efectos a largo o corto plazo del NO2 en la función pulmonar.

         A partir de un nivel básico de 15 µg/m3 (0,008 ppm) y del hecho
    de que con un nivel adicional de 28,2 µg/m3 (0,015 ppm) o más se
    producen efectos negativos considerables en la salud, se propone un
    valor orientativo anual de 40 µg/m3 (0,023 ppm). Este valor evitará
    las exposiciones más graves. Hay que subrayar el hecho de que no se
    haya determinado todavía un nivel sin efectos para concentraciones de
    exposición al NO2 subcrónicas o crónicas.

    3.4  Valores orientativos basados en la salud para el dióxido de
         nitrógeno

         A partir de los estudios de exposición humana controlada, se
    recomienda como valor orientativo en períodos breves una concentración
    máxima diaria media de NO2 de 200 µg/m3 (0,11 ppm) durante una hora.
    El valor orientativo para períodos prolongados, basado en estudios
    epidemiológicos del aumento del riesgo de enfermedades respiratorias
    en niños, es de un promedio anual de 40 µg/m3 (0,023 ppm).
    


    See Also:
       Toxicological Abbreviations
       Nitrogen, oxides of (EHC 4, 1977)