United States
Environmental Protection
Agency
Environmental Criteria and
Assessment Office
Research Triangle Park, NC 27711
EPA/60Q/8-84/Q2QbF
August 1 986
Research and Development
Air Quality
Criteria for
Ozone and Other
Photochemical
Oxidants
Volume II of V

-------

-------
                             EPA/600/8-84/020bF
                                   August 1986
    Air Quality Criteria
  for Ozone and Other
Photochemical Oxidants

      Volume It of V
    Environmental Criteria and Assessment Office
    Office of Health and Environmental Assessment
       Office of Research and Development
      U.S. Environmental Protection Agency
       Research Triangle Park, IM.G. 27711

-------
                                  DISCLAIMER


     This document has been reviewed ih accordance with U.S. Environmental
Protection Agency policy and approved for publication.  Mention of trade'
names or commercial products does not constitute endorsement or recommendation
for use.           "                              '      L
                                     IT

-------
                                   ABSTRACT


     Scientific information is presented and evaluated relative to the, health
and welfare effects associated with exposure to ozone and other phbtqqhemical
oxidants.   Although it is not intended as a complete and detailed'literature
review, the document covers pertinent literature through early 1986.       ;

     Data on health and welfare effects are emphasized, but additional infor-
mation is provided for understanding the nature of the oxidant pollution pro-
blem and for evaluating the reliability of effects data as well as their
relevance to potential exposures to ozone and other oxidants at concentrations
occurring in ambient air.  Information is provided on the following exposure-
related topics:  nature, source, measurement, and concentrations of precursors
to ozone and other photochemical oxidants; the formation of ozone and other
photochemical oxidants and their transport once.formed; the properties, chem-
istry, and measurement of ozone and other photochemical oxidants; and the
concentrations of ozone and other photochemical oxidants that are typically
found in ambient air.

     The specific areas addressed by chapters on health and welfare effects
are the toxicological appraisal of effects of ozone and other oxidants; effects
observed in controlled human exposures; effects observed in field and epidemi-
ological studies; effects on vegetation seen in field and controlled exposures;
effects on natural and agroecosystems; and effects on nonbiological materials
observed in field and chamber studies.
                                     m

-------
                        AIR QUALITY CRITERIA FOR OZONE
                       AND OTHER PHOTOCHEMICAL OXIDANTS
                                                                           Page

VOLUME I
   Chapter 1.    Summary and Conclusions 	     1-1

VOLUME II
   Chapter 2.    Introduction	     2-1
   Chapter 3.    Properties, Chemistry, and Transport of Ozone and
                 Other Photochemical Oxidants and Their Precursors 	     3-1
   Chapter 4.    Sampling and Measurement of Ozone and Other
                 Photochemical Oxidants and Their Precursors .	.....     4-1
   Chapter 5.    Concentrations of Ozone and Other Photochemical
                 Oxidants in Ambient Air	     5-1

VOLUME III
   Chapter 6.    Effects of Ozone and Other Photochemical Oxidants
                 on Vegetation 	     6-1
   Chapter 7.    Effects of Ozone on Natural Ecosystems and Their
                 Components 	     7-1
   Chapter 8.    Effects of Ozone and Other Photochemical Oxidants
                 on Nonbiological Materials	     8-1

VOLUME IV
   Chapter 9.    Toxicological Effects of Ozone and Other
                 Photochemical Oxidants	     9-1

VOLUME V
   Chapter 10.   Controlled Human Studies of the Effects of Ozone
                 and Other Photochemical Oxidants	     10-1
   Chapter 11.   Field and Epidemic!ogical Studies of the Effects
                 of Ozone and Other Photochemical Oxidants 	     11-1
   Chapter 12.   Evaluation of Health Effects Data for Ozone and
                 Other Photochemical Oxidants	     12-1
                                     iv

-------
                               TABLE OF CONTENTS
LIST OF TABLES 	    xi
LIST OF FIGURES	     xiv
LIST OF ABBREVIATIONS AND SYMBOLS 	    xvii
AUTHORS, CONTRIBUTORS, AND REVIEWERS	    xxii

2.   INTRODUCTION	 ,    2-1
     2.1  PURPOSE AND LEGISLATIVE BASIS OF THIS DOCUMENT 	     2-1
     2.2  THE OXIDANT PROBLEM	     2-2
     2.3  SCOPE AND ORGANIZATION OF THIS DOCUMENT 	     2-4
     2.4  REFERENCES 	     2-7

3.   PROPERTIES, CHEMISTRY, AND TRANSPORT OF OZONE AND OTHER
     PHOTOCHEMICAL OXIDANTS AND THEIR PRECURSORS	:	     3-1
     3.1  INTRODUCTION	     3-1
     3.2  DESCRIPTIONS AND PROPERTIES OF OXIDANTS AND THEIR
          PRECURSORS	     3-2
          3.2.1  Ozone and Other Photochemical Oxidants 	     3-2
                 3.2.1.1  Ozone	     3-2
                 3.2.1.2  Peroxyacetyl Nitrate 	     3-3
                 3.2.1.3  Hydrogen Peroxide	     3-4
                 3.2.1.4  Formic Acid	     3-7
          3.2.2  Organic Precursors 	     3-7
                 3.2.2.1  Hydrocarbons	     3-8
                 3.2.2.2  Aldehydes	     3-10
                 3.2.2.3  Other Organic Compounds	     3-11
                 3.2.2.4  Volatility and Reactivity	".     3-11
          3.2.3  Nitrogen Oxides 	     3-14
     3.3  ATMOSPHERIC CHEMICAL PROCESSES:  FORMATION AND TRANSFORMATION
          OF OZONE AND OTHER PHOTOCHEMICAL OXIDANTS 	     3-15
          3.3.1  Inorganic Reactions	     3-16
                 3.3.1.1  Formation of Ozone:  The NO-N02-03 Cycle ....     3-16
                 3.3.1.2  Formation of Radical Intermediates 	     3-17
                 3.3.1.3  Termination Reactions 	     3-21
                 3.3.1.4  Reactions Involving Nitrous Acid 	     3-22
                 3.3.1.5  Reactions Involving Nitric Acid and
                          Dinitrogen Pentoxide 	     3-23
          3.3.2  Organic Reactions 	     3-24
                 3.3.2.1  Reactions with Hydroxyl Radicals 	     3-25
                 3.3.2.2  Reactions with Ozone	     3-34
                 3.3.2.3  Reactions with Nitrate Radicals 	     3-41
          3.3.3  Atmospheric Lifetimes of Organic Compounds 	     3-44
          3.3.4  Atmospheric Reactions of Peroxyacetyl Nitrate 	     3-45
          3.3.5  Role of Ozone in Aerosol Formation 	     3-47
                 3.3.5.1  Formation of Sulfate Aerosol 	     3-47
                 3.3.5.2  Formation of Nitrate Aerosol 	     3-48
                 3.3.5.3  Formation of Organic Aerosols 	     3-49

-------
                          TABLE.OF CONTENTS
                             (continued)
                                                                      Page
     3.3.6  Role of Ozone and Other Photochemical  Oxidants in the
            Acidification of Rain 	     3-50
            3.3.6.1  Reactions of Ozone in Aqueous Droplets 	     3-50
            3.3.6.2  Reactions of Hydrogen Peroxide in Aqueous
                     Droplets	     3-51
            3.3.6.3  Reactions of Formic Acid in Aqueous
                     Droplets 	     3-54
3.4  METEOROLOGICAL AND CLIMATOLQGICAL PROCESSES 	     3-54
     3.4.1  Atmospheric Mixing	     3-54
     3.4.2  Wind Speed and Mixing	*•	     3-61
     3.4.3  Effects of Sunlight and Temperature 	     3-66
     3.4.4  Transport of Ozone and Other Oxidants and Their
            Precursors	     3-68
     3.4.5  Surface Scavenging in Relation to Transport ......	     3-74
     3.4.6  Stratospheric-Tropospheric Ozone Exchange	     3-75
     3.4.7  Stratospheric Ozone at Ground Level	     3-80
     3.4.8  Background Ozone from Photochemical Reactions 	     3-84
3.5  PRECURSORS TO OZONE AND OTHER PHOTOCHEMICAL OXIDANTS 	     3-88
     3.5.1  Sources and Emissions	     3-88
            3.5.1.1  Manmade Sources and Emissions	     3-88
            3.5.1.2  Natural Sources and Emissions	     3-98
     3.5.2  Representative Concentrations of Ozone Precursors
            in Ambient Air	     3-104
            3.5.2.1  Concentrations of Nonmethane Organic
                     Compounds in Ambient Air	     3-104
            3.5.2.2  Concentrations of Nitrogen Oxides in
                     Ambient Air	     3-108
3.6  SOURCE-RECEPTOR (OXIDANT-PRECURSOR) MODELS	     3-110
     3.6.1  Definitions, Descriptions, and Use	     3-112
            3.6.1.1  Statistical Models 			     3-112
            3.6.1.2  Trajectory Models	     3-113
            3.6.1.3  Fixed-Grid Models	     3-119
            3.6.1.4  Box Model s	     3-121
     3.6.2  Validation and Sensitivity Analyses for Dynamic
            Models	     3-121
3.7  SUMMARY	     3-124
     3.7.1  Descriptions and Properties of Ozone and Other
            Photochemical Oxidants	     3-124
     3.7.2  Nature of Precursors to Ozone and Other Photochemical
            Oxidants	     3-125
     3.7.3  Atmospheric Reactions of Ozone and Other Oxidants
            Including Their Role in Aerosol Formation 	     3-127
            3.7.3.1  Formation and Transformation of Ozone
                     and Other Photochemical Oxidants 		     3-127
            3.7.3.2  Atmospheric Chemical Processes Involving
                     Ozone	.'		...     3-128
                                 VI

-------
                               TABLE  OF  CONTENTS
                                  (continued)
                 3.7.3.3  Atmospheric Reactions  of PAN,  H202,  and
                          HCOOH	     3-129
          3.7.4  Meteorological and Climatological  Processes  	     3-130
                 3.7.4.1  Atmospheric Mixing	     3-131
                 3.7.4.2  Wind Speed and Direction	     3-132
                 3.7.4.3  Effects of Sunlight and Temperature  	     3-133
                 3.7.4.4  Transport of Ozone and Other Oxidants
                          and Their Precursors		     3-133
                 3.7.4.5  Stratospheric-Tropospheric Ozone  Exchange ...     3-135
                 3.7.4.6  Stratospheric Ozone at Ground  Level  	     3-135
                 3.7.4.7  Background Ozone from  Photochemical
                          Reactions		     3-136
          3.7.5  Sources,  Emissions, and Concentrations  of  Precursors
                 to Ozone  and Other Photochemical Oxidants  	     3-137
                 3.7.5.1  Sources and Emissions  of Precursors	     3-137
                 3.7.5.2  Representative Concentrations  in  Ambient
                          Air	     3-138
          3.7.6  Source-Receptor (Oxidant-Precursor) Models	     3-140
                 3.7.6.1  Trajectory Models	     3-140
                 3.7.6.2  Fixed-Grid Models		     3-141
                 3.7.6.3  Box Models	     3-142
                 3.7.6.4  Validation and Sensitivity Analyses  for
                          Dynamic Models	     3-142
     3.8  REFERENCES	     3-142

4.   SAMPLING AND MEASUREMENT OF OZONE AND OTHER PHOTOCHEMICAL
     OXIDANTS AND THEIR PRECURSORS 	    4-1
     4.1  INTRODUCTION	    4-1
     4.2  QUALITY ASSURANCE AND OTHER SAMPLING FACTORS IN
          MONITORING FOR OZONE	    4-2
          4.2.1  Quality Assurance in Ambient Air Monitoring for
                 Ozone	    4-2
          4.2.2  Sampling Factors in Ambient Air Monitoring for
                 Ozone 	    4-3
                 4.2.2.1  Sampling Strategies and Air Monitoring
                          Needs	:	    4-4
                 4.2.2.2  Air Monitoring Site Selection	    4-4
          4.2.3  Measurement Methods for Total Oxidants  and Ozone	    4-6
                 4.2.3.1  Total Oxidants	    4-6
                 4.2.3.2  Ozone 	    4-7
          4.2.4  Generation and Calibration Methods for  Ozone 	    4-13
                 4.2.4.1  Generation	    4-13
                 4.2.4.2  Calibration.	    4-14
          4.2.5  Relationship between Methods for Total  Oxidants
                 and Ozone 	    4-21
                 4.2.5.1  Predicted Relationship 	    4-22
                                    Vll

-------
                               TABLE OF CONTENTS
                                  (continued)
                                                                           Page
                 4.2.5.2  Empirical  Relationship Determined from
                          Simul taneous Measurements	 —	    4-24
          4.2.6  Methods for Sampling and Analysis of Peroxyacetyl
                 Nitrate and Its Homologues 	    4-33
                 4.2.6.1  Introduction	    4-33
                 4.2.6.2  Analytical Methods for PAN 	    4-34
                 4.2.6.3  Generation and Calibration of PAN	    4-39
                 4.2.6.4  Methods of Analysis of Higher Homologues  .....    4-42
          4.2.7  Methods for Sampling and Analysis of Hydrogen
                 Peroxide			    4-43
                 4.2.7.1  Introduction	    4-43
                 4.2.7.2  Sampling 	    4-44
                 4.2.7.3  Measurement	,	    4-45
                 4.2.7.4  Generation and Calibration Methods	    4-48
     4.3  SAMPLING,  MEASUREMENT, AND CALIBRATION METHODS FOR
          PRECURSORS TO OZONE AND OTHER PHOTOCHEMICAL OXIDANTS .....	    4-49
          4.3.1  Nonmethane Organic  Compounds 	    4-50
                 4.3.1.1  Nonmethane Hydrocarbons		    4-50
                 4.3.1.2  Aldehydes	,	    4-61
                 4.3.1.3  Other Oxygenated Organic Species 	    4-66
          4.3.2  Nitrogen Oxides		..........    4-66
                 4.3.2.1  Measurement Methods for N02 and NO 	    4-67
                 4.3.2.2  Sampling Requirements	    4-70
                 4.3.2.3  Cal ibration	    4-70
     4.4  SUMMARY	    4-71
          4.4.1  Sampling and Measurement of Ozone and Other
                 Photochemical Oxidants	    4-71
                 4.4,1.1  Quality Assurance and Sampling 	    4-72
                 4.4.1.2  Measurement Methods for Total Oxidants
                          and Ozone  	    4-72
                 4.4.1.3  Calibration Methods	    4-74
                 4.4.1..4  Relationships of Total Oxidants and Ozone
                          Measurements	    4-77
                 4.4.1.5  Methods for Sampling and Analysis of Peroxy-
                          acetyl Nitrate and Its Homologues	     4-78
                 4.4.1.6  Methods for Sampling and Analysis of
                          Hydrogen Peroxide 	    4-80
          4.4.2  Measurement of Precursors to Ozone and Other
                 Photochemical Oxidants		    4-82
                 4.4.2.1  Nonmethane Organic Compounds	    4-82
                 4.4.2.2  Nitrogen Oxides	    4-84
     4.5  REFERENCES	    4-86

5.   CONCENTRATIONS OF OZONE AND OTHER PHOTOCHEMICAL OXIDANTS
     IN AMBIENT AIR	    5-1
     5.1  INTRODUCTION 	    5-1
     5.2  TRENDS IN NATIONWIDE OZONE CONCENTRATIONS	    5-3


                                    vi ii

-------
                          TABLE OF CONTENTS
                             (continued)


                                                                      Page


5,3  OVERVIEW OF OZONE CONCENTRATIONS IN URBAN AREAS 	    5-7
5.4  OVERVIEW OF OZONE CONCENTRATIONS IN NONURBAN AREAS	    5-16
     5.4.1  National Air Pollution Background Network (NAPBN) 	    5-16
     5.4.2  Sulfate Regional Experiment Sites (SURE)	,	    5-25
5.5  VARIATIONS IN OZONE CONCENTRATIONS:  DATA FROM SELECTED
           AND          SITES	    5-25
     5.5.1  Temporal Variations in Ozone Concentrations	    5-28
            5.5.1.1  Diurnal Variations in Ozone Concentrations ...    5>-28
            5.5.1.2  Seasonal Variations in Ozone Concentrations ..    5-42
            5.5.1.3  Weekday-Weekend Variations in Ozone
                     Concentrations 	    5-48
     5.5.2  Spatial Variations in Ozone Concentrations 	    5-49
            5.5.2,1  Urban Versus Nonurban Variations 	    5-49
            5.5.2.2  Intracity Variations	    5-50
            5.5.2.3  Indoor-Outdoor Concentration Ratios 	    5-55
            5.5.2.4  Altitudinal and Latitudinal Variations  	    5-60
            5.5.2.5  Vertical Gradients at Ground Level 	    5-69
5.6  CONCENTRATIONS OF PEROXYACETYL NITRATE AND PEROXYPROPIONYL
     NITRATE IN AMBIENT AIR	    5-71
     5.6.1  Introduction	    5-71
     5.6.2  Historical Data	    5-72
     5.6.3  Ambient Air Concentrations of PAN and Its
            Homo!ogues i n Urban Areas 	    5-* 74
     5.6.4  Ambient Air Concentrations of PAN and Its
            Homol ogues i n Nonurban Areas	    5^80
     5.6.5  Temporal Variations in Ambient Air
            Concentrations of Peroxyacetyl Nitrate	    5-83
            5.6.5.1  Diurnal Patterns	    5-83
            5.6.5.2  Seasonal Patterns 		    5-89
     5.6.6  Spatial Variations in Ambient Air Concentrations
            of Peroxyacetyl Nitrate 	    5-89
            5.6.6.1  Urban-Rural Gradients and Transport of  PAN ...    5-89
            5.6.6.2  Intracity Variations 	    5-93
            5.6.6.3  Indoor-Outdoor Ratios of PAN Concentrations ..    5-93
5.7  CONCENTRATIONS OF OTHER PHOTOCHEMICAL OXIDANTS IN
     AMBIENT AIR	    5-93
5.8  SUMMARY	    5-95
     5.8.1  Ozone Concentrations in Urban Areas .....	......	    5-98
     5.8.2  Trends  in Nationwide Ozone Concentrations	...,..,    5-101
     5.8.3  Ozone Concentrations i n Nonurban Areas	    5" 103
     5.8.4  Diurnal and Seasonal Patterns in Ozone
            Concentrati ons	    S-'IOS
     5.8.5  Spatial Patterns in Ozone Concentrations 	    5-^107
            5.8.5.1  Urban-Nonurban Differences in Ozone
                     Concentrati ons 	    5-107
            5.8.5.2  Geographic, Vertical, and Altitudinal
                     Variations in Ozone Concentrations 	    5-109

                                ix

-------
                               TABLE OF CONTENTS
                                  (continued)
                                                                           Page
                 5.8.5.3  Other Spatial Variations in Ozone
                          Concentrations 	    5-112
          5.8.6  Concentrations and Patterns of Other Photochemical
                 Oxidants	    5-112
                 5.8.6.1  Concentrations 	    5-112
                 5.8.6.2  Patterns..	    5-114
          5.8.7  Relationship Between Ozone and Other Photochemical
                 Oxi dants	    5-114
5.9  REFERENCES	    5-116

-------
                                LIST OF TABLES


Table                                                                    Page

 3-1  Physical properties of ozone	,	    3-3
 3-2  Physical properties of peroxyacetyl nitrate		    3-5
 3-3  Infrared absorptivities of peroxyacetyl  nitrate 	    3-6
 3-4  Physical properties of hydrogen peroxide	    3-6
 3-5  Physical and chemical properties of nitric oxide and
      nitrogen dioxide			    3-15
 3-6  Calculated lifetimes of selected organic compounds resulting
      from atmospheric loss by reaction with 03 and OH and N03
      radicals 	".		    3-44
 3-7  Published episodes of transport of stratospheric ozone to
      ground level	,....    3-82
 3-8  Emissions of VOC by decade, 1940 through 1980 	    3-90
 3-9  Emissions of NO  by decade, 1940 through 1980	    3-93
3-10  Yearly quantities of motor vehicle fuels sold in the
      United States for highway use, 1980 through 1983 	    3-97
3-11  Summary of NO  emissions from mobile sources	    3-98
3-12  Area-wide biogenic emission fluxes	    3-101
3-13  Global estimates of nitrogen transformation	    3-103
3-14  Nonmethane hydrocarbon concentrations measured between 6:00
      and 9:00 a.m. in various United States cities 	    3-106
3-15  Nonmethane hydrocarbon concentrations measured in nonurban
      atmospheres			    3-107
3-16  Average 6:00 to 9:00 a.m. NO  concentrations and HC/NO
      ratios in urban areas	    3-109

 4-1  Performance specifications for automated methods of ozone
      analysis	    4-9
 4-2  List of designated reference and equivalent methods of
      ozone analysis 	—	    4-10
 4-3  Factors for intercomparison of data calibrated by UV
      photometry versus KI colorimetry	,	    4-16
 4-4  Response of NBKI reagent and Mast  meter to various
      oxidants	    4-23
 4-5  Comparison of corrected instrument readings to colorimetric
      oxidant readings during atmospheric sampling 	    4-27
 4-6  Summary of parameters used in determination of PAN by
      GC-ECD			    4-35
 4-7  Infrared absorptivities of peroxyacetyl nitrate
      (Base 10) 	    4-38
 4-8  Measurement methods for hydrogen peroxide  	    4-46
 4-9  Percentage difference from known concentrations of
      nonmethane hydrocarbons obtained by sixteen users	    4-52
4-10  Problems associated with gathering NMOC data with
      automated analyzers and recommendations for reducing
      these effects	    4-53
4-11  Summary of advantages and disadvantages of primary
      collection media for NMOC analysis	    4-58
                                     XI

-------
                                LIST OF TABLES
                                  (continued)
Table
4-12  GC/continuous NMOC analyzer comparisons, least-squares
      regressions 	    4-62
4-13  Summary of ozone monitoring techniques ...	    4-73
4-14  Ozone calibration techniques 	    4-76

5-1   Second-highest 1-hr ozone concentrations reported for
      Standard Metropolitan Statistical Areas having
      populations > 0.5 million, 1981 through 1983 	    5-13
5-2   Annual ozone summary statistics for three sites of the
      National Air Pollution Background Network	    5-18
5-3   Concentrations of ozone during 6-day period of high
      values at NAPBN site in Mark Twain National Forest,
      Missouri, 1979	    5-20
5-4   Percent!* le distributions of ozone concentrations at
      sites of National Air Pollution Background Network,
      aggregated by quarter across several years	    5-22
5-5   Summary of ozone concentrations measured at Sulfate
      Regional Experiment (SURE) nonurban stations, August
      through December 1977	    5-27
5-6   Number of consecutive-day exposures or respites when ,the
      daily 1-hr maximum ozone concentration was >_ 0.06 ppm,
      in four cities (April through September, 1979 through 1981) 	    5-38
5-7   Number of consecutive-day exposures or respites when the
      daily 1-hr maximum ozone concentration was >_ 0.12 ppm, in
      four cities (April through September, 1979 through 1981) 	    5-39
5-8   Number of consecutive-day exposures or respites when the
      daily 1-hr maximum ozone concentration was > 0.18 ppm, in
      four cities (April through September, 1979 through 1981) 	    5-40
5-9   Number of consecutive-day exposures or respites when the
      daily 1-hr maximum ozone concentration was >_ 0.24 ppm, in
      four cities (April through September, 1979 through 1981) 	    5-41
5-10  Ozone concentrations at sites in and around New Haven,
      Connecticut, 1976	,    5-52
5-11  Quarterly maximum 1-hour ozone values at sites in and
      around New Haven, Connecticut, 1976 ............................    5-52
5-12  Peak ozone concentrations at eight sites in New York
      City and adjacent Nassau County, 1980	    5-53
5-13  Summary of reported indoor-outdoor ozone ratios 	    5-57
5-14  Comparison of ozone concentrations at three different
      elevations, High Point Mountain, NJ, and at Bayonne, NJ,
      July 1975			    5-68
5-15  Means and standard errors of ozone concentrations measured
      over 4 years at two sampling heights at three stations in
      the rural, upper-midwestern United States 	    5-70
5-16  Summary of concentrations of peroxyacetyl nitrate in
      ambient air in urban areas of the United States 	    5-75
                                    XII

-------
                                LIST OF TABLES
                                  (continued)


Tab! e                                                                    Paige

5-17  Relationship of ozone and peroxyacetyl nitrate at urban
      and suburban sites in the United States 	   5-79
5-18  Ambient air measurements of peroxypropionyl nitrate (PPN)
      concentrations by electron capture gas chromatography
      at urban sites in the United States	   5-81
5-19  Concentrations of peroxyacetyl and peroxypropionyl nitrates
      in Los Angeles, Oakland, and Phoenix, 1979		   5-82
5-20  Concentrations in ambient air of peroxyacetyl and
      peroxypropionyl nitrates and ozone at nonurban remote sites
      in the United States	   5-84
5-21  PAN and ozone concentrations in ambient air, New
      Brunswick, N.J., for September 25, 1978, to May 10, 1980 	   5-91
5-22  Intracity variations in peak ozone and PAN concentrations
      in Houston, October 26 and 27, 1977	   5-94
5-23  Concentrations of hydrogen peroxide in ambient air at
      urban and nonurban sites			   5-97
5-24  Second-highest ozone concentrations among daily maximum
      1-hr values in 1983 in Standard Metropolitan Statistical
      Areas with populations >^ 1 million, given by census divisions
      and regions			   5-99
                                     xm

-------
                                  LIST OF FIGURES


  F1gure                             .                                        Page

  3-1   Experimental time-concentration profiles for propene, NO,
        N02s OB, HCHO, and PAH for an irradiated NO -propene-air
        mixture 	'.....	     3-18 .
  3-2   Reaction scheme for OH radical-initiated oxidation of propene
        in the presence of NO	     3-28
  3-3   Rate of aqueous-phase oxidation of S(IV) by 03 (30 ppb) and
        H202 (1 ppb), as a function of solution pH	     3-52
  3-4   Isopleths (m x 102) of mean summer morning mixing heights, AGL ..     3-59
  3-5   Isopleths (m x 1Q2) of mean summer afternoon mixing
        heights, AGL			     3-59
  3-6   Percentage of summer 2315 GMT (6:15 p.m. EST, 3:15 p.m. PST)
        soundings with an elevated inversion base between 1 and
        500 m above ground 1 eve!	     3-60
  3-7   Mean resultant surface wind pattern for the United States
        for July.j  Direction and length of arrows indicate monthly
        resultant wind	     3-62
  3-8   Percentage of summer 1115 GMT (6:15 a.m. EST, 3:15 a.m. PST)
        soundings with an inversion base at the surface and wind
        speeds at the surface £2.5 m/sec	"...'..	     3-64
  3-9   Isopleths (m/sec) of mean summer wind  speed averaged
        through the morning mixing layer .....		     3-65
  3-10  Isopleths (m/sec) of mean summer wind  speed averaged
        through the afternoon mixing layer .............		     3-65
 --3-11  Schematic cross section, looking downwind along the jet
        stream, of a tropopause folding event  as modeled by
        Daniel sen	     3-76
  3-12  Measured vertical cross sections of (A) 03, (B) dewpoint, and
        (C) the 500 mb chart and the flight track for October 5, 1978 ...     3-78
  3-13  Hypothesized models of the process that mixes tropopause
        folding events into the troposphere	     3-79
  3-14  National trend in estimated emissions  of volatile organic
        compounds, 1970 through 1983 		     3-91
  3-15  Comparative trends in highway vehicle  emissions of nitrogen
        oxides (NO ) and volatile organic compounds (VOC) versus
        vehicle mites traveled, 1970-1983	     3-92
  3-16  National trend,in estimated emissions  of nitrogen oxides,
        1970 through 1983	.				     3-94
  3-17  Schematics of the. three types, of dynamic models	     3-114
-  3-18  Example of EKMA diagram for high-oxidant urban area	     3-116

  4-1   Comparison of ozone and total oxidant  concentrations in the
        Pasadena area, August 1955	     4-26
v 4-2   Comparison of ozone and total oxidant  concentrations in the
        Los Angeles area, August 1955	     4-26
  4U3   Measurements of ozone and total oxidants in Los Angeles,
        September 4 through September 30, 1971	     4-29
  4-4   Measurements of ozone and total oxidants in St. Louis,
        October 14 through December 21, 1971	;.		....     4-30
                                       xiv

-------
                                 LIST OF FIGURES
                                   (continued)                  .


 Figure                                                                     Page

 4-5   Measurements of ozone and total oxidants, Houston Ship
       Channel,  August 11,  1973		     4-32

 5-1   National  trend in composite average of the second-highest
       value among daily maximum 1-hour concentrations at
       selected  groups of sites, 1975 through 1983 ...	:		     5-4
'5-2   Comparison of the 1979-1980, 1981-1982, and 1983
       composite average of the second-highest daily maximum
       1-hour ozone concentrations across EPA regions 	     5r8
 5-3   Average daylight (6:00 a.m. to 8:00 p.m.) concentrations
       of ozone  in the second and third quarters (April through
       September), 1981	     5-9
 5-4   Average daylight (6:00 a.m. to 8:00 p.m.) concentrations
       of ozone  in the first and fourth quarters '(January through
       March and October through December), 1981 	..	     5-10
 5-5   Distributions of the three highest 1-hour ozone
       concentrations at'valid sites (906 station-years)
       aggregated for 3 years (1979, 1980, and 1981) and
       the highest ozone concentrations at NAPBN sites aggre-
       gated for those years (24 station-years) .........	...	     5-12
 5-6   Locations of the eight national forest (NF) stations
       constituting the National Air Pollution Background Network
       (NAPBN) ...... .'...;.,.......	,.-. .•	     5-17
 5-7   Trajectory analysis plots for the NAPBN site at Mark Twain
       National  Forest, .MQ, July 21, 1979	     5-21
 5-8   (A) Second-highest value among maximum 1-hr ozone
       concentrations at five NAPBN monitoring stations, 1979
       through 1983.  (B) Composite averages of the second-highest
       value among daily maximum 1-hr ozone concentrations at
       five NAPBN stations, 1979 through 1983 .......:....;		     5-24
 5-9   Location, of Sulfate Regional Experiment (SURE) monitoring
       stations	.'.....		..-..'	     5-26
 5-10  Diurnal pattern of 1-hr ozone concentrations on July 13,
       1979, Philadelphia, PA	     5-29
 5-11  Diurnal patterns of 1-hour ozone concentrations, September 20
       and 21, 1980, Detroit, MI			     5-30
 5-12  Diurnal and 1-month composite diurnal  variations in ozone
       concentrations, Washington, DC, July 1981 ...,	     5-31
 5-13  Diurnal and 1-month composite diurnal  variations in ozone
       concentrations, St: Louis County, MO,  September 1981	     5-31
 5-14  Diurnal and 1-month composite diurnal  variations in ozone
       concentrations, Alton, IL, October 1981 (fourth quarter) ...	.     5-32
 5-15  Composite diurnal patterns by quarter of ozone
       concentrations, Alton, IL", 1981 ..	....'		......	...     5-33
 5-16  Three-day sequence of hourly ozone concentrations at
       Montague, MA, SURE station'showing locally generated
       midday peaks and transported late, peaks	•;	     5-36
                                       xv

-------
                                LIST OF FIGURES
                                  (continued)
Figure
5-17  Quarterly composite diurnal patterns of ozone concentrations
      at selected sites representing potential for exposure of
      major crops, 1981		,		     5-43
5-18  Composite diurnal ozone pattern at a rural NCLAN site in
      Argonne, IL, August 6 through September 30, 1980	     5-45
5-19  Daily 7-hour and 24-hour average ozone concentrations at
      a rural NCLAN site in Argonne, IL,  1980	     5-45
5-20  Seasonal variations in ozone concentrations as indicated
      by monthly averages and the 1-hour maximum in each month
      at selected sites, 1981	     5-46
5-21  New York State air monitoring sites for Northeast Corridor
      Monitoring Program (NECRMP)		.	.     5-54
5-22  Altitudinal sequence of monitoring sites in the San Bernardino
      Mountai ns	     5-64
5-23  Relationship between elevation and diurnal patterns of total
      oxidant concentrations, temperature, and vapor pressure at
      four sites (A-D) in the San Bernardino Mountains, CA, July-
      August 1969	     5-66
5-24  Comparison of monthly daylight average and maximum PAN
      concentrations at Riverside, CA, for 1967-1968 and 1980 	     5-78
5-25  Variation of mean 1-hour oxidant and PAN concentrations,
      by hour of day, in downtown Los Angeles, 1965	     5-86
5-26  Variation of mean 1-hour oxidant and PAN concentrations, by
      hour of day, Air Pollution Research Center, Riverside CA,
      September 1966	     5-87
5-27  Diurnal profiles of ozone and PAN at Claremont, CA,
      October 12 and 13, 1978, 2 days of a multi-day smog episode  	     5-88
5-28  Monthly variation of oxidant (Mast meter, continuous 24-hr)
      concentrations and PAN (GC-ECD, sequential, 6:00 a.m. to
      4:00-5:00 p.m.) concentrations, Air Pollution Research
      Center, Riverside, CA, June 1966-June 1967 	     5-90
5-29  Average daily profile by month (July 7-October 10) for
      PAN and ozone in New Brunswick, NJ, 1979 	     5-92
5-30  Diurnal profile of HCOOH, along with other oxidants and
      smog constituents, on October 12 and 13, 1978, at
      Claremont, CA	     5-96
5-31  National trend in composite average of the second highest
      value among daily maximum 1-hour ozone concentrations at
      selected groups of sites, 1975 through 1983	     5-102
5-32  Distributions of the three highest 1-hour ozone
      concentrations at valid sites (906 station-years) aggregated
      for 3 years (1979, 1980, and 1981) and the highest ozone
      concentrations at NAPBN sites aggregated for those years
      (24 station-years)	     5-104
                                      xvi

-------
                       LIST OF ABBREVIATIONS AND SYMBOLS

~                                approximately
A                                wavelength
APR                              air:fuel ratio
APHA                             American Public Health Association
aq                               aqueous
AGL                              above ground level
ASL                              above sea level
atm                              atmosphere
avg                              average
b,p.                             boiling point
bz                               benzene
C                                carbon
°C                               degrees Celsius
CA                               chromotropic acid
CAMP                             Continuous Air Monitoring Program
CARB                             California Air Resources Board
cc                               cubic centimeter
CH4                              methane
CO                               carbon monoxide
C02                              carbon dioxide
cm                               centimeter
concn                            concentration
DBH                              tree diameter at breast height
DNPH                             2,4-dinitrophenylhydrazine
DOT                              Department of Transportation
EA                               normal electrode potential
ECD                              electron-capture detector
EKMA                             Empirical Kinetic Modeling Approach
EPA                              U.S. Environmental Protection Agency
EST                              eastern standard time
FID                              flame ionization detector
FRM                              Federal Reference Method
ft                               foot
                                    XVI1

-------
                       LIST OF ABBREVIATIONS AND SYMBOLS
                                  (continued)
FTIR
9
g/mi
GC
GMT
GPT
hr
hv
HC
HCN
HCOOH
HFET
Hg
H202
H02
MONO
HON02
HPLC

HPPA
HRP
H20
H2S04
in
IR
k
KI
km
L
LAAPCD
LCV
In
LSI
Fourier-transform infrared
gram(s)
grams per mile
gas chromatography
Greenwich mean time
gas-phase titration
hour(s)
photon
hydrocarbons
hydrogen cyanide
formic acid
Highway Fuel Economy Driving Schedule
mercury
hydrogen peroxide
hydroperoxy
nitrous acid
nitric acid
high-pressure liquid chromatography; also,
high-performance liquid chromatography
3-(p_-hydroxyphenyl)propionic acid
horseradish peroxidase
water
sulfuric acid
inch(es)
infrared
constant
potassium iodide
kilometer
liter(s)
Los Angeles Air Pollution Control District
leuco crystal violet
natural logarithm (base e)
local standard time
                                    xv 1.11

-------
                       LIST OF ABBREVIATIONS AND SYMBOLS
                                  (continued)
M
tn
mb
MBTH
rag
mg/m3
MGE
min
ml
mm
mM
MMC
m.p.
mph
MS
MSL
MTBE
NA
NAAQS
NADB
NAMS
NAPBN
NAS
NBS
NECRMP
NEDS
NEROS
NH3
NH4N03
NF
nm
NMHC
NMOC
molar
meter(s)
mi llibar(s)
3-methyl-2-benzothiazolinone hydrazone
milligram(s)
milligrams per cubic meter
modified graphite electrode
minute(s)
milliliter(s)
millimeter(s)
millimolar
mean meridional circulation
melting point
miles per hour
mass spectrometry
mean sea level
methyl tertiary butyl ether
not available
National Ambient Air Quality Standard
National Aerometric Data Bank
National Aerometric Monitoring Stations
National Air Pollution Background Network
National Academy of Sciences
National Bureau of Standards
Northeast Corridor Regional Modeling Project
National Emissions Data System
Northeast Regional Oxidant Study
ammonia
ammonium nitrate
National Forest
nanometer(s)
nonmethane hydrocarbons
nonmethane organic compounds
                                     xix

-------
                       LIST OF ABBREVIATIONS AND SYMBOLS
                                  (continued)
NO
NOX
N02
N03
N20
NR
NYCC
02
03
PAN
PBzN
PNA
PPN
ppb
ppm
ppt
PSD
psig
PST
PUFA
RAPS
RTI
S.D.
SAROAD
SBR
SCAB
sec
SLAMS
SMSA
SRM
SSET
STA
STP
nitric oxide
nitrogen oxides
nitrogen dioxide
nitrogen trioxide
nitrous oxide
natural rubber
New York City Driving Schedule
oxygen
ozone
peroxyacetyl nitrate
peroxybenzoyl nitrate
peroxynitric acid
peroxypropionyl nitrate
parts per billion
parts per million
parts per trillion
Prevention of Significant Deterioration
pounds per square inch gauge
Pacific standard time
polyunsaturated fatty acids
Regional Air Pollution Study
Research Triangle Institute
standard deviation
Storage and Retrieval of Aerometric Data
styrene-butadiene rubber
South Coast Air Basin
second(s)
State and Local Air Monitoring Stations
Standard Metropolitan Statistical Area
Standard Reference Material
small-scale eddy transport
seasonal tropopause adjustment
standard temperature and pressure
                                     xx

-------
                       LIST OF ABBREVIATIONS AND SYMBOLS
                                  (continued)
SURE
TEL
Tenax GC
TF
Tg/yr
THC
TML
TNMHC
TWC
Mg/m3
MM
U
UHAC
U.S.
UV
V
v/v
VHAC
VOC
vol %
w/w
WCOT
XAD-2
XO
Sulfate Regional Experiment Sites
tetraethyl lead
adsorbent used in NMOC analysis
tropopause-folding events
teragrams per year
total hydrocarbon
tetramethyl lead
total nonmethane hydrocarbons
three-way catalyst
microgram(s) per cubic meter
micromolar
uranium
uranium hydroxamic acid chelates
United States
ultraviolet
vanadium
volume-volume
vanadium hydroxamic acid chelates
volatile organic compounds
volume percent
weight-weight
wall-coated open tubular (column)
absorbent used  in NMOC analysis
xylenol orange
year(s)
                                    xxi

-------
                     AUTHORS, CONTRIBUTORS, AND REVIEWERS


Chapter 3:


Pri nci pal Authors

*Mr. Thomas B. McMullen
Environmental Criteria and Assessment Office
MD-52
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

*Mr. Elmer Robinson
Director, Mauna Loa Observatory
National Oceanic and Atmospheric Administration (NOAA/CMCC)
Hilo, HI  96720

*Ms. Beverly E. Tilton
Environmental Criteria and Assessment Office
MD-52
Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Halvor Westberg
Director, Laboratory for Atmospheric Research, and
Professor, Civil and Environmental Engineering
Washington State University
Pullman, WA  99164-2730

*Dr. Arthur M. Winer
Assistant Director
Statewide Air Pollution Research Center
University of California
Riverside, CA  92521


Contributing Authors

*Dr. A. Paul Altshuller
Atmospheric Sciences Research Laboratory
MD-59
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Basil Dimitriades
Atmospheric Sciences Research Laboratory
MD-59
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                    xxn

-------
Chapter 3 Contributing Authors.(cont'd.)

*Dr. Marcia C.  Dodge
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. James M. Kawecki
TRC Environmental Consultants, Inc.    .
2001 Wisconsin Avenue, N.W.
Suite 261
Washington, DC  20007

*Dr. Harold G.  Richter
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
The following people reviewed Chapter 3 at the request of EPA:

Dr. Joseph J. Bufalini
Atmospheric Sciences Research Laboratory
MD-84                                                        •
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Donald Fox
Associate Professor
Department of Environmental Science and Engineering
School of Public Health, 201H
University of North Carolina
Chapel Hill, NC  27514

Mr. Bruce Gay
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Eric Ginsburg
Office of Air Quality Planning and Standards
Control Programs Development Division                     '
MD-15
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                   xxi 11

-------
Chapter 3 Reviewers (cont'd.)

Mr. Gerald Gipson
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Robert Hall
Industrial Environmental Research Laboratory
MD-65
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Michael R. Kuhlman
BatteH e, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Mr. William A. Lonneman
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Chuck Mann
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. E. L. Martinez
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Edwin L. Meyer
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Johnnie Pearson
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                   xxiv

-------
Chapter 3 Reviewers (cont'd.)

Mr. Kenneth L. Schere
Atmospheric Sciences Research Laboratory
MD-80
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Stanley Sleva
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Dlvslon
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Chester W. Spicer
BatteHe, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Mr. Bruce Tichenor
Industrial Environmental Research Laboratory
MD-54
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
^Authors also reviewed portions of this chapter.
                                    xxv

-------
Chapter 4:  Measurement of Ozone and Other Photochemical Oxidants and Their
            Precursors


Principal Authors


*Dr. Jimmie A. Hodgeson
U.S. Environmental Protection Agency
Environmental Monitoring Systems Laboratory
26 W. St. Clair
Cincinnati, OH  45268)

*Mr. Michael W. Holdren                                             •
Battelle, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Dr. M. Rene Surgi
Department of Chemistry
Louisiana State University
Baton Rouge, LA  70803


ContributingAuthors

Dr. Sandor Freedman
Piedmont Technical Services
Hillsborough, NC  27278

*Ms. Beverly E. Tilton
Environmental Criteria and Assessment Office
MD-52
U.S. Environmental Protection Agency
Research Triangle Park, NC  -27711


The following peoplereviewed Chapter 4 at the requestof EPA:

Dr. A. Paul Altshuller
Atmospheric Sciences Research Laboratory
MD-59
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Joseph J. Bufalini
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                    xxvi

-------
Chapter 4 Reviewers (cont'd):

Dr. Donald Fox
Associate Professor
Department of Environmental Science and Engineering
School of Public Health, 201H
University of North Carolina
Chapel Hill, NC  27514

Mr. Bruce Gay
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. William A. Lonneman
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Kenneth Rehme
Environmental Monitoring Systems Laboratory
MD-77
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Harold G. Richter
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Stanley Sleva
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Chester W. Spicer
Battelle, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Dr. Arthur M. Winer
Assistant Director
Statewide Air Pollution Research Center
University of California
Riverside, CA  92521
 *Authors also  reviewed portions of this chapter.

                                   xxv i i

-------
Chapter 5;  Concentrations of Ozone and Other Photochemical Oxidants in Ambient
            Air


Principal Authors

*Mr. Thomas B. McMullen
Environmental Criteria and Assessment Office
MD-52
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

*Mr. Elmer Robinson
Director, Mauna Loa Observatory
National Oceanic and Atmospheric Administration (NOAA/CMCC)
Hilo, HI  96720

*Ms. Beverly E. Tilton
Environmental Criteria and Assessment Office
MD-52
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711


Contributing Author

Dr. Sandor Freedman
Piedmont Technical Services
Hillsborough, NC  27278


The following people reviewed Chapter 5 at the request of EPA:

Mr. Gerald Akland
Environmental Monitoring Systems Laboratory
MD-56
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Joseph J. Bufalini
Environmental Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Gary Evans
Environmental Monitoring Systems Laboratory
MD-56
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                  xxvi

-------
Chapter 5 Reviewers (cont'd):

Dr. Donald Fox
Associate Professor
Department of Environmental Science and Engineering
School of Public Health, 201H
University of North Carolina
Chapel Hill, NC  27514

Mr. Gerald Gipson
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Jimmie A. Hodgeson
Professor, Department of Chemistry
407 Choppin Hall
Louisiana State University
Baton Rouge, LA  70803

Mr. Michael W. Holdren
Battelle, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Dr. Michael R. Kuhlman
Battelle, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Mr. William A. Lonneman
Atmospheric Sciences Research Laboratory
MD-84
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Thomas McCurdy
Office of Air Quality Planning and Standards
Strategies and Air Standards Division
MD-12
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Edwin L. Meyer
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                    xxix

-------
Chapter 5 Reviewers (cont'd):

Dr. Harold 6.  Richter
Office of Air Quality Planning and Standards
Monitoring and Data Analysis Division
MD-14
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Chester W. Spicer
BatteHe, Columbus Laboratories
505 King Avenue
Columbus, OH  43201

Dr. Arthur M.  Winer
Assistant Director
Statewide Air Pollution Research Center
University of California
Riverside, CA  92521
^Authors also reviewed portions of this chapter.
                                    xxx

-------
                            SCIENCE ADVISORY BOARD
                    CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE
     The substance of this document was reviewed by the Clean Air Scientific
Advisory Committee of the Science Advisory Board Jn public sessions.
                             SUBCOMMITTEE ON OZONE
                                   Chairman

                              Dr. Morton Lippmann
                                   Professor
                     Department of Environmental Medicine
                      New York University Medical Center
                            Tuxedo, New York  10987
                                    Members
Dr. Mary 0. Amdur
Senior Research Scientist
Energy Laboratory
Massachusetts Institute of Technology
Cambridge, Massachusetts  02139

Dr. Eileen G. Brennan
Professor
Department of Plant Pathology
Martin Hall, Room 213, Liptnan Drive
Cook College-NJAES
Rutgers University
New Brunswick, New Jersey  08903

Dr. Edward D. Crandall
Professor of Medicine
School of Medicine
Cornell University
New York, New York  10021

Dr. James D. Crapo
Associate Professor of Medicine
Chief, Division of Allergy, Critical
   Care and Respiratory Medicine
Duke University Medical Center
Durham, North Carolina  27710
Dr.  Robert Frank
Professor of Environmental Health
  Sciences
Johns Hopkins School of Hygiene
  and Public Health
615 N. Wolfe Street
Baltimore, Maryland  21205

Professor A. Myrick Freeman II
Department of Economics
Bowdoin College
Brunswick, Maine  04011

Dr.  Ronald J. Hall
Senior Research Scientist and Leader
Aquatic and Terrestrial Ecosystems
  Section
Ontario Ministry of the Environment
Dorset Research Center
Dorset, Ontario
Canada POA1EO

Dr.  Jay S. Jacobson
Plant Physiologist
Boyce Thompson Institute
Tower Road
Ithaca, New York  14853
                                    xxxi

-------
Dr. Warren B. Johnson
Director, Atmospheric Science Center
SRI International
333 Ravenswood Avenue
Menlo Park, California  94025

Dr. Jane Q. Koenig
Research Associate Professor
Department of Environmental Health   /
University of Washington
Seattle, Washington  98195

Dr. Paul Kotin
Adjunct Professor of Pathology
University of Colorado Medical School
4505 S. Yosemite, #339
Denver, Colorado  80237

Dr. Timothy Larson
Associate Professor
Environmental Engineering and
  Science Program
Department of Civil Engineering
University of Washington
Seattle, Washington  98195

Professor M. Granger Morgan
Head, Department of Engineering
  and Public Policy
Carnegie-Mellon University
Pittsburgh, Pennsylvania  15253

Dr. D. Warner North
Principal
Decision Focus Inc., Los Altos
  Office Center, Suite 200
4984 El Garni no Real
Los Altos, California 94022

Dr. Robert D. Rowe
Vice President, Environmental and
  Resource Economics
Energy and Resources Consultants, Inc.
207 Canyon Boulevard
Boulder, Colorado  80302
Dr. George Taylor
Environmental Sciences Division
P.O. Box X
Oak Ridge National Laboratory
Oak Ridge, Tennessee  37831

Dr. Michael Treshow
Professor
Department of Biology
University of Utah
Salt Lake City, Utah  84112

Dr. Mark J. Utell
Co-Director, Pulmonary Disease Unit
Associate Professor of Medicine and
  Toxicology in Radiation Biology
  and Biophysics
University of Rochester Medical
  Center
Rochester, New York  14642

Dr. James H. Ware
Associate Professor
Harvard School of Public Health
Department of Biostatisties
677 Huntington Avenue
Boston, Massachusetts  02115

Dr. Jerry Wesolowski
Air and Industrial Hygiene Laboratory
California Department of Health
2151 Berkeley Way
Berkeley, California  94704

Dr. James L. Whittenberger
Director, University of California
  Southern Occupational Health Center
Professor and Chair, Department of
  Community and Environmental Medicine
California College of Medicine
University of California - Irvine
19772 MacArthur Boulevard
Irvine, California  92717

Dr. George T. Wolff
Senior Staff Research Scientist
General Motors Research Labs
Environmental Science Department
Warren, Michigan  48090
                                    xxxi i

-------
                         PROJECT TEAM FOR DEVELOPMENT
                                      OF
        Air Quality Criteria for Ozone and Other Photochemical Oxldants
Ms. Beverly E. Tilton, Project Manager
  and Coordinator for Chapters 1 through 5, Volumes I and II
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. Norman E. Chi Ids
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. J.H.B. Garner
Coordinator for Chapters 7 and 8, Volume III
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr; Thomas B. McMullen
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr. James A.  Raub
Coordinator for Chapters 9 through 12, Volumes IV and V
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. David T.  Tingey
Coordinator for Chapter 6, Volume III
Environmental Research Laboratory
U.S. Environmental Protection Agency
200 S.W. 35th Street
Corvallis, OR 97330
                                    XXXI11

-------

-------
                               2.   INTRODUCTION
2.1  PURPOSE AND LEGISLATIVE BASIS OF THIS DOCUMENT
     The Clean Air  Act  specifies that the Administrator of the United States
Environmental Protection Agency  (EPA)  issue,  and revise on a periodic basis,
air quality  criteria for certain air pollutants.  Air quality criteria may be
defined as qualitative and quantitative information that describes the effects
of a pollutant on public health and welfare in terms of the respective exposures
that elicited them.   According to section 108 of the Clean Air Act, as amended
in 1977, criteria shall

     ...accurately reflect the latest scientific knowledge useful in indicating
     the  kind  and extent  of all identifiable effects  on  public health or
     welfare which may  be  expected  from  the presence of such pollutant  in the
     ambient air, in varying quantities.
                                                        (U.S.Code, 1982)

Air quality  criteria  provide the Agency with a scientific basis for deciding
whether regulations controlling given pollutants are necessary and for deriving
such ambient air quality standards as may be needed.
     Among  the  air pollutants designated  by  the Administrator  as  criteria
pollutants  are  those known  as photochemical  oxidants.   This document is a
revision  of Air Quality Criteria  for  Ozone and  Other Photochemical Oxidants
(U.S.  Environmental Protection  Agency, 1978a).   Its purpose is  to review and
evaluate the scientific literature on ozone and related oxidants and to document
their effects on public health and welfare.
     The  term  "photochemical oxidants" has historically been defined  as those
atmospheric  pollutants  that are  products of photochemical reactions and that
are capable  of oxidizing  neutral iodide ions (U.S.  Environmental  Protection
Agency,  1978a).   Research  has  established that photochemical  oxidants in
ambient air  consist mainly of ozone, peroxyacetyl nitrate, and nitrogen dioxide,
and  of considerably  lesser  amounts of other peroxyacyl nitrates,  hydrogen
peroxide,  alkyl  hydroperoxides,  nitric and nitrous  acids,  and formic acid.
Other  oxidants  suspected  to occur  in  ambient air  but  only in trace amounts
include peracids and ozonides.
                                    2-1

-------
     Although it is by definition a photochemical oxidant, nitrogen dioxide is
not included  among the oxidants  discussed  in this document.  The formation of
nitrogen dioxide clearly  precedes  the formation of ozone  and  other related
oxidants in the  ambient  air.   Thus, nitrogen dioxide is the dominant oxidant
early in the day, while ozone and other related oxidants predominate from late
morning or midday through much of the afternoon.  Nitrogen dioxide is known to
exert deleterious effects on  human health and  welfare.   This  fact, coupled
with temporal and spatial  variations  in concentrations  that differ  from those
of ozone and related oxidants, underlies the listing of nitrogen dioxide under
section 108(a)(l) of  the  Clean Air Act as a criteria pollutant separate from
                                                                          «
ozone and other photochemical  oxidants.  The second criteria document prepared
by EPA  on  the oxides  of  nitrogen was  completed in  1982 (thS.  Environmental
Protection Agency,  1982a).  That document discussed nitric and nitrous oxides,
nitrogen dioxide, nitric and nitrous acids, and nitrosamines.   As used in this
document, the term  "photochemical  oxidants"  refers to  ozone,  the peroxyacyl
nitrates, hydrogen peroxide, and formic acid.   The oxides of nitrogen are dis-
cussed,  but  only in the  context of their role as precursors  to  ozone and
related oxidants.
2.2  THE OXIDANT PROBLEM
     Ozone  (Og),  a reactive allotrope of  oxygen  (02),  occurs as a  natural
component of the atmosphere.  It is found in its highest concentrations in the
stratosphere, where  it is formed through cyclic reactions resulting from the
photolysis  of oxygen into atomic oxygen  and  the subsequent reaction  of atomic
oxygen with other oxygen molecules.
     Incursions of stratospherically produced ozone into the lower troposphere
occur through meteorological  and atmospheric exchange phenomena,  resulting  in
a global background  of ozone.  To this global background of ozone of stratospheric
origin are added ozone formed in the free troposphere and the contributions of
ozone produced  in the  ambient air  from photochemical  reactions involving
manmade emissions  and natural products  (e.g., natural  emissions  of  volatile
organic compounds).  Manmade emissions of nitrogen oxides and volatile organic
compounds are the chief contributors  to  the  ozone  burden  found  in the ambient
air of urban areas.  The presence of ozone in ambient air is the  net result of
various formation,   stratospheric-tropospheric exchange,  transport,  and de-
struction processes.
                                    2-2

-------
     Other photochemical oxidants  occur  in ambient air, but  the  nature and
source of their  global  backgrounds are not well established.   The additional
photochemical oxidants  of concern  in this  document, namely hydrogen peroxide,
peroxyacetyl nitrate and  its  homologues,  and formic acid, have,  except for
the latter, been detected in remote environments thought to be free of manmade
influences; and  all  have  been detected in the  ambient  air of urban areas.
     The toxicity of ozone is well known.   It is a strong oxidizing agent that
is highly reactive with a wide spectrum of chemical moieties.  Since ozone  is
a gas,  studies  of its   health-related toxicity  have centered  largely on its
capacity for affecting pulmonary function and the morphology of the respiratory
tract, which  is  now  well documented.   In addition, its  effects on extrapulmo-
nary tissues  and systems  are also of  concern and  are the subject of some of
the research discussed  in this document.   Studies of toxic effects of ozone on
vegetation  are  also  well  documented and have focused  on foliar injury and
reduction in  growth  and yield.  The toxicities  of  the peroxyacyl nitrates,  of
hydrogen peroxide, and  of formic acid are  less well documented than the toxicity
of ozone,  having been   the focus  of considerably less  research because the
levels  at which  these  oxidants occur  in the  ambient air, even in urban  areas,
appear to warrant much  less concern.
     Ozone, but  not  the other oxidants mentioned  above,  is  regulated under
provisions  and  procedures  spelled out in  the Clean Air Act.   Its concentra-
tions in ambient air are controlled through the promulgation and attainment of
primary  and secondary   national ambient  air quality standards (NAAQS).  As
described  in the Clean Air  Act,  criteria  pollutants  are those atmospheric
pollutants  that  are  ubiquitous and are emitted  into the air from  numerous and
diverse sources.  While widespread, ozone  and the other photochemical oxidants
found  in ambient air  are  not emitted into  the air as  primary pollutants.
Rather, they are formed as secondary pollutants in the atmosphere from ubiqui-
tous primary  organic and inorganic precursors  that are emitted by a multi-
plicity  of sources.   Consequently, photochemical  oxidant  pollution in this
country  is  the result of a combination of  many  factors, such  as local meteoro-
logical conditions and  the concentrations, composition, and patterns of occur-
rence of the  primary pollutants that give  rise  to  the oxidants.
                                     2-3

-------
2.3  SCOPE AND ORGANIZATION OF THIS DOCUMENT
     The atmosphere  does  not easily lend itself to the partitioning required
for documentation.   Nevertheless,  certain boundaries are  logical for purposes
of discussion  as well as  for purposes  of regulatory decisions.  Ozone and its
organic precursors  are known*  to give rise to  secondary  organic  aerosols.
Likewise,  ozone and hydrogen peroxide both appear to participate in the atmos-
pheric oxidation of  nitrogen dioxide (N02) and sulfur dioxide (S02) to those
inorganic aerosols leading  to  visibility degradation in the atmosphere and to
acidic deposition.   The  contributions  of ozone and hydrogen peroxide to the
oxidation of NO, and SQ2  cannot  be quantified at present, but are  known to be
minor compared to  the  oxidation of these compounds by hydroxyl  radicals.   In
addition,  ozone, the principal photochemical  oxidant in  ambient air, has no
direct effects on  visibility since, unlike N02 and S02,  it does  not absorb
energy in the  visible  region of the spectrum.  Thus, this document includes
brief discussions, in  Chapter  3, of the atmospheric chemistry  of ozone and
hydrogen peroxide  relative  to  the formation of inorganic nitrogen and sulfur
aerosols but does  not  include information on the  actual  effects  associated
with visibility degradation or acidic  deposition.   Since N02 and S02 are the
immediate, direct  precursors  to  the aerosol  species  involved, visibility
degradation and acidic deposition  are  discussed in the respective  air quality
criteria documents on oxides of  nitrogen and on particulate matter and sulfur
oxides (U.S.  Environmental  Protection Agency,  1982a,b).
     This document has  been divided into five  volumes  for ease of review,
printing,  and  distribution.  Volume I  consists of  the summary and  conclusions
for the entire document.  Volume II contains the introduction to the document
(Chapter 2) and  all  chapters dealing with the formation, transport, and fate
of photochemical oxidants (Chapter 3); the measurement of oxidants and their
precursors (Chapter  4);   and the  concentrations  of oxidants in ambient air
(Chapter 5).   Volume III contains the documentation of the effects of photo-
chemical  oxidants  on  vegetation,  ecosystems,  and nonbiological  materials
(Chapters 6, 7, and 8, respectively).  Volume IV reviews the extensive body of
data available on  the  toxicological effects of ozone  and other oxidants in
experimental animals and on in vitro effects on human cells and body fluids
(Chapter 9).    In  Volume  V,  effects observed  in human controlled  exposures
(Chapter 10) and  in  field  and epidemiological  studies (Chapter 11) are pre-
sented.  In addition, th'at  volume  contains an evaluation of the health effects
data of probable consequence for regulatory purposes (Chapter 12).
                                    2-4

-------
     Neither control techniques  nor  control  strategies for the abatement of
photochemical oxidants are  discussed in this document, although some of the
topics included are relevant to abatement strategies.   Technology for control-
ling the emissions  of  nitrogen oxides and of  volatile organic compounds is
discussed in documents issued by the Office of Air Quality Planning and Stand-
ards (OAQPS) of the U.S.  Environmental Protection Agency (e.g., U.S.  Environmen-
tal Protection Agency, 1978b, 1983).   Likewise, issues germane to the scientific
basis for control  strategies,  but not pertinent  to the development  of cri-
teria, are addressed in numerous documents issued by OAQPS.
     In addition,  certain  issues of direct relevance to standard-setting are
not explicitly  addressed in this document, but are addressed  instead  in docu-
mentation prepared  by  OAQPS as part  of  its regulatory  analyses. Such  analyses
include:  (1) discussion of what constitutes an "adverse effect," that is,  the
effect or effects  the  standard is  intended to protect  against; (2) assessment
of risk; and (3) discussion of  factors to be considered in providing an ade-
quate margin of safety.   While  scientific data  contribute significantly to
decisions regarding  these  three  issues, their resolution  cannot be achieved
solely on the  basis of experimentally acquired information.  Final decisions
on items (1) and (3) are made by the Administrator.
     A fourth  issue directly pertinent to standard-setting is identification
of the population at risk, which is basically a selection by the Agency of the
population  to  be protected by the promulgation  of a given standard.  This
issue is addressed only partially in this document.   For  example,  information
is  presented in Chapter 12  on factors,  such as pre-existing disease, that
biologically may predispose individuals  and  subpopulations to adverse effects
from exposures  to ozone.  The identification of a population at risk, however,
requires information above  and beyond data on  biological  predisposition, such
as  information  on  levels of exposure,  activity patterns,  and  personal habits.
Such  information is included in a staff paper developed by OAQPS.   Thus, the
identification  of  the  population at risk relative to standard-setting is the
purview of OAQPS and is not addressed in this document. For information on the
standard-setting process,  see  Padgett and Richmond (1983)  and McKee et al.
(1985).
     This document  consists  of the review and evaluation of relevant  literature
on  ozone  and other photochemical oxidants through  early  1986.   The  material
selected for review and comment  in the text generally comes from the more recent
                                    2-5

-------
literature, with emphasis on studies conducted at or near pollutant concentra-
tions found  in  ambient air.   Older literature that was cited in the previous
criteria document for ozone and other photochemical oxidants (U.S.  Environmental
Protection Agency, 1978) has often been summarized and presented briefly.  An
attempt has been made, however, to discuss at greater length in the text older
studies  (1)  judged significant because of their usefulness  in  deriving the
1979  standards; (2) open  to  reinterpretation because  of newer data;  or
(3) potentially useful  in  deriving  subsequent standards.  The newer informa-
tion on  oxidants  now available may in some  instances make possible a better
understanding of the earlier studies, such that a more detailed and comprehen-
sive picture of health effects  is emerging on  several issues.  An attempt has
been made  to discuss key literature in the  text and present  it in tables as
well.  Reports  of lesser  importance  to  the purposes of this document  may
appear in tables only.
     Generally, .only published material  that  has  undergone  scientific  peer
review is included.   In the interest of admitting new and important information,
however, some material  not published in the  open literature but meeting other
standards  of scientific reporting may be ihcluded.  Emphasis has been placed
on  studies in   which  exposure  concentrations were <1 ppm.   On  this basis,
studies  in which  the lowest concentration employed exceeded  this level have
been included only if they contain unique data, such  as documentation of a
previously unreported effect  or of mechanisms of  effects;  or if they were
multiple-concentration studies designed to provide information on concentration-
response relationships.  Application of a concentration  cutoff ,of  1 ppm to
health effects  studies eliminates  discussion of  studies on mortality and
sublethal  effects.   In the areas  of mutagenesis, teratogenesis, and reproduc-
tive effects, however, results of studies conducted at much higher than ambient
levels have  been  included  because of the potential importance of these long-
term effects to public health and welfare.
     In  selecting studies  for  consideration, each paper  or other publication
was  reviewed in detail.  Technical  considerations  for inclusion of a  specific
study  on health or  welfare effects,  for example, included, but were  not
restricted to,   an analysis of the exposure method; specificity or appropriate-
ness  of  the analytical  method used to monitor  the oxidant concentration;
information on  oxidant monitoring practices  such as location, calibration, and
sampling time;  and the appropriateness of the technique used to measure  the
                                    2-6

-------
effect.   In  addition,  for  health  effects studies technical considerations
included the characteristics  of  the subjects studied and the techniques used
for obtaining or  selecting  the study cohorts.   Interpretation of the results
included consideration  of the following  factors:  the end  results  of the
statistical analysis; the degree to which the results  are  plausible in the
context of other extant data; the appropriateness of the hypothesis developed;
and the agreement  between the hypothesis and the results  reported.   Unless
otherwise  stated,  all  results cited in the  text for health and vegetation
studies are statistically significant at p £0.05.
     The general policy of  EPA is  to express concentrations  of air pollutants
                                                           3
in metric  units,  e.g.,  in micrograms per cubic  meter (M9/m  ); as well  as  in
the more  widely used units,  parts per million  (ppm) or parts  per billion
(ppb), which  are  neither metric nor English units.  That  policy  has been
followed in those  chapters  in which most of the  data have been obtained from
laboratory studies  done  at  room temperature (e.g., Chapters 9 and 10).  Data
reported in ppm for studies conducted outdoors,  such as  field and open-top
chamber vegetation studies, ambient air monitoring, and research on atmospheric
chemistry, have not been  converted.  Conversion  of reported  ppm  and  ppb units
is highly  questionable  in these cases because it assumes standard or uniform
temperatures and pressures.   For data in the health chapters, the conversion
                                        3                             3
units  used are 1  ppm ozone = 1962 pg/m  , and 1 ppm PAN  = 4945 Mfl/m ;  at
1 atmosphere pressure and 25°C.
2.4  REFERENCES

McKee,  D.;  Johnson,  P.;  Richmond, H.; Jones, M.;  McCurdy,  T.;  Walton, T.
     (1985) Work  plan for the ozone  National  Ambient Air Quality Standards.
     Presentation  to the  Clean  Air  Scientific  Advisory Committee, March;
     Research  Triangle  Park,  NC:  U.S." Environmental  Protection Agency,  Office
     of Air Quality  Planning and Standards.
Padgett,  J.;   Richmond, H.  (1983)  The process of establishing and revising
     national  ambient air quality standards.  J. Air Pollut. Control  Assoc.
     33:14.
U.S.  Code.  (1982)  Clean  Air Act,  §108,  air quality  criteria and control
     techniques. U.S.C. 42: §7408.
                                    2-7

-------
U.S. Environmental  Protection  Agency.  (1978a) Air quality criteria for ozone
     and other photochemical oxidants. Research Triangle Park, NC: U.S. Environ-
     mental Protection Agency,  Environmental  Criteria and Assessment Office;
     EPA report  no.  EPA-60Q/8-78-004.  Available from: NTIS, Springfield, VA;
     PB80-124753.

U.S. Environmental Protection Agency. (1978b)  Control techniques for volatile
     organic emissions from stationary sources.  Research Triangle Park, NC:
     U.S. Environmental Protection Agency, Office of Air Quality, Planning and
     Standards;  EPA report no. EPA-450/2-78^022.   Available from:  NTIS,
     Springfield, VA; PB-284804/2.

U.S. Environmental  Protection  Agency.  (1982a)  Air  quality  criteria  for oxides
     of  nitrogen.  Research  Triangle Park, NC:  U.S.  Environmental Protection
     Agency, Environmental  Criteria and  Assessment Office; EPA  report  no.
     EPA-600/8-82-026.  Available  from:  NTIS,  Springfield, VA; PB83-163337.

U.S. Environmental  Protection  Agency.  (1982b)  Air  quality  criteria  for parti-
     culate matter and sulfur oxides. Research Triangle Park, NC: U.S. Environ-
     mental Protection Agency,  Environmental  Criteria and Assessment Office;
     EPA report  no.  EPA-60Q/8-82-Q29. .Available from: NTIS, Springfield, VA;
     PB84-156777.

U.S. Environmental  Protection  Agency.  (1983)   Control  technology for  nitrogen
     oxides  emissions  from stationary sources.   Revised second edition.
     Research Triangle Park, NC: U.S.  Environmental  Protection Agency,  Office
     of  Air Quality, Planning  and  Standards;  EPA report  no.  EPA-450/3-83-002.
     Available from: NTIS,  Springfield, VA; PB84-118330/REB.
                                     2-8

-------
       3.   PROPERTIES, CHEMISTRY, AND TRANSPORT OF OZONE AND .OTHER
               PHOTOCHEMICAL OXIDANTS AND THEIR PRECURSORS
3.1  INTRODUCTION
     Ozone and other oxidants found in ambient air, such as the peroxyacyl
nitrates and hydrogen peroxide, are formed as the result of atmospheric
physical and chemical processes involving two classes of precursor
pollutants, volatile nonmethane organic compounds (NMOC) and nitrogen
oxides (NO ).  The formation of ozone and other oxidants from these
          f\
precursors is a complex, nonlinear function of many factors, including the
intensity and spectral distribution of sunlight; atmospheric mixing and
related meteorological conditions; the concentrations of the precursors in
ambient air and, within reasonable concentration ranges, the ratio between
NMOC and NO  (NMOC/NO ); and the reactivity of the organic precursors.
           X         X
     This chapter describes the physical and chemical properties of
ozone and other photochemical oxidants (Section 3.2).  It also character-
izes the nature of the precursors in terms of their physical and chemical
properties, their sources and emissions into the atmosphere, and their
concentrations in ambient air (Section 3.5).  In addition, a brief description
is provided (Section 3.3) of the complex atmospheric chemical processes by
which ozone and other photochemical oxidants are formed from their precursors.
A brief discussion is also included of the relationship of ozone and other
oxidants to atmospheric phenomena that result from the formation of secondary
organic and inorganic aerosols.
     In addition to the information on the chemistry of oxidants and their
precursors, the chapter includes a discussion of meteorological processes
(Section 3.4) that contribute to the formation of ozone and other oxidants
and that govern their transport and dispersion once formed.  Finally, an
overview is given (Section 3.6) of models of source-receptor relationships
between precursor emissions and ozone formation in the atmosphere, which
either implicitly or explicitly include the relevant emissions, atmospheric
chemistry, and meteorological processes.
                                   3-1

-------
3.2  DESCRIPTIONS AND PROPERTIES OF OXIDANTS AND THEIR PRECURSORS

3.2.1  Ozone and Other Photochemical 0x1dants
3.2.1,1  Ozone.  Ozone (03) is a triangularly shaped molecule consisting
of three oxygen atoms arranged in four basic resonance structures:
           * fi *    *€"    • fi •      • fi«    *fc~    »fi •      * fi •    *^~    * n *      *
           * u«    -^.    *u*      *u*    _^.    «u»      «u«    ->.    • u»      i

       (I)  ~~             (II)"~         ""(III)            ""  (IV)
The first and fourth structures, which predominate, are characterized by
the presence of a terminal oxygen atom having only six electrons.  The
resonance forms depicted above have no unshared electrons.  As the result
of the presence of only six electrons on one of the oxygen atoms in ozone,
the chemical reactions of ozone are electrophilic; that is, ozone removes
electrons from or shares electrons with other molecules or ions.  By
definition, then, ozone is an oxidant since the term "oxidant."
characterizes an ion, atom, or molecule that is capable of removing one or
more electrons from another ion, atom, or molecule, a process called
"oxidation."  A "reducing agent" adds one or more electrons to another
ion, atom or molecule, a process called "reduction."  Oxidation and
reduction reactions occur in pairs and the coupled reactions are known as
"redox reactions."  In redox reactions, the oxidizing agent is reduced and
the reducing agent is oxidized.  The two components of such redox
reactions are known as "redox pairs."  The significance of redox reactions
involving ozone is discussed in Chapters 6 and 9.  The capability of
a chemical species for oxidizing or reducing is termed "redox potential"
(positive or negative standard potential) and is expressed in volts.
Ozone is, in fact, a strong oxidant having a standard potential of +2.07
volts in aqueous systems  (Weast, 1977).
     The physical properties of ozone are given in Table 3-1 (U. S.
Department of Health, Education, and Welfare, 1970, modified).
                                   3-2

-------
                 TABLE 3-1.  PHYSICAL PROPERTIES OF OZONE
Physical  state
Chemical  formula
Molecular weight
Melting point
Boiling point
Specific gravity relative to air
Vapor density
  At 0°C, 760 mm Hg
  At 25°C, 760 mm Hg
Solubility at 0°C
  (Indicated volume of ozone at
   0°C, 760 mm Hg)
Henry's Law constant,
  37°C and pH = 7
Conversion factors
  At 0°C9 760 mm Hg

  At 25°C, 760 mm Hg
Colorless gas; blue violet liquid

%
48.0
-192.7 ± 0.2°C
-111.9 ± 0.3°C
1.658
2.14 g/liter
1.96 g/liter
0.494 ml/100 ml water
8666 atm/mole fraction3
                 o
1 ppm = 2141 (j,g/m
  1  g/nr = 4.670 x 10"4
1 ppm = 1962
        3 = 5.097 x 10~4 ppm
Calculated by formula of Roth and Sullivan (1981).
Source: U. S. Department of Health, Education, and Welfare (1970), modified,
3.2.1.2  Peroxyacetyl Nitrate.  Peroxyacetyl nitrate (PAN) has been
observed as a constituent of photochemical smog in many localities, though
its concentrations and its ratio to ozone differ as a function of time at
a given location as well as from place to place (Chapter 5).  Peroxyacetyl
nitrate, which has the formula CH3C(0)02N02, exists in a temperature
dependent equilibrium with its decomposition products, N02 and acetyl-
peroxy radicals.  It can persist for substantial periods of time in the
atmosphere, depending upon temperature and the N02/NO ratio (Cox and
Roffey, 1977).
                                   3-3

-------
     The chief property of interest regarding PAN is its oxidizing
ability.  A second property of PAN of interest is its thermal  instability.
In the laboratory, this thermal instability necessitates that  precautions
be taken in synthesizing, handling and storing PAN, since improper
handling and storage have resulted in explosions (Stephens et  a!.,
1969).  The ready thermal decomposition of PAN results in a notable
temperature dependence in the rate of PAN decomposition in ambient air.
     Partly because of the thermal instability of PAN, its properties have
not been as well characterized as those of 03 or H202.  Recent work on the
physical properties of PAN, however, has confirmed data reported earlier,
and results of the earlier and more recent work are shown in Tables 3-2
and 3-3 (Stephens, 1969; U. S. Dept. of Health, Education, and Welfare, 1970;
Kacmarek et al., 1978; Bruckmann and Willner, 1983; Holdren et a!.,  1984).
     The infrared (1R) spectrum of PAN is important since most researchers
rely on it for establishing concentrations of PAN for calibration.  Bruck-
mann and Willner (1983) reported the IR spectrum of pure PAN and the Raman
spectrum of liquid PAN at -40°C in an argon matrix.  Their work confirmed
effects that correlate with the ultraviolet (UV) spectrum published
earlier by Stephens (1969); that is, PAN was shown to be stable at x>300
nm but was efficiently photolyzed at x<300 nm (Bruckmann and Willner,
1983).  Actinic radiation falling upon the surface of the earth has
wavelengths >295 nm, and it is light at wavelengths between ~295 and ~430
nm which is involved in photochemical air pollution formation.
3.2.1.3  Hydrogen Peroxide.  Hydrogen peroxide (H202) is an oxidant that
occurs in ambient air as a component of photochemical smog.  It is
believed to be formed through the recombination of two hydroperoxy
radicals (H02) in the presence of a third, energy-absorbing molecule
(section 3.3.1.3).  In aqueous media, H202 is an inorganic actd that has a
dissociation constant of 2.4 x 10~12 and a pK of 11.62 (at 25°C) (Weast,
1977).  Hydrogen peroxide has a standard potential  of +1.776 in the redox
pair, H202/H20.  The physical properties of H202 are given in  Table 3-4.
     Additional properties should be noted here that are of interest
relative to whether effects of H202 in biological receptors are of
significance.  First, H202, though classed as a reasonably strong oxidant
on the basis of its standard potential for the redox system H202/H20, has
                                   3-4

-------
          TABLE 3-2.   PHYSICAL PROPERTIES OF PEROXYACETYL NITRATE
Physical state, @25°C

Chemical formula

Molecular weight

Boiling point, °C


Triple point, °C
Vapor pressure,
  @room temperature

Vapor pressure curve
Hydrolysis
  In alkaline solution
  In acidic solution
  @22°C, pH 5.6

  <325°C, pH 5.6

Henry's Law constant,
  
-------
        TABLE 3-3.   INFRARED ABSORPTIVITIES OF PEROXYACETYL NITRATE
         (RELATED TO 295°K and 973 mb)  (ppm"1 m"1 x 104)  (base 10)
Frequency
Reference
Bruekmann and Wi liner (1983a)
Stephens (1964, 1969b)
1842
12.4
10.0
1741
32. 6C
23. 6d
1302
13.6
11.2
, cm'1
1162.5
15.8
14.3

791.5
13.4
10.1
aAt 4 mbar; no diluent; resolution 1.2 cm""1.
bAt 7 mbar in N2 diluent at 973 mb total pressure and 295 K; grating
 instrument.
CQ branch resolved (1.5 torr).
 Q branch not resolved (1.4 mbar).
           TABLE 3-4.  PHYSICAL PROPERTIES OF HYDROGEN PEROXIDE
Physical state, @25°C
Chemical formula
Molecular weight
Melting point, °C
Boiling point, °C, @760 mm Hg
Density, @25°C, 760 mm Hg
Vapor pressure, §16.3°C
Conversion factors
  §0°C, 760 mm Hg
  @25°C, 760 mm Hg
Colorless liquid
H£02
34.01
-0.41
150.2
1.4422
~1 mm Hg
1 ppm = 152U
                      -4
          = 6.594 x 10""* ppm
1 ppm = 1390
        3 = 7.195 x lO'4 ppm
Source:  Weast (1977).
                                   3-6

-------
been reported to be a positive interference in measurements of total
oxidants made by the Mast meter but to give a very slow response (slow
color development) in the NBKI method for total oxidants (Chapter 4).
This difference should be borne in mind when effects attributed to
oxidants, as opposed to ozone, are evaluated.  Second, #2®2 occurs
normally as a substrate in biological systems and is involved in several
redox pairs of biological importance (see, for example, West et al.,
1966).  It should also be noted that enzymes are present, at least in
mammalian systems, that catalyze the breakdown of H20£.
3.2.1,4  FormicAcid.  Formic acid is a stable product formed in photo-
chemical air pollution from, for example, the reaction of HO? radicals
                                           . -  ,             . C, „
with HCHO and from the reactions of the Criegee biradical CHpOO with water
vapor (Atkinson and Lloyd, 1984).  It has been detected in polluted
ambient atmospheres by longpath infrared spectroscopy on the basis of its
characteristic Q-branch absorption at 1105 cm   (Hanst et al., 1975;
Tuazon et al., 1978a, 1980, 1981a).
     Formic acid has the structure of both an acid and an aldehyde and
hence it differs in chemical behavior from other carboxylic acids in which
the carboxyl group is linked to a hydrocarbon residue rather than to a
lone hydrogen atom.  In concentrated form, HCOOH is a pungent-smelling,
highly corrosive liquid with a boiling point of 100.5°C.

3.2.2  Organic Precursors
     This section briefly describes and defines those hydrocarbons and
other volatile organic compounds commonly found in the ambient air of the
United States and provides relevant information about their chemical and
physical properties.
     The term "hydrocarbon" has been used since the initial investigations
of tropospheric photochemistry to represent those compounds of carbon and
hydrogen that exist as gases in the ambient air and that participate along
with oxides of nitrogen in reactions that form ozone and other photochemi-
cal oxidants.  As knowledge of atmospheric chemistry has increased, some
carbon compounds containing elements such as oxygen and the halogens have
also been shown to be important in photochemical air pollution.  Thus, the
term "volatile organic compounds" (VOC) has come to be used to describe
                                    3-7

-------
stable organic compounds that exist as gases under normal atmospheric
conditions, most of which can participate in the formation of photo-
chemical oxidants.  Recognition that methane (CH^) is virtually unreactive
in the photochemical formation of ozone and other oxidants has given rise
to the more accurate term, "nonmethane organic compounds" (NMOC), for
describing those gas-phase organic compounds in ambient air that serve as
precursors to ozone and other photochemical oxidants.  While these three
terms may sometimes appear to be used interchangeably in this chapter, the
terminology used reflects that reported in the specific literature cited
in this chapter, though in some instances differentiations may have been
made for purposes of discussion.
     As discussed in Chapter 4, methods for measuring total  gas-phase
hydrocarbons are not specific for hydrocarbons but may also detect other
gas-phase organic compounds, though they will not measure them quantita-
tively.  Where methods are used that permit speciation of the compounds
measured, organic compounds other than hydrocarbons can be and usually are
excluded from the summation of individual species used to arrive at a
total nonmethane hydrocarbon (TNMHC) concentration.  Where researchers
have used methods that do not permit speciation, an indefinite and vari-
able fraction of the reported TNMHC concentration may, in fact, be the
result of the presence of nonhydrocarbon organics and such concentration
data are more properly reported as total nonmethane organic compounds
(NMOC).
     The discussion that follows is aimed at presenting basic facts on
nomenclature and those characteristics of photochemically reactive
volatile organic compounds that are relevant to the information given
subsequently in this and later chapters.
3.2.2.1  Hydrocarbons.  Hydrocarbons are compounds consisting of hydrogen
and carbon only.  For a given homologous series the volatilities of hydro-
carbons are related generally to the number of carbon atoms in each
molecule, as well as to temperature.  Hydrocarbons with a carbon number of
one to four are gaseous at ordinary temperatures, while those with a
carbon number of five or more are liquid or solid in pure state.  Liquid
mixtures of hydrocarbons such as gasoline may include some compounds in
pure form that are gases, as well as those that are liquids.  Likewise,
                                   3-8

-------
gas-phase mixtures In ambient air will  usually include compounds that are
liquid or solid in their pure form.  Hydrocarbons with a carbon number of
about eight or less are abundant in ambient air, but those with a carbon
number greater than about 12 have generally not been reported in the gas
phase at significant concentrations, probably because of the inability of
analytical techniques to detect these high molecular weight organics.
     A saturated hydrocarbon has each of its carbon atoms bonded to four
other atoms; whereas an unsaturated hydrocarbon has two or more carbon
atoms bonded to fewer than four other atoms.
     Alkanes.  Alkanes, also known as paraffins, are saturated hydro-
carbons having the general formula CnH2n+2*  The first compound in the
series is methane, CH^» which, because of its low reactivity, does not
contribute significantly to photochemical air pollution in urban atmo-
spheres.  Alkanes as a class are the least reactive of the photochemically
important hydrocarbons (U» S, Environmental Protection Agency, 1978a,b).
Alkanes may be straight- or branched-chain compounds, and comprise the
open-chain (acyclic) hydrocarbons known as aliphatic hydrocarbons.
     Alkenes.  Alkenes, also known as olefins, have at least one
unsaturated bond.  The number of hydrogen atoms in the general formula of
alkenes is decreased by two with respect to the alkanes for each double
bond between carbon atoms; the general formula for alkenes with one double
bond, for example, is CnH2n.  The first compound in the alkene class is
ethene, also known as ethylene; the second is propene, also known as
propylene.  For alkenes containing more than three carbon atoms the
position of the double bond is specified by a numerical prefix (e.g., 1-
butene).  Compounds with carbon numbers three or higher can have two
double bonds between the carbon atoms and are called dienes.  The complete
name of a diene is formed by including a prefix with numbers that indicate
the location of the double bonds.  Like alkanes, alkenes are aliphatic
hydrocarbons and may exist as straight or branched chains.  As a class,
alkenes are among the most reactive hydrocarbons in photochemical systems
(see Section 3.2.2.4).
     Terpenes.  Terpenes are a naturally occurring subgroup of alkenes,
many of which have the formula C1QH16.  Among the terpenes identified in
ambient air a- and p-pinene have been most frequently studied.  Both a-
                                   3-9

-------
and p-pinene contain six-membered rings, as do several other terpenes; but
at least one commonly occurring member of this group, myrcene, is an
acyclic or open-chain compound.  Isoprene, also a naturally occurring
alkene, is a hemiterpene having the formula CgHg.
     Alkynes.  Alkynes are open-chain hydrocarbons that contain one or
more triple bonds.  Acetylene, C2H2, is the simplest member of the class,
which as a whole is often referred to as the acetylenes.  The general
formula for the acetylenes is CnH2n_2, and for each additional triple bond
in the molecule four hydrogen atoms must be removed from the general
formula.  Acetylene is commonly present in ambient air, is thought to be
emitted largely from mobile sources, and has often been taken to be an
indicator of auto exhaust emissions, since it is relatively unreactive in
ambient air and persists in the atmosphere longer than most other exhaust
components.
     Aromatics.  Aromatic hydrocarbons include various compounds having
atoms arranged in six-membered carbon rings with only one additional atom
(of hydrogen or carbon) attached to each atom in the ring.  Benzene is the
simplest compound in the series, having no side chains but only six carbon
atoms and six hydrogen atoms, linked by three conjugated double bonds.
     Compounds containing the aromatic ring and elements other than carbon
and hydrogen are included with aromatic hydrocarbons in the general class-
ification "aromatics."  The double bonds in aromatics are not nearly as
chemically active as those in alkenes because of an effect called
"resonance stabilization."  As a class, aromatics exhibit a wide range of
photochemical reactivity, with benzene having a low photochemical reac-
tivity and 1,3,5-trimethylbenzene, for example, showing high reactivity.
3.2.2.2  Aldehydes.  Aldehydes probably constitute the single most
abundant group of volatile organic compounds other than hydrocarbons in
ambient air.  They are photochemically. important compounds because they
photolyze to form free radicals that will react with oxygen in ambient air
to form alkylperoxy or hydroperoxy radicals (National Research Council,
1977a) (see below).
     Aldehydes are characterized by the presence of the formyl functional
group (CHO).  A carbonyl group having a carbon-oxygen double bond, C=0, is
part of the formyl group.  The carbonyl group is not unique to aldehydes,
                                   3-10

-------
since it is found also in ketones and carboxylic acids; but it forms the
basis for one of the analytical methods used for measuring aldehydes in
ambient air (Chapter 4).
3.2.2.3  Other Organi cCompounds.  Other organic compounds found in
ambient air are known to be photochemically reactive in the formation of
ozone and other photochemical oxidants.  These other organic compounds do
not occur in ambient air collectively, much less singly, at concentrations
that approach the total concentrations of nonmethane hydrocarbons.  Some
of them are suspected of having potentially adverse health effects,
however, and are therefore under scrutiny by the U. S. Environmental
Protection Agency.  These compounds are mentioned here only because they
are photochemically reactive, can serve as precursors to oxidants, and
because they contribute a small but indeterminate fraction of the total
NMOC concentrations reported when continuous hydrocarbon analyzers
(Chapter 4} are used to determine ambient levels of volatile organic
compounds.
     Many of the volatile organics in ambient air that are not hydro-
carbons are organic halides, in which one or more hydrogen atoms of a
hydrocarbon have been replaced by a halogen atom such as chlorine,
fluorine, or iodine.  An enormous number of relatively simple organic
halides are possible, since a halogen atom can be attached to an organic
compound in many different positions.
3.2.2.4  Volatility and Reactivity.  The physical and chemical properties
of nonmethane organic compounds that are most pertinent to their role as
precursors to ozone and other oxidants are those properties that govern
their emission into the atmosphere (volatility) and their lifetime in the
atmosphere, the latter being determined by photochemical reactions
(reactivity) and other removal processes (e.g., gas-to-particle conversion
and dry deposition).
     To be significant in atmospheric reactions, an organic.compound must
have a sufficiently high volatility.  Based upon a review of the available
                                                                   o
literature, Singh et al. (1984) have chosen a vapor pressure of 10   atm
as the criteria for deciding whether an organic compound should be de-
scribed as volatile.  Those compounds with vapor pressures less than 10
atm were considered by Singh et al. (1984) to occur predominately in
                                   3-11

-------
the condensed phase and therefore not to participate in atmospheric
reactions.  Clearly, a rigid cutoff for vapor pressures will not
necessarily be applicable to all organic compounds.
     The photochemical reactivity of subclasses and individual species of
hydrocarbons and of other volatile organic compounds is relevant to mech-
anistic studies in atmospheric chemistry, to modeling, and to other
oxidant-control-related research; but it is not pertinent to the deriva-
tion of criteria.  A major discussion of these properties therefore lies
outside the scope of this document, but a brief discussion of the concept
of hydrocarbon reactivity and its application is presented here.
     Differences in reactivities among volatile organic compounds have
been the focus of considerable attention and research for nearly three
decades.  In early research on photochemical air pollution, many different
definitions or criteria were used to evaluate the reactivity of organic
compounds.  Examples of such criteria include:  rate of NO-to-NOg conver-
sion, maximum ozone concentration formed, initial rate of disappearance of
the organic compound, eye irritation, damage to vegetation, and aerosol
formation.
     Historically, reactivity classifications have been based on environ-
mental chamber measurements of these criteria, as observed in the photo-
oxidation of hydrocarbon-oxides of nitrogen mixtures under conditions
approximating those of polluted ambient atmospheres.  Many of the reac-
tivity data that had been accumulated through 1969 for each of these
manifestations (except plant damage) were critically reviewed by
Altshuller and Bufalini in 1971.  They noted general agreement in reac-
tivity trends from studies employing different reactivity criteria, but
they also cited a number of significant discrepancies in the specific
assignments of reactivity to individual compounds and even to whole
classes of compounds.
     In recent years assessments of the reactivity of volatile organic
compounds have focused almost exclusively on the ability of an organic
compound to produce ozone and other photochemical oxidants.   This focus
arises from interest in regulating most stringently the emissions of
organic compounds having the highest potential for forming ozone and other
photochemical oxidants.
                                   3-12

-------
     The 1978 criteria document for ozone and other photochemical  oxidants
summarized reactivity data acquired from the mid-1960s to the mid-1970s
(U. S. Environmental Protection Agency, 1978a).  Reference to tables of
reactivity schemes given in the 1978 document shows the relatively higher
reactivities of internally double-bonded alkenes, of aliphatic aldehydes
and other carbonyl compounds (such as branched alkylketones and
unsaturated ketones), of dienes, of 1-alkenes, of partially halogenated
alkenes, and of alkylbenzenes (primary and secondary monoalkylbenzenes,
and di-, tri- and tetraalkylbenzenes).  Other compounds also have
relatively high reactivity but are not expected to be as abundant in
ambient air as the compounds cited above.
     For more information the reader is referred to the 1978 criteria
document (U. S. Environmental Protection Agency, 1978a) and the references
therein [e.g., Dimitriades (1974) and Pitts et al. (1977)] for further
information on reactivities of specific compounds.  The reader is also
referred to a more recent and comprehensive assessment of the present
literature on the reactivity and volatility of 118 organic chemicals by
Singh et al. (1984) in which a three-tiered classification scheme was
developed based on the potential for involvement of a given chemical in
photochemical air pollution formation.
     Since a key reaction of volatile organic compounds in ambient air,
regardless of their class, is their oxidation via attack by hydroxyl
radicals (Atkinson et al., 1979, 1982c; Atkinson, 1985), the basis of
several proposed reactivity classifications is the rate of reaction
between an organic compound and the OH radical.  This reaction is the
first step in a chain reaction that is propagated by various organic
peroxy radicals.  As discussed in section 3.3.2.1, reaction  with the OH
radical is thought to be the predominant loss process for most organ!cs in
the troposphere (Atkinson et al., 1979; Atkinson, 1985).  On this basis, a
five-class reactivity scale was proposed by Pitts and coworkers (Darnall
et al., 1976; Pitts et al., 1977) based on the rate of reaction of more
than 100 VOC with OH radicals.   In this scale, each class spanned an order
of magnitude in reactivity relative to methane, with Class I corresponding
to an atmospheric half-life of greater than 10 days and Class V a half-
life of less than 2.5 hours.  The scale has the advantage that any
                                   3-13

-------
compound whose OH rate constant has been measured can be placed in a
precise position in the scale.  It has a number of limitations (Pitts et
al.» 1977), however, since it makes the implicit assumption that OH
radical reaction is the sole loss process for an organic, and that the
subsequent atmospheric chemistry is identical for all organic compounds.
Both these limitations and the advantages of the OH radical reactivity
scale are discussed in detail elsewhere (Pitts et a!., 1977, 1985).

3.2.3  Nitrpgen Oxides
     The physical and chemical properties of the nitrogen oxides that
serve as precursors in the formation of ozone and other photochemical
oxidants have been documented in a recent air quality criteria document
(U. S. Environmental Protection Agency, 1982a).  The most pertinent
properties are briefly summarized here.  The role of nitrogen oxides in  -
the formation of oxidants in the troposphere is discussed in Section 3,3
and in the document cited above.
     The three most abundant oxides of nitrogen in ambient air are nitric
oxide (NO), nitrogen dioxide (N02), and nitrous oxide (N20).  The latter,
a product of soil microbiology, is not known to participate in photo-
chemical reactions in the troposphere.  The two important oxides of
nitrogen relative to photochemical processes in the troposphere are NO and
N0«, which are abundant in ambient air and participate in cyclic reactions
leading to the production of ozone and other oxidants, as described later
in this chapter.
     The basic reactions of importance are (1) the photolysis of NQ2 (K <
430 nm); (2) subsequent formation of ozone from the reaction of atomic
          o
oxygen [0('5P)] produced from the photolysis of N02 with $2 (^n tne
presence of a third, energy-absorbing molecule); and (3) the subsequent
regeneration of N02 by the reaction of NO with Oj.  Coupled with these
basic reactions are reactions between NO and free radicals in the
atmosphere (hydroperoxy, alkylperoxy and acylperoxy) that oxidize NO to
N02, disturbing the N0-N02 equilibrium that would otherwise exist, and
leading, then, to the buildup of Og.  These reactions, and further
information on the source of the free radicals, are given below.  Basic
physical and chemical properties of NO and N02 are given in Table 3-5.
                                   3-14

-------
               TABLE .3-5.  PHYSICAL AND CHEMICAL PROPERTIES
                   OF NITRIC OXIDE AND NITROGEN DIOXIDE
 Property
Other
  properties
  of note:
        NO
Odor
Taste
Color
Absorption
X, nma
None
-
None
<230
Pungent
"
Reddish-brown
Broad range,
both >400
and <400
Uneven number of
  valence electrons
Corrosive, strong oxidant.
Photolyzes at x <430 nm.
Low partial pressure in ambient air.
Uneven number of valence electrons.
Forms dimers
aVisible light x >40Q nm; ultraviolet x <400 nm.  Solar UV radiation in
the troposphere extends from about x290 nm to about x4UO nm.

Source:  Derived from National Research Council (1977b) and U. S. Environ-
         mental Protection Agency (1982a).
3.3  .ATMOSPHERIC CHEMICAL PROCESSES:   FORMATION AND TRANSFORMATION OF
     OZONE AND OTHER PHOTOCHEMICAL OXIDANTS

     The photochemistry of the polluted atmosphere is exceedingly
complex.  Even if one considers only a single hydrocarbon pollutant, with
typical concentrations of nitrogen oxides, carbon monoxide, water vapor,
and other trace components of air, several hundred chemical reactions are
involved in a realistic assessment of the chemical evolution of such a
system.  The, actual urban atmosphere contains not just one but hundreds of
different hydrocarbons, each with its own reactivity and oxidation
products.
                                         (National Research Council, 1977a)

     Despite the complexities of the chemistry of polluted atmospheres, it

is sufficient to understand certain basic processes involved in the formation

of photochemical oxidants from precursor compounds in the presence of sun-

light.  The concentrations of ozone and other oxidants found in urban areas

and in downwind and rural receptor regions are the net result of at least

three general processes:  (1) the initial emission, dispersion, and transport

of precursors; (2) the photochemical reactions that occur in the atmosphere

                                   3-15

-------
as the dispersion and transport take place; and (3) the scavenging processes
along the trajectory that reduce the concentrations both of precursors and
the resulting oxidants.
     Ozone (03) is formed in ambient air through the addition of an atom of
oxygen (0) to a molecule of oxygen (Op).  The breakdown by sunlight (photoly-
sis) of nitrogen dioxide (N02) into nitric oxide (NO) and atomic oxygen provide
the atoms of oxygen involved.  The NO formed in this reaction then reacts with
the 03 produced from the reaction between atomic and molecular oxygen.  In
these cyclic reactions, no net increase in 0, occurs, with the result that an
equilibrium is set up among 03, N02, and NO.   Any reactions that produce N02
without destroying 0- will upset this equilibrium, however, and will result in
a net increase in Oo.  In ambient air, the oxidation of photochemically reactive
hydrocarbons and other nonmethane organic compounds (NMOC) provides a source of
reactive species (radicals) that convert NO to N02 without destroying 03, thus
upsetting the equilibrium.  The reactions of these radicals with NO also consti-
tute a cyclic process.  Since terminating reactions occur between N02 and these
radicals, as well, which remove both N02 and the radicals from the photochemical
reaction system, the cycles described above would gradually end, even in the
presence of sunlight, unless fresh NO  emissions were injected into the atmos-
                                     X
phere.  The complexity of these cyclic, coupled reactions is such that ozone
concentrations in ambient air are a nonlinear function of the NMOC and NO
                                                                         /\
concentrations and, within realistic ranges of precursor concentrations, of the
NMOC:NOV ratio.
       f\
     In the following sections, the processes just described in a simplistic and
summary manner are presented in more detail.   For a complete and thorough dis-
cussion of the many complex reactions thought to take place in polluted atmos-
pheres, the primary literature should be consulted (e.g., Demerjian et al., 1974;
Finlayson and Pitts, 1976; Logan et al., 1981; Whitten, 1983; Atkinson and Lloyd,
1984; and Atkinson, 1985).

3.3.1  Inorganic Reactions
3.3.1.1  Formation of Ozone:  The NO-NOo-0^ Cycle.  Many aspects of the
inorganic reaction systems in the atmosphere are now well understood.  The
photodissociation of N02 by near-ultraviolet solar radiation is a critical
process:

            N02 + hv (295^\<430 nm) ^ NO + 0(3P)                      (3-1)
                                   3-16

-------
The subsequent reaction of the resulting 0( P) atom with molecular oxygen
produces an ozone molecule:

                  0(3P) + 02 + M > 03 + M                           (3-2)
where M is a "third body" molecule (e.g., N2) which can carry away excess
energy of reaction.   In the absence of any competing reactions, the rapid
reaction of NO with Og completes this reaction cycle, regenerating ah N02
molecule:
                       NO + 03 ->• N02 + 02                            (3-3)

As a result of the above three reactions, an equilibrium or steady-state
condition is established among NO, N02 and 0^, and the concentration of Og
in the atmosphere is  governed by the expression,
                                [NO ]
                                                                    (3-4)

where K = ^3-1/^3-3 which depends on the sunlight intensity.  Typically, K
in the lower troposphere is less than or equal to 0.025 ppm.
     Because reaction  (3-3) is rapid, ozone concentrations in urban
atmospheres cannot rise until most of the NO has been converted to f^.
This accounts in part  for the fact that 03 levels may be lower on average
in city centers where  high NO emissions occur, but higher in downwind
suburban areas to which the resulting N02 is transported and then photo-
dissociated, leading  to 03 formation.  The characteristic behavior of
irradiated NMOC-NOX systems in producing Og and other photochemical
oxidants is shown in  Figure 3-1, which depicts data obtained from an
environmental chamber irradiation of a propene-NQ-N02 mixture (Pitts et
al., 1979).
3.3.1.2  Formation of RadicalIntermediates.  Reactions (3-1) through
(3-3), however, cannot by themselves explain the buildup of ozone, since
for each molecule of  NO oxidized to N02 in reaction (3-3)  a molecule of
ozone is also destroyed.  An alternate pathway of conversion of NO to N02
that does not destroy 03 is needed to explain the high ozone levels
observed in urban environments.  Such an alternate pathway is available
through the oxidation  of reactive organic compounds.  In the atmosphere,
these compounds can be oxidized by ozone (03) and/or hydroxyl radicals  (OH)

                                  3-17

-------
0.6
            60
120
180
240
300
360     420
                           ELAPSED TIME, minutes
   Figure 3-1. Experimental time-concentration profiles for propene, NO, NO2,
   O3, HCHO, and PAN for an irradiated NOx-propene-air mixture.
   Source: Pitts et al. (1979).
                              3-18

-------
3.3.1.2.1  Hydroxyl  and Hydroperoxyl  Radicals.  There are at least three
significant formation routes leading to the production of hydroxyl
radicals in the atmosphere.  A pathway for OH radical formation, which
becomes important in the afternoon as ozone concentrations rise, is the
photolysis of 03:

            03 + hv (\<319 nm) * 0(1D) + 02(1Ag)                    (3-5)

The electronically excited O^D) atoms may be quenched to ground state
O(^p) atoms or may react with water vapor to yield OH radicals with an
approximately 20 percent efficiency at 298 K and 50 percent relative
humidity:

                     O^D) + H20 * 2,OH                      "      (3-6)

     Nitrous acid, which has been shown to accumulate to concentrations of
~1 to 8 parts-per-billion  (ppb) during the night in the Los Angeles basin
(Platt et al., 1980a; Harris et al.f 1982; Pitts et al., 1984c) will
photolyze at sunrise, producing a "pulse" of OH radicals (Harris et al.,
1982):

               HOMO + hv (X<400 nm) •* OH + NO                       (3-7)

This photolytic reaction represents a major sink for HONO during daylight
hours.   (Other aspects of the formation and atmospheric chemistry of this
important species are discussed in Section 3.3.1.4.)
     A third significant source of OH radicals is the photolysis of HCHO-.
          HCHO + hv (X<370 nm)
                                            H + HCO                (3-8a)
                                                 CO                (3-8b)
Formaldehyde is both a primary (e.g., from motor vehicle exhausts) and
secondary pollutant that may occur in significant concentrations in the
morning hours as well as in the afternoon (Tuazon et al., 1978a, 1981a;
Grosjean, 1982).

                                   3-19

-------
     The H atoms formed in reaction (3-8a) or from reactions such as
(3-9):
                      OH + CO -> C02 + H                             (3-9)

can react with oxygen to produce hydroperoxyl radicals:
                              M
                      H + 02  -> H02 + M                            (3-10)

These can then react with NO to form hydroxyl radicals:

                     H02 + NO * OH + N02                           (3-11)

Reaction (3-11) then completes a chain reaction involving reactions (3-9),
(3-10) and (3-11) and is a major pathway for the oxidation of nitric oxide
in ambient air.
     The formyl radical in reaction (3-8a) may also serve as a precursor
to H02 radicals and hence OH radical formation:

                     HCO + 02 -> H02 + CO                           (3-12)

Other sources of formyl radicals include the photolysis of higher
aldehydes:

               RCHO + hv U<350 nm) ->• R + HCO                      (3-13)

     In addition to reactions (3-11) and (3-12), H02 radicals are produced
by H-atom abstraction from alkoxy radicals as discussed below:

                    RCH20 + 02 + RCHO + H02                        (3-14)

These H02 radicals will also oxidize NO to N02 via reaction (3-11).
     Based upon environmental chamber data, computer modeling studies and
measured ambient concentrations of unreactive organics such as fluoro-
chlorocarbons, concentrations of OH and H02 radicals in polluted
atmospheres are believed to be in the ranges 5 x 10  to 5 x 10^ and 10^ to
109 radicals cm~3, respectively.
                                   3-20

-------
3.3.1.2.2  Mitrate Radicals.  Ozone can react with N02 to produce the
nitrate radical  and an oxygen molecule:

                     03 + N02 •» N03 + 02;                          (3-15)

however, because of its large photolytic cross section (Graham and
Johnston, 1978; Magnotta and Johnston, 1980) the N03 radical photolyzes
rapidly in sunlight:

                   N03 + hv •»• N02 + 0(3P)                         (3-16a)

                            -»• NO + 02                             (3-165)

Direct spectroscopic measurements (Platt et al., 1980b, 1984; Noxoh.  et al.,
1980; Pitts et al., 1984c) have confirmed that N03 radical concentrations only
rise above part-per-trillion (ppt) levels after sunset.  The atmospheric
reactions of this important radical intermediate are discussed in Sections
3.3.1.5 and 3.3.2.3, including the fact that at night the N03 radical will
participate in a rapid equilibrium between N02» N03 and dinitrogen
pentoxide (N20g).
3.3.1.3  Termination Reactions.  Although the photochemical  reactions
described above require sunlight, the presence of sunlight does not mean
that the reactions continue indefinitely.  Terminating reactions gradually
remove NO and N02 from the  reaction mixtures such that the cycles would
slowly come to an end unless fresh NO  emissions were injected into the
atmosphere.  Specifically, the inorganic chemistry system includes termin-
ation reactions for OH and H02 radicals with NO and N02 to form nitrogen
acids such as nitrous acid  (HOMO), nitric acid  (HN03) and peroxynitric
acid (H02N02):
                               M
                       OH + NO •» MONO                              (3-17)

                               M
                      OH +  N02 + HN03                              (3-18)

                               M
                     H02 +  N02 •* H02N02                            (3-19)

                                   3-21

-------
Pernitric acid, however, thermally back-dissociates rapidly U]y2 ~10 sec
at 298 K) and HONO photolyzes, so that under typical  atmospheric
conditions OH + N02 is the major sink of NOX.
     Nitrous acid and nitric acid have now both been  reliably measured in
ambient air by longpath spectroscopic techniques and, in the case of
nitric acid by other techniques as well.  Peroxynitric acid has not yet
been observed in the atmosphere, although it has been predicted to be
present at fractional ppb concentrations and has been extensively studied
in laboratory systems (Graham et al., .1977; Hanst and Gay,  1977; Howard,
1977; Niki et al., 1977; Graham1 et al., 1978).
     In the presence of NO, radical-radical reactions are generally not of
major importance in the atmosphere, because concentrations  of the radicals
are low.  In the absence of NO, however (for example  at night), the
reactions of peroxy radicals with H02 and with other  peroxy and acylperoxy
radicals, and the self-combination of H02 radicals, can become
important.  For example, the reaction of H02:

                    H02 + H02 -> H202 + 02                          (3-20)

could be an important route to the oxidation of S02 to sulfate in solution
(see section 3.3.4).
     Clearly, termination of the chain reactions can  lead to the formation
of other oxidants as well as relatively stable organic nitrates in the
atmosphere.  In addition to HONO, HN03, H02N02 and H202, these oxidants
include other peroxyacyl nitrates, organic hydroperoxides and organic
peracids which have been observed either in polluted  atmospheres or in
irradiated laboratory mixtures (National Research Council,  1977a).  These
compounds almost always occur in low concentrations in ambient air, but
they may play a significant or even critical role in  atmospheric chemistry
(Pitts et al., 1983) as shown in the following sections.
3.3.1.4  Reactions Involving Nitrous Acid.  An alternative  pathway for
reaction (3-19) is the formation of HONO:

                    H0  + N0  + HONO + 0                           (3-21)
                                   3-22

-------
This reaction pathway has been shown, however (Graham et al . , 1977, 1978;
Howard, 1977), to be negligible compared to reaction (3-19).
     The equilibrium between NO, NCL, and hLO results in still another
potential source of HONO:

                   NO + N0  + H0 -> 2 HONO                          (3-22)

                               2
                           2 HONO ^ NO + NQ2 + H20                  (3-23)

Reactions (3-22) and (3-23) may proceed both homogeneously and hetero-
geneously, but they appear to be too slow to be of atmospheric importance
at part-per-million concentrations of NO .
                                        X,
     Similarly, the reaction of N02 with water may proceed in the gas
phase or on surfaces:
                  2 N02 + H20 -> HONO + HN03                 .        (3-24)

Recent work (Sakamaki et al., 1983; Pitts et al., 1984a) has shown that
HONO is produced in environmental chambers from the reaction of NQ2 with
water vapor, almost certainly via heterogeneous processes (although HNOg
was not observed in these studies).  This process may be a minor source
of HONO in the atmosphere and in the exhaust plumes from combustion sources
(Pitts et al. , 1984b).
3.3.1.5  Reactions Involving Nitric Acid and Dinitrogen Pentoxide.
Another equilibrium reaction of importance is that involving NQ2, NO*,, and
N205:                                                              ;
                                  M                            ••...--.
                       N02 + N03  $  N205                        (3-25a,b)

The equilibrium constant for this system has been measured by several   ,
groups (Graham and Johnston, 1978; Kircher et al., 1984; Tuazon et al.,
1984; Perner et al., 1985) and appears to be 3 (±0.5) x ID"11 cm  molecule'1
at 298 K, which corresponds to a time to reach equilibrium of about one
minute.  The equilibrium constant remains somewhat uncertain, however.
     Dinitrogen pentoxide is a potentially important precursor to HNO,
(and hence acid deposition) through its reaction with water either in the
gas phase (Tuazon et al., 1983) or on surfaces (Heikes and Thompson, 1983):
                                    3-23

-------
                      N2°5 + H2° "*" 2 HN03                         (3-26)

Thus, for N0£ and NQg radical concentrations representative of receptor
sites downwind from major urban areas such as Los Angeles (Pitts et al.,
1983, 1984c; Platt et al.s 1984), and using an N00 + NO, $ N,0C
                              -11   3         -1
equilibrium constant of 3 x 10    cm  molecule  , an HNOQ formation rate
                 -1
of several ppb hr   is obtained at ^-50 percent relative humidity [assuming
                                         -?1   3         -1    -1
the upper limit rate constant of 1.3 x 10    cm  molecule   sec   for
reaction (3-26)].
     This estimated nighttime formation rate of HNCL via reaction (3-26)
                                                                   -1
can be compared to a calculated daytime formation rate of *1 ppb hr   from
reaction (3-18) for ^20 ppb of N02 and 1 x 106 molecule cm"3 of OH radicals.
Reaction (3-26) could potentially be an important loss process for NO  and a
                                                                     J\
significant nighttime pathway for HNQg formation in urban atmospheres.

3.3,2  Organic Reactions
     It is now well recognized that all organic compounds emitted into the
atmosphere may be degraded by one or more of the following four pathways:
reaction with hydroxyl radicals, reaction with ozone, reaction with
nitrate radicals, or photolysis.  Indeed, knowledge of the rates and
mechanisms of these processes has advanced to the point that the process that
will predominate for a given compound can be predicted with a reasonable
degree of certainty.
     This progress notwithstanding, there remain substantial differences
in the degree to which the detailed atmospheric chemistry is understood
for the principal classes of hydrocarbons found in polluted atmospheres:
alkanes (paraffins), alkenes (olefins), and aromatics.  Thus, the photo-
oxidation reactions of the smaller alkanes and the simple alkenes, such
as ethene, propene, and trans-2-butene, are fairly well understood.  There
is much less certainty, however, about the detailed reactions undergone by
the higher alkanes, the higher alkenes, and the aromatics subsequent to
their initial reactions with OH radicals, ozone, or N03 radicals.
     The following sections briefly summarize the basic features of the
four reaction pathways identified above for organic compounds emitted into
the atmosphere.
                                  3-24

-------
3.3.2.1  Reactjons with Hydroxyl Radlcals.  The following sections treat
separately the mechanisms of reaction of OH radicals with the major
classes of organic compounds including alkanes, alkenes, aromatics and
oxygenated compounds, as well as nitrogen- and sulfur-containing
compounds.  The treatment here of relevant reactions is necessarily an
overview.  For a comprehensive and current detailed description of the
kinetics and mechanisms of the atmospheric reactions of OH radicals with
organic compounds, the reader is referred to a review by Atkinson (1985).
3.3.2.1.1  Alkanes.  It is now well established that the only significant
atmospheric chemical loss process   for the alkanes is reaction with OH
radicals.  These reactions proceed by hydrogen abstraction (Atkinson,
1985) to produce alky! radicals (R)» which then add 02 to form alkyl
peroxy radicals (R02):

                     OH + RH t- R* + H20                            (3-27)
                             M
                     R' + 02 + R02'                                (3-28)

     In polluted atmospheres R02 radicals rapidly oxidize NO to N02»
forming alkoxy radicals (RO); or add N02 to form alkyl peroxynitrates:
                    RU2* + NO f RO* + N02                          (3-29)

                   R0£* + N02 + R02N02                          (3-30a,b)

The latter, however, are not expected to be present in ambient air at
significant concentrations because of their short (<1 sec at 298 K)
lifetimes with respect to thermal decomposition (reaction 3-30b).
     Alkoxy radicals may also undergo hydrogen abstraction by molecular
oxygen to form aldehydes or ketones (Baldwin et al., 1977):

                  R'RCHO* + 02 * R'OR + HQ2*                       (3-31)

or they may decompose to form oxygenates (Baldwin et al., 1977; Batt,
1979):
                    R'RCHO"
RCHO + R"                          (3-32)
R'CHO + R*                          (3-33)
    3-25

-------
In both of these reaction sequences, however, H02 radicals are formed, and
hence OH radicals are regenerated.  The carbonyl compounds thus formed may
subsequently react with OH radicals or may photodecompose (see Section
3.3.2.1.4).
     A further important reaction pathway for acyl radicals is the
addition of 02, followed by reaction with N02 to form peroxyacyl nitrates:

                        0          0
                       RC* + 02 •* RCOO*                            (3-34)

                     0             0
                    RCOO' + N02 -» RCOON02                          (3-35)

The simplest member of this class of compounds, peroxyacetyl nitrate, has
been measured in polluted atmospheres throughout the world (see Chapter 5).
     The reactions described above suggest the importance of simple (C^) alkanes, namely alkoxy
radical isomerization (Carter et al., 1976; Baldwin et al., 1977; Hendry
et al., 1978; Batt, 1979; Batt and Robinson, 1979; Carter et al., 1979)
and alky! nitrate formation from the reaction of R02* with NO (Atkinson et
al., 1982a, 1983, 1984f):
                                M
                      R02* + NO •* RON02                            (3-36)

Discussion of these processes is beyond the scope of this chapter and the
reader is referred to the original literature and to appropriate reviews
(e.g., Atkinson, 1985) for summaries of these aspects of the atmospheric
chemistry of longer-chain alkanes.
3.3.2.1.2  Alkenes. In polluted atmospheres, unsaturated hydrocarbons
react primarily with OH radicals and with Og (Herron and Huie, 1977, 1978;
Dodge and Arnts, 1979; Akimoto et al., 1980; Kan et al., 1981; Niki et
al., 1981; Atkinson et al., 1982c; Whitten, 1983; Atkinson, 1985).  For
most alkenes studied to date, the reaction with OH radicals proceeds
                                   3-26

-------
almost entirely by addition to the double bond.  In the case of propene,
for example, addition of the OH radical to the double bond is expected to
be followed by 02 addition, with the oxidation of NO to N02;by the
resulting peroxy radical to form an alkoxy radical  (Atkinson et al., 1985)

                OH* + CH3CH=CH2 •* CH3CHCH2OH                       (3-37)

                                     00*
                                     I
                CH3CHCH2OH + 02 •* CH3CHCH2OH                       (3-38)

                   00'               0 *
                CH3CHCH2OH + NO + CH3CHCH2OH + N02                 (3-39)

     Decomposition of the alkoxy radical and subsequent reactions lead to
the formation of acetaldehyde and formaldehyde, both of which can be
detected in polluted atmospheres:
                 CH3CHCH2OH + CH3CHO +  "CH^H                      (3-40)
                              HCHO + H0                        .   • (3-41)
                                    OH + N02

The overall reaction resulting from reactions  (3-37) to  (3-41) is:
 OH +  CH3CH=CH2 +  202 + 2 NO + CH3CHO + HCHO + 2 N02 + OH            (3-42)
                                                1
                                                0,
hv, 02
Thus, the reaction of OH radicals with alkenes increases the  rate of NO-
to-N02 conversion and hence increases the yield of ozone.  The specific
reaction sequence for the OH  radical-initiated oxidation of propene
(Atkinson,  1985) is shown in  Figure 3-2.
                                   3-27

-------
                                OH
               (-65%)
  (-35%)

CH-CHCH OH CH,C1
J £-i -J

NO —
\
°2
-»* N02 NO -
^
                                                    N0n
            f
       _CH CHO_ + CH OH
                   1'
               HO   + HCHO
                 £.      ~::
       OH
CH-CHOH + HCHO

        l°2
        f
  CH^CHO + HO^
Figure 3-2. Reaction scheme for OH radical-initiated oxidation of propene
in the presence of NO.
Source:  Atkinson (198i).
                           3-28

-------
     The reaction schemes presented for propane illustrate the major role
that organic compounds (not only alkenes but also alkanes and aromatics)
play in producing photochemical air pollution, namely acceleration of the
conversion of NO to NQ2 and the resulting formation of ozone.
     Al k e n e s Emi 11 e d F rom Ve get at\onf  ^ special class of alkenes
receiving considerable attention over the past decade are those
unsubstituted compounds emitted from vegetation.  Examples include iso-
prene and monoterpenes such as a- and p-pinene, d-limonene, and myrcene.
Much research and discussion have been devoted to assessments of the
potential for such compounds, which are emitted from vegetation in large
quantities, to contribute to photochemical air pollution (Coffey, 1977;
Westberg, 1977; Arnts and Gay, 1979; Tingey and Burns, 1980; Bufalini and
Arnts, 1981; Dimitriades, 1981; Altshuller, 1983), but a detailed treat-
ment of this topic is beyond the scope of this document.  Presented here,
and in other appropriate parts of section 3.3, are relevant aspects of
current knowledge of the atmospheric chemistry of organic compounds known
to be emitted from vegetation.
     Although reliable rate constants for the reaction of these naturally
emitted organic compounds with OH radicals are now available (Winer et
al., 1976; Kleindienst et a!., 1982; Atkinson, 1985), with the exception
of isoprene their detailed atmospheric chemistry is still not well
characterized.  Summarized briefly here are the OH radical-initiated
photooxidation reactions for isoprene and a-pinene.
     Isoprene.  Based on the relative atmospheric concentrations of OH
radicals and ozone, and their  rate constants for reaction with isoprene,
the dominant atmospheric reaction pathway for isoprene is expected to be
OH radical addition to the olefinic double bonds (Lloyd et al., 1983):
                                           CH0
                                           I 3
                                 —> HOCH2-C-CH=CH2               (3-43a)
                  CH
         OH + CH2=C-CH=
I  3
   H=CH2 	
                       1-
                                        CH0
                                        I  3
                                     CH2C-CH=CH2                  (3-43b)
OH
 CH,
                                           3
                                     CH2=C-CHCH2OH                (3-43c)
                                  -> CH2=C-CH-CH2                 (3-43d)
                                   3-29     OH

-------
The hydroxyalkyl radicals formed in reactions (3-43a)-(3-43d) react
rapidly with 02 to form peroxy radicals, which can then rapidly oxidize NO
to N02 (Atkinson and Lloyd, 1984).  The resulting hydroxyalkoxy radicals
decompose to form methyl vinyl ketone and methacrolein.  The subsequent
atmospheric reactions of these products are described in detail by Lloyd
et al. (1983) and by Killus and Whitten (1984).
     a-Pinene.  Based on the rate constant for its reaction with OH
radicals (Atkinson et al., 1979; Kleindienst et al., 1982), a-pinene is
expected to react exclusively by OH radical addition at the least
substituted carbon atoms, with the resulting radical rapidly adding 02
(Lloyd et al., 1983):
                                   °2
                   OH +    >      -»     >                      (3-44)
A mechanism for the subsequent reaction pathways of this hydroxyperoxy
radical has been reported (Lloyd et al., 1983), but it was of necessity
largely parameterized because of the lack of data on the reaction products
resulting from the photooxidation of a-pinene under atmospheric
conditions.  Even less information is available for other monoterpenes and
any detailed consideration of their atmospheric chemistry would be highly ,
speculative.
     For further descriptions of the OH radical-initiated photooxidations
of isoprene and monoterpene, the reader should consult the,,primary
literature (Lloyd et al., 1983; Killus and Whitten, 1984).
3.3.2.1.3  Aromatics.  The aromatic fraction in gasolines has increased in
recent years, partly as a result of the reduction of lead in gasoline.
Given that there are also substantial emissions of aromatic compounds
(e.g., benzene, toluene, xylenes, etc.) from a wide range of industrial
processes, the importance of aromatics in the hydrocarbon distribution in
ambient atmospheres has grown.
     Reactions with OH radicals constitute the sole atmospheric loss
process for aromatic compounds.  The available kinetic and mechanistic
                                   3-30

-------
data concerning these reactions have been reviewed critically (Atkinson,
1985).  It is clear from these data that two reaction pathways are
possible.  The first of these is OH radical addition to the aromatic ring
to form an initially energy-rich QH-aromatic adduct which can either
decompose back to the reactants or be collisionally stabilized.  This is
illustrated for toluene, an abundant aromatic constituent in urban
atmospheres:
                OH
                          H,
(plus other isomers)  (3-45)
For alkyl-substituted benzenes, the second reaction pathway  involves
H-atom abstraction from the substituent group:
                OH
                                                                   (3-46)
     This latter is a minor (<1Q$) process at room temperature (Atkinson,
1985).  The reaction pathways subsequent to the H-atom abstraction
reaction pathway (3-46) are reasonably well understood (Atkinson and
Lloyd, 1984).  Thus, under atmospheric conditions the benzyl radical is
expected to react via the following sequence of reactions:
                                         CH200'
                                                                   (3-47)
                                   3-31

-------
                              + NO
                                   M
+ NO,
(3-48a)
                                                                  (3-48b)
with reaction (3-48b) occurring approximately 10 percent of the time at
atmospheric pressure and room temperature (Hoshino et al., 1978).
The CgHgCHgO" radical then reacts with 02 to yield benzaldehyde and an H02
radical:
                                               + HO,
                    (3-49)
Analogous reaction pathways are expected to be applicable to the other
aromatic hydrocarbons, after H-atom abstraction from the substituent alkyl
groups (Atkinson and Lloyd, 1984).
     The rates of reaction of OH radicals with aromatics are now well
characterized (Atkinson, 1985), as is the relative importance of the
alternative pathways (3-45) vs (3-46), at least for selected aromatics.
In the case of toluene, for example, OH radical addition is expected to
occur ~80 percent of the time at the ortho position (Kenley et al., 1981).
     The fate, however, of the addition adducts formed from OH radical-
aromatic reactions remains unclear (Atkinson, 1985), although several
mechanisms have been proposed in the case of toluene (Atkinson et al.,
1980; Leone et al., 1985).  Although substantial progress has been made in
understanding reaction mechanisms for toluene and certain other aromatic
hydrocarbons (Atkinson and Lloyd, 1984; Leone et al., 1985), much
additional research is needed before a complete understanding of the
complex N0x-photooxidation chemistry of aromatic compounds is obtained.
                                   3-32

-------
3.3.2.1.4  Aldehydes.  Aldehydes are consumed in the atmosphere both by
photolysis and attack by OH radicals.  Photolysis of acetaldehyde, for
examples leads to methyl and formyl radicals that then react as discussed
above:
                   CHgCHO + hv ->• CH3 + HCO                          (3-50)

Attack by OH radicals forms acetyl radicals which can successively  add Q£
and N02 to form PAN:

                         + OH * CHjCO + HgO                         (3-51)
         CHgCO + 02 * CHgCOO* —>• CHgCOONOg                       '  (3-52)

3.3.2.1.5  Nitrogen-Containing Compounds.  Only  limited  information  is
available for the reactions of OH radicals with  nitrogen-containing
compounds and their subsequent reactions under atmospheric  conditions.
The OH radical reactions with the aliphatic amines are rapid, with  room
                                                          1 1   *3           -1
temperature  rate constants being in the  range (2-6)  x  10   cm  molecule"1
sec"-'- (Atkinson et al.» 1985).   For the methyl-substituted  amines,  the
trend of the room temperature rate constants suggests  that  these  reactions
proceed via  abstraction from the C-H bonds and,  where  possible, the  N-H
bonds.  Product studies of several irradiated amine-air  systems have been
reported in which plausible reaction,pathways following  OH  radical  attack
have been proposed (Pitts et a!., 1978;  Tuazon et al., 1978b; Lindley et
al., 1979).
     To date, kinetic data are available only for OH radical reactions
with hydrazine and methylhydrazine (Atkinson, 1985), and only limited
product data are available for these reactions (Tuazon et al.,  1981b,
1982).  Reactions of nitrites with OH  radicals are expected to  proceed via
H-atom abstraction from the C-H  bonds  but, since no  product data  are
presently available, no reliable assessment of the initial  reaction
pathway can  be made.
     No product or direct mechanistic  data are available for organic
nitrates.  Howevers the reactions of OH  radicals with  at least  the  smaller
                                   3-33

-------
alky! nitrates (for which isomerization of the alkoxy radicals cannot
occur) will probably ultimately yield N0£ together with the corresponding
aldehydes  (Atkinson et al., 1982d).  These reactions may be of importance
in long-range transport and acid deposition, since alky! nitrates are
formed in significant yields from the atmospheric photooxidation of
certain alkanes (Atkinson and Lloyd, 1984).
3.3.2.1.6  Sulfur-Containing Compounds.  The literature concerning reac-
tions of sulfur-containing compounds with OH radicals has been reviewed
recently by Atkinson (1985).  Reactions of thiols with OH radicals must
proceed via either H-atom abstraction from the weak S-H bonds or, more
likely, by the formation of an QH-thiol adduct.  The reaction of OH
radicals with the sulfides, RSR» can proceed via either H-atom abstraction
from the C-H bonds or OH radical addition to the sulfur atom.  Product
data for the reaction of OH radicals with dimethyl sulfide under
atmospheric conditions have been obtained from numerous studies (Srosjean
and Lewis, 1982; Hatakeyama et al., 1982; Hatakeyama and Akimoto, 1983;
Niki et al., 1983; Grosjean, 1984); the major stable products are HCHO,
S02 and CHgSOgH, together with CHgSNO as an intermediate product.
     Only for dimethyl disulfide have kinetic (Cox and Sheppard, 1980;
Wine et al., 1981) and product (Hatakeyama and Akimoto, 1983) data been
reported.  On the basis of these data, it appears that the initial
reaction proceeds via OH radical addition to form an adduct, followed by
rapid decomposition of this adduct to CH3S and CH3SQH radicals.
Subsequent reactions of these CHgSOH and CH3S radicals then lead to the
observed products:  S02, HCHO, and CH3S03H.
3.3.2.2  Reactions with Ozone.  The atmospheric reactions of ozone are
complex and result in products and processes that have significant
environmental implications, including effects on biological systems,
visibility, and materials.  Ozone, for example, is highly reactive towards
certain classes of organic compounds (e.g., alkenes) and certain of those
reactions lead to the formation of secondary organic aerosols.  Ozone may
also play a role in the oxidation of S0£ to H^SO*, both indirectly in the
gas phase  (via formation of OH radicals and Criegee biradicals) and
directly in aqueous droplets.
     In the following sections, the atmospheric reactions of ozone with
organic compounds are summarized in some detail, including the mechanisms
                                   3-34

-------
of certain of these reactions.  Emphasis is placed, whenever possible, on
those reactions that lead to products or processes suspected or known to
have effects on biological or other important receptors.
     In discussing the reactions of ozone with organic compounds in the
troposphere, it is important to recognize that organics undergo competing
reactions with OH radicals during daytime hours (Atkinson and Lloyd, 1984;
Atkinson, 1985) and, in certain cases, they can photolyze or react with
N03 radicals at night (Japar and Niki, 1975; Carter et al., 1981a;
Atkinson et al., 1984a,b,c,d; Winer et al., 1984).  All organics except
the perhaloalkanes exhibit room temperature OH radical rate constants "of
>5 x 10"15 cm3 molecule"1 sec"1 (Atkinson, 1985).  Since the ratio of 03
to OH radical concentrations in the unpolluted troposphere during daylight
hours is believed to be of the order of 106 (Singh et al., 1978; Crutzen,
1982), only for those organics whose 03 reaction rate constants are
                 pi   O          11
greater than ~10    cm  molecule   sec   can consumption by 03 be
considered atmospherically; important.  These ozone reactions of interest
are summarized below.
3.3.2.2.1  Alkenes.  Ozone reacts rapidly with the acyclic mono-, di- and
trialkenes and cyclic mono-, di-, and tri-alkenes.  The rate constants for
these reactions range from ~1Q"18 to ~10"^ cm  molecule"  sec"  (Atkinson
and Carter, 1984), corresponding to atmospheric lifetimes ranging from a
few minutes (for the more reactive cyclic alkenes such as the
monoterpenes) to several days.  In polluted atmospheres, especially in the
afternoons during photochemical oxidant episodes, the principal
consumption of the more reactive alkenes will therefore occur via reaction
with 03, rather than with OH radicals.
     It is now reasonably well established that the initial step in the
Q-j-alkene reaction involves the formation of a "molozonide" that rapidly
decomposes (Harding and Soddard, 1978; Herron et al., 1982) to a carbonyl
compound and a biradical  (which is also initially energy rich):
                                   3-35

-------
         OO
                .0-0
                           V   V  3
                            >-<
                           V           R4

i>
                                                                 (3-53)
                                               0-0'
                               X
                          Ci-C
where C 3  denotes an energy-rich species.
     Based  on an analysis of reported product and mechanistic studies of
the reactions of 03 with ethene  (Herron and Huie, 1977; Su et al.,  1980;
Kan et al., 1981; Niki et al.» 1981) and propene  (Herron and Huie,  1978;
Dodge and Arnts, 1979), and the  much less extensive studies of the  higher
alkenes (Martinez et al.» 1981), Atkinson and Lloyd (1984) have suggested
that these  initially energy-rich biradicals react under atmospheric
conditions  as shown below:
             M
                CH200
                 CH
CH
                               /'
                                 0*.
  [HCOOH]
,/
 X
                                                    H + HC02
                                    (40%)
C02 -I- H2
CO + H90
(12%)
(42%) (3-54)
and
                                                          20,
                                                                  (6%)
                                 3-36

-------
 [CH3CHOO]q
                  CHjCHOO
                    CH
                              (40%)
    0        CHg + CO + OH    (19%)
[CH^COH]*^
         ^  CH3 + C02 + H    (24%)
                                      CH4 + C02
                                                                          (3-55)
                               (5%)
                              (12%)
where CH^OO and CHoCHOO denote thermalized biradicals.  These thermal lied
biradicals have been shown (Calvert et'al., 1978; Herron et a!., 1982;
Atkinson and Lloyd, 1984) to undergo bimolecular reactions with aldehydes,
S0o> CO, and HoO, and it is believed that they will also react with HQ and
W$  (Calvert et al., 1978; Herron et al., 1982; Atkinson and Lloyd, 1984):
                   RCHOO + HQ + RCHO + HQ,
                  RCHOO + HQ2 > RCHO +
                             (3-56)

                             (3-57)
                 RCHOO + S02—> RCHO
                  RCHOO + H20 > RCOOH +
                   RCHOO + CO > products
                             (3-58)

                             (3-59)

                             (3-60)
                RCHOO + R'CHO
                             (3-61)
     Under atmospheric conditions, the reactions with HQt N02> or HgO are
expected to be the dominant loss processes of these thermalized
biradicals, with the precise major reaction pathway depending on the
relative concentrations of NO, N0  or   Q (Atkinson and Lloyd, 1984).
                                   3-37

-------
Hence, Og-alkene reactions in the atmosphere can lead ultimately to the
formation of aldehydes and acids, as well as to the conversion of SO^ to
H«S(L» although the latter is probably a minor process in the overall
oxidation of SO- during long-range transport (Fin!ayson-Pitts and Pitts,
1982).
     Isoprene and monoterpenes are alkenes emitted from vegetation.  The
role of these compounds in photochemical air pollution has been a subject
of discussion, and sometimes controversy, for more than a decade.  The
case of ozone reactions with compounds such as isoprene and the monoterpenes
is particularly interesting since these reactions can represent a sink for
ozone as well as for the hydrocarbons themselves.  This adds complexity
to an overall assessment of the role of hydrocarbons emitted from
vegetation since, depending upon the specific atmospheric conditions,
they may be both sources and sinks for ozone (Dimitriades, 1981; Altshuller,
1983).
     Again, as in the case of their OH radical reactions, the detailed
reaction sequences following reaction of 0, with iosprene and the
monoterpenes are not well understood although substantial kinetic data are
available (Atkinson and Carter, 1984); thus only a brief summary of
available information is presented here.
     Isoprene.  The reaction of Q3 with isoprene (Kamens et a!,, 1982;
Lloyd et a!., 1983; Kill us and Whitten, 1984) leads to molozonides that
are presumed to decompose subsequently into stable products and radical
intermediates analogous to those produced in other ozone-alkene reactions
described earlier in this section.  The major products, methlvinylketone
and methacrolein, undergo further reactions with OH radicals and ozone
(Lloyd et al., 1983; Killus and Whitten, 1984).
     The reaction of ozone with isoprene can also lead to aerosol
formation (Kamens et al., 1982) and this is discussed in Section 3.3.4.
     Honoterpenes.  The detailed pathway for the reaction of 0, with the
monoterpenes under atmospheric conditions is unknown.  Lloyd et al. (1983)
have proposed a reaction sequence for crpinene involving addition of 0, to
the double bond to form a molozonide, with subsequent ring opening.  They
note, however, that this reaction sequence is entirely speculative, and
                                  3-38

-------
many additional kinetic and mechanistic data will be required to elucidate
the detailed reactions of 03 with the monoterpenes.
                                                                     9"\
3.3.2.2.2  ATkanes and A1 kynes.  Given reported rate constants of 10~" to
10"^ cm3 molecule"1 sec"1  (Atkinson and Carter, 1984), there appears to
be no convincing evidence in the literature for an elementary reaction
between 03 and the alkanes.  Similarly, although there is presently
substantial uncertainty concerning the rate constants for the reactions of
ozone with the simple alkynes (e.g., acetylene, propyne and 1-butyne),
most of the available room temperature data for these 50^ alkynes indicate
03 rate constants in the range of ~10"20 to i(T19 cm3 molecule'1 sec"1
(Atkinson and Carter, 1984).  Thus, those reactions will also be
unimportant in the atmosphere since the corresponding OH radical rate
                                    1Q                          _i p   -a
constants are, for example, ~8 x 10    for acetylene and ~7 x 10    cm
molecule"  sec"1 for propyne and 1-butyne (Atkinson, 1985).
3.3.2.2.3  Aromatics.  As in the case of the alkanes, the aromatic
hydrocarbons react only very slowly with 03 (Atkinson and Carter, 1984)
and these reactions are not expected to be important in the atmosphere^
Although the cresols are significantly more reactive than the aromatic
hydrocarbons (Atkinson and Carter, 1984), under atmospheric conditions
their reactions with 03 are minor compared to their reactions with OH
radicals (Atkinson et a!., 1978) or N03 radicals (Carter et a!., 1981a).
3.3.2.2.4  Oxygen-Containing Organlcs.  For those oxygen-containing
compounds that contain no unsaturated carbon-carbon bonds (e.g.,
formaldehyde, acetaldehyde, glyoxal, and methylglyoxal) the reactions with
ozone are very slow, and, by analogy, this is expected to be the case for
all ethers, alcohols, aldehydes, and ketones containing no unsaturated
carbon-carbon bonds.  For the carbonyls and ethers (other than ketene)
that contain unsaturated carbon-carbon bonds, however, much faster
reactions are observed (Atkinson and Carter, 1984).
     Few data are available, however, concerning the mechanisms of the
reactions of 03 with such oxygen-containing organics, the only published
information being that of Kamens et al. (1982).  From a study of the
reactions of 03 with methacrolein and methylvinylketone, methylglyoxal was
observed as a product, along with other minor products  (Kamens et al.,
1982), as anticipated from the reaction schemes:
                                   3-39

-------
              CH2=CHCOCH3+
       HCHO + [CH,CQGHOQ]*
                 CH2-CHCOCH3
                                   (3-62)
                  ChLCOCHO + [CH-OOT
                    •5     ,      £
                                   (3-63)
and
        HCHO
                     CHO
   CHO
   I  •
CH3COO
                       'CHO
                                      \
                                                   (3-64)
CH3COCHO +  [CH200
Kamens et al. (1982) discuss the possible subsequent reactions.
3^3,2.2.5  N11rogen-Containing Organics.  Studies of the kinetics of the
reactions of 03 with a variety of nitrites, nitriles, nitramines, nitroso-
amines, amines, and hydrazines (Atkinson and Carter, 1984) indicate that
only for the hydrazines are reactions with 03 sufficiently rapid to be of
atmospheric importance (Carter et al., 1981b; Tuazon et al., 1982).
3.3.2.2.6  Sul f ii r-Contai ni ng Organi cs.  Based upon the kinetic data
available for dimethyl sulfide, thiirane (C2H^S) and thiophene, it appears
at the present time that the rates of reaction of Og with sulfur-
containing organics can be considered to be unimportant under atmospheric
conditions (Atkinson and Carter, 1984).
3.3.2.2.7  Organometallies.  Rate constants have been reported only for
tetramethyl- and tetraethyl-lead (Harrison and Laxen, 1978), and no
mechanistic or product data are available for these reactions.
3.3.2.2.8   Radical Species.  Because of the low concentration of 03 and
of both alky! and most alkoxy radicals in the atmosphere, and because
these radicals react at significant rates with 02 (which is present at a
concentration >105 higher than 03 in ambient atmospheres), the reactions
of ozone with such radical species can be considered to be of negligible
importance in the atmosphere.  Of course, the reaction of ozone with the
hydroxyl radical must be considered, as discussed earlier.
                                  3-40

-------
3.3.2.3  Reactions with Nitrate Radicals.  The pioneering work of Nikl and
coworkers (Morris and Niki, 1974; Japan and Niki, 1975) showed that
gaseous N03 radicals react with alkenes, with the rate constants
increasing markedly with the degree of substitution on the double bond.
Carter et al. (1981a) showed that the hydroxy-substituted aromatics
(phenols and the cresols) also react rapidly with N03 radicals.
     More recently, Atkinson and coworkers have investigated the kinetics
of the reactions of N03 radicals with a wide range of organics at room
temperature (Atkinson et al,, 1984a-e) and from these and the earlier
studies, information concerning the mechanisms of these reactions has been
forthcoming.
     In the remainder of this section, current understanding of the
mechanisms of reaction of NQ3 radicals with the various classes of
organics is briefly summarized.
3.3.2.3.1  Alkanes.  The relevant kinetic data (Atkinson et al . , 1984e) in-
dicate that these reactions proceed via H-atom abstraction from the C-H
bonds, almost certainly predominantly from secondary or tertiary C-H bonds:

             .  . .       N03 + RH * HN03 + R'                         (3-65)

Hence, these reactions lead directly to HN03 formation.  The measured room
                                                                  17   "\
temperature rate constants for the alkenes range between 3.6 x 10    cm
molecule"1 sec'1 for jrj-butane to 2.2 x 1Q"16 cm3 molecule""1 sec"1 for 2,3-
di methyl butane.
3.3.2.3.2  Alkenes.  The reactions of N03 radicals with the alkenes have
been shown from both kinetic (Japar and Niki, 1975; Atkinson et al.,
1984a) and product  (Bandow et al., 1980) studies to proceed via initial
                                                                       't '
addition of the N03 radical to the olefinic double bond:

                               ON00
                               I  2
             + 2CH3CH=CH2 * CH3CHCH2 + CH3CHCH2ON02,                (3-66)

with addition at the terminal carbon expected to dominate  (Atkinson and
Lloyd, 1984).  Possible reaction sequences  beyond this initial reaction
                                                                      ' I
have been discussed (Bandow et al., 1980; Atkinson and Lloyd, 1984) but
are highly uncertain at the present time  (Atkinson and Lloyd, 1984).
                                   3-41

-------
Thermally unstable nitro-peroxynitrates such as CH3CH{ON02)CH2OON02» and
stable dinitrates such as CH3CH(ON02)CH2ON02 have been reported as
products in NC^-NC^-propene-air systems (Bandow et al., 1980).
     Monoterpenes.  While no mechanistic information is available
concerning the reactions of N03 radicals with the monoterpenes, N03
radical reaction rate constants have recently been reported for a
substantial number of these compounds (Atkinson et al., 1984c, 1985).
Based in part on these kinetic data, Winer et al. (1984) have proposed
that reaction with N03 radicals at night may be an important reaction
pathway for certain naturally occurring organics such as the monoterpenes
and dimethyl sulfide.  Conversely, these kinetic data also show that
reactions with the more reactive alkenes, including isoprene and certain
of the monoterpenes, as well as dimethyl sulfide and the hydroxy-
substituted aromatics, can be important loss processes for N03 radicals at
night (Winer et al., 1984).  The importance of N03 radical reactions in
determining the atmospheric lifetimes of the monoterpenes is discussed in
section 3.3.3.
3.3.2.3.3  Aldehydes.  Based upon the product data of Morris and Niki
(1974), i.e., the observed formation of HN03 from the reaction of N03
radicals with CH3CHO, it is expected that these reactions proceed via H-
atom abstraction from the relatively weak H-CO bonds:

                   N03 + RCHO * RCO + HN03                         (3-67)

Thus the reaction of N03 radicals with acetaldehyde could be a nighttime
source of peroxyacetyl nitrate (PAN):
                 NO, + CH^CHO -» HNO.+ CH-CO                        (3-68)
                   O     O         *5    O

                   CH3CO + 0£ * CH3CO*                             (3-69)


                                    ?
                 CH-CO " + NO, * CHoCOON09                         (3-70)
                   O  O      £     «3     £

Reaction of N03 radicals with the higher aldehydes will lead, by analogous
reaction schemes, to the higher peroxyacyl nitrates, RC03N02.  Reaction

                                   3-42

-------
with formaldehyde, however, will lead to H02 radical formation, since HCO
reacts rapidly with 02 (Atkinson and Lloyd, 1984):
HCO + 0  -> H0
                                      CO
(3-71)
3.3.2.3.4  Aromatics.  As discussed by Atkinson et al. (1984d), the
reactions of NOo radicals with the monocyclic aromatic hydrocarbons and
the hydroxy-substituted aromatics appear, based upon kinetic evidence, to
proceed via H-atom abstraction from the C-H or 0-H bonds on the        .«
substituent groups.  This conclusion is based upon the observation that
for the xylenes and the eresols the meta-isomer reacts more slowly (by a
factor of ~2) than the ortho- and para-isomers.  This is in contrast to
the addition reactions of 0( P) atoms and OH radicals, in which the meta-
isomer is the most reactive (Atkinson, 1985).  Furthermore, ^-nitrophenol
has been tentatively identified as a product of NoOg-NOg-phenol-air
reaction mixtures, presumably formed by the reaction sequence:
                               HN0
                                              (3-72)
followed by  (Nik1 et al., 1979):
                                 OH          OH
                                      ,NQ0
                                           -  NO,
                                                                   (3-73)
Thus, these reactions can also be a direct source of nitric acid as well
as  forming low-volatility organic nitro compounds.
      In summary, it is now clear that reaction with NQ3 radicals at night
is  a major atmospheric reaction pathway for many organic pollutants.   It
must therefore be considered, along with the reactions of OH radicals and
03  and photolysis, as one of the dominant loss processes for organics in
the atmosphere.
                                   3-43

-------
3.3.3  _Atmosjgher1 c LIfetImes of Organi c Compounds                       ;
     Table 3-6 compares the atmospheric lifetimes (i.e., the time to reach
1/e of the initial concentration) for selected organic compounds, arising
from manmade and natural sources, as the result of reaction with 03 over a
24-hour period, with OH radicals during the day, and with N03 radicals at
night.  It can be seen that under the atmospheric conditions assumed reac-
tions with Og are important for the higher alkenes, including the monoter-
penes during daylight when N03 radical concentrations are low, and for the
hydrazines. 'For the other organics for which kinetic data are available,
including alkanes and aromatics, reactions with 63 are generally of negli-
gible or minor importance in determining their atmospheric lifetimes.
 TABLE 3-6.  CALCULATED LIFETIMES OF SELECTED ORGANIC COMPOUNDS RESULTING
     FROM ATMOSPHERIC LOSS BY  REACTION  WITH Do  AND  OH  AND  NO, RADICALS
                                             o    ,           %*
Organic lifetimes3*
Organic compound
Alkenes from manmade sources
Ethene
Propene
trans-2-Butene
2-Methyl-2-butene
2»3-Di methyl -2-butene
Naturally emitted alkenes
Isoprene
a-Pinene
p-Pinene
A -Carene
d-Limonene
24 hr
2.7 days
11 hr
35 min
17 min
6 min
10 hr
1.4 hr
5.5 hr
1.0 hr
11 min
OH,
daytime ,
16 hr
5.6 hr
2.0 hr
1.6 hr
1.3 hr
1.4 hr
2.3 hr
1.8 hr
1.7 hr
1.0 hr
N03,
nighttime
79 days
1.1 days
33 min
1.3 min
0.2 min
22 min
2 min
5 min
1.2 min
0.9 min
aT1me to reach 1/e of the initial concentration.
Assuming 100 ppb of 03 (24 hr average), 2 x 106 molecule cm"3 (0.08 ppt)
 of OH radicals during daylight hours, and 100 ppt of N03 radicals during
 nighttime hours, and at room temperature.
                                   3-44

-------
3.3.4-   Atmospheric  Reactions  of  Peroxyacetyl  Nitrate           ,
     With  the  recognition  in  recent  years that PAN is a ubiquitous
nitrogenous  species in  the troposphere  (Singh and Hanst, 1981; Aikin et
al., 1983;  Penkett, 1983;  Singh  and  Salas, 1983; Spicer et al., 1983) and
in the  lower stratosphere  (Aikin et  al.,  1983), there has been renewed
.focus on the atmospheric  role of this  organic compound.
     Smog  chamber studies  have shown that, once formed, PAN can be
relatively  stable under atmospheric  thermal  conditions (Pitts et al.,
1979; Akimoto  et al,.,  1980).   Since  PAN is in equilibrium, however, with
acetylperoxy radicals  and  N02:


                       8           8
                    CH3COON02  j CH3COO + N02                        (3-74)

any process  that removes  either  acetyl  peroxy radicals or N02 will lead to
the decomposition ,of PAN.   One such  process is the reaction of NO with
CH3C(0)02  radicals.  Because  PAN has been shown to persist through the
night in urban atmospheres (Tuazon et  al., 1980, 1981a) the reaction of
PAN with NO  during  the morning traffic peak can lead to the formation of
OH radicals  via the foil owing,mechanism (Cox and Roffey, 1977; Garter et
al., 1981c):                              -       .   .

                     0             0
                 CH-COO*  + NO ->  CH,CO"  + N09                       (3-75)
                    *3        _      *3         c*.          '•     •'        .

                          o
                       CH3C-0* +  CH3* + C02                         (3-76)

                               M
                     CH3*  + °2 *  CH3DO                              (3-77)

                  CH300*  + NO -»  CH30* + N02                        (3-78)

                    CH^O-  + 09 *  HO,  + HCHO                         (3-79)
                      *5      <•.,     <—         '                       ,     -

                      H02  + NO *  OH + N0£                           (3-80)


                                   3-45

-------
                              M
                      OH + NO -> HONO                               (3-81)
                              M
                     OH + N0£ -> HN03                               (3-82)
     Thus, the reaction with NO of PAN carried over from previous air
pollution episodes will lead to enhanced smog formation on subsequent
days.  This enhancement in reactivity results both from the fact that
these reactions form radicals that initiate the transformations occurring
in photochemical smog and from the fact that these reactions convert NO to
N02, which allows earlier formation of 03 and higher levels to be
attained.  It should be noted that this enhancement will result even if
all of the PAN reacts with NO emitted at nighttime, since the NO
conversion does not require sunlight and since at least some of the
radicals formed will be "stored" as nitrous acid, to be released when
photolysis begins at sunrise (Harris et al., 1982).
     These results could have important implications regarding multiday
photochemical pollution episodes in which significant buildup of PAN is
observed.  Under such conditions, the carry-over of PAN may be a
significant factor in promoting ozone formation on subsequent days and
may, in part, contribute to the progressively higher 03 levels often
observed during such episodes (Tuazon et al., 1981a).
     A second important role of PAN is its ability to contribute to the
long-range transport of NOX.  In the absence of significant levels of NO
(i.e., in the cleaner troposphere) and in regions of lower temperature and
in the upper troposphere, when the thermal decomposition of PAN becomes
unimportant, the atmospheric lifetime of PAN will be determined by its
reaction with OH radicals.  This reaction is sufficiently slow (Wellington
et al., 1984) that PAN will probably be long-lived and hence serve as a
reservoir for odd nitrogen in a manner analogous to HN03 (Aikin et al.,
1983).  Also analogous to the case for HN03, dry deposition of PAN may be
a significant loss process in cleaner atmospheres.
                                   3-46

-------
3*3.5  Role of Ozone In AerosolFormation
     In addition to having direct effects on human health and on vegeta-
tion, ecosystems, and nonbiological materials, ozone can contribute
indirectly to visibility degradation and to acidic deposition through its
participation in the formation of both organic and inorganic aerosols
(National Research Council, 1977a; U. S. Environmental Protection Agency,
1982b).
3.3.5.1  Formation of SulfateAerosol.   It is well established that the
source of the vast majority of manmade sulfate aerosol in the atmosphere
is the oxidation of sulfur dioxide (S02), ultimately to sulfuric acid
(H2SQ,|).  The correlations between elevated levels of ozone and of sul fate
aerosol in ambient air have been noted by several investigators in field
studies concerned with visibility reduction by aerosols.  Wilson (1978)
and Gillani et al, (1981) have pointed out that atmospheric mixing
intensity and the background 03 concentration are the two most important
factors in determining SOo oxidation at relative humidities lower than 75
percent.  It is also clear, however, that the rate of reaction of Do with
S02 is far too slow to account for observed formation rates of sulfate
aerosol (U. S. Environmental Protection Agency, 1982b).
     Of the many possible gas-phase reactions of S0£, only a few appear to
have any significance in the production of sulfate aerosol and the
reaction of OH radicals with SQ2 appears to be the dominant pathway for
the oxidation of S02 (Calvert and Stockwell, 1983,  1984;  Calvert and Mohnen,
1983).  A recent analysis by Stockwell and Calvert (1983)  shows that the
formation of HOS02 radicals from the reaction of OH radicals with S02,
followed by reaction with 02, is the reaction mechanism for the formation
of S0q:
     s>

                    OH + S02 (+ M) -> HOS02 (+ M);                  (3-83)

                        HOS02 + 0£ * H02 + S03.                    (3-84)

From the reaction of H02 radicals with NO, OH radicals are regenerated and
the cycle begins again as discussed earlier in Section 3.3.1.2,1.
                                   3-47

-------
     The importance of the reaction of OH radicals with S02 in the
atmosphere is supported by observations of power plant plumes, in which no
aerosol is formed at night when the OH radical  concentration in ambient
air is negligible; and none is formed during the day before the plume is
well mixed with ambient air (the ambient air contains much higher
concentrations of OH radicals and 03 than the plume) (Davis et al., 1979;
Blumenthal et al., 1981).
     Though it does not react directly with S02 at an appreciable rate, by
virtue of its role in OH radical production 03 plays an important indirect
role in the transformation of S02 to sulfate aerosol via the homogeneous
oxidation of S02 in both clean and polluted atmospheric systems.  Ozone
plays a further role in the oxidation of sulfur in aqueous droplets as
discussed later in this section.
3.3.5.2  Formation of Nitrate Aerosol.  Despite limited relevant data,
the possible contribution of nitrate aerosol to visibility reduction
should not be neglected and the role of 03 in the formation of this
aerosol species is briefly considered here.
     The principal manmade nitrogen emissions in this case are NO and
N02.  Nitric oxide is relatively insoluble in aqueous systems (Section
3.2) and does not react with water in any significant manner.  Thus, NO
must be converted to a more highly oxidized form, for example N02, in
order to participate in the formation of particulate nitrate.
     The oxidation of NO to N02 can occur through thermal oxidation at
high concentrations of NO such as those in and  very near the stacks of
power  plants  (U.  S.  Environmental  Protection Agency,  1982a).    This
generates only a small portion of the N02 formed in the atmosphere,
however.  As previously discussed (Section 3.3.1), the most important
reactions leading to formation of N02 in ambient air are the reaction of
NO with 03 and the oxidation of NO to N02 by hydroperoxyl radicals and
other peroxy radicals (reactions 3-3 and 3-11,  respectively).  Thus, if
N02 is a precursor of nitrate aerosol, 03 plays a significant direct role
in its formation by oxidizing NO, and an indirect role by leading to
formation of OH radicals (Section 3.3.1).
     As discussed earlier, N02 can be converted in the gas phase to nitric
acid (HN03) vapor by reaction with OH radicals  during the day or by
                                   3-48

-------
reaction with Q3 to form NO, radicals, which at night are in equilibrium with
NgOg.  As shown in Section 3.3.1.5, homogeneous or heterogeneous hydrolysis, or
both, of NgOg is an important nighttime pathway to nitric acid formation.   Once
it has been produced in the gas phase, HNQ3 is sufficiently volatile to remain
in the atmosphere as a vapor.  The available laboratory and ambient air data
indicate, however, that HN03 vapor reacts with ammonia in a reversible reaction
to form NH4NQ3 (Doyle et al., 1979; Stelson et al., 1979; Appel et al., 1980),
which, because of its low vapor pressure, will form nitrate aerosol particles:

                          HN03 + NH3  -»  NH4N03                       (3-85)

If acidic sulfate is present, however, it will react with NH,N03 to form HN03
again.  Consequently, reaction (3-85) is not a major sink for nitric acid in
areas with high sulfate loading, such as the eastern United States.  Evidence
also indicates that HN03 vapor will .react with NaCl aerosol in the following
way:                   '          .  "   -

                      HN03 + N.aCl •* NaN03 + HC1                    (3-86)

This second reaction (equation 3-86) may account for the fact that much of the
observed particulate nitrate in Los Angeles is. found in the coarse mode (Farber
et al., 1982).  Obviously, the importance of this mechanism for nitrate aerosol
formation is determined by the availability of sea salt particles.
3.3.5.3  Formation of Organic Aerosols.  Sulfate and nitrate aerosols are
present at significant levels in the atmosphere in the form of just a few
compounds.  In contrast, secondary organic aerosols are composed of a large
number of species, but there is no clear consensus concerning which ones
contribute most to the mass concentration.  For all the species that are found
in the secondary organic aerosol,. however, the fundamental formation mechanism
is the same.  The vapor-phase precursor undergoes some reaction that results
in formation of a product having an equilibrium vapor pressure sufficiently
low  that condensation, nucleation, or both are possible at the gaseous concen-
tration achieved.  From the available data, it seems clear that the more highly
oxygenated, larger-carbon-number species generally are those precursors likely
to form secondary aerosols in the atmosphere.
                                  3-49

-------
     For an earlier but thorough review of the formation of secondary
aerosol, the reader is referred to the 1977 monograph on ozone and other
photochemical oxidants prepared by the National Research Council
(1977a).  This monograph reviews the reactions of manmade volatile organic
compounds that produce aerosol.  Biogenic as well as manmade volatile
organic compounds, however, can participate in aerosol formation
(Altshuller and Bufalini, 1971; Arnts and Gay, 1979).  Direct experimental
evidence of aerosol formation, along with product analysis, is only avail-
able, though, for a limited number of natural compounds, including iso-
prene (Kamens et al., 1982) and a-pinene and p-pinene (Schwartz, 1974;
National Research Council, 1977a; Hull, 1981), mainly because the
analysis and characterization of these kinds of products at ambient con-
centrations is extremely difficult.  Hull (1981) has conducted experiments
with a- and p-pinene at high concentrations in a small tube reactor.
Analysis of the products showed, on a weight basis, that almost all of the
reacted a-pinene carbon was found in the condensed materials extracted
from the walls.  Although the products identified from these experi-
ments were either in the condensed phase or on the walls, Hull suggested
that at the a-pinene levels found in ambient air these products have a
high enough vapor pressure to exist both in the gas phase and in aerosols
(Hull, 1981).  In a recent review of the role of biogenic volatile organic
compounds, Altshuller (1983) has discussed at length the contribution of
these compounds to ambient air aerosols.

3.3.6  Role of Ozone and Other Photochemical Oxidants in the Acidification
       of Rain
     Two recent criteria documents prepared by the U. S. Environmental
Protection Agency (1982a; 1982b) contain thorough discussions of the
contributions of ozone and of hydrogen peroxide, also an oxidant, to the
oxidation of S02 and the roles of S02 as a precursor to acidic deposition..
3.3.6.1  Reactions of Ozone in Aqueous Droplets.   While  the thermal  oxidation
of SCL by ozone in the gas phase appears to be too slow to be important in
acid deposition phenomena, the role of ozone in oxidizing SC^ dissolved in
water droplets (e.g., cloud, fog or rain) may be of considerable
significance.  At 25°C ozone has a Henry's law constant of 10"^ mol L~^
       (Kirk-Othmer, 1981).  Given ambient concentrations ranging from 30
                                   3-50

-------
to ~300 ppb, Oj would be expected to have concentrations in aqueous
droplets in the atmosphere of approximately 3-30 x 10"10 mol L"1.  The
rate of reaction between 03 and S02, when both are dissolved in aqueous
droplets, has been shown in laboratory studies to be relatively fast
(Penkett et al., 1979; Kunen et al., 1983; Brock and Durham, 1984; Hoffman
and Jacob, 1984; Martin, 1984; Schwartz, 1984), but the rate of this
reaction is pH dependent and decreases as the acidity of the solution
increases.
     Figure 3-3 shows data reported by Schwartz (1984) for the rate of the
aqueous-phase oxidation of S(IV) by 30 ppb of Og (and also by 1 ppb of
I^C^) as a function of solution pH,  The aqueous-phase oxidation rate, R,
per part-per-billion S02 partial pressure decreases with decreasing pH by
roughly a factor of 20 per pH unit.  This pH dependence reflects the
solubility of S(IV), as well as a slight pH dependence of the second-order
rate constant for the oxidation of S(IV) by 03 (Erickson et al,, 1977;
Larson et al., 1978; Penkett et al., 1979).  Schwartz (1984) concluded,
from consideration of these data and uptake times for S0g» that the
oxidation of SOg by Og cannot produce solution pH values below ~4.5.
Schwartz (1984) has also, however, interpreted the field data of Hegg and
Hobbs (1981) for sulfate production rates at the inflow and outflow
regions of lenticular clouds as being consistent with the aqueous phase
oxidations of S(IV) by 03.
     An additional aspect of the role of Oj in the chemistry of aqueous
droplets concerns its photolysis to yield OH radicals in solution (Graedel
and Weschler, 1981; Chameides and Davis, 1982):

              (03)aq + hv -» 0(1D)aq + 02 aq                        (3-87)

        OtXq + (H2°)aq * 2(°H)aq                                (3-88)

and its reactions with aqueous OH" ions and #2®2 to yield aqueous H02
radicals (Chameides and Davis, 1982).  The OH radicals formed by this
aqueous process can result in the oxidation of S(IV).
3.3,6.2  Reactions of Hydrogen Peroxide in Aqueous Droplets.  Although
hydrogen peroxide formed in the gas phase from the reactions of
                                   3-51

-------
   10"
   10"'
o
CO
Q.
a.

"L
£
to
o.
QC
   10'8
                  H2O2, 1 ppb
                                              1000
                                              100
10  ^
    o
    X
    a
                                              0,1
                                               0.01
      Figure 3-3. Rate of aqueous-phase oxidation
      of S(IV) by O3 (30 ppb) and H2O2 (1 ppb), as a
      function of solution pH. Gas-aqueous equilibria
      are  assumed for all reagents. R/pSO2 rep-
      resents aqueous reaction rate per ppb of gas-
      phase SO2; p/L represents rate of reaction
      referred to gas-phase SO2 partial pressure per
      cm3—nr3 liquid water volume fraction.
      Source: Schwartz (1984).
                          3-52

-------
hydroperoxyl radicals plays a role in HOX chemistry in the troposphere,
and especially in the stratosphere (Crutzen and Fishman, 1977; Cox and
Burrows, 1979), the major importance of hydrogen peroxide arises from its
high solubility in water.  The latter ensures that a large fraction of
gaseous H2Q2 will be taken up in aqueous droplets.  Over the past decade,
evidence has accumulated that H202 .dissolved in cloud, fog, and rainwater
may play an important, and, in acidic droplets (i.e., pH <5), even a
dominant role in the oxidation of S02 to H2SO/j. (Hoffman and Edwards, 1975;
Penkett et a!., 1979; Dasgupta 1980a,b; Graedel and Weschler, 1981; Martin
and Damschen, 1981; Chameides and Davis, 1982; Calvert and Stockwel1,
1983, 1984; Brock and Durham, 1984; Hoffman and Jacob, 1984; Schwartz,
1984).  Discussion of several proposed mechanisms for previous rate
studies of the oxidation of S(IV) by H202 are beyond the scope of this
document, but have recently been reviewed by several authors (see for
example, Calvert and Stockwell, 1983, 1984).  Hydrogen peroxide may also
play a role in the oxidation of NQ2 dissolved in aqueous droplets,
although relevant data are limited (Halfpenny and Robinson, 1952a,b; Anbar
and Taube, 1954; Gertler et al., 1984) and additional research is
required.  In addition to the direct oxidation of S02 and N02 dissolved in
aqueous droplets, the photolysis of H202 to produce aqueous OH radicals:

                   (H202)aq + hv •+• 2(OH)aq                         (3-89)

can lead to oxidation rates of S(IV) that can be competitive with calcu-
lated oxidation rates of S(IV) by (H202)   and (03)a(. (Chameides and
Davis, 1982).
     It should be emphasized, however, that substantial uncertainties
remain concerning the quantitative role of H202 in the acidification of
aqueous particles and droplets (Richards et al., 1983).  This is further
complicated by the lack of reliable measurements of gas-phase H202
concentrations in the atmosphere (see Chapter 4).  Moreover, it has also
been suggested recently that H202 may be formed in situ in aqueous
droplets as the result of absorption of OH and H02 radicals and other
precursors into solution from the gas phase (Graedel and Weschler, 1981;
Chameides and Davis, 1982; Heikes et al,, 1982).
                                   3-53

-------
3.3.6.3  Reactions of Formic Acid in Aqueous Droplets.  As a gas-phase species,
formic acid (HCQQH) cannot strictly be defined as a photochemical oxidant.   It
can be scavenged  rapidly into water droplets, however,  and can potentially
function therefore as  an oxidant in cloud water and rain water.  It can also
be  differentiated from  other  acids  in  that it is  formed  readily  from the
reactions of  the  Criegee intermediates  discussed earlier and of hydroperoxyl
radicals with  formaldehyde  (Calvert  and Stockwell, 1983).  The  formation of
other acids  may  be  orders  of magnitude slower as  the  result of both  the
apparently lower  rates  of reaction of H02 radicals with the higher aldehydes
and the much lower atmospheric concentrations of the higher aldehydes (Grosjean,
1982).  Thus, formic acid is an example of a compound that is a non-oxidant or
weak  oxidant  in  the  gas phase but that  is  transformed  upon incorporation in
aqueous solutions into an effective oxidizer of S(IV).
     Formic acid  (as well as acetic acid) has been identified among the acidic
components of rain (Galloway et al., 1982).   Although much uncertainty remains
concerning their  quantitative roles, HCOOH and the higher organic acids potentially
play a minor but  still significant role in the acidification of rain.
3.4  METEOROLOGICAL AND CLIMATOLOGICAL PROCESSES
     As discussed  in  Section  3.3,  ozone  and  oxidants  are  formed by the action
of sunlight  on  the precursors, HQy and hydrocarbons.  The accumulation of the
products to  form an appreciable concentration is also dependent,  however, on
the prevailing  meteorology  in the vicinity  of the  precursor emissions.   To
understand the details of the effects of meteorology on air quality requires a
thorough knowledge of meteorology  and climatology,  but an appreciation of the
general factors important  in  the  formation  of  elevated  concentrations  of
oxidants is  relatively easy to acquire.   Following  is a brief  presentation of
some features of atmospheric  mixing and  transport  that will provide a basic
understanding of the  meteorological  factors  that affect the  concentrations of
ozone and other  oxidants in urban  and rural  areas.

3.4.1  Atmospheric Mixing
     The concentration of an air pollutant depends significantly on the degree
of mixing  that  occurs between  the time a pollutant  or  its  precursors are
emitted and  the arrival  of the pollutant at the  receptor.  Since, to a first
                                   3-54

-------
approximation, the diurnal cycle of weekday urban emission patterns for ozone
and oxidant precursor  pollutants  is  generally uniform, it is  reasonable  to
ascribe a significant  proportion of the large day-to-day changes in pollutant
concentrations to  changes  in meteorological  mixing processes.  The  rate  at
which atmospheric  mixing processes occur and the extent of the final dilution
of the pollutants  depends on the amount of turbulent mixing and on wind speed
and wind direction.  Moreover,  the transport of pollutants and precursors  from
a source region to a distant receptor is also dependent on wind speed and  wind
direction.
     The degree of turbulent mixing  can be characterized by atmospheric sta-
bility.  In an atmospheric layer with relatively low turbulence,  pollutants do
not spread as rapidly as they do in an unstable layer.   Also, because a stable
layer has a  relatively low rate of mixing, pollutants  in a  lower layer will
not mix through  it to higher altitudes.  The stable layer can act as a trap
for air pollutants lying beneath it.   Hence, an elevated inversion  is ofte.n
referred to as a "trapping" inversion.  Also, if pollutants are emitted into a
stable layer aloft, such as from a stack, the lack of turbulence will keep the
effluents from reaching the ground while the inversion persists.
     In air pollution  considerations,  a stable atmospheric layer or  situation
is  usually  equated with a  temperature inversion,  which is  a  layer  of the
atmosphere in which the temperature increases with increasing altitude, because
inversions are common and also represent the most stable atmospheric conditions.
The lowest part of an  inversion layer is called the base and is defined as the
altitude at which the temperature begins to increase.  The top of the inversion
is  the point at  which the temperature begins to  decrease  with  increasing
altitude.  The distance between the base and top of the inversion layer is the
"depth" or  "thickness" of the  inversion.  Inversion layers  may begin at  the
ground surface  (i.e.,  the  altitude of the base is  zero), or  the entire inver-
sion layer may be  above the surface.   The former is known as a "surface inver-
sion"  and  the latter as an  "elevated  inversion."  The  two types are usually
caused by different sets of weather conditions, but it is not unusual for both
types  of  inversions to be present at  a  given  location  at the  same time.   In
the  United  States,  surface  inversions are  characteristic of nighttime and
early morning hours except when heavy cloud cover or windy and stormy conditions
prevai1.
     Surface  and  elevated  inversion  layers are both important in determining
pollutant concentration  patterns  since,  as noted  above, mixing and  dilution
                                   3-55

-------
processes proceed  at  a relatively slow rate in such layers.   Thus, if pollu-
tants are emitted  into an  inversion  layer,  relatively  high concentrations can
persist for  a  considerable period of time or over a considerable distance of
wind travel  from the  source.   For example,  a surface inversion  in  the morning
could cause  automotive exhaust pollutants released at the surface during the
morning rush hours to persist with minimum dilution near the ground surface
for an extended period of time, probably for 1 or 2 hours after sunrise, until
solar radiation  heats the ground  and causes the  inversion  to disappear or
"break" (Hosier, 1961;  Slade,  1968).   High concentrations may  occur  at the
ground even  when an elevated inversion is  present  and the  layers below the
inversion are  unstable and are  undergoing  good mixing.   Such a persistent
elevated inversion  layer  is a  major meteorological  factor that  contributes to
high pollutant concentrations and photochemical  smog  situations  along the
southern California coast (Holzworth, 1964; Hosier, 1961;  Robinson,  1952).
     The vertical  mixing  profile through  the lower layers of the  atmosphere
follows a typical  and predictable cycle on a generally clear day.   In such a
situation a  surface  inversion  would be expected to form during the  early
morning and  to persist until  surface heating becomes significant,  probably 2
or 3 hours after sunrise.   Pollutants initially trapped in the surface inversion
may cause relatively high, local concentrations, but these concentrations will
decrease rapidly when the surface inversion is broken by surface heating,
usually about  midmorning.   The  surface inversion  will  begin to form again
during the early evening  hours and pollutants  from near-surface sources such
as automobiles  or  home fireplace  chimneys will experience progressively less
dilution as the surface inversion  develops.
     Elevated  inversions,  when the base is above the ground, are also common
occurrences  (Hosier, 1961; Holzworth, 1964).  Since these conditions,  however,
are  identified with specific  synoptic conditions,  they are much less frequent
than the nighttime  radiation inversion.   Because it may persist throughout the
day and thus restrict vertical mixing, an elevated inversion is nevertheless a
very significant air  pollution feature.   Smog-plagued southern California is
adversely affected by persistent elevated  inversions (Robinson, 1952).   When
compared to  a  source  near the  surface and the effects of a radiation (surface)
inversion, the pollutant dispersion pattern is quite different for an elevated
source plume trapped  in a layer  near the base of an elevated inversion. This
plume will  not be   in  contact  with the  ground  surface in the early morning
                                   3-56

-------
hours because there is no mixing downward through the surface radiation inver-
sion.  Thus, the  elevated  plume will not affect surface pollutant concentra-
tions until the mixing processes become strong enough to reach the altitude of
the plume.  At  that time,  the plume may be mixed downward quite rapidly in a
process called  "fumigation."  During fumigation, surface ozone concentrations
will increase if  the morning ozone concentration is higher aloft than at,the
ground, and if  insufficient scavenging by NO occurs at ground-level.  After
this initial mixing, surface concentrations will decrease as the usual daytime
mixing processes  continue  to develop.   If the  daytime mixing becomes  strong
enough to  break the upper inversion, the  pollutants may be mixed through  an
increasingly deep  layer  of the  atmosphere.  When surface heating decreases in
the late afternoon and early evening, both the surface and elevated inversions
will form  again.   The surface inversion will again  prevent pollutants from
elevated  sources  from reaching the ground and  surface  scavenging processes
will gradually  reduce the concentrations  of  pollutants trapped during the
formation of the surface inversion.
     Geography  can have a significant impact  on  dispersion of pollutants
(e.g., along the coast of  an ocean or one of the Great Lakes).  Near the coast
or  shore,  the  temperatures of land and water masses can be different, as can
the  temperature of the air above such land and water masses.  When the water
is warmer than  the land, there is a,tendency toward reduction in the .frequency
of  surface  inversion conditions inland over a relatively narrow coastal strip
(Hosier,  1961).   This in turn tends to increase pollutant dispersion  in such
areas.  Such conditions  may occur frequently on the Gulf  Coast and  near  the
Great  Lakes in  the winter.  The opposite condition also occurs if,the water is
cooler than  the land, as  in summer or fall.  Cool air near the water  surface
will tend to increase the  stability of the boundary layer in the coastal zone,
and  thus decrease  the mixing processes that act on pollutant emissions.  These
conditions  occur frequently, along  the New England  coast  (Hosier,  1961).
Similarly, pollutants from, the Chicago area have been observed repeatedly  to
be  influenced by  a stable  boundary  layer over Lake Michigan (Lyons and Olsson,
1972).  This  has been observed especially in summer and fall when the lake
surface is most likely to be cooler than the  air that  is carried  over it  from
the  adjacent land.
     Since  the  diurnal  mixing conditions  are such  an  important part  of the
meteorological  parameters  for understanding pollutant mixing and diffusion, it
                                   3-57

-------
is useful to  have some knowledge of the mixing cycles that prevail over the
United States,  Figures 3-4 and 3-5 show the average summer morning and after-
noon mixing heights  (AGL»  above ground level) as calculated on the basis of
upper air  temperature data and  an  estimated midmorning urban temperature.
Since Holzworth  (1972)  attempted to include the influence  of  an  urban heat
island in this  estimated  temperature,  the morning  results  in  Figure 3-4 are
probably most applicable to larger urban areas.  Rural or nonurban areas would
be expected to have lower mixing heights.
     Summer conditions  are  useful  to consider because of  the  prevalence of
high photochemical  oxidant  concentrations during this season.  As  shown in
Figure 3-4, morning mixing heights are estimated to be greater than 300 meters
except for the  central  part of  the  Great  Basin, where a 200-meter isopleth
includes parts  of Oregon,  Idaho, Utah, Arizona, and most of Nevada.   By mid-
afternoon (Figure  3-5), the estimated mixing  height  at  the time  of maximum
temperature has  increased markedly,  and  only a few  coastal  areas  have an
average afternoon maximum mixing height of less than 1000 meters.   In contrast
to the morning data, the central Great Basin area becomes the area of greatest
mixing in the afternoon.  This would be expected since this is a hot,  arid,
desert region,  and the driving  force generating the  surface mixing layer is
the solar heating of the ground  surface.
     The magnitude  of the afternoon mixing height is generally an indication
of the potential  for recurring  urban air pollution problems.  If a trapping,
elevated inversion  does not rise high enough  in the afternoon to release the
generated pollutants  that are  trapped, an accumulating episode is likely.
From the average summer afternoon data shown  in Figure 3-5, where the lowest
average mixing  height is  600 meters and  almost  all of the area has a value
greater than  1500 meters, it  would  appear  that urban  air pollution should not
be severe.   On the average this  is probably correct; however, there are several
departures from the average, which result in relatively  low mixing heights and
adverse dispersion  over many  areas  of  the  United States  on a  recurring  basis.
     Figure 3-6  (Holzworth and Fisher, 1979) shows the frequency of occurrence
of elevated  inversions in summer having a  base between I and  500 meters  (1600
feet) at the  time of the afternoon  upper  air temperature measurement, 6:10
p.m. EST or  3:15 p.m. PST.   The California coastal conditions, in which  low
inversions occur with a frequency of nearly 90 percent, are clearly evident.
The northeastern  coastal area from New Jersey  north to Maine, where cool ocean
                                   3-58

-------
                                                             11
 Figure 3-4. Isopleths (m x 10^} of mean summer morning mixing heights, AGL.
 Source: Holzworth (1972).
                                                                18
Figure 3-5.  Isopleths (m x 10^) of mean summer afternoon mixing heights, AGL.
Source:  Holzworth (1972).
                             3-59

-------
 30
Figure 3-6.  Percentage  of summer  2315 GMT (6:15 p.m. EST, 3:15 p.m. PST)
soundings with an elevated inversion base between 1 and 500 m above ground level.
Source: Adapted from Holzworth and Fisher (1979).
water prevails,  also has  a  relatively high percentage, above  20  percent,
compared to most  of the rest of the country.  Stations bordering one of the
Great Lakes—Green  Bay,  Sault  St.  Marie,  and Buffalo—reflect a stabilizing
lake effect with percentages above 5 percent.   Except along the  Pacific  Coast,
these ocean and  lake coastal  situations are probably  limited to  relatively
narrow coastal zones  (Hosier,  1961).   Examples  are evident in Figure 3-6,  in
which it may be noted that inversion frequencies of 21 to 28 percent occur in
coastal  New England compared to only 2 percent at Albany in upstate New York.
A similar situation is  evident in a comparison of  the 3 percent inversion
frequency at Washington,  D.C.,  with the 16 percent frequency on the Delaware
coast.  A non-coastal region having summer afternoon low-level  elevated  inver-
sions more than  5 percent of the time  is  the  Southeast, where  an  area  from
Louisiana and Arkansas  to the Atlantic coast shows frequency values between 5
and 10 percent.   Other seasons differ in details,  but the general  patterns  are
similar.
                                   3-60

-------
     This means that,  for  most of the United States, low-level  stable layers
that persist through  the  afternoon hours are rare events, occurring on less
than 1 day  in  20.   Thus,  air pollution situations in areas such as Kansas or
Iowa, although they can be the result of transport, may commonly be related to
the periods when the  expected  morning  surface inversion persists later in the
morning than usual  and when winds are  not  strong  enough  to carry pollutants
rapidly away  from the  local  area.  Along  both  the  Pacific  Coast and the
Northeast Coast, low-level  afternoon inversions  are frequent enough to be a
significant contributor to local and regional air pollution episodes.

3.4.2  Wind Speed and Mixing
     Another major  meteorological factor in the  urban pollutant dispersion
problem  is  low-level  or surface-layer  wind.  As  would be expected, strong
winds across a source area will dilute  pollutant  concentrations even though
there is  a  strong,  low-level  inversion base.  San  Francisco is  one example of
such  a  location where  strong  winds  frequently  provide good ventilation  in
spite of  a  low inversion.   Conversely,  light and variable or calm wind condi-
tions over  an  area can lead to excessive pollutant accumulations even though
the  afternoon  mixing  depth is quite large.  Thus,  it is necessary to include
wind  direction and  wind speed frequencies  in any evaluation of air pollution
potential for  a given area.   It must  also  be  recognized that both elevated
inversion conditions  and  surface  wind  patterns are governed to  a major degree
by  the  synoptic,  or large-scale,  weather patterns.  Both  wind  and inversion
factors tend to favor pollutant buildup when a deep, slow-moving high-pressure
system  dominates  the  weather  across an  area  (Korshover, 1967; Korshover,
1975).
      Figure 3-7 shows the wind climatology across the United  States in the
month of  July  by depicting the monthly  resultant vector wind at  major weather
stations  (U.S.  Department of Commerce,  1968),   Note that the flow across the
West  Coast  is  generally directed  inland, from west to  east.  This  contributes
to  a typical  situation for major  California cities:  significant  urban pollu-
tant  plumes  are found east of the urban core source  areas while the  immediate
coastline or  beach areas  are relatively pollutant-free.   In the  Northeast
States,  the  average wind flow  is  from southwest to  northeast more or less
parallel  to the coastline.  As  a  result, pollutant plumes from the major urban
areas along  this  coast are frequently  additive  along  the trajectory of the
                                    3-61

-------
CT>
                  SAN FRANCISCO
                                                                                   *^w:e—53"V      --"-^"j|
                                                                                   wS  N>  ^/TA
               ROI.     j
               t       f\
               t  1       \
                  '
                                                                                     /   M I
                                                                               MILWAUKEE GRAND RAPIDS
                                                 ALBUQUEHOUE AMARILLO  OKLAHOMA CITY     MEMPHIS

                                                                         LITTLE ROCK
                                                                              rrii      -r	 £

                                                                               /T  [ATLANTA*""*
      .   -,      ^n^uno  SHREVEPORT JACKSON muniuumcni   y I


^™°v       V   \   /r  i    n      7
 \       \        \    »   f  .. MOBILE—TALLAHASSEE-JljACKSl
  \        N   AUSTIN \  LAKE CHARLES-TjO^^ JC   Wj'AUK,

  V  /*>w     1 GALVESTONg     iPTT    ^^ X.  \\
   X/   A     \    «T~^^^V > GULFPORT       A    J
          \     \   J^L  I   ^>            TAMPA V
          A      \ rSr                      IA  \
                                                                                                             NOTE: BASED ON

                                                                                                             HOURLY OBSERVATIONS

                                                                                                             1951-1960.
                                                                                   SCALE IN mph
                        Figure 3-7. Mean resultant surface wind pattern for the United States for July. Direction and

                        length of arrows indicate monthly resultant wind.

                        Source: U.S. Dept. of Commerce (1968).

-------
wind.  Polluted air  moving toward the coast from major  inland urban sources
may  also  be  a factor in this Northeast  region.   Along the Gulf Coast, the
average winds  form  southerly,  onshore flow.  Under some weather situations,
however,  there  is  often an offshore flow  in one  area (e.g., Texas) and an
onshore flow  in an adjacent area.  Thus,  because of  this  recirculation, the
onshore Gulf  air  masses are not always pollutant-free (Price, 1976; Wolff et
al., 1981).   Before  the situation was examined carefully, the recirculating
pollutants were  sometimes confused with  natural  background  concentrations.
     Wind climatology provides  an average wind flow pattern, but it does not
provide a complete  assessment  of the influences of the wind on air pollution
dispersion. Wind  speed  and,  in  particular,  the  frequency of weak winds are an
important aspect  to  be  considered.   Figure 3-8, adapted from Holzworth  and
Fisher  (1979),  shows the  frequency with which early morning  (6:15 a.m. EST or
3:15 a.m.  PST) surface  inversions occur with calm or  weak  surface winds; that
is,  wind  speeds  equal to or less than 2.5 m/sec or 6 mi/hr.   There is consi-
derable variation between stations because terrain and geography (e.g., coastal
locations) influence both the wind flow and inversion frequency.   It is clear,
however,  that over large areas of the United  States, especially in heavily
industrialized  inland areas  east of the Mississippi  River,  calm  amd stable
summer  mornings are  a frequent  occurrence:  50  percent or  more in many areas.
This means  that there will  be  frequent incidents of morning  pollutant accumu-
lation; but  afternoon heating,  as shown by Figure 3-5,  will  usually mix the
pollutant accumulations through a deep mixing layer and disperse them.   Figure
3-9  (Holzworth and Fisher, 1979) shows  the average  wind speed through the
depth of  the  summer  morning mixing layer. Note that the area east of the Rocky
Mountains, except for the Appalachians,  can on  the average,  expect  winds of 4
m/sec  (about 10  mi/hr)  or higher through  the  morning mixing layer.   This
probably  would provide  acceptable midmorning dilution of  accumulated pollu-
tants.   In  summer afternoons,  as shown in  Figure 3-10, the average  wind speed
within  the  mixing layer increases in all areas and may even double over some
of the western mountain states.   It should be noted,  however, that  since winds
normally  increase with  altitude above the  ground, much of  the increase in the
average afternoon mixing  layer wind is probably the result of the considerable
increase  in  the depth of the mixed  layer,  as shown by the  differences  between
Figures 3-4  and 3-5.
                                    3-63

-------
                                                                            16
                                                                    41
Figure 3-8. Percentage of summer 1115 GMT (6:15 a.m. EST, 3:15 a.m. PST) soundings
with an inversion base at the surface and wind speeds at the surface ^2.5 m/sec.

Source: Adapted from Holzworth and Fisher (1979).
                                    3-64

-------
Figure 3-9. Isopleths (m/sec) of mean summer wind .speed averaged
through the morning mixing layer.
Source; Holzworth and Fisher (1979),
Figure 3-10. Isopleths (m/sec) of mean summer wind speed averaged
through the afternoon mixing layer.
Source; Holzworth and Fisher (1979),
                            3-65

-------
     In summary,  atmospheric  mixing parameters of stability and wind in the
pollutant transport  layers  can exert controlling effects on 0,  and oxidant
concentrations.   The  effects  include the amount of dilution occurring in the
source area,  as  well  as along the trajectory followed by an urban or source-
area plume.   Regions  having steady prevailing winds,  such  that  a given air
parcel can pass over a number of significant source areas, can develop signifi-
cant levels of pollutants even in the absence of weather patterns that lead to
the  stagnation type of  air pollution episodes.  The  Northeast  states  are
highly susceptible to pollutant plume transport effects, although some notable
stagnation episodes  have also affected this area (e.g.,  Lynn et al., 1964).
Along the  Pacific Coast, especially along the  coast  of California, coastal
winds and  a  persistent  low inversion layer contribute  to  major pollutant
buildups in urban source areas and downwind along the urban plume trajectory
(Robinson, 1952;  Neiburger  et al., 1961).  In the southern Appalachians, the
weather favors longer-term air pollution episodes (Korshover, 1967; Korshover,
1975).  Generally,  low  pollution   potential  results from the conditions that
occur in the  Great Plains area and  south to the Texas-Louisiana Gulf Coast;
and between the Mississippi River and the crest of the Rocky Mountains.
     It should be clear  even  from  this brief discussion  that there  are funda-
mental differences  in  regional  meteorological  conditions that cause the air
pollution potential  applicable generally  to California to be more  severe than
in other parts of the United States.  When adverse meteorology is coupled with
high population  and source  concentrations, it  is quite evident why  California
areas have severe photochemical air pollution problems.  This very significant
difference  in the magnitude of the photochemical  air pollution problem in
California regions  compared to non-California locations can serve as a basis
for  separating air pollution  statistics  into two sets  for evaluation, namely,
California and non-California groups.

3.4.3  Effects of Sunlight and Temperature
     The significance of sunlight in photochemistry is related to its intensity
and  its  spectral distribution, both of which  have  direct effects  upon  the
specific chemical  reaction  steps that  initiate and  sustain  oxidant  formation.
Sunlight intensity varies with season and geographical latitude but the latter
effect is strong  only during the winter months.  During the summer, the maximum
light intensity  is fairly constant throughout the contiguous United States and
only the  duration of the solar day varies  to  a small degree with  latitude.
                                   3-66

-------
     The effects of  light  intensity on individual photolytic reaction steps
and on  the overall process of  oxidant formation have been  studied  in  the
laboratory (Peterson, 1976; Demerjian et a!., 1980).   All  of the early studies,
however, employed  constant light intensities, in contrast  to  the diurnally
varying intensities that occur  in the  ambient atmosphere.  More recently, the
diurnal variation  of light intensity  has been recognized  and  studied as a
factor  in  photochemical  oxidant formation (Jeffries et al., 1975;  Jeffries
et al., 1976).  Such studies showed that the effect of this factor varies with
initial reactant  concentrations.   Most  important was  the observation that
similar NMOC/NO   systems  showed  different  oxidant potential  depending on
               /\
whether studies of these were conducted using constant or diurnal  light.   This
has led to  incorporation  of the effects of  diurnal  or variable  light into
photochemical models (Tilden and Seinfeld, 1982).
     While  the  effect of  sunlight  intensity is  direct and  has  been amply
demonstrated  (Leighton, 1961; Winer et al.,  1979), the effect of wavelength
distribution on the overall oxidant formation process is subtle.   Experimental
studies have  shown the photolysis of  aldehydes  to be  strongly dependent on
radiation wavelength in the near UV region (Leighton, 1961); and there is some
indication  (Bass et  al., 1980)  that the  photolysis rates for aldehydes may be
temperature-dependent.  Since aldehydes  are  major products  in the atmospheric
photooxidation of  NMOC/NO   mixtures, it is inferred that the radiation wave-
                         s\
length  should have an effect on  the overall photooxidation process.  This
inference was  directly verified,  at least for the propylene/NO  and  rrbutane/
                                                              x\      """"""'
NO  chemical  systems, in  smog  chamber studies (Jaffee et al., 1974; Winer
  X
et al., 1979).   In the ambient atmosphere,  some  variation  in the wavelength
distribution  of  sunlight does occur as a result  of variations  in  time of day,
stratospheric Oo,  ambient  aerosol (Stair, 1961),  and cloud cover.
     It has been  observed  that days on  which significant  ozone-oxidant  con-
centrations occur  are  usually days with warm, above-normal temperatures  (Bach,
1975).  The possibility that photochemical  reactions  show some  temperature
dependence  has been raised by smog chamber studies (e.g., Carter  et  al., 1979;
Countess et al., 1981), but is  usually thought to be the production  of chamber
artifacts  concomitant with increases  in temperature.   In ambient air,  the
observed  correlation  between temperature and oxidant  concentrations can be
explained,  at least in part,  as a  synoptic meteorological correlation rather
than  as a temperature-photochemical rate constant effect,  in that periods of
                                   3-67

-------
 clear skies  and warm temperatures  are  periods  of  high  air  pollution potential,
 as  discussed above.  Because  of  the close correlation between above-normal
 temperatures and high photochemical air  pollution  potential,  a maximum  daily
 temperature  forecasting procedure is  often  useful  as  a substitute for a more
 elaborate  and  specific  program for forecasting  possible  photochemical  air
 pollution.   The correlation between temperature  and,  thus,  synoptic weather
 conditions  and photochemical air  pollution  intensity  has  been observed  in a
 number of  areas.  Evaluation of photochemical  air pollution in Los  Angeles as
 early as 1948  showed a  correlation  with  temperature.   Recent studies of  0-
 patterns in  St.  Louis,  Missouri,  have also shown  a  correspondence  between
•daily maximum CL concentration and  temperature  (Shreffler and Evans, 1982).
 Wolff and Lioy (1978) have shown  high  correlations  between ozone  concentration
 and  both the temperature for the current  day and  the temperature for  the
 previous day.

 3.4.4  Transport of Ozone and  Other  Qxidants and  Their Precursors
      The 1978 air quality criteria document for ozone  and  photochemical  oxidants
 made a convincing case  for the fact  that  ozone and  other photochemical oxidants
 are  transported from urban  source areas,  other  than those in California,  to
 downwind regions in concentrations  of 0.1  ppm or greater  (U.S. Environmental
 Protection Agency,  1978a, and  references  therein).  The 1972  study by Research
 Triangle Institute at McHenry, Maryland, was  an  early examination of rural
 oxidant in  the  eastern  United  States  (Ripperton  et al.,  1977).   Bach (1975)
 examined the meteorology  of these observed conditions and showed details  of
 the  influence  of transient high  pressure systems.   The transport  of large
 urban plumes in  the  northeastern  states,  especially from New York City  into
 Connecticut, was the subject  of  an  EPA field  study in 1975  (Westberg et al.,
 1976; Wolff et  al.,  1977d;  Westberg et al., 1978a).   Transport of ozone on  a
 regional basis in the northeast was  also  described  by  Cleveland et al. (1976a,
 1976b), while Wolff  et  al.  (1977a)  presented  details of  several east coast
 urban oxidant plumes.   The  Northeast  Regional Oxidant Study (NEROS) carried
 out  by EPA  in 1979 and  1980 in the corridor from Washington, D.C.,  to Boston,
 was  designed specifically to support  urban plume model development (Clarke et
 al., 1982).   Plume models have been  based on other  urban plume investigations,
 as  well (e.g.,  Wolff et al.,  1977c),  and reactive plume  modeling for other
 urban areas  has progressed in  recent years  (e.g., U.S. Environmental  Protection
                                    3-68

-------
Agency, 1981a).  These several early research programs and analyses served to
answer a  number of perplexing problems that had been  identified previously in
studies of ozone and other photochemical oxidants  in nonurban areas  (e.g.,
Ripperton et al.,  1977; Bach, 1975).  These questions included reasons for the
occurrence of  high ozone or  total  oxidant concentrations  in  areas  remote  from
identifiable sources.
     Studies of the transport of ozone and other photochemical oxidants (0~-0 )
                                                                          *5  X
are  classified into  three regimes, depending upon  transport distance (U.S.
Environmental  Protection Agency, 1978a).   In the first, urban-scale transport,
the  occurrence of  transport  of photochemical pollutants  can be detected in
most large urban areas if there is sufficient Q--Q  monitoring  information.
                                                *3   X
It has been  identified as a significant, characteristic feature of the 0~-Q
                                                                         •C*  X
problem in the Los Angeles basin (Tiao et al., 1975), as  well  as  in San  Fran-
cisco, New York, Houston, Phoenix,  and St.  Louis (Altshuller, 1975; Coffey and
Stasiuk,  1975; Shreffler and Evans, 1982; Wolff  et al., 1977a).  As  noted
above, the  recognition and  assessment of the probable  magnitude of urban
oxidant problems in  locations other than in California has been a major research
topic  since  the mid-1970s  (e.g.,  Bach, 1975; Cleveland et al.,,1976a, 1976b;
Westberg  et al., 1976; Wolff et al., 1977a,b,c; and others).
     Urban-scale transport patterns result  from one  or more of a combination
of factors.  First is the simple advection of the photochemically reacting air
mass  and  -the  development of maximum  0~-0   after 1 or 2 hours of downwind
                                       O  A
travel.   Maximum concentrations may be displaced up to 20 or so miles from the
center of the  major  source area.   It has been noted (U.S. Environmental Protec-
tion Agency, 1978a)  that pollutant  concentrations in  air parcels in the central
core area of major source areas may not be the most conducive  for 0~-0  forma-
                                                                   *3  X
tion because of the  tendency toward occurrence there  of more effective scavenging,
especially scavenging  related to NO and  its reactions.
     The  distance of.the peak 0~-0  concentrations  from the  urban core area  is
                               O   X
dependent on the  local wind pattern and is, in general,  inversely related to
the  peak 0~-0 concentration.   Stronger winds will  carry the air parcels
,           «3   X
farther during the reaction  period, increasingly diluting pollutant concentra-
tions  along  the trajectory.   Weak  winds and  very restricted mixing heights
will  tend to  cause  higher Q»-0  concentrations  closer to the central  source
                             *3  X
area.   The diurnal  wind cycle will also  be  an  important  factor, since in some
situations calm conditions may prevail until late in  the  morning but  in  others
                                    3-69

-------
a steady wind  may be present throughout  the  emission and reaction process,
Wolff and  Lioy (1978)  were able to  model  urban ozone concentrations on the
basis of meteorological observations, especially temperature.
     The second,  or  mesoscale,  kind  of  transport  of  CL-0   is in many  respects
                                                      *5  X
an extension of the  urban-scale transport and is  characterized by  urban  plume
development.   A  report by  Bell  (1960) described November  1959 OVO incidents
                                                               O   JTX
in northern coastal San Diego County, California.   It showed conclusively that
these were caused by the 0,,-Q  and precursors formed and emitted,  respectively,
                          3  X
the previous day in  the Los Angeles  basin.  The transport in these situations
was over the coastal Pacific Ocean, and  the  Q.j-0  arrived at the San Diego
                                               *j  X
receptor site  as a  contaminated sea breeze  after overnight travel   (Bell,
1960).  The studies of Cleveland et al.  (1976a, 1976b) are early documentation
of a similar scale of transport in the New York-Connecticut area.  The results
of extensive aircraft  measurements and  modeling assessments  of studies in  the
Washington, D.C., to Boston corridor have also  been  reported by Wolff and  his
colleagues (Wolff et al., 1977a,c; Wolff and  Lioy, 1978).
     In the 1978 0--0   criteria document, more than 30 references were cited
                   *3  X
relating to urban plume observations and  investigations.   Since that  document
was published,  the  results of the 1975 New  England oxidant study have been
published in detail  (e.g., U.S. Environmental Protection Agency, 1977; Westberg
et al.,  1978a),  and results  of a more comprehensive  2-year field program
carried out along the  Washington, D.C.-Boston corridor in 1979 and 1980 have
appeared in  the  literature  (Clark and Clarke, 1982; Clarke et  al.,  1982;
Vaughan et  al,,  1982).   A major field  program  supported  by  local  industries
has been conducted in Houston, Texas, although the D.,-0  downwind plume phases
                                                    O  f\
of that study were not as extensive as in NEROS.  Chicago, Detroit, and adjacent
shoreline areas  of  Lake Michigan have  also been  the subject of a number of
ground-level and airborne studies  over distances of 70  to  300 kilometers
(Lyons and  Olsson,  1972; Sexton and Westberg,  1980;  Westberg et  al., 1981;
Kelly et al.,  1986).   As described above,  Oo~0  plumes  from  major  urban  areas
                                           O   )\
can  extend  about 100  to 200  miles,  with widths  of  tens  of  miles (Sexton,
1982), frequently up to half the length  of  the plume.   Other field  studies
conclusively demonstrating mesoscale transport over New England  have been
reported (Spicer  et  al., 1979; Clarke et al., 1982; Cleveland et al., 1976a,b;
Rubino et  al., 1976; Westberg et al., 1976;  Westberg et al., 1978a; Wolff et
al.,  1977a).   Although urban plumes are  frequently  thought  of  as a problem
                                   3-70

-------
related only to large source areas such as New York and other major metropolitan
areas, measurements  in  plumes  from smaller urban areas have shown that these
sources cannot be ignored (Sticksel et al., 1979; Sexton, 1982; Spicer et al.,
1982; Wolff et al., 1977c).
     In the third  kind  of pollutant transport, synoptic-scale, the transport
of 0~-0  and precursors is characterized by the general and widespread elevated
    «5  y\
concentrations of pollutants that can occur on an air-mass scale under certain
favorable weather patterns.   These weather situations are generally slow-moving,
well-developed high-pressure,  or anticyclonic  systems.   This type of deep
high-pressure area was  considered  by  Korshover (1967,  1975) as a prerequisite
for stagnating air pollution episodes.
     A major  criterion  of these synoptic systems is the reinforcement of the
surface high-pressure area  by  a warm high-pressure  circulation in the upper
air.   The  surface weather  is  then  frequently characterized by weak winds,
stable surface layers, and  high pollution potential over regional  or air-mass-
sized  areas.  This  is the  generalized meteorological  model  pattern  that is
involved in synoptic-scale  pollutant transport (Korshover, 1967;  Korshover,
1975).  As  with all  generalized models, there  have been specific oxidant
episodes that departed  from this model.  In  particular,  weak winds  are not
always a prerequisite;  and  relatively strong  winds  have  been  observed to be
associated with some oxidant transport episodes (Mueller and Hidy, 1983; Wolff
and  Lioy,  1980), when  such  winds  did  not also produce  rapid plume  dilutipn as
is usually expected.
     While surface  highs  that  are  reinforced  aloft by  a warm  high  can lead to
air  pollution episodes, there are  other high-pressure systems that usually
cause  few  or no  widespread air pollution problems.   These  are  the strong
surface highs found  behind  rapidly moving storm systems,  in which  the surface
high underlies a cold low-pressure  system aloft.  These systems characteristi-
cally  have good vertical  mixing (instability),  brisk winds, and low air
pollution potential.
     Bach  (1975)  assessed in detail the  relationship  between  elevated ozone
levels and synoptic  systems.  Other investigators have since described specific
instances  of  large-scale  ozone transport and  associated  meteorological condi-
tions  (e.g.,  Wolff and Lioy, 1980; Wolff et  al.,  1980;  Wolff et  al.,,1982).
For  example, Wolff et al. (1982) have described synoptic meteorological systems
and  the occurrence of haze  and elevated  ozone  levels in an area extending from
                                   3-71

-------
the Midwest  to  the Gulf Coast.  In  a  separate study, Wolff and Lioy (1980)
examined the  spatial  and temporal  distributions of ozone during three photo-
chemical smog episodes in July 1977.  A stagnating high-pressure system formed
over the Gulf of Mexico and two high-pressure systems originating  in Canada
were described  as  the respective meteorological systems associated with the
three episodes.  In  all  three cases, elevated  ozone  concentrations (~120 to
130 ppb) were found to extend, in a virtual "ozone river",  from the Gulf Coast
to New England (with 328 ppb ozone measured in Connecticut).   The high-pressure
system originating in the Gulf area affected  the  entire southeastern United
States and extended  from western Texas, northeastward through  Illinois, and
east to the  Atlantic Ocean.   Elevated ozone  levels affected anywhere from  a
few hundred  square miles to  a  thousand square  miles during the  1-week period.
     The importance  of  synoptic-scale or  air-mass pollutant  situations has
been recognized  for  many years,  probably  much  longer than the  importance of
major plumes  has been apparent.   The Donora,  Pennsylvania, smog episode in
1948 (Schrenk et al., 1949), while not a photochemical smog situation, involved
the occurrence  over  a wide area  of  a regional  air  mass having relatively high
pollution levels simultaneous  with  the occurrence of a stagnating warm high-
pressure area  over the Ohio  Valley  and the northern Appalachian area,   Donora
was an especially adversely affected pocket within this larger system; in that
case, however, ozone was probably not one  of the important pollutants.
     The synoptic-scale  high-pressure air pollution  system is  not charac-
terized by well-defined  urban plumes.  Rather, a warm, slow-moving or stagnant
anti-cyclone  provides  a  synoptic-scale weather system that, because of weak
winds and limited  vertical mixing,  favors  the  accumulation of relatively high
concentrations of  air pollutants.   On a climatological basis,  these systems
are most common  in the  summer  and fall months  over most of the  United States,
as shown by the work of  Korshover (1967; 1975).  In many cases,  an anti-cyclone
will stagnate or recurve and intensify over the Midwest or East as circulation
patterns in  the  upper air change and  become  more  supportive  of the surface
anti-cyclonic pattern  (Schrenk et al., 1949;  Lynn  et  al., 1964).  The typical
paths followed  by  the air masses forming  these slower-moving anticyclones in
the  summer  and  fall  months, as  described  by Bach (1975) and Wight et al.
(1978), is southeastward from Canada into  the upper Mississippi  Valley, eastward
across  the  Ohio Valley  and  then across the East  Coast, either toward the
northeast into New England, east into the  central  Atlantic States, or southeast
                                   3-72

-------
into the southeastern states.  The actual stagnation of a pressure system for
one or more  days,  as occurred in the  Donora  case,  is, of course, the most
severe example of a slow-moving system.
     Along the  West Coast,  air  pollution problems are also  the  result  of
persistent high-pressure system  influences.   In this case,  however,  the  high
is the persistent  subtropical  anti-cyclone of the eastern Pacific rather than
the series of transitory anti-cyclonic systems characteristic of the  area east
of the  Rocky Mountains.   The persistent  or semipermanent  subtropical  anti-
cyclone in the Pacific is linked to the large-scale general  circulation of the
atmosphere rather  than  to  moving wave systems (Neiburger et a!.,  1961).   The
effect  is  much the  same, except  that the area of  limited mixing  and  more
adverse air  pollutant effects  is  found on the eastern  edge of the subtropical
anti-cyclone rather  than on the trailing western edge as  in the transitory
systems.
     It is worthwhile to point out that the typical pattern of ozone  concentra-
tions in the slow-moving midwest anticyclone,  in contrast to the conditions in
California,  shows  the  lowest values  in  the eastern portion of the  system,
behind the advancing cold front.  The central  area of the anticyclonic system,
where dispersion  is  usually at a minimum, shows a  broad expanse of  elevated
ozone concentrations; but  the highest ozone values in the typical system are
usually found  in  the western parts of the system, the so-called "back side,"
where dispersion  conditions,  although better  than  in  the central portion  of
the  system,  combine with  increased residence  time  and  longer  exposure  to
emission sources  to  cause  the maximum ozone  concentrations  in  a given high-
pressure system.   Investigators describing this ozone pattern over the midwestern
and  eastern  parts of the United  States  include Bach (1975),  Vukovich et al.
(1977), Wolff et al. (1977b), Wight et al. (1978), and Westberg et al.  (1978a,
1981).
     The identification and understanding of photochemical Q~-Q  and precursor
                                                            •5  X
transport by weather systems has provided a significant advance in comprehending
photochemical air  pollution and the potential  extent of its effects.   Considerable
progress has been  made in the  development of long-range photochemical modeling
techniques so  that the likely impact  of  synoptic systems can be anticipated.
Such  tools  are very much in  the  research stage because the  local impact of
0,-Q  results from a complex  interaction  of distant  and local precursor, sources,
  «5   X
                                   3-73

-------
urban plumes, mixing  processes,  atmospheric chemical reactions, and general
meteorology.

3.4.5  Surface Scavenging in Relation to Transport
     Major scavenging processes  for CU in the atmospheric boundary layer are
adsorption and subsequent destruction at the ground surface (i.e.,  dry deposi-
tion),  and reactions with  boundary  layer  pollutants,  especially  NO  and
alkenes.  In dry deposition, eddy diffusion moves air parcels downward through
the turbulent boundary layer to the laminar sub-layer where individual  molecules,
such as  ozone,  can  move by Brownian motion to the underlying surface.   There
reactive molecules,  such as ozone, can be removed by reactions at the surface.
Chemical reactions between ozone and NO and reactive hydrocarbons are described
in other sections of this  document.   When they occur  in  the near-surface
layers,  these  reactions can  play an important role in the  boundary layer
scavenging process.   Dry deposition  and boundary layer chemical  reactions
result in  a  vertical  concentration gradient, with the  lowest concentrations
occurring at the surface of the ground.
     Because of this  surface-scavenging  process,  ozone will  persist in an
atmospheric parcel  in the  absence of ozone-forming reactions  only if the
parcel  is  dispersed  such that contacts  with  the  ground surface or  surface
pollutant sources are minimized.   It  is  likely that  only in those  air parcels
moving above the  surface layer will  ozone  escape  the  surface reactions and
persist  long enough to  undergo long-distance transport or persist overnight.
It should be noted here that ozone transported aloft, judging from the limited
data available, may be 20 to 70 percent additive in urban areas, as determined
by simulations  using  the EKMA chemical  kinetics model  and several  different
rates of vertical mixing (U.S. Environmental Protection Agency,  1977b).   Thus,
if 0.1  ppm ozone were to be transported aloft and was 40 percent additive at
ground level, the contribution of transport to the  peak ozone  concentration
downwind would be 0.04 ppm.
     Aircraft  observations  have  documented frequently the  occurrence  of
relatively high  ozone concentrations  above  lower-concentration  surface  layers
(e.g.,  Westberg et a!.,  1976).   This is a clear  indication  that  ozone is
essentially preserved in layers  above the surface  and can be  transported over
relatively  long  distances even when continual  replenishment through precursor
                                   3-74

-------
reactions is not  a  factor,  such as at night.   Boundary-layer scavenging pro-
cesses are also  responsible  for the fact that ozone concentrations in urban
areas rapidly fall to zero during the night.

3.4.6  Stratospheric-Trppospheric Ozone Exchange
     The  fact  that  0,  is formed in the stratosphere,  mixed downward,  and
incorporated into the  troposphere,  where it forms a more  or less uniformly
mixed background  concentration,  has been known in various degrees of detail
for many years (Junge, 1963).
     It  is widely accepted that  the long-term average tropospheric background
concentration at  the  surface ranges from about 30 ppb  to 50 ppb  (Fabian and
Pruchniewicz, 1977; Oltmans,  1981);  and  that it results primarily, though  not
exclusively, from the transfer of stratospheric _ozone into the upper troposphere,
followed  by  subsequent  dispersion throughout the troposphere (e.g.,  Singh et
a!., 1980;  Kelly  et al., 1982).  Ozone residence time in the troposphere has
been estimated at 1 to 2 months  by Junge  (1963); and at 1  to 2  months for
spring and  summer but  at 2.5 to  3.5 months  for  fall and winter, respectively,
by Singh  et al. (1980).
     The  mechanisms by  which stratospheric air is mixed into the troposphere
have  been examined by  a number  of authors.  Danielsen conducted extensive
analyses  of major synoptic weather events that injected stratospheric air into
the  troposphere  (Danielsen,   1968;  Danielsen and Mohnen,  1977;  Danielsen,
1980).   Reiter has been especially active  in  describing the atmospheric
mechanisms by  which stratospheric air injection  takes  place and  in relating
these processes  to  the global circulation  of the atmosphere (Reiter,  1963;
Reiter and Mahlman, 1965; Reiter, 1975).  As a result of such research, exchange
between  the  stratosphere and troposphere  in  the  middle latitudes has  been
determined to occur to  a  major extent in events called "tropopause folds."  In
a  tropopause fold (TF), the  jet stream  in  the upper  troposphere plays  a major
role  in  directing stratospheric air and high ozone  concentrations  into the
troposphere.   Figure 3-11 is  a schematic presentation of the  intrusion process
as  described by  Danielsen (1968).   The  subsidence occurs  along the poleward
side  of  the polar jet  stream in the area  where the  jet is associated  with a
cold  front  at ground  level.   The result is  downward  transport in  the cold  air
behind the cold front.
                                    3-75

-------
Figure 3-11. Schematic cross section, looking downwind along the jet stream,
of a tropopause folding event as modeled by Danielsen (1968).

Source: Johnson and Viezee (1981).
                             3-76

-------
     Since 1978, a  considerable  amount of research on TF and ozone injection
has been  done,  especially by  SRI-International  (Johnson and Viezee, 1981;
Ludwig et al,, 1977; Singh et a!., 1980; and Viezee et a!., 1979).   Figure 3-12
from Johnson and Viezee (1981) shows one example of the probing by SRI of a TF
event In  the  midwestern  United States.  Concentrations of ozone in excess of
90 ppb were found as  low  as 13,000  feet  (ca. 2,5 miles or  3.9  kilometers), as
shown in the upper part of Figure 3-12.  These authors found that ozone intrusion
was lower during  this fall  study (October 5, 1978) than in a number of other
spring TF events.   The dew point measurements in the second part of the figure
confirm the stratospheric injection.  The weather pattern accompanying this TF
is shown  at the bottom of Figure 3-12  by a  500 millibar (about 6  kilometer)
chart; the surface  cold front  is  also  indicated.  Note that the intrusion was
detected well behind  the  cold front and appears to have assumed a  layered
formation in the altitude range of 8,000 to 12,000 feet (2.4 to 3.6 kilometers).
     From their analysis  of measurement  flights in  a  number of TF  situations,
Johnson and Viezee  (1981) concluded that the  ozone-rich intrusions studied
sloped downward toward the south.  In terms of dimensions,  the average crosswind
width (north  to south) at an  altitude  of 5.5  kilometers (ca.  18,000  feet or
3.4 miles) for  six  spring intrusions averaged  226 kilometers  (364  miles), and
for four fall  TF systems, 129  kilometers (208 miles).   Ozone concentrations at
5.5 kilometers  (ca.  18,000  feet or 3.4 miles) averaged 108 ppb in the spring
systems and 83  ppb  in the fall systems.  Previously it had been assumed that
only a few fairly intense systems would produce a TF event and trans-tropopause
mixing.   From their data, however, Johnson  and Viezee  (1981) drew the very
important conclusion  that all  low-pressure trough systems,  such as that
illustrated in  Figure  3-12, may induce a TF event and cause the trans-tropopause
movement  of ozone-rich air into the troposphere.
     On the basis  of their field  studies and  the earlier  models and work of
Danielsen (1968), Johnson and  Viezee (1981) proposed a set of model mechanisms
or types  of  TF  injection, which  are illustrated in Figure 3-13 and described
in the following general  manner:

     1-   Type 1.  The  intrusion  is broken up and dispersed by mixing and
                  diffusion in the middle or free troposphere.
     2.   Type 2.  The  intrusion  persists down to the planetary boundary
                  layer  or  the top of the  mixed  layer,  where the  lower
                  part of the  intrusion may be incorporated  into the
                  mixed  layer and  may subsequently  reach  the ground.
                                    3-77

-------
   28

   24

   20

   16

   12

_j   R
(/)   O
5
*   4

1   0
a
3
      3   '8.g'|i«§8l3  4
          50
                               I   I
                                                        1    I
         (A) CROSS-SECTION OF OZONE, ppb



       I    I    I    I   I   1   I
                                     CNG
                                  I   I    I
                                             J	I
                                     40-
                                    ,   MEM
28

24

20

16

12

 8
I
                         I   I   I    I    I    I
            -30
         (BI CROSS-SECTION OF DEWPOINT TEMPERATURE, °C

                                        CNG
       I    I    I
                   J	I
                           I
             I
             I   I   I    L
                       MEM
                    I   I    I
                                        as

                                        7,3

                                        6.1

                                        4.9

                                        3.7

                                        2.4

                                        1.2

                                        0

                                        8.5

                                        7.3

                                        6.1

                                        4.9

                                        3.7

                                        2.4

                                        1.2
    CMI
       -160
       (-296)
          -120
          I-222)
 -80
(-148)
-40
(-74)
0
(0)
                                               +40
                                                (74)
+80
(148)
+120
 (222)
+ 160
 (296)
                        DISTANCE FROM CNG, nautical mi and (km)
                       ****** FLIGHT TRACK .
                       	 CONTOURS (DYNAMIC METERS)
                        V V f V   SURFACE WEATHER FRONT

         1C) 500-mb CHART AND FLIGHT TRACK

      Figure 3-12.  Measured vertical cross-sections of (A)  ozone, (B)
      dew-point,  and (C) the 500-mb chart and the flight track for
      October 5,1978. CMI = Champaign, IL; CNG = Cunningham, KS;
      MEM = Memphis, TN.

      Source: Johnson and Viezee (1981).
                                    Cfl
                                    5

                                    j<
                                    ul
                                    a
                                    3
                                 3-78

-------
                               Q
                              *Z
                               3
                               O
                               DC
                               O

                    LATJRAL   £
                    MIXING    O
                               CO

                               I-

                               2
                               UJ
 TYPE1
                                       MIXING LAYER
TYPE 2
7///////////////////W/fl7/////W//7/////////////W
WMm/m//W//Mwm/w//^^^^
  Figure 3-13. Hypothesized models of the process that mixes tropopause
  folding events into the troposphere.

  Source: Johnson and Viezee (1981).
                           3-79

-------
     3-  Type3.  The intrusion occurs close behind the cold front, where
                  the air parcels are caught by the downdrafts behind the
                  cold  front;  and is brought  to  the ground by direct
                  circulations associated with the front.
     4-  Type 4.  The ozone-rich parcels are incorporated into convective
                  cells  and  brought to the ground  in  association with
                  rain-showers and  thunderstorm downdrafts; similar to
                  Type 3.

     Johnson and Viezee  (1981)  summarized the possible impacts of these four
types of TF  events  by noting that Types  1 and 2 should produce "relatively
moderate effects" at  the ground in comparison  to  those to be expected from
Types 3 and  4.   The latter two could cause "substantial" effects in terms of
high surface ozone  concentrations.   The action described by Types 3 and 4 is
supported  by meteorological  theory (Bjerknes,  1951) and  by observations of
surface ozone such  as those made  by Daniel sen  and Mohnen  (1977),  Lamb (1977),
and Davis and Jensen (1976).

3.4.7  Stratospheric Ozone at Ground Level
     After a detailed review of  background tropospheric ozone, Viezee  and
Singh (1982) concluded,  as also concluded by other  investigators  (e.g.,  Kelly
et al., 1982),  that the stratosphere is  a major  but not the sole source of
background ozone in the unpolluted troposphere.  This stratospheric ozone is
brought to the surface mixed layer by vertical mixing processes that have been
known for many years (Junge,  1963).  In the northern hemisphere, at midlatitudes
between 30°N and 50°N,  annual average background surface  ozone  concentrations
generally range between 30 and 50 ppb, but in the tropics, lower concentrations,
15 to  20 ppb, prevail  (e.g.,  Fabian and  Pruchniewicz,  1977; Oltmans, 1981).
The  stratospheric  ozone reservoir has  a  strong seasonal variation, with a
maximum in the  spring and a minimum in fall and winter months, especially at
middle latitudes.  This seasonal cycle is reflected at ground-level background
observation stations, where the average spring background ozone at some stations
may be as high as 80 ppb and the average fall values range between 20 and 40 ppb
(e.g., Singh et al., 1977; Mohnen, 1977; U.S. Environmental Protection Agency,
1978).  In the troposphere, concentrations generally increase gradually to the
tropopause,  but  the seasonal  pattern is  the  same  (Viezee and Singh, 1982).
     Using data  acquired in  their studies of  TF events,  researchers at  SRI-
International examined the frequency with which stratospheric intrusions occur.

                                   3-80

-------
According to Singh  et  al.  (1980), an intrusion can be expected to be present
over the  United States  on  about 90 percent  of  the days.   This  frequency
diminishes somewhat  in summer,  However (Singh et  al.,  1980),  which is the
season when most ground-level smog episodes occur.
     Viezee and Singh  (1982) concluded that relatively high ozone concentra-
tions can occur for  short periods of time, minutes  to a few hours, over local
areas as  a result  of stratospheric intrusions.  They were  able to document
from published literature ten situations of probable intrusion of stratospheric
ozone.   These  instances  are  shown in Table 3-7,  reproduced from Viezee et al.
(1983).  The concentrations reported in Table 3-7 were measured at ground-level
stations.  Note that all • of the short-term situations  in which peak concen-
trations exceeded 80 ppb occurred in fall, winter, or spring months and not in
the photochemically  active  summer season.  Of the  three summer  instances that
were reported,  two  at  Whiteface Mountain, New York, and one at Pierre, South
Dakota, the highest  reported concentration was 56 ppb for a 1-hr average.   The
lower  incidence  in summer of reported  ground-level  impact  by stratospheric
ozone may be  attributable in part to  the reduced  frequency of  intrusions in
summer, as reported  by Singh et.aT.  (1980).   In addition, however, the potential
for ground-level  impact  by stratospheric ozone  in summer is lessened because
of the  stability  provided by the  upper-level, warm anticyclone  present in the
weather  systems  characteristic of  summertime photochemical smog  episodes.
     There have been a number of attempts to quantify  the proportion  of the
surface  ozone  attributable  to stratospheric  sources.  The most common method
used is based  on the assumption that  Be  is a  unique tracer for air parcels of
stratospheric  origin.   Both ozone and  Be are measured and the proportion of
the surface ozone that might be of stratospheric origin is calculated by using
a  derived ratio between ozone and  Be.   Some  of the investigators using this
technique include Ferman and Monson (1978), Wolff et al. (1979), Husain et al.
(1977),  Dutkiewicz  and Husain (1979),  Singh  et  al.  (1980),  and Johnson and
Viezee  (1981), among others.
     Calculated  correlations between surface  ozone  and  Be  show that their
relationship is variable.  The results  of Kelly  et  al. (1982), from  studies of
 Be, 0Q,  and  air mass  classification in South Dakota,  showed  that continental
                                                          7
tropical  (cT)  air masses  frequently seemed to have  higher  Be and 0^ values on
the western  side  of the traveling cT anticyclone.   A similar  relationship was
not prominent  in  the polar air masses studied, however,  and maritime tropical
                                    3-81

-------
                                      TABLE 3-7.  PUBLISHED EPISODES OF TRANSPORT OF STRATOSPHERIC OZONE TO GROUND LEVEL
Case
no.
1
2
3
4
5
c*>
CO
rog
4
10
6
7
8
Date
3 March 1964
26 February 1971
19 November 1972
6 March 1974
8,9 January 1975
11, 12 July 1975
19 March 1977
24, 25, 28 June and
1 July 1977
4 March 1978
July 1978
15 March 1978
Ground-level 03
Geographic location concentration, ppb
Quincy, Florida (near Tallahassee)
Observatory Hohenpeissenberg
(1000 m MSL), SW of Munich, Germany
Santa iosa, California
Harwell, Oxon, United Kingdom
Zugspitze Mountain, near Garaisch-
Partenkirchen, Germany (3000 in MSL)
Whiteface Mountain, New York
(1150 m MSL)
Sibton, Suffolk, United Kingdom
Whiteface Mountain, New York
Denver, Colorado
Pierre, South Dakota
Kisatchie National Forest, Louisiana
100 to 300
415
250
200 to 230
110 to 115
160 to 193
< 37
100 to 110
< 47
82
< 56
< 46
100 to 105
Duration of
observed event
3 hr
10 rain
50 rain
1 hr
2 hr
4 hr
24-hr average
2 hr
24- hr average
1 hr
1 hr
24-hr average
2 hr
Length of data
record examined
July 1963 through
July 1973
December 1970 through
May 1971
November 1972
4 to 5 yr discontinuous
August 1973 through
February 1976
July 1975
4 to 5 yr discontinuous
June and July 1977
1975 to 1978
July through September
1978
Spring 1978
Source
Davis and Jensen (1976)
Atmannspacher and
Hartmannsgruber (1973)
Lamb (1977)
Derwent et al. (1978)
Singh et al. (1980)
Husain et al. (1977)
Derwent et al. (1978)
Dutkiewicz and Husain
(1979)
Haagenson et al. (1981)
Kelly et al. (1982)
Viezee et al. (1982)
Source:   Viezee et al.  (1983).

-------
air masses did  not reach the South Dakota site.  Kelly et al. (1982) found a
linear correlation, r, equal to 0,65,  for 1978  summer  measurements of  Be and
ozone in South  Dakota.   Ferman and Monson (1978) reported r = 0.60 at McKee,
Kentucky, for 27 daily samples taken during August and September 1976.   Johnson
and Viezee (1981)  reported an r value of about 0.50 for  Be and ozone data
from Dodge City, Kansas,  during April and May 1978.   A much lower correlation,
r = 0.15, was reported by Husain et al. (1977) for July and August 1975 measure-
ments at Whiteface Mountain, New York.  These data  imply  that the variability
of the surface  ozone  concentration that can be explained by  Be varies  from
about 40 percent to less than 5 percent.
     Singh et al.  (1980) and Viezee and Singh (1982) have pointed out a number
of problems  with this  technique in their detailed analyses of the application
of  Be measurements to the quantification of the amount of stratospheric ozone
in surface air.  Some  of the problems  encountered when applying  Be/0,, ratios
over short sampling periods (in contrast, for  example, to seasonal averages)
include the  following:

     1.   Because  7Be  is an aerosol and 03 is  a gas, they have fundamentally
          different atmospheric scavenging mechanisms and thus respond differ-
          ently  to tropospheric meteorological  conditions.
     2.   Although  the stratosphere is its dominant source,  on  the  average,
          7Be  is synthesized in both  the troposphere  and the stratosphere.
          The 7Be  source cannot be assumed to  be equivalent  to the 03 source
          without  meteorological verification.
     3.   7Be sampling data are primarily averaged over  24 hours, which does
          not give sufficient time  resolution for disaggregating short-period,
          rapidly  moving  stratospheric  intrusion  events from longer-term
          processes.
     4.   The lower stratosphere is a  highly variable  and poorly characterized
          region for which  no 7Be/03 ratio has  been firmly established.

As a result  of  these and  other  factors,  Singh et al.  (1980) concluded that
"the experimental  technique involving a  Be/0-  ratio  to  estimate the daily
                                              O
stratospheric component  of ground-level  0~  is  unverified  and considered  to be
inadequate for  air quality applications" (Singh et al.,  1980, p. 1009).  The
investigators at SRI-International have suggested, however,  that   Be  may be
used,  under  the  appropriate meteorological conditions,  as  a qualitative tracer
for  air  masses  of stratospheric origin  (Johnson and Viezee,  1981;  Viezee et
al., 1979).  ,
                                    3-83

-------
     Other methods have been used to estimate the quantitative contribution of
stratospheric ozone  to  ozone concentrations measured at ground-level.   These
include:  (1) aircraft  observations  of TF events (Johnson and Viezee,  1981),
coupled with  calculations  of the downward flux  of  ozone  in the troposphere
                                        90
(Viezee et a!., 1983);  (2)  use of mean   Sr  surface measurements, from radio-
active  fallout studies,  to calculate a mean stratospheric ozone contribution
(Reiter, 1978); and  (3) examination of data, especially  multiyear  data,  on
surface ozone concentrations  at  remote sites (Viezee and Singh, 1982;  Viezee
et a!., 1983).  Using surface measurement data and their own data from aircraft
measurements and  calculated ozone fluxes, Viezee et al. (1983)  concluded that
direct ground-level impacts by stratospheric ozone may be infrequent, occurring
<1 percent  of the time; that such  ground-level  events are short-lived and
episodic; and that they are most likely to be associated with ozone concentra-
tions <0.1 ppm.   Viezee et al.  (1983) recommended further study, however,  on
the possible contribution at ground-level of stratospheric ozone.
     On  a  qualitative basis, as mentioned earlier,  there is no doubt that
stratospheric ozone  is present in the  atmospheric  surface layers,  and the
meteorological mechanisms  that bring  this  about have  been  described  by  a
number  of  investigators,  including  Wolff et al.  (1979),  Johnson and Viezee
(1981), and others.  Most investigators cite the basic meteorological analyses
of Danielsen (1968) as a basis for their exchange model.
     The downward transfer of air parcels.and ozone from the stratosphere  into
the troposphere has been described above.  There is, of course, a compensating
transfer of  tropospheric air  upward .into the  lower stratosphere.   Reiter
(1975)  has examined  various mechanisms  that  contribute to this  transfer.  Air
parcels  moving  out of the  troposphere  will  carry with them the background
concentrations of  ozone that they had  in the  troposphere, and, as the air
parcels mix in the stratosphere, these ozone molecules will become part of the
stratospheric background  ozone.  Since  the ozone concentrations are  very much
lower in the  troposphere  compared to the stratosphere,  however,  this exchange
of tropospheric and  stratospheric air parcels will  not result  in a net upward
transport  of  ozone and is  not  considered  to be: a  factor.in air pollution
situations.                                        ,

3.4.8   Background Ozone from Photochemical  Reactions
     The apportionment of  the  background  ozone concentrations  measured at
remote  and other  nonurban  locations to stratospheric versus tropospheric
                                   3-84

-------
processes and sources  Is  the subject of continuing discussion and research.
In addition,  the apportionment  of ozone produced  photochemically  in the
troposphere to natural  versus  manmade sources of precursors, and,  in turn,
their respective  contributions  to  background levels of ozone, also remain  a
focus of discussion and research.  The preceding discussion has  emphasized the
thesis  that the  major source  of tropospheric  background ozone is  the
stratosphere.   It must  be recognized, however, that some investigators argue
for  a  much larger  role for the formation of  significant  amounts of  ozone
within  the  troposphere  (e.g., Fishman et a!.,  1980; Fishman and Seiler,  1983;
Fishman and Carney,  1984; Fishman et a!., 1985;  Fishman  and Crutzen, 1978;
Chameides and Walker,  1976).   The question of the potential relative impacts
of the  stratospheric  source of ozone and tropospheric photochemical  sources
has been evaluated critically by Singh et al. (1978, 1980).
     Singh  et al. (1980) concluded that background ozone in the troposphere is
"principally  of  stratospheric origin," and  supported  this position with a
number  of arguments,  including  the following.  First,  the NO  concentrations
                                                             X
in the  free troposphere appear to be very low, probably 0.01 to 0.1  ppb; at
NO   concentrations  <0.03  ppb a tropospheric ozone reservoir cannot be gener-
  X
ated,  using photochemical  simulation models.   Second,  Singh et al.  (1980)
pointed out that the tropospheric concentration pattern of  ozone,  with its
spring  maximum  and  fall  minimum,  is out of phase  and  inconsistent with a
photochemical source,  which would be at a maximum during the summertime peak
of solar  radiation  flux.   Also, at  low  latitudes,  where  the solar flux  is
relatively  constant,  the  ozone  seasonal  cycle is  quite  pronounced and is
consistent  with  the  seasonal  cycle  of stratospheric-tropospheric  exchange
processes.  In addition,  Singh  et  al. (1980) noted  that the  vertical  gradient
of ozone in the  troposphere  is contrary to a dominant photochemical  source  for
background  tropospheric ozone.                                ;
     Other  evidence,  however, especially more  recent modeling studies, indicates
that  a  substantial  part of  the  03 measured  during the warmer months of the
year over the United  States and Western  Europe is of photochemical  origin
(Altshuller,  1986).   Attempts to obtain agreement with observed 03 concentrations
with  a  general  circulation model  including  only a stratospheric source of  03
have been  unsuccessful, especially over  continents  (Levy  et al., 1985).   From
their  tropospheric  modeling studies, Fishman and  eoworkers have  predicted the
photochemical formation of near-surface, summer ozone at  concentrations of:
                                    3-85

-------
(1) 40  ppb,  using surface NO  emissions  corresponding  to  a  concentration  of
                             X
~0.25 ppb  (Fishman  and Seiler, 1983; Fishman  and  Carney,  1984);  (2) 15 ppb,
assuming zero  NO  emissions  (Fishman and  Seiler, 1983);  and  (3) 80 ppb, using
                X
NO  emissions  corresponding  to a concentration of  about 5 ppb and ignoring
  J\
contributions  from  individual  urban  or  industrial  plumes (Fishman et al.,
1985).   Using  another model,  Dignon and Hameed (1985)  have predicted the
photochemical  production  of  ozone at midlatitudes of the Northern Hemisphere
at annual  average concentrations  of:  (1) <60  ppb,  using a 1980 NO  emissions
                                          /v                       X
inventory; or  (2) ~25  ppb, assuming  zero  manmade NO   emissions.   On the basis
                                                    s\
of his review of  these and other studies, Altshuller  (1986) has estimated that
average stratospheric contributions to near-surface ozone concentrations range
from 10 to 15 ppb.
     As these data and the preceding discussion in Section 3.4.7  indicate, the
portion  of background ozone concentrations attributable  to stratospheric
sources  versus  tropospheric  photochemistry  remains uncertain, especially  in
the absence  of a  quantitative technique  for determining whether  ozone is  of
stratospheric origin.
     Investigations on the contributions of photochemistry to background ozone
have focused on (1) the  role  of  transport  from urban into  nonurban areas,
which was discussed in Section 3.4 (see also Chapter  5); and (2)  the respective
roles  of biogenic  VOC and  of  NO ,  from  all sources, in the photochemical
                                 s\
formation of ozone  in nonurban areas.
     Altshuller (1983) has evaluated the specific question of whether naturally
emitted VOC (i.e.,  from biogenic VOC sources) could be a significant source of
background ozone.   Since  biogenic emissions  are released into  the atmospheric
boundary  layer,  this potential source of ozone is expected to affect only
ground-level or  boundary-layer concentration  patterns, in contrast to the
whole troposphere as described in the preceding discussion.  While a matter of
controversy in recent years, the data on the role of  biogenic VOC in boundary-
layer photochemistry  now  appear more conclusive.   Inventories  of  biogenic  VOC
emissions indicate  that they are of  the same order of magnitude as manmade VOC
emissions, as  described  in  Section 3.5.1  (below).   On the  other hand, concen-
tration  data indicate  that  biogenic  VOC  occur,  even in  nonurban areas, at  low
levels  in  ambient air relative to  VOC  from manmade  sources (Section 3.5.2
below).  In  an extensive review of  the  literature,  Altshuller (1983)  noted
that the  concentrations  of biogenic hydrocarbons  are very low, constituting
                                   3-86

-------
even at rural sites probably 10 percent or less of the total nonmethane hydro-
carbons.  He concluded that the contribution of biogenic hydrocarbons to ozone
formation, as a result, does not appear to be significant under most atmospheric
conditions.
     The role of NO  in the photochemical formation of ozone in nonurban areas
                   X.
has been the subject of a number of modeling and experimental studies, as well
as a  recent  review (Altshuller,  1986).   As  summarized in Section  3.5  (below),
and as  reviewed by Altshuller  (1986),  NO concentrations in nonurban  areas  of
                                         /\
the United States-are appreciably lower  than those in urban areas, sometimes
by  an order  of magnitude.   Likewise,  concentrations  of NO  in the western
                                                           }\
United  States  are  usually lower by about a factor of ten  than those in the
northeastern  states (Section  3.5.2.2.2).   At nonurban  locations  inside  of
populated areas, however, the  concentrations of NO  are  much higher than those
                                                  f\:
measured  in  clean  air within  the continental and maritime  boundary layer and
in the  free troposphere (Altshuller, 1986).
     Global  background concentrations of NO  were  previously  thought to be
                                            /s.
lower than more recent measurements show  them to be.  For example, the analysis
cited earlier  in this section (Singh  et a!.,  1980),  on  the contributions of
photochemistry  versus  stratospheric  intrusions,  assumed  NO   concentrations  in
                                                          /x
the free troposphere of 0.01 to 0.1 ppb  and an NO  reservoir of <0.03 ppb;  and
                                                 f\
the global models  described  in Kelly et  al. (1984) assume an NO   background of
0.1 to  0.2 ppb.  In contrast,  the mean NO concentrations tabulated by Altshuller
                                          f\
(1986)  for remote  surface sites, while <1 ppb and often  <0.5 ppb,  are still in
excess  of the ranges described above.   At  nonurban sites  east of the Rocky
Mountains  in the United States, mean NO   concentrations  range from 1 to 10  ppb
(Altshuller, 1986).
      Martinez  and  Singh (1979) analyzed  the  role  of NO  in nonurban  ozone
formation  using theoretical  approaches and aerometric data from  the  Sulfate
Regional  Experiment.   They concluded  that the impact of NO   on  nonurban  ozone
is  a  function of geographical  location since ozone production was NO  -limited
at  some sites they examined but not at  others.  Kelly et al. (1984) examined
the role  of  photochemistry  in nonurban  ozone  at three  very different rural
sites (in South Dakota,  Louisiana,  and Virginia,  respectively)  by ambient air
analyses,  captive  air irradiations,  and  photochemical modeling.   Ambient air
.analyses  indicated the formation of about 6 ppb ozone per ppb of  NO .  Irradia-
                                                                   /v
tion  and  modeling  results  indicated similar, but slightly higher,  ozone forma-
tion  per ppb of initial  NO  .   They concluded that photochemistry in  these
                                   3-87

-------
rural areas was  not  NOX~"limited  but  depended,  as well,  on  hydrocarbon  concen-
trations.  In  a  study that included modeling and a comparison of the modeled
predictions with observed ozone concentrations, Liu et al.  (1984) arrived at a
similar  conclusion;  i.e., that  nonurban  ozone formation in the areas  they
studied  was  not exclusively controlled by  the equilibrium between NO   and
                                   •                                   x^
ozone (Equation 3-4, Section 3.3.1) but by the hydrocarbon reactions, as well.
Consistent with  the  results  of the Kelly et al. (1984) study, Parrish et al.
(1986) developed empirical relationships  between ozone  and NOV concentrations
                          :       -                             "
(June to  September)  at a high-altitude location (Niwot Ridge, Colorado) and
found the photochemical  formation of about 17 ppb ozone for each ppb of NO  .
                                                                           s\
     The studies and analyses cited in this and the preceding section indicate
that research  is still  needed  to permit the apportionment  of  background ozone
to stratospheric and respective tropospheric sources.  From his review of data
pertinent to  the role  of NO   in boundary-layer photochemistry, Altshuller
                            /\
(1986) concluded that  photochemically generated ozone should equal or exceed
ozone transported down from the stratosphere to relatively remote locations  at
lower elevations.  At more polluted rural  locations, photochemically generated
ozone from manmade  emissions  are predicted to  constitute  most  of the ozone
measured during  the  warmer months of  the  year.  Thus, the  evidence  supporting
the  stratosphere as  being the  major  (but  not exclusive)  source for  background
tropospheric ozone is relatively  strong but still  not conclusive.  Research  in
these topic areas is continuing,  and the number of uncertainties and necessary
assumptions needed for the respective conclusions cited above should be reduced
as new data become available.
3.5  PRECURSORS TO OZONE AND OTHER PHOTOCHEMICAL OXIDANTS
3.5.1  Sources and Emissions
3.5.1.1  Manmade Sources and Emissions.   This  section  presents  information on
the manmade sources and emissions of precursors to ozone and other photochemical
oxidants.  Estimates  of  annual  emissions  are presented by  source  category for
volatile organic compounds  and nitrogen oxides.  In addition, information is
presented on  emission rates and the composition  of  emissions from principal
stationary sources and from mobile sources.  The annual estimates are based on
emission inventories  prepared  according to procedures  developed and  published
by  EPA (U.S.  Environmental  Protection Agency,  1980a,b;  1981a,b,c,d,e,f).
                                   3-88

-------
3.5.1.1.1  Trends  inemissionsof volatile organic compounds.   Emission data
on volatile organic compounds  (VOC) include data on hydrocarbons as well-as
other organic compounds  found  in ambient air.  Because  of their negligible
reactivity toward  the  photochemical  production  of ozone, methane, ethane,
methylene chloride,  and several halogenated  methanes  and ethanes are  also
excluded  from  emission  inventories  (U.S.  Environmental Protection Agency,
1980b).    Estimates  of  total  emissions of volatile organic compounds  (VOCs)
provide a gross measure of compounds .available for photochemical production of
ozone and other  photochemical  oxidants.   Emissions of VOCs are reported here
as the  collective  mass of reactive VOC.  (See the footnote to  Table 3-8  for
the calculation of emissions-based NMOC/NO  ratios.)
                                          /v
     Retrospective  estimates  of total manmade VOC  emissions in the United
States, based on  records of economic activity (e.g.,  fuel  usage,  industrial
production) have been  prepared by-decade, beginning with 1940  (U.S., Environ-
mental  Protection  Agency,  1986).   From a level of 18.5 Tg/yr  in 1940, VOC
emissions increased about  14  percent each decade through 19704 then began to
decline (Table 3-8).
     Figure 3-14 shows national  trends in yearly emissions of  manmade VOC by
general source category for the period 1970 through 1983 (U.S. Environmental
Protection Agency,  1984a).   Emissions in 1983 of manmade  VOCs  in  the.  United
States  have  been  estimated  at 19.9  Tg/yr  (U.  S.  Environmental Protection
Agency, 1984a);  total  manmade  VOC emissions nationwide were 26 percent Jower
in 1983 than  in  1970.    The observed  decrease is  attributable  largely,  to  a
decrease of 30 percent  in estimated VOC emissions from highway  vehicles during
this  period.  This  decrease occurred in spite of a  42 percent increase in
vehicle miles traveled.   Trends  in  manmade  VOC emissions versus vehicle miles
traveled, urban  only  and total, for  1970 through  1983,.are shown in Figure
3-15  (U.S. Environmental Protection Agency, 1984a; Motor Vehicle Manufacturers
Association, 1984).  The main  sources nationwide of manmade VOC are industrial
processes, which  emit  a wide  variety of  VOCs such as  chemical  solvents;  and
transportation, which  includes the  emission of VOCs  in gasoline vapor  as  well
as in gasoline combustion products.            .
3.5.1.1.2   Trendsin emissions of nitrogen  oxides.   Emissions  of  nitrogen
oxides  (NO  )  are reported here as  the  sum. of NO and  NO,,,  all  expressed  as
          ^\                                              c.
equivalent N09.  Retrospective estimates of total manmade  NO  emissions in the
             c.                                              X,
                                   3-89

-------
           TABLE 3-8.  EMISSIONS OF VOC BY DECADE, 1940 THROUGH 1980
                                    (Tg/yr)a
Source category
Transportation
Stationary fuel combustion
Industrial processes
Solid waste
Miscellaneous
Total
1940
5.2
4.7
3.2
0.9
4.5
18.5
1950
7.9
3.1
5.2
1.0
3.6
20.8
1960
11.1
1.9
6.1
1.4
3.1
23.6
1970
12.3
1.0
8.7
1.8
3.3
27.1
1980
8.2
2.1
8.9
0.6
2.9
22.7
Source:  U.S. Environmental Protection Agency (1986).

aTo calculate emissions-based NMOC/NO  molar ratios, as used in control-
 related data, a multistep procedure ^s required (Novak, 1986) as described
 by the equation:

                             4
      Total NMOC (Tg/yr) x SUM[Pct. x no../mw-]/[Total NO  (Tg/yr)/46]
                            i=l   111           x

     where Pet = percent of a compound class present in the NMOC emissions;
           no. = average carbon number for a compound class;
            mw = average molecular weight for a compound class;
               i represents the four compound classes of interest; and
             NO  emissions are expressed as N02 (mw = 46).
               /\
Values for terms in the equation vary in the published literature.  Example
values, where no. and mw are derived from emissions inventories (Novak, 1986)
and pet from ambient air concentration data (Richter, 1986), are:  paraffins
= [63.5 x 4.56/95]; olefins = [15.5 x 3.57/43]; aromatics = [15.5 x 7.56/97];
and aldehydes = [2.0 x 1.5/37].


United States have been prepared by decade, beginning with 1940 (U.S. Environ-
mental  Protection  Agency, 1986).   From  a level of 6.8  Tg/yr  in 1940, NO
                                                                          r\
•emissions  increased  almost 40 percent each decade  through 1970  (Table 3-9).
The increase from 1970 to 1980 was 13 percent.
     Total NO   emissions  in the United States  in 1983 (19.4 Tg/yr) were some
             /\
17 percent above the 1970 level,  but  appear  to have declined slightly from
about  1980 through 1983 (Figure 3-16) (U.S. Environmental Protection Agency,

1984a).  The net increase over the period 1970 through  1983 may  be attributed

primarily  to two causes:   (1)  increased  fuel  combustion  in stationary sources

such  as power  plants;  and (2) increased  fuel  combustion in highway motor
vehicles,  as the result of the increase  (42 percent) in vehicle miles driven
over the 14 years  in question.  Emissions  associated with  industrial  processes

                                   3-90

-------
   30
   25
   20
05


O
   15
ui
>
=>
O
   10
                       i      i       r
                  TRANSPORTATION
                  INDUSTRIAL PROCESSES, STATIONARY SOURCES
               <' SOLID WASTE
                  NON-INDUSTRIAL SOLVENTS
               it [MISCELLANEOUS
         1971   1972   1373  1974   1975   1976  1977   1978   1979  1980   1981   1982  1983

                                         YEAR


Figure 3-14. National trend in estimated emissions of volatile organic compounds, 1970 through 1983.


Source: U.S. Environmental Protection Agency (1984a).
                                       3-91

-------
  160
  150
  140
  130
2120
Q.
GO"
HI
D
< 110
  100
o
oc
u.
LU
1  go
o
   80
   70
   60
   50
           1      T
 .VEH. MILES (URBAN)

 •VEH. MILES (TOTAL)
 •NOX
                	voc
                          \
                            s
            I	I
              J	I
                                                   J	I
          1971
1973
1975
   1977

YEAR
1979
1981
1983
    Figure 3-15. Comparative trends in highway vehicle emissions of nitrogen oxides (NOx) and
    volatile organic compounds (VOC) versus vehicle miles traveled, 1970-1983.


    Source: Motor Vehicle Manufacturers Association (1984); U.S. Environmental Protection
    Agency (1984a).
                                          3-92

-------
           TABLE 3-9.   EMISSIONS OF NO  BY DECADE,  1940 THROUGH 1980
                                    (Tff/yr)
Source category
Transportation
Stationary fuel combustion
Industrial processes
Solid waste
Miscellaneous
Total
1940
2.2
3.4
0.2
0.1
0.9
6.8
1950
3.5
4.7
0.3
0.2
0.6
9.3
I960
4.9
6.7
0.5
0.3
0.4
12.8
1970
7.6
9.1
0.7
0.4
0.3
18.1
1980
9.2
10.2
0.7
0.1
0.2
20.4
Source:   U.S.  Environmental Protection Agency (1986).

remained relatively  constant,  but  solid  waste and miscellaneous emissions
decreased slightly.
     The national  trends shown  do  not reflect the considerable  local  and
regional differences that  exist  in  the relative amounts of NO  emitted  in the
                                                             s\
major source categories.   For example, motor vehicle emissions in Los Angeles
County,  California,  increased sixfold from 1940 to 1970 (Los Angeles County,
1971), compared to  a threefold  national   increase.  Figure 3-15 compares the
relative trends in mobile  source NO  emissions versus  vehicle miles  traveled,
                                    f\
urban only and  total,  from 1970 through  1983  (U.S. Environmental Protection
Agency,  1984a; Motor Vehicle Manufacturers Association, 1984).
3.5.1.1.3   Sources  and emissions of volatile organic compounds.   The source
category contributing  the  largest  percentage of VOC emissions in 1983, 37.7
percent, is Industrial Processes (U.S. Environmental Protection Agency,  1984a).
The category  consists  almost entirely of point sources.  The composition of
these emissions varies widely, depending on the process or product and the use
of emission reduction equipment and operating practices.
     The second largest VOC source category, Transportation, accounting for
36.2 percent of the annual total in 1983, is discussed below.
     The third  largest VOC source  category, Miscellaneous, accounts for 12.6
percent of  the  annual  total, over  half of which consists of the subcategory,
Miscellaneous Organic  Solvents  (U.S.  Environmental  Protection Agency, 1984a).
These emissions generally qualify  as area-source  emissions.   Some  of these
solvents are  widely used in domestic products such as furniture polish, shoe
polish, shaving soap,  perfumes, cosmetics, shampoo, hair spray, hand lotion,

                                    3-93

-------
   30
   25
   20
I

W-

O
55
   15
UJ



I


i
=3
O
   10
                   TRANSPORTATION
                   FUEL COMBUSTION (STATIONARY)
                   INDUSTRIAL PROCESSES, SOLID WASTE. MISCELLANEOUS      I

                  	I	I	I	i      i	I	I	I
   1970  1971   1972  1973  1974   1975  1976   1977  1978   1979  1980   1981   1982  1983


                                             YEAR



     Figure 3-16. National trend in estimated emissions of nitrogen oxides, 1970 through 1983.



     Source: U.S. Environmental Protection Agency (1984a).
                                      3-94

-------
rubbing alcohol, and  nail  polish remover.   The predominant compounds emitted
are isopropyl alcohol  and ethyl alcohol (Bucon et a!., 1978).
     Emissions of volatile organic compounds from the production and marketing
of gasolines and motor  oils are classed in the Industrial Processes category
of VOC emissions.  Emissions of  VOC from these petroleum  products after their
sale to vehicle owners are included in the Transportation category.
     A significant portion of VOC emissions from gasoline-fueled mobile sources
arises from  evaporation, but the most  conspicuous mobile  source emissions are
the combustion  products.   In a  study  by   Black  et al. (1980), evaporative
emissions were  found  to constitute one-third to one-half  of total hydrocarbon
emissions from  all  of  the vehicles tested  (n = 60).   Based upon  actual
surveillance tests of in-use vehicles, Fisher (1980) found emission rates of
2.53 g/mi in 1968 and 0.15 g/mi  in 1980 for composite evaporative and crankcase
emissions from gasoline-fueled cars.
     Exhaust emissions  from gasoline-fueled vehicles typically contain fuel
components and  low-molecular-weight  hydrocarbons  that are not present in the
fuel.   Typically, exhaust  from a catalyst-equipped automobile contains about
62 percent alkanes,  17  percent  aromatics, 18  percent alkenes,  and  3 percent
acetylene.   This may  be compared with the  corresponding  typical  values for
automobiles  without  converters:   40,  24, 26,  and  11 percent, respectively.
Methane levels  generally  range from about  10 to 30  percent (Black and  Bradow,
1975; Black,  1977).   Exhaust gases from  gasoline-fueled  vehicles also  contain
organic compounds  such as  aldehydes,  ketones, ethers,  esters, acids, and
phenols, amounting to as much as one-tenth of the total VOC content.
     Factors  other  than gasoline  composition  influence  the  composition of
exhaust from internal combustion engines.  These include  driving patterns, the
specific configuration  of  emission control devices,  ambient  temperature and
humidity, and,  of course,  individual  automobile parameters  such  as tuning,
make,  and  model year.   Fuel additives can also  influence emissions.   For
example,  in  one study,  tetraethyl  lead increased  hydrocarbon  emissions by
about 5 percent but did not change the type of emissions  (Leihkanen and Beckman,
1971).
     Evaporative emissions  from diesel vehicles are negligible because of  low
fuel volatility (Linnell  and Scott,  1962;  McKee  et  al.,  1962).  In  studies of
exhaust emissions from diesel automobiles (Black and  High, 1979; Gibbs et al.,
1983), 15 to 40 percent of  the hydrocarbons emitted were  found to be associated
                                   3-95

-------
with particles  by  the time the exhaust stream exited the tailpipe.  Gibbs et
al. (1983)  reported  THC emissions from 19  in-use  diesel automobiles,  repre-
senting 1977 to 1979 model years, that were tested periodically over a 28-month
period.  Emissions of THC at the end  of  the  period  ranged  from 0.17 to 0.88
g/mi for individual  vehicles  and  averaged 0.65 g/mi.   It should be noted  here
that the population of diesel-powered passenger cars is not growing as rapidly
as expected.  Sales of diesel-powered cars peaked at 6 percent for 1981 models
and dropped to less than 3 percent by 1983 (Automotive News, 1982a,b,; 1983a,b;
Plegue, 1983).
     A brief  summary of both  exhaust and  evaporative emissions characteristic
of a variety of engines, fuel types, and control  devices has been presented by
filton and  Bruce  (1981).   Their summary indicates, for example,  that exhaust
emissions of  aldehydes and ethylene increase in cars  fueled with  ethanol-
gasoline blends (gasohol)  as  opposed to gasoline.  Exhaust  emissions of total
hydrocarbons  (THC)  decrease but  evaporative  emissions increase,  for a net
increase in emissions from gasohol-fueled cars.   In diesels, evaporative THC
emissions are virtually nonexistent, and net exhaust  emissions  of THC are
lower.  The percentage of  carbonyl  compounds (aldehydes,  ketones) in the
exhaust of  diesels  is  higher than  from  gasoline  combustion,  but  the net
photochemical reactivity  per  gram of emissions is lower nevertheless (Tilton
and Bruce,  1981).
     In view  of changes in emissions with  fuel  type,  it is of interest  to
examine fuel  usage  in the United States  in  recent years.   Yearly  sales  of
vehicle fuels fluctuate in consequence of several factors,  including  retail
fuel price, general  economic conditions,  and the  age  and fuel efficiency of
the composite vehicle population.   In  addition to the year-to-year influence
of these factors  on  the sales of the  principal  vehicle fuels, gasoline and
diesel fuel, a change in fuel composition is  emerging with the introduction in
the United  States  of gasolines containing a  percentage of ethanol.  As shown
in Table 3-10, sales  of gasoline containing up to 10 percent ethanol (gasohol)
were first  reported  in 1981 and  had increased almost  sixfold 2 years later
(NPN: National Petroleum News, 1981-1985 summary issues).
3.5.1.1.4   Sources and  emissions  of  nitrogen  oxides.   Fuel  combustion  is  the
dominant source of  NO  emissions nationally.  Stationary sources contributed
                      /\
50 percent  and  mobile sources contributed 45.4 percent in 1983 (U.S.  Environ-
mental Protection Agency,  1984a).  In  contrast to their contributions to VOC
                                   3-96

-------
         TABLE 3-10.   YEARLY QUANTITIES OF-MOTOR VEHICLE FUELS SOLD IN
             THE UNITED STATES FOR HIGHWAY USE, 1980 THROUGH 1983
Gallons x 109
Fuel
Gasoline
Gasohol
Diesel
1980
106.45
->-
13.60
1981
104.14
0.72
14.30
1982
100.70
2.16
14.67
1983
101,55
4.29
15.87
Source:   NPN: National Petroleum News (1981-1985 summary issues).

emissions, industrial process  sources contributed only about  3 percent to the
national total of manmade NO  emissions in 1983 (U.S. Environmental Protection
                            r\
Agency,  1984a).
     Ratios  of NO/NO  in emissions  vary  depending  upon the  source.  Nitric
                     "
oxide (NO) is  the dominant oxide of nitrogen  emitted by most sources;  NOp
generally comprises less than 10 percent of the total NO  emissions. More than
                                                        A
30 to 50  percent, however, of the  total  NO   emissions from  certain  diesel
                                           /\
(Braddock and  Bradow,  1976;  Springer and Stahman,  1977  a,b)  and jet  turbine
engines (Souza and  Daley,  1978) can be NO,,  under  specific load conditions.
Likewise, tail gas from nitric acid plants, if uncontrolled, may contain about
50 percent NO, (Gerstle and Peterson, 1966).  Variations in NO/NO  ratios by
              £.                                                   X
source  type  can be  significant in  local  situations,  as,  for  example,  in the
immediate vicinity  of a high-volume roadway carrying a significant number of
diesel-powered vehicles.
     Emissions  of NO   from the principal categories of stationary combustion
                     
-------
in the summer  than in the spring (U.S. Department of Energy, 1978).  Greater
degrees of variation  and different seasonal patterns have been  reported  for
stationary sources  in  different regions of the  country  (California Board of
Sanitation, 1966).
     Emissions of NO  from mobile sources, gasoline- and diesel-fueled vehicles,
                 *   if\
are affected by  a number of  variables  such  as  speed,  load, and air:fuel ratio
(APR), as  reported by  Billiard  and Wheeler  (1979).  Seasonal  variations  in NO
                                                                            J\
emissions from mobile  sources will  occur in relation to temperature (about a
35 percent decrease in emissions  per vehicle mile with an ambient temperature
increase from  20 to 90°F) (Ashby et a!., 1974), and  number  of vehicle  miles
traveled (about 18 percent higher in summer than in winter,  nationwide) (Federal
Highway Administration,  1978).  Emissions of NO  also vary with  vehicle miles
                                               ^\
traveled in urban versus rural  areas and among states in different  regions of
the country (Federal Highway Administration, 1978).   Diurnal  variations in NO
                                                                             -A.
emissions  associated  with motor  vehicle  traffic are  especially important
because of their potential  impact on ambient air quality.   Table 3-11 summa-
rizes data on NO  emissions from mobile sources.
           TABLE 3-11.  SUMMARY OF NOX EMISSIONS FROM MOBILE SOURCES
Vehicle type
Gasoline-fueled
passenger cars
Emissions, g/mi
0.41 to 1184
Comments
In-use; equipped
with three-way
Reference
Smith and
Black, 1980
Diesel passenger
 cars
Heavy-duty trucks
 Gasoline (leaded)
 Diesel
0.84 to 3.15
12.29 to 15.58
17.47 to 42.40
catalysts; four
test cycles used
In-use, 1977 to 1979
model years; tested
at ca. 35,000 mi
One test cycle
One test cycle;
several diesel fuels
Gibbs et al.,
1983
Dietzman et al.,
1980, 1981
3.5.1.2  Natural Sources and Emissions
3.5.1.2.1  Natural sources and emissions of volatile organic compounds.  This
section presents a  brief overview of the  nature  and quantity  of  hydrocarbon
emissions  from  biogenic sources.   For  detailed information,  the reader is
                                   3-98

-------
referred to several excellent review articles  (Altshuller,  1983; Bufalini and
Arnts, 1981; DimitHades, 1981).
     To date,  isoprene and the monoterpenes are the only biogenic hydrocarbons
identified  as  emissions from vegetation  (e.g.,  Sanadze and Dolidze, 1962;
Rasmussen, 1964, 1970;  Evans et al., 1982).  Other volatile organic  emissions
from  vegetation  have  been  documented,  but consist  of oxygenated organic
compounds, e.g., camphor and 1,8-cineole (see, e.g.,  Graedel, 1978).   Isoprene
and the monoterpenes are of interest to atmospheric scientists largely because
they are volatile enough to be released under normal  environmental conditions,
because they are abundant relative to other biogenic  VOCs, and  because  they
have  been  shown  to  be potential ozone precursors.   The commonly identified
monoterpenes are a-pinene,  p-pinene,  camphene, A3-carene,  limonene, myrcene,
and  p-phellandrene.   As a  general  rule,  coniferous  trees  emit primarily
monoterpenes,  and deciduous trees emit isoprene.
     Biogenic emission rates.   Biogenic  emission rates have been determined
almost exclusively  by techniques that  involve  enclosing the entire plant or a
portion of it,  such as  a branch of  a tree,  in  a bag  or chamber constructed of
light, transparent  material.   Isoprene and monoterpene emission  rates given
here  [as  ug (g dry biomass)    hr   ]  were measured  by the enclosure method.
Because biogenic  emission rates are species-  and  temperature-dependent, the
species tested and  the enclosure temperature are also  given.
      Isoprene  emission  rates reported by  Evans et  al.  (1982)  for  various
species of trees range  from 3 pg g   hr    for  spruce (Picea sitchensis,  28°C)
to  233  ug g"1 hr"1 for  willow  (Salix  babylonica,  30°C).    Reported  emission
                                —I  —1
rates for  oak  range from 9 |jg  g   hr    (Quercus virginiana, 30°C; Zimmerman,
1979) to 49 ug g"1  hr"1  (Quercus agrifolia, 30°C; Winer, 1983).   The reader is
referred  to the studies by Zimmerman  (1979),  Evans  et al.  (1982),  and Winer
(1983) for additional rate data and experimental details.
      Monoterpene  emission  rates are probably  best-documented for pine.   Rates
obtained by Winer (1983) ranged from 0.6 ug g    hr     (Pinus radiata and Pinus
halepensis, 30°C) to  2 jjg g   hr    (Pinus  canariensus,  30°C).  For Pinus taeda,
Arnts et al. (1978) reported a  rate of 4 ug g    hr     (at  30°C)  and  Knoppel et
al.  (1982) reported a rate of 1 ug  g   hr   (at  30°C).  Among the higher monoter-
                                                           _1   —i
pene  emission  rates  are those for Pinus clausa,  11  ug  g   hr   at 30°C
(Zimmerman, 1979)  and for Douglas  fir  (Pseudotsuga taxifolia),  15 ug g    hr
at  30°C  (Knoppel et al., 1982).  See  also the  studies by Rasmussen  (1972) and
by  Evans  et al.  (1982)  for additional  monoterpene emission rate  data.
                                    3-99

-------
     Besides temperature, biogenic emission rates are affected by other environ-
mental  factors.   Rasmussen  (1972) reported that emission  rates  varied with
species, plant maturity,  resin  gland  integrity, and  leaf temperature.  Dement
et al. (1975) found that the emission rate of monoterpenes from Sal via mellifera
(California Black  Sage)  is  dependent on the vapor pressures of the terpenes,
the humidity, and  the amount of  oil  present on the  surface of the  leaf; but
found that  the emission  rate is not directly dependent on the photosynthetic
activity or on the stomatal  openings of the plant.
     Tingey and  his  coworkers,  in extensive studies  on  factors effecting
isoprene emission  rates  in  live-oak  seedlings (Tingey et  al.,  1981), found
light intensity  to be a chief  determinant  of emissions,  which decreased to
near zero levels in the dark.  In  contrast  to isoprene, monoterpene emission
rates did not appear  to  be  influenced by  light intensity but were affected by
temperature (Tingey et al.,  1980); a  log-linear increase in emission  rates of
monoterpenes with temperature was observed in studies of slash pine.
     The validity of  emission rate data obtained by the bag enclosure technique
has been widely  discussed because of uncertainties associated with (1) isolating
the vegetation in  an  artificial environment; (2) possible damage to isolated
vegetation; (3)  representativeness of emission  rates  measured from just one
branch; and (4)  relationship of emission rates  to ambient air concentrations
of biogenic hydrocarbons.
     Attempts to validate the bag enclosure method have focused on comparing
enclosure emission estimates with those obtained by alternate procedures, such
as micrometeorological gradient procedures  (e.g.,  Lamb et al.,  1983;  Knoerr
and Howry,  1981).   In a study  of Pennsylvania hardwood forest,  the gradient
                                            -2   -1
profile procedure  gave a flux of 8,000  [jg m  hr   ,  while  the  enclosure tech-
                         -2    -1
nique yielded 7,300 |jg m   hr   (Lamb et  al., 1983).   Good agreement  has been
reported, also,  for orpinene emission fluxes measured by a micrometeorological
procedure and by the  enclosure  method (Knoerr and Mowry, 1981).
     Although the  micrometeorological  approach  yields mass fluxes similar to
the enclosure  method, it,  too, has  certain limitations.   For example, the
measurement of small  vertical gradients above a forest canopy and the applica-
tion of surface  layer theory to non-ideal sites can lead to erroneous results.
Such  difficulties  can largely  be  avoided by simulating the forest  emissions
with an inert tracer  release  (such as SFg) and measuring ambient concentrations
of the  tracer and  biogenic  gases along  downwind sample lines  (e.g., Arnts and
                                   3-100

-------
Meeks, 1981). Isoprene fluxes obtained using the tracer procedure in a central
Washington oak grove  compared well  with flux estimates determined simultane-
ously with the enclosure technique (Allwine et a!., 1983).
     Biogenic emission Inventories.   Development of a biogenic emission inven-

tory  requires  knowledge  of  (1)  emission  rates for  individual  species;  (2) the
vegetation coverage of an area, by species; (3) leaf biomass per tree (derived
through allometric  equations);  and  (4) a biomass factor  for the  forested area

(derived from (2) and (3).
     Table 3-12  contains a  listing  of area-wide biogenic emission  fluxes that
have  been  reported  for the United  States  and portions thereof.  Accuracy of

the estimates  reported  in Table 3-12 depends upon the size of the area for
which the inventory has  been prepared.   Many of the problems and uncertainties
encountered  in preparing inventories have  been discussed in detail in the
literature (Altshuller,  1983;  Zimmerman, 1979,  1980; Wells, 1981; Box, 1981;

Dimitriades, 1981).


               TABLE 3-12.  AREA-WIDE BIOGENIC EMISSION  FLUXES
      Location
 Emission
   flux,
jq  m~2hr"1      Comment
                      Reference
South Coast Air Basin,      <780
 California
Lake Tahoe, California      1950
Lake Tahoe, California      2438

San Francisco Bay Area,     1388
 California
San Francisco Bay Area,     2265
 California
San Francisco Bay Area,      777
 California
Tampa/St.  Petersburg,       2540
 Florida
Southeastern Virginia       8890
Pennsylvania                1660
Houston, Texas              1170
United  States               1712

United  States               1099
United  States                884
Entire basin
Forested area
 of basin
Daytime

Nighttime



Forested area only
                                Winer  (1983)

                                JSA, Inc.  (1978)
                                JSA, Inc.  (1978)

                                Sandberg  et al.
                                  (1978)
                                Hunsaker  (1981)

                                Hunsaker  (1981)

                                Zimmerman (1979)

                                Salop  et  al.  (1983)
                                Flyckt et al.  (1980)
                                Zimmerman (1980)
                                Marchesant et al.
                                  (1970)
                                Zimmerman (1977)
                                Zimmerman (1978)
                                    3-101

-------
     Considering all the  variables  that affect biogenic emissions and their
inventories,  it  is somewhat  surprising  that the area-wide emission fluxes
listed in Table 3-13 show no  more variation  than they do.  With the exception
of the Southeastern Virginia area, which is a forested region with high biomass
coverage, most of  the  values in Table 3-12  differ by less than a factor of
three.
3.5.1.2.2  Natural  sources and emissions of nitrogen oxides.   Natural  emissions
of nitrogen  oxides (NO )  originate from the oxidation of nitrogen  gas  by
                       J\
electrical discharge  in  the atmosphere, from the  ammonifi cation  of organic
nitrogen during biological  decomposition,  and from the oxidation of organic
nitrogen during forest fires.  Nitrogen  fixation and electrical discharge are
normal processes of the nitrogen cycle  that convert  inert nitrogen gas to
biologically  useful  nitrate or ammonia.   For  a discussion of the nitrogen
cycle relative to NO  emissions, the reader is referred to Air Quality Criteria
                    
-------
           TABLE 3-13.  GLOBAL ESTIMATES OF NITROGEN TRANSFORMATION
                                   (Tg N/yr)
                                    Range of estimates
                         References'
Biological fixation
  (N2  •*  NH4 )
54 to 270
al. Delwiche (1970).
 2. Burns and Hardy  (1975).
 3. Soderlund and Svensson  (1976).
 4. Robinson and Robbins  (1975).
 5. Liu et al. (1977).
 6. Crutzen and Ehhalt  (1977).
 7. Noxon (1976).
 8. Sze and Rice (1976).
 9. Council for Agricultural  Science  and Technology (1976).
10. Chameides et al.  (1977).
1-5
Electrochemical fixation
lightning (N2 -> NO )
atmospheric (N2 -* N02)
Biological denitrification
(N03" -*• N2)
(N03~ -> N20)
combined
Industrial denitrification
(Organic -N •» NO )
(Other -»• N0x) x
Atmospheric denitrification
(NH3 -» NOX)
Natural NO emissions
from lana and sea
NHs emissions to atmosphere
from land and sea

10 to 40
14 to 20
96 to 190
20 to 340
83 to 270

14 to 19
30 to 36
3 to 30
40 to 210
110 to 850
1
2, 6, 7, 8, 10
2, 4
2, 3
2-4
5, 8, 9 -. '

2-4
2, 3, 5
2, 3
3, 4
2-4
                                    3-103

-------
are virtually nonexistent.  In addition, scaling of emissions from such sources
as bacterial  nitrification and denitrification  for  use  in preparing area-wide
emission inventories  is  not  possible.   Thus, the emissions  reported in this
section should  be  taken  as very gross  approximations that serve to identify
natural sources of NO  and to present  the estimated relative magnitude of
                      X
emissions from such sources.

3.5.2  Representative Concentrations of Ozone Precursorsin Ambient Air
     As discussed  earlier in  this chapter (Section  3.3), nonmethane organic
compounds (NMOC) and  the  oxides  of  nitrogen  (NO ) in the presence of sunlight
                                                X
react to form ozone and other photochemical oxidants. The reaction sequence is
complex, so  that dependable  precursor-oxidant  relationships are difficult to
establish.    Factors  such  as  absolute NMOC and  NO   concentrations,  relative
                                                 X
NMOC and NOV concentrations (NMOC/NOV ratios),  NMOC reactivity, and NOV compos!-
           X                        X                                 X
tion are known  to  affect the photochemical  reactions that produce ozone and
other oxidants  in  ambient atmospheres.   Concentration-based NMOC/NO   ratios
                                                                    X
are  used  in some  precursor-oxidant models.   Ratios  of NMOC/NO  that are
                                                                X
concentration-based are calculated  directly,  using  measured  concentrations of
NO   expressed as N0?  (ppm or ppb) and measured  concentrations  of hydrocarbons
  f\                **»
expressed as  carbon  (ppm  or  ppb C).  The latter is easily obtained from gas-
chromatographic  measurements, since the chromatograph yields a known response
per  carbon  atom.   This section  provides summaries of NMOC and NO  concentra-
                                                                 A
tions recorded  at  various urban and nonurban locations in the United States.
In  addition,  HC/NO   (or  NMOC/NO )  ratios are  given for  some  urban  areas.
                   X-             X
3.5.2.1   Concentrations  of Nonmethane OrganicCompoundsin  AmbientAir.   The
NMOC data  in this section are  segregated into  (1)  nonmethane  hydrocarbons
(NMHC) and  (2)  oxygenated hydrocarbons.   The concentrations  reported here for
NMOC were obtained by gas chromatographic methods for the identification and
quantification  of individual  NMOC species (see  Chapter 4).  A fairly substantial
data base exists for characterizing urban nonmethane hydrocarbon concentrations.
Measurements  of nonurban hydrocarbon levels, as well  as both  nonurban and
urban oxygenated hydrocarbons, are much more  limited.  Among oxygenated hydro-
carbons, aldehydes have received the most attention.  Insufficient information
exists for establishing ambient air concentrations of other classes of oxygenated
hydrocarbons such as alcohols, ketones, acids,  and ethers.
                                   3-104

-------
3.5.2.1.1  Urban nonmethane hydrocarbon concentrations.  Most of  the data on
ambient air concentrations of nonmethane hydrocarbons (NMHC) have been obtained
during the 6:00  to  9:00 a.m.  time period.   Since urban hydrocarbon emissions
peak during  that period of the day  and atmospheric dispersion is  limited,
these  concentrations  generally reflect maximum  diurnal  levels.   Table 3-14
lists the mean and  range  of NMHC concentrations  recorded in a number of urban
areas  throughout the  United States.   For most  urban areas included in the
table,  a mean NMHC  value between 400 and 900 ppb C was observed, though mean
concentrations in some  cities  (e.g., Houston, Las Vegas, and Los Angeles) are
in excess of 1000 ppb C.   The data in Table 3-14 are not meant to serve as a
comparison of NMHC  levels in various cities but rather are shown to indicate
the mean and range of concentrations that have been reported.   Comparisons are
invalid because  of  major  differences in sample  numbers,  site classifications,
and seasonal sampling periods.  In many cases, the range of values reported in
Table 3-14 might not  reflect  the true maximum and minimum  concentrations  that
occur  in a  particular urban area.   Most of the hydrocarbon sampling programs
were of  short duration (~1 month) and  in  some  cases were not operated on a
daily  basis.  For example,  the relatively high  mean values reported in Las
Vegas  are undoubtedly the result of the fact  that  ambient air samples were
only analyzed for hydrocarbons on days when conditions  were appropriate  for
oxidant formation.  It is probably safe to assume, however, that NMHC levels
during the  6:00  to  9:00 a.m.  time period  in  major  urban areas will usually
exceed 50 ppb C  but seldom surpass 10,000 ppb C.
     Species  in  the C2-C-,g molecular-weight range  dominate the hydrocarbon
composition  of urban  atmospheres,  with the alkanes generally constituting 50
to  60  percent  of the  hydrocarbon  burden, aromatics  20  to 30 percent, and
alkenes  and acetylene making up the remaining  5 to 15 percent (Sexton and
Westberg, 1984).  The alkane fraction  is usually dominated by species in the
C.-Cfi molecular-weight  range.  Predominant aromatics include benzene, toluene,
ethyl benzene, and the xylenes.   The most abundant  alkenes are ethylene and
propene.  The studies cited in Table 3-14 provide information on the individual
species of NMHC  found  in  urban atmospheres.
3.5.2.1.2  Nonurban nonmethane hydrocarbon concentrations.  Nonurban nonmethane
hydrocarbon  concentrations  are generally one  to two orders of  magnitude  lower
than those  measured in urban areas (Ferman,  1981;  Sexton and Westberg,  1984).
Concentrations of individual  species seldom exceed 10 ppb  C.  Total hydrocarbon
                                   3-105

-------
              TABLE 3-14.   NONMETHANE HYDROCARBON CONCENTRATIONS
      MEASURED BETWEEN 6:00 and 9:00 a.m.  IN VARIOUS UNITED STATES CITIES
City
(Date)
Atlanta (1981)
Baltimore (1980)
Boston (1980)
Cincinnati (1981)
Detroit (1981) •
Houston (1976)
Houston (1978)
Las Vegas (1980)
Las Vegas (1983)
Los Angeles (1968)
Los Angeles (1982)
Milwaukee (1981)
Newark (1980)
New York (1969)
Philadelphia (1979)
St. Louis (1973)
Tulsa
Washington, DC (1980)
Mean
NMHC
concn. ,
ppb C
491
659
569
840
330
1414
1679
2506
2762
3388
2920
324
732
830
669
817
426
671
Range
113 to 1677
51 to 2798
83 to 4750
260 to 1870
60 to 1720
356 to 16,350
400 to 4500
689 to 4515
1835 to 4590
N.A.a
390 to 6430
24 to 3116
89 to 6946
N.A.
305 to 1710
N.A.
103 to 3684
210 to 2953
Reference
Westberg and Lamb (1983)
Sexton and Westberg (1984)
Sexton and Westberg (1984)
Holdren et al. (1982)
Kelly et al. (1986)
Sexton and Westberg (1984)
Lonneman (1979)
Nay lor et al. (1981)
Naylor et al. (1984)
Lonneman (1977)
Grosjean and Fung (1984)
Sexton and Westberg (1984)
Sexton and Westberg (1984)
Lonneman (1977)
Sexton and Westberg (1984)
Lonneman (1977)
Eaton et al. (1979)
Sexton and Westberg (1984)
 Data are not available.

concentrations range up to ~15Q ppb C, but usually fall  in the range of about
5 to  100 ppb C.   Alkanes  comprise the bulk  of species present,  with C«-Cr
compounds most abundant.   Ethylene  and propene are occasionally reported at
concentrations of 1 ppb C or less, and toluene is usually present at ~1 ppb C.
     Table 3-15 provides  a summary of  the range of hydrocarbon concentrations
measured at  various  nonurban  locations in the United  States.   Samples  were
carefully selected at  most  of the sites in order to guarantee their nonurban
character.  At the coastal and near-coastal  sites, only those samples collected
upwind of manmade  sources (onshore  advection) were included.  The nonmethane
                                   3-106

-------
              TABLE  3-15.   NONMETHANE  HYDROCARBON  CONCENTRATIONS
                       MEASURED  IN  NONURBAN ATMOSPHERES
Location
Belfast, ME
Benicia, CA
Miami, FL
Glascow, IL
Janesville, WI
Houston, TX
Robinson, IL
Smoky Mtns. , TN
Northern Idaho
Virginia
Atlanta (urban)
Whiteface Mtn. ,
NY
Elkton, MO
Eastern TX
North Carolina
Colorado
Species
analyzed
C2 - Cg
^2 ~ £5
C2 ~ C5
£-2 ~ £3.0
^2 ~ CIG
£•2 ~ ClO
^2 " CIQ
^2 ~ CIQ
Terpenes
Isoprene
Isoprene
Terpenes
Isoprene
ot-pinene
a-pinene
Terpenes
Concentration
range, ppb C
10
7
2
60
9
2
13
38
0.1
4
0
6
0
0.1
0.6
0
to 22
to 14
to 23
to 150
to 24
to 24
to 113
to 149
to 18
to 150
to 8
to 84
to 28
to 8
to 13
to 8
Reference
Sexton and Westberg (1984)
Sexton and Westberg (1984)
Sexton and Westberg (1984)
Chatfield and Rasmussen (1977)
Sexton and Westberg (1984)
Sexton and Westberg (1984)
Sexton and Westberg (1984)
Cronn (1982)
Holdren et al. (1979)
Ferman (1981)
Westberg and Lamb (1983)
Whitby and Coffey (1977)
Rasmussen et al. (1976)
Seila (1981)
Seila (1981)
Roberts et al. (1983)
hydrocarbon concentrations reported at coastal  sites (Belfast,  Benicia,  Miami,
and Houston) are  definitely  lower than those measured at most of the inland
sites.   It should be pointed out, however, that the numbers  of samples measured
for each  of the  nonurban locations listed in Table 3-15 is  small.   This,
coupled with the fact that only a limited range of hydrocarbons were monitored
                                                                           i
in some cases,  makes intersite comparisons tenuous at best.
     Ambient air concentrations  of naturally  emitted hydrocarbons  (e.g.,
isoprene, a-pinene,  p-pinene,  A-carene,  and limonene) are generally reported
only in  nonurban  hydrocarbon sampling programs.   Because they are present at
very low  concentrations,  natural  hydrocarbons  are extremely  difficult to
identify  unequivocally when  they mix with manmade  emissions in an urban area.
The one  exception  is  isoprene,  which  has been reported  in both  urban and
                                   3-107

-------
nonurban sampling  programs.   Monoterpene (ClnH.,g) concentrations  in  ambient
air  seldom  exceed 20 ppb C.   Average  concentrations of orpinene, the  most
commonly reported  monoterpene,  are  usually  below 10  ppb  C.   During the  summer
months, isoprene concentrations as high as 150 ppb C have been measured (Ferman,
1981), but maximum concentrations in the 30 to 40 ppb C range are more common.
Ambient concentrations of the naturally emitted  hydrocarbons are site-dependent,
with the highest concentrations observed in or immediately adjacent to forested
areas.  Concentrations  vary with  season, as well,  because  natural  hydrocarbon
emission fluxes  are  directly related  to the  amount of biomass present  and
increase with  temperature.  In a  recent  review article,  Altshuller (1983)  has
provided a more  detailed discussion of natural  hydrocarbons and their effect
on air quality.
3.5.2.2  Concentrations of Nitrogen Oxides  in Ambient Air.  Ambient air levels
of  nitrogen  oxides have been monitored  throughout the United States for  a
number of years.   Since nitrogen dioxide (NOp)  is the only oxide of nitrogen
for which an NAAQS has been promulgated, it has  received the greatest attention.
The  emphasis here  is on NO  measurements that can be  related  to  the  diurnal
                           Jr\,
photochemical processes that produce ozone.
3.5.2.2.1  Urban NO,, concentrations.   Concentrations of  NO ,  like hydrocarbon
            "" -•^===-:T-;I-- ^'J"——"--'-   J---1:::::::::: ^^,                         ^
concentrations, tend  to peak in urban areas  during the  early  morning period
when  atmospheric dispersion is limited and  automobile  traffic  is  dense.  Most
of the NO  is emitted as nitric oxide  (NO), but  the  NO is converted rapidly to
         /%
N0«  by ozone  and peroxy radicals produced  in atmospheric photochemical reac-
tions.  Since  ozone  levels  and photochemical  activity  vary diurnally  and from
day  to day,  the  relative concentrations of NO and N0? can fluctuate signifi-
cantly.  Generally,  urban NO  concentrations peak during  the  6:00  to 9:00 a.m.
period, followed by a rapid decrease caused by the photochemical conversion of
NO and NOg  and increased atmosphere mixing.  Nitric oxide levels remain low
during the  daytime and then  usually  build up again through the nighttime
hours.  Nitrogen  dioxide concentrations typically increase  during the mid-
morning hours  and then abate as the  afternoon  progresses.   Levels  of NOg
increase again following the late afternoon rush-hour period, often continuing
to increase during the  nighttime.
     The average  NO   concentration in  urban  areas of the United  States is
                    s\                                                        •
about  70 ppb,  with NO  and  N02  contributing  about  equally  (Logan, 1983).
Monitoring data for 1975 through 1980  showed that  peak 1-hr NOp concentrations
                                   3-108

-------
equalled or  exceeded 400 ppb  in  Los Angeles "and  several  other California
locations,  as well as at sites in Kentucky (Ashland) and Michigan (Port Huron).
Cities with one peak hourly concentration exceeding 270 ppb during those years
include Phoenix; St. Louis; New York City; Springfield (Illinois); Cincinnati;
Saginaw and  Southfield  (Michigan); and more than a  dozen sites  in California.
Reported hourly concentrations in excess of 140 ppb were quite common nationwide
during the years  between 1975  and 1980 (U.S. Environmental  Protection Agency,
1982a).
     Urban NO   concentrations during  the 6:00 to  9:00 a.m.  period are of
             f\
primary importance in terms of oxidant production.   Average NO  levels recorded
                                                              f\
in several urban  areas  during this morning period  are listed in  Table  3-16,
which shows mean 6:00 to 9:00 a.m.  NO  concentrations in the range of about 50
                                     ^\
to 150 ppb.     Concurrent  6:00 to 9:00 a.m. hydrocarbon  samples  were also
obtained in the studies reported in Table 3-16, and the hydrocarbon-NQ  ratios
                                                                      f\
in each of these urban areas are included.
         TABLE 3-16,  AVERAGE 6:00 to 9:00 a.m. NO  CONCENTRATIONS AND
                         HC/NO  RATIOS IN URBAN AREAS
City
Atlanta
Baltimore
Boston
Houston
Detroit
Linden, NJ
Los Angeles
Milwaukee
St. Louis
Tulsa
Washington, DC
Average NO ,
ppb x
57
85
63
125
67
59
147
66
77
46
94
Average
HC/NOV
f\
9
10
10
13
5
16
10
5
8
13
14
References
Westberg and Lamb (1983)
Richter (1983)
Ri enter (1983)
Westberg et al. (1978b)
Kelly et al. (1986)
Richter (1983)
U.S. Environmental
Protection Agency (1978a)
Westberg and Lamb (1983)
U. S. Environmental Protection
Agency (1978a)
Eaton et al. (1979)
Richter (1983)
                                   3-109

-------
     Hydrocarbon concentrations (ppb C) exceeded the NO  levels by a factor of
                                                       *\
5 to 16 during the 6:00 to 9:00 a.m. period.  Smog chamber experiments indicate
that significant quantities of ozone can be produced when HC/NO  ratios are in
                                                               /\
this range.   Indeed,  ozone production has  been  observed  in  the vicinity of
most of the cities referenced in Table 3-16.
3.5.2.2.2   Nonurban NO  concentrations.   Concentrations of  NO  in  "clean"
                      jrST"                                       /\
remote  environments are usually below 0.5  ppb  (Logan,  1983).   For example,
median concentrations measured on Niwot Ridge in Colorado are about 0.3 ppb in
the summer and 0.24 ppb in winter.  In exceptionally clean air, NO  concentra-
                                                                  }\
tions as  low  as  0.015 have been recorded (Bellinger et al.,  1982).  Slightly
higher NO   concentrations  have  been reported  at other  remote locations  in  the
         s\
western United States and Canada.  Kelly et  al.  (1982) deduced a mean NO
                                                                           A
concentration of about  1 ppb  from measurements  in  South Dakota.  At  the South
Dakota site,  nitric oxide  generally contributed less  than 20  percent of the
total NO .  Measurements of NO  during the 1970s at rural locations in Montana
        S\                     f\
(Decker et al., 1978) and Saskatchewan (McElroy and Kerr, 1977) yielded average
concentrations similar to those recorded in South Dakota.
     At a rural site in Louisiana,  Kelly et al.  (1984) found mean concentrations
of ~1 ppb NO and ~4 ppb NOg.  The same investigators observed mean concentrations
of ~0.7 ppb NO and ~1.6 ppb  N02  at a rural  site in Virginia  (Kelly et al.,
1984).
     In the  northeastern United States,  nonurban NO  concentrations  appear to
exceed those in the west by about a factor  of ten.   A median NO  concentration
                                                               jf\
of 6.6 ppb was derived from data collected  at nine rural sites utilized in the
Sulfate Regional Experiment (SURE)  program (Mueller and Hidy, 1983).  Median
concentrations at  the individual  stations,  which  extended eastward  from the
Ohio River Valley to the Atlantic Coast, varied from 2 to 11 ppb.  Measurements
at  nonurban sites  in Pennsylvania  and  Louisiana, during the  summer  of  1975
showed mean hourly NO  concentrations of 4.7  and 4.1 ppb, respectively (Decker
                     /\                                                   " .
et al., 1978).   Nitric  oxide composed approximately 40 percent  of the  total
NO  at these latter two nonurban sites.
  .A
3.6  SOURCE-RECEPTOR (OXIDANT-PRECURSOR) MODELS
     In  order  to apply knowledge of  the  atmospheric chemistry of ozone and
other photochemical  oxidants and their precursors during their dispersion  and
                                   3-110

-------
transport, models describing these phenomena have been developed in a variety of
forms over the past 15 years.   Most of these models relate the rates of precursor
emissions from mobile and stationary sources, or precursor atmospheric concen-
trations, to the resulting ambient concentrations of secondary pollutants that
impact receptors at downwind sites.  For this reason they have been described
as source-receptor or oxidant-precursor models.
     A wide variety  of source-receptor models exists, ranging in  complexity
from empirical  relationships based on air monitoring or smog chamber data to
complex  computer-based grid or trajectory  airshed models that may contain
detailed  emission  inventories, sophisticated dispersion  and transport sub-
models, and lengthy chemical reaction mechanisms.  Moving-box models represent
an approach of intermediate complexity.
     All presently available source-receptor models require a degree of simpli-
fying  assumptions  to  deal  with  practical  limitations  imposed  by existing
computer capabilities, time and cost constraints, or lack of knowledge concern-
ing  inputs  such as boundary conditions, emissions, wind  fields, or detailed
reaction  mechanisms.   The  reliability  and  applicability of any particular
model  therefore  depend upon its  specific limitations, data requirements, and
degree of validation  against  experimental data from ambient air measurements
or environmental chamber runs.
     A detailed discussion  of  the range of available source-receptor models,
and  their validation  and applications, is beyond the scope of this document.
Instead,  brief  conceptual  descriptions are provided of  the  major  classes  of
source-receptor  models.   It is important to  recognize that  such models are
continually undergoing evolution,  revision, and refinement, particularly  as
knowledge  of  atmospheric chemistry grows and, for example, as more sophisti-
cated  approaches to  dealing with  boundary conditions become available.  Even
under  the best  circumstances,  however, the  present  generation of source-
receptor  models should be viewed  as  being  most useful  for  investigating the
relative  effects on  air quality of particular  emission  sources or emission
control  strategies,  rather than  for predicting absolute concentrations  of
secondary pollutants resulting from specific  precursor emissions.
                                    3-111

-------
3.6.1  Definitions, Descriptions, and Use
     Current  air-quality,  or source-receptor,  models  can be classified as
either statistical or computational/dynamic.  Statistical models are generally
based on an analysis of historical air quality data.  An example of a statistical
model is the linear rollback concept.
     Computational, or  dynamic,  models attempt  to describe mathematically the
atmospheric chemical and physical processes Influencing air pollution formation
and  impacts.   Examples  of computational  models  include  trajectory and grid
airshed models.   The basis of these  models  is the solution of the  atmospheric
diffusion equation (Bird et al., 1960; Liu and Seinfeld, 1975).
     Two phenomenologically different approaches have been employed in dynamic
models with  respect  to  the coordinate systems  chosen.   A coordinate system
fixed with respect to the earth is termed Eulerian,  while in Lagrangian models
the reference frame moves with the air parcel whose behavior is being simulated.
     In the  following section  these and other  models are briefly  described.
3.6.1.1  Statistical Models.  Two  widely used models of this kind are linear
rollback and the Appendix J approach, both of which were employed by EPA prior
to the advent  of  more sophisticated  dynamic  modeling approaches.   The  concept
of linear  rollback is  based on the assumption that ambient concentrations of
air pollutants are directly proportional  to emissions;  and that a given reduc-
tion in emissions will  result in a proportional  decrease in the maximum ambient
concentrations of that pollutant.  In principle, linear rollback models should
be applied only  to inert primary pollutants and their  original use was for
unreactive pollutants such as  carbon monoxide  (Larsen,  1969).   Such models
have been  applied,  however,  in modified form to secondary pollutants such as
ozone (Barth, 1970).
     A prominent example of a statistical model  was the Appendix J relationship
developed  by EPA to relate  maximum  1-hour average ozone concentrations in
several United States cities to 6:00-to-9:00 a.m. average nonmethane hydrocarbon
concentrations in  those cities  (F.R., 1971).   This relationship was used to
calculate the  amount of NMHC control needed to achieve  the  Federal  standard
for photochemical oxidants.
     Two important  limitations  of  past statistical  methods were their  failure
to take  into account the transport  of primary  and  secondary pollutants  from
source areas  to  downwind receptor sites, and lack of recognition of the role
of oxides of nitrogen in the formation of ozone and other photochemical oxidants.
                                   3-112

-------
These and other weaknesses led EPA to abandon the use of statistical models in
state implementation plans (F.R., 1981).
3.6.1.2  Trajecto ry Mode1s.   Figure 3~17(a) contains a schematic representation
of the trajectory  model  approach, in which a  hypothetical  air parcel moves
through the  area  of interest along a path calculated from wind trajectories.
Thus, a moving-coordinate system (Lagrangian)  describes pollutant  transport
under the influence of local meteorological conditions.  Emissions are injected
into the  air parcel  and  undergo  vertical  mixing  and  chemical  transformations.
     The data requirements for trajectory models include:  (1) initial concen-
trations of  all  relevant pollutants arid  species;  (2)  rates  of emissions of
NMOC and NO  precursors into the parcel along its trajectory; (3) meteorological
           s\
characteristics such  as  wind speed and direction; and (4) solar  ultraviolet
radiation.   Various  trajectory models  exhibit a range  of sophistication and
complexity with  regard  to  such  elements  as  chemical  mechanisms (Atkinson
et al., 1982c),  emission inventories (Braverman  and  Layland,  1982),  treatment
of vertical mixing (Whitten and Hogo, 1978; Drivas, 1977; Meyers et al., 1979;
Lloyd et  al.,  1979;  Lurmann et al., 1979),  and trajectory determination  (U.S.
Environmental  Protection  Agency,  1980c; Whitten  and  Hogo, 1978).   Basic  limi-
tations of trajectory models include the amount and density of data required
for  precise  calculations of emissions  input,  chemical transformations, and
dilution; neglect  of horizontal  wind shear; neglect  of cell  volume changes
resulting from convergence and divergence of the wind  field; and uncertainties
in boundary  conditions,  including conditions aloft (Liu and Seinfeld, 1975).
Conversely,  moving-cell  models  provide a dynamic  description  of atmospheric
source-receptor  relationships  that is simpler and less expensive to derive
than that obtained from  fixed-cell models.
     The  simplest  form of trajectory model is the empirical  kinetic modeling
approach  (EKMA).   This modeling approach was developed from earlier efforts
(Dimitriades,  1972)  to use  smog chamber data to  develop  graphical relationships
between morning NMOC  and NO  levels and afternoon  ozone  maximum concentrations.
                           J\
Dodge (1977a,b) presented an approach in  which smog chamber data  (Dimitriades,
1970;  1972)  were used to test  and validate a photochemical  kinetics  model
(Durbin et  al.,  1975).   Dimitriades (1977) used  the resulting 03 isopleths to
define  a method  for obtaining  the relative  degree of precursor emissions
control  needed to achieve a given percentage reduction in  ozone.   In the EKMA
approach, which  has been extensively utilized (U.S. Environmental  Protection
                                    3-113

-------
                                                 SUNLIGHT GIVEN AS
                                                 FUNCTION OF TIME
                                                                 TIME-DEPENDENT MIXING AND
                                                                 REACTION COMPUTED FOR AIR
                                                                PARCEL UP TO MIXING HEIGHT, h
SPACE/TIME TRACK THROUGH SOURCE
  GRID; DERIVED FROM WIND DATA
        POLLUTANT INFLUXES AT ANY
    ELEVATION (INCLUDING THE GROUND)
  IMPOSED BY EMISSION SOURCE FUNCTIONS
                         (a) TRAJECTORY MODEL
                                                           TRANSPORT
                                                                          TOP OF MODE LING
                                       	1	,	REGION

                                        «•	1—»•  CHEMISTRY-ELEVATED EMISSIONS t—|	fc
                                       . TRANSPORT            *	 TRANSPORT TOP °£

                                            •^^^^         » TRANSPORT     I     LAYER
                                                CHEMISTRY-ELEVATED EMISSIONS "—1	"
                                       5 TRANSPORT            1	   TRANSPORT

                                                           • TRANSPORT     1
                                      3 *	1"~* CHEMISTRY-ELEVATED EMISSIONS ••—I	••
                                      S TRANSPORT	»_	TRANSPORT
                                                           j TRANSpoRT     I

                                               CHEMISTHY-ELEVATEO EMISSIONS ••—I	••
                                       TRANSPORT            ,          TRANSPORT
                                                          t
                                                    SURFACE T REMOVAL

                                                   	1 TO 10 km	
                                                                         GROUND SURFACE
 GRID SPECIFICATION
                                                    GRID CHEMISTRY AND TRANSPORT
                             (b) GRID MODEL
                                                 RISING MIXED
                                                    HEIGHT  ENTRAINMENTOF
                                                           POLLUTANTS ALOFT
                    WIND DIRECTION
                         ADVECTIVE
                           INFLOW
                                                                              ADVECTIVE
                                                                               OUTFLOW
                             (c) BOX MODEL
                 Figure 3-17. Schematics of the three types of dynamic models.
                                         3-114

-------
Agency, 1978a,b; 1980c), the Ozone  Isopleth Plotting Package (OZIPP)  (Whitten
and Hogo,  1978) is used to  generate ozone isopleths at  various  levels of
sophistication corresponding to "standard" EKMA, "city-specific"  EKMA,  or the
"simplified trajectory" model  (F.R.,  1979).   These models are designated as
Levels IV,  III,  and  II,  respectively, with Level  IV being the least  sophis-
ticated and Level  II  the most sophisticated.   Substantial documentation and
guidance concerning the  use  of OZIPP and a more flexible modified version of
the Program (OZIPM)  are available  (U.S.  Environmental  Protection Agency,
1981a, 1984b,c).
     The sensitivity of this method to variables such as the input hydrocarbon
composition and  the choice of chemical kinetics mechanisms has been reported
(Carter et  a!.,  1982;  Jeffries et a!., 1981;  Shafer and  Seinfeld, 1985) and
further refinements in the EKMA approach to accomodate these factors have been
made.   For  example, site-specific versions of EKMA allow the  user to select
particular  dilution rates,  emissions, and solar intensity applicable to the
city  or airshed of interest. Another  version allows  the  user to  employ an
alternative mechanism  for making the  EKMA calculations (U.S.  Environmental
Protection Agency, 1984b).
      Because  it is often recommended  by  EPA  for use in  determining  needed
precursor  reductions  and  is  in widespread use, EKMA  is  discussed here in
detail.  An example  of an EKMA diagram  is presented  in Figure 3-18,  which
shows  ozone isopleths  for  sites downwind  of  an  urban source area in  which
morning precursor  emissions  are high.  The isopleths  in  this  diagram depict
downwind, peak 1-hour  ozone  concentrations as an explicit function of initial
(i.e., morning)  concentrations of nonmethane  hydrocarbons (NMHC)  and  nitrogen
oxides (NO  );  and  as indirect  functions of (1) NMHC and NO  emissions  occurring
          A                                              P\
later in  the  day;  (2) specified meteorological  conditions; (3) reactivity of
the precursor mix; and (4) concentrations of  ozone and precursors transported
from  upwind areas  (U.S. Environmental  Protection Agency,  1977).   The  relation-
ships between ozone  and its precursors  that  are depicted in  Figure  3-18  are
based on  empirical data  and the application of a chemical  kinetics  model
(Dodge, 1977a; Whitten and Hogo, 1977)  that  has been adjusted by comparing
model  predictions  against  smog-chamber data obtained by irradiating automobile
exhaust (Dimitriades,  1972).   Alternatively,  EKMA  diagrams  can be constructed
using more recent chemical  mechanisms  that  have  been tested against  smog
chamber data  (Gipson,  1984; Whitten and  Gery, 1986).
                                    3-115

-------
0.28
          0.2
0.4      0.6      0.8      1.0     1.2      1.4     1.6
NONMETHANE HYDROCARBON CONCENTRATION, ppm
1.8
                                                               0.28
                                                                                 0,24
                                                                                 0.20
                                                                               - 0.16
                                                                               - 0.12
                                                                               - 0.08
                                                                                 0.04
2.0
            Figure 3-18 Example of EKMA diagram for high-oxidant urban area.
                                    3-116

-------
     For the EKMA  diagram  given here as Figure 3-18, several general, inter-
related features are  of interest.   First, the  isopleth  lines in the lower
right quadrant are  more or less parallel to  the  abscissa (the NMHC concen-
tration).   Second,  the  isopleth lines in the upper left quadrant are slanted
with respect to the ordinate (the NO  concentration).  Because of the shape of
                                    J\
the isopleth lines, varying the NMOC  or  NO  concentration will have different
                                          X
effects on ozone at different NMHC/NQ  ratios.
                                     yv   •           .              ,    .
     These two features of the diagram are related to the underlying photo-
chemistry (see, e.g., Whitten,  1983).  For extremely low NO   concentrations
                                                             A
(i.e.,  high  NMHC/NO  ratios),  where  the lines parallel  the abscissa,  the
                    J\.
formation of 0, is  (1)  insensitive  to changes in  NMHC concentrations, and  (2)
NO -limited; that  is,  changes  in  NO  concentrations  cause  co-directional
  X                                  X
changes in peak Og concentrations.   As described early in Section 3.3 and in
Section 3.3.1,  the atmospheric oxidation of hydrocarbons produces an abundance
of peroxy radicals  (RQp*), more than enough to oxidize NO to NQp rapidly !and
completely.   In  addition,  the  NO   present,  before being  removed from the
                                 A.
cyclic  reactions via  termination reactions with various  radicals, completes  a
               RO
number  of NO 	^—»  NQp —	»»  NO (+ GO) cycles, thus producing a number
                         £m       .          O
of Oq molecules.   Small-to-moderate changes in NMHC concentrations will there-
fore have little impact, since there will still be  an abundance of R02* radicals
in the  atmosphere.  On  the other hand,  changing the already  low  concentration
of NOX  does  not  have an appreciable  impact  on the  ROp  radicals,  but  it dres
change  co-directionally the number  of 0, molecules  produced,   since the photolysis
                                       •3      ,         •           . ~!
of NOp  and the oxidation of NO  to NO  are essential for 0^ production.
     For  moderately higher concentrations  of NO   (lower NMHC/NO   ratios),
                                                 X               X
there  is  no  longer a large abundance of ROp- relative to the NO present; and
changes in NMHC, therefore, have a  co-directional effect on RQ2* and, hence,on
Og.  Thus, for moderate NMHC/NO  ratios, the  effects on 03 formation of varying
the concentrations  of NMHC or NO  are similar in  direction.
                                 X
     Finally,  for  much  higher  NOV  concentrations  (i.e., very low  NMHC/NO
                                 X            .                             X
ratios),  the dominant effects  are the relative depletion of  radicals,  through
the  reactions  of  N02 with these radicals  (Equations 3-18 and 3-35, Section
3.3);  with  a resulting decrease in the  rate of  reaction of  peroxy radicals
with  NO (Equation 3-29, Section 3.3).   The  consequence of these effects  is
that  the  conversion of NO to NOp is  very slow.   At such  low  NMHC/NOX ratios,
therefore,  increasing  the NMHC  concentration enhances  Qg  formation but
increasing NO  concentrations  inhibits  it.
                                    3-117

-------
     As examination  of  Figure 3-18 reveals, for an NMHC concentration of 0.6
ppm C, for example, increasing NO  leads to increased 0~ until NMHC/NO  ratios
                                 X                     «3              X
of about  5:1 to 6:1 are reached;  further NO   increases,  leading to lower
                                             X
NMHC/NO   ratios,  inhibit 0- formation.  Thus,  in  this  example,  there is a
       X                   O
"critical" ratio (approximately  5,6:1) at which the NO  effect on 0« changes
                                                       X           * O
direction.  Besides  this  "critical"  ratio, an "equal control" NMHC/NO  ratio
                                                                      X
also exists,  above which the reduction of NO  is more beneficial in terms of
                                             X
0,, reduction  than  an equal  percentage  reduction  in  NMHC.   This ratio, for the
isopleths shown in Figure 3-18, is roughly 8:1 to 9:1 for low levels of control,
and as high as 20:1 for the levels of control needed to reduce 03 to 0.12 ppm.
Thus, for this particular case (Figure 3-18), the chemical mechanism modeling
evidence  suggests  that  (1)  NO  control will  increase the peak downwind 0.,
                              X                                           O
concentration at NMHC/NO   ratios  of 5.6:1 or lower; (2) both NO  control and
                        X                                      X
NMHC control  will  be beneficial  at somewhat  higher ratios,  with control of
NMHC  being  more effective;  and  (3)  for ratios  above 20:1,  NO   control  is
                                                              X
relatively more effective in reducing 03<
     The calculation of precursor controls necessary to reduce 03 to 0,12 ppm,
from the  isopleths given in Figure 3-18, shows that NO  control, although at
                                                       X
first beneficial,  is ultimately detrimental  because it  makes  the  reduction of
03 to 0.12  ppm  more difficult;  This  can  be demonstrated through use of the
EKMA isopleths in  Figure  3-18 as follows.   For a high-oxidant atmosphere with,
for example,  a  peak 0,, concentration  of 0.30 ppm and an  NMHC/NO   ratio of
                      O                                           X
12:1, a 74  percent control  of NMHC (1.84 ppm C to 0.48 ppm C) will be needed
in order to reduce downwind, 1-hour peak 03 to 0.12 ppm through the unilateral
control of NMHC (line AB  in Figure 3-18).   If, however, NO  is first controlled
                                   •                       /\   *•"
by, for example, 29 percent (from 0.152 ppm to 0.108 ppm, line AA1), this will
cause a 15  percent reduction in 03  (line  AA1)  but  it will also  increase the
NMHC control  requirement  (to reduce  03 to  0.12  ppm) from  74  percent (line AB)
to 81 percent (line A'B1).   Since it  is not ordinarily feasible to reduce 0,
to 0.12 ppm  in  a  high-oxidant area through the unilateral control of NO  (it
                                                                        X
would take,  in  this case, an almost 85 percent control of NO ,  line AC), it
                                                              X
follows that  29  percent control  (or any control) of NO , would ultimately be
                                                       X
detrimental for  the situation represented by these isopleths.   There may be
situations, however,  in which the control of NO will  not increase the  NMHC
                                                 J\
control required to achieve a given percentage  reduction in ozone.   It  must
be emphasized that the preceding discussion of the  implications of controlling
VOC and NO   is  not  applicable to  all  situations.   These  diagrams may differ
   »
                                   3-118

-------
when different chemical mechanisms or other model input data are used.  Further-
more, conclusions  drawn  from Figure 3-18 may  differ  if a  different  starting
point on the diagram is used.
     As noted early in Section 3.3, the effects on 0- formation of controlling
NO  emissions are a matter of continuing discussion and research.  The inhibition
  
-------
models (Roth et al., 1976).  In addition to atmospheric data, various emissions
sources must be  considered in each model.  At  the most sophisticated  level,
emission models  and inventories  are  constructed to  provide estimates,  by
category,  of  vehicular, aircraft,  power plant,  refinery,  and distributed
source emission  rates,  including  their  temporal and seasonal  variations  (Roth
et al., 1976; Braverman and Layland, 1982; F.R., 1979).
     Shortcomings in grid  models  stem from the theoretical and computational
complexities that are  necessary in this type of simulation.  Inaccuracies in
grid-model  predictions arise from:  (1) theoretical  deficiencies in the mathe-
matical representation  of  atmospheric processes;  (2) numerical inaccuracy in
the solution of  the atmospheric diffusion equation (Liu and Seinfeld, 1975);
and (3) inadequate  input  data  resulting from  incomplete  data bases.   An
incomplete understanding of  advection and turbulent diffusion, necessitates
the use  of estimates  or parameter!'zations to  provide appropriate  values
(Seinfeld  and Wilson,  1977).   Atmospheric chemical  kinetics descriptions are
continually updated as  new information  is obtained, but uncertainties  associ-
ated with  these  mechanisms may be propagated during  solution.  In addition,
sparse and often unrepresentative  data are utilized  to  derive continuous
fields (wind fields, turbulence,  and  mixing depths) over the region  (Seinfeld
and Wilson, 1977), a problem that is common to all dynamic models.   In general,
wind and turbulence data are rarely collected  aloft;  surface data are much
more abundant but still vary widely in terms of number, frequency,  and quality
of measurements  (Roth  et  al.,  1976).   This implies  that  critical  values,
especially aloft, must  often be estimated to provide  initial,  boundary,  and
operating  conditions  (Seinfeld and Wilson, 1977).   Finally, uncertainty in
grid-model  solutions  also  arises  from  emission inventories that  are poorly
resolved,  either spatially,  temporally, or  with respect  to hydrocarbon
reactivity specifications  (Braverman and Layland,  1982).
     The structure and complexity of grid models also account for their utility.
The increased temporal  resolution afforded by grid models  can provide  minute-
by-minute  concentration estimates  in  each  cell,  and there  can be  as much
spatial resolution  as  the  data will allow.  Since  it accounts for specific
atmospheric processes within  the  system, the model  allows explicit insertion
of new information  (e.g.,  on meteorological or chemical processes)  into  the
structure  of  the model.   In addition,  the impact  of individual  precursor
sources may  be  analyzed with  this type of model (Association of  Bay Area
Governments, 1979a,b).
                                   3-120

-------
3.6.1.4  Box Models.   Box models (Hanna, 1973; Demerjian  and Schere,  1979;
Derwent and HOv,  1980) are the simplest of  dynamic  models.   They treat the
atmosphere as a single cell, bounded by  the  mixing layer,  with an area  on the
order of 100 square miles  [see Figure.3-15(c)].  The chemistry within the box
is affected by:   (1)  instantaneously  mixed regional  emissions,  (2), dilution
from lifting of the inversion, (3)  ventilation and transport  resulting  from a
characteristic wind field, and (4) entrainment of species from aloft.   Because
the only consideration of spatial resolution occurs when the modeling boundaries
are chosen, data  requirements  are minimal.  Results can only be interpreted
temporally, however,  for a mass  average of a species, and  results  can be
strongly affected by uncertainties in boundary conditions.

3.6.2  Validation andSensitivity Analyses for Dynamic Models
     Dynamic models are mathematical representations of atmospheric processes.
They are based on many assumptions, however, and can only be considered approxi-
mations  of real  processes.  Therefore,  it is  important to .investigate  the
extent to which model  predictions disagree with actual measurements,   Deviations
occur for  two basic reasons:   (1)  a completely valid mathematical description
of natural  systems  does not presently exist; and (2) input data and data for
comparison with predictions  are  often unresolved and imprecise  (Seinfeld and
Wilson, 1977).  It is  therefore difficult to determine numerically the overall
accuracy of model  calculations.   Rather, attempts are made to validate model
predictions by comparing them  with real  observations and operating parameters
are often  varied  to determine  the  sensitivity  of  the model (Gelinas and Vajk,
1979).   In addition,   the extent of agreement  between the results from two
simulations can be  tested.   "In this way,  completely different models may  be
compared,  or  an  internal component, such as the chemical  kinetics mechanism,
may be substituted and a model re-run to .ascertain the effects of such  substi-
tutions.                                       .
     It  should be noted that the  validity of  all  dynamic .models depends,  in
part, on  the  quality  of the chemical  kinetics mechanisms  used to define the
Og-HC-NO   relationship.   These mechanisms have the advantage of being  cause-
and-effect  descriptions derived  from actual experimental  data.  The data are
subject,  however,  to  the effects  of  smog  chamber artifacts  (Carter et al.,
1982), which  may  or may not occur in  the atmosphere.   Also,  there remain sub-
stantial uncertainties in the detailed chemistry of certain classes of  organics
                                   3-121

-------
such as the  aromatics  (Section 3.3),   In addition, only  recently  have data
become available  so that mechanisms can  be tested  against data bases in which
the hydrocarbon composition has been systematically varied or in which dynamic
dilution and  injection of  new  reactants  has occurred  (Jeffries et a!,, 1981).
Therefore, dynamic models using existing chemical mechanisms may not accurately
describe all  of the conditions that apply in the atmosphere  (Jeffries  et al.,
1981).
     Evaluations  of complex dynamic models have  been  of two  forms:  numerical
sensitivity  analyses  and simulation  performance studies  using ambient air
data.  Sensitivity analyses have considered the effects of varying meteorolog-
ical factors,  initial  and boundary air  quality  data and emissions inputs,
model structure and computational  factors,  and reactions within the chemical
kinetics mechanism  (U.S.  Environmental  Protection Agency,   1981g;  Liu and
Seinfeld, 1975; Gelinas  and  Vajk,  1979; Til den  and  Seinfeld,  1982; Dunker,
1980, 1981; Falls et a!., 1979).
     A comparative study of the Photochemical Box Model (PBM) of Demerjian and
Schere (1979), the  Lagrangian  Photochemical  Model (LPM) developed by Lurmann
et al.  (1979),  and the  Urban  Airshed Model  (UAM) (Ames  et a!.,  1978) was
performed in which the models were compared in "off-the-shelf" use (Schere and
Shreffler, 1982a,b).  That is, no  effort was made  to  adjust  the model  predic-
tions, although  great care was taken in preparation of  data  and  in  model
execution.   Based upon  application of these models to St.  Louis air quality
data, a fourth model, LIRAQ, was shown to be unsatisfactory  and was not eval-
uated further (Schere and Shreffler, 1982a,b).
     The remaining three models yielded  adequate source-receptor information,
provided knowledgeable interpretation of the output was applied.   For instance,
the  PBM  averaged  23 percent overprediction of ozone  concentration  over all
test days, but  in the 5 stagnation days for which the maximum ozone observed
occurred within the PBM  domain the average overprediction was  only  8 percent.
The  LPM  showed the largest variance  in  the ozone concentration residuals,
possibly because  the input data were not precise enough to fulfill  the temporal
and  spatial  demands  of  the model.   As in previous studies (Whitten and Hogo,
1981; Reynolds  et al.,  1982;  Cole et al., 1982), the  UAM  predicted  ozone
maximum concentrations with little bias  (about 4 percent  overprediction),  but
had  difficulty  placing the  "ozone cloud" at  the correct time and place.
Again, this  suggests  uncertainty in specifying the wind field  data.  The user
                                   3-122

-------
of any  of these models  must  have a strong scientific  background and must
exercise  extreme  care in  implementing  air quality simulations (Schere and
Shreffler, 1982a,b).
     Because of their  wide use,  Levels II, III, and IV of EKMA have been the
subject of  many evaluation efforts.   These  include  parameter sensitivity
studies,  comparison between the  various levels of the  approach,  and  studies
comparing EKMA with other dynamic models (Meyer et al., 1981).
     Jeffries and coworkers (1981) evaluated the performance of EKMA Levels II
and III using  10  days  of data from the 1976 St. Louis Regional Air Pollution
Study.  To  evaluate the effects  of  chemistry and meteorology inputs, four
different chemical kinetics mechanisms and three methods of calculating mixing-
height  profiles  were  employed.   The choice  of mechanism,  trajectory, and
mixing-height  profile  proved to  have a large  effect  on  the prediction of
absolute  ozone levels.  No one mechanism or mixing-height profile was superior,
however,  at  producing  the  "best fit" over all days.   When the EKMA procedure
was used  in  a relative sense to  estimate  needed  reductions  in NMHC,  results
obtained  were  not  consistent.   Improvements needed in  the simple trajectory
model (Level III) were identified as:  (1) improved chemical kinetics descrip-
tions,  (2)  smoother and more defined  trajectories,  (3) better treatment of
point sources, and (4) improved mixing-height profiles.
     In another  sensitivity study,  standard  and city-specific versions  of
EKMA were used to simulate 100 pre-selected test days  (Maxwell and Martinez,
1982).  A statistical  analysis was performed to determine how  accurately these
models, using  three different  chemical  mechanisms,  could predict absolute
ozone levels.  As  above (Jeffries et al., 1981),  none of the models  was  a
consistently good predictor of ozone and ozone  levels were usually overpredicted
by more than 20 percent.
     Finally,  all  three levels of EKMA were  compared on  a  limited number of
test days with respect to:   (1) level of EKMA,  (2) chemical  kinetics mechanisms,
and  (3) isopleth  diagram entry parameters (Hayes  and Hogo,  1982).   Again it
was found that substituting chemical mechanisms produced significant differences
in the  shape of the isopleth curves.  Also, and partly because of, this, EKMA
predictions were found to  be quite sensitive to low NMHC/NO  precursor ratios.
                                                           Jf\.
Precursor-ozone  relationships  derived  for various levels of  EKMA were  not
particularly  different,  nor did they appear to be sensitive to the choice of
trajectory  (for Level  II analysis).
                                   3-123

-------
     Studies  comparing  EKMA performance with  that of more complex dynamic
models have  also been carried  out  (Whitten and Hogo,  1981; U.S. Environmental
Protection Agency, 1981a),  Input to the EKMA model was often generated by the
comparison model so that specific features could be compared.   In all  studies,
the most sensitive difference, in terms of absolute prediction of ozone levels,
was found to be the choice of chemical mechanism.  It was shown, however, that
application of the EKMA produced precursor-ozone descriptions similar to those
from the more complex models when the same chemical mechanism was used in each
of the models.  Finally,  it should  be  noted that the  non-linear relationships
between 03 and its precursors means that good model performance in replicating
base case  conditions  does not ensure accurate emission control calculations.
To address  this problem,  emission control estimates  obtained  with EKMA  have
been compared with trends  and  emission reduction estimates obtained with grid
models (Meyer et a!., 1981).  Further information on the comparative performance
of EKMA is found in DimitHades and Dodge (1983).
     Selection  of  a modeling approach  for  determining ozone concentrations
that is acceptable and appropriate for given circumstances necessitates making
many interrelated decisions.  All models considered should be able, of course,
to simulate the physical  and chemical processes  known  or  suspected  to be
important.  Potential users must then weigh the advantages of greater credibi-
lity and  capability against  the disadvantages  of  greater  cost,  time, and
personnel  requirements  (Association of  Bay Area  Governments,  1979a).  In
addition to  the technical  aspects of potential modeling approaches, specific
selection constraints also include:  (1) extent of data requirements;  (2) costs
of data  collection,  model  implementation,  and operation;  (3) computer con-
straints;  (4) personnel  requirements; and  (5) schedule constraints.   These
criteria are not independent of one another, and time spent defining a selection
plan can  result in  substantial  benefits throughout the modeling  exercise.
3.7
3.7.1   Descriptions and Properties of Ozone and Other Photochemical Oxidants
     Ozone  (Q~)  and other photochemical oxidants occurring at low concentra-
tions  in  ambient air,  such as peroxyacetyl  nitrate  (PAN),  hydrogen  peroxide
(H202)5 and formic  acid (HCOOH), are characterized chiefly by their ability to
remove electrons from, or to share electrons with, other molecules or ions (i.e.,
                                   3-124

-------
oxidation).   The capability  of  a chemical species for oxidizing or reducing
other chemical  species is termed "redox potential" (positive or negative stan-
dard potential) and  is  expressed in volts.  A  reactive  allotrope  of oxygen
that is only about one-tenth as  soluble as oxygen  in water, ozone has a stan-
dard potential of +2.07 volts  in aqueous  systems  for the redox pair, Q-XH-O
(Weast, 1977).  Hydrogen peroxide, which  is highly soluble  in water and other
polar solvents, has a standard potential of +1.776 in the redox pair, 1-LOp/hLO
(Weast, 1977).  No  standard  potential  for peroxyacetyl  nitrate in neutral or
buffered  aqueous  systems,  such  as those  that  occur  in biological systems,
appears in the  literature.   In acidic solution  (pH  5  to 6), PAN hydrolyzes
fairly rapidly (Lee  et  al.s  1983; Holdren  et a!.,  1984); in alkaline solution
it decomposes with the production of nitrite ion and molecular oxygen (Stephens,
1967; Nicksic et al., 1967).  An important property  of PAN, especially  in the
laboratory, is its thermal  instability.  Its explosiveness dictates its synthesis
for  experimental  and calibration purposes by  experienced  personnel  only.
     Formic acid is formed as a  stable product in photochemical air pollution.
It has the structure  of both an  acid and  an aldehyde and in concentrated  form
is a pungent-smelling, highly corrosive liquid.
     The  toxic effects  of oxidants are attributable  to their oxidizing  abili-
ty.  Their oxidizing properties  also  form the  basis  of  several  measurement
techniques for Q-,  and PAN.   The  calibration of ozone  and  PAN measurements,
however,  is achieved via their spectra in the ultraviolet and infrared regions,
respectively.  All three pollutants of most concern in this document (0~, PAN,
and  H-O-) must  be generated i_n  situ for the calibration of measurement tech-
niques.   For  ozone  and  H~Q?, generation  of calibration  gases  is reasonably
straightforward.

3.7.2  Nature of Precursors to Ozone and Other Photochemical Oxidants
     Photochemical oxidants  are  products  of atmospheric reactions  involving
volatile  organic  compounds  (VOC) and  oxides of nitrogen (NO ),  as well  as
                                                             s\
hydroxyl  (OH)  and other radicals, oxygen, and sunlight (see, e.g., Demerjian
et al., 1974;  National  Research  Council,  1977a;  U.S.  Environmental  Protection
Agency, 1978a; Atkinson, 1985).   The oxidants are  largely secondary pollutants
formed in the atmosphere from their precursors by  processes that are a complex,
nonlinear function of precursor  emissions  and meteorological factors.
                                    3-125

-------
     The properties  of organic compounds that are most  relevant to their role
as precursors to ozone and other oxidants are their volatility, which governs
their emissions  into the atmosphere;  and  their chemical reactivity, which
determines their lifetime  in the atmosphere.  Although vapor-phase hydrocar-
bons  (compounds  of  carbon and  hydrogen only)  are the  predominant organic
compounds in ambient air that serve as precursors to photochemical oxidants,
other volatile organic compounds  are also photochemically reactive  in those
atmospheric processes that give rise to oxidants.  In particular,  halogenated
organics (e.g., haloalkenes)  that  participate in photochemical reactions are
present in ambient air,  although at lower concentrations than the hydrocarbons.
They are oxidized  through  the same initial step involved in the oxidation of
the hydrocarbons;  that is,  attack by hydroxyl radicals.   Alkenes, haloalkenes,
and aliphatic  aldehydes  are, as  classes,  among the most reactive organic
compounds found in ambient  air  (e.g.,  Altshuller and Bufalini, 1971; Darnall
et al.,  1976; Pitts et al., 1977; U.S.  Environmental  Protection Agency,  1978a,
and references  therein).  Alkenes and haloalkenes are  unique among VOC in
ambient air  in that they are susceptible both  to attack by OH radicals (OH)
and by  ozone  (Niki et al.,  1983).   Methane, halomethanes, and certain haloe-
thenes  are of  negligible reactivity in ambient air and have  been  classed as
unreactive by the U.S. Environmental Protection Agency  (1980a,b).  Since methane
is considered only negligibly reactive in ambient air, the volatile organic
compounds of  importance  as  oxidant precursors are usually  referred to as
nonmethane hydrocarbons  (NMHC)  or,  more properly, as  nonmethane organic
compounds (NMOC).
     The oxides of nitrogen that are  important as  precursors to ozone and
other photochemical oxidants are nitrogen dioxide (N0?) and nitric oxide (NO).
Nitrogen dioxide is  itself an oxidant that produces deleterious effects, which
are the subject of a separate criteria document (U.S.  Environmental Protection
Agency,  1982a).  Nitrogen  dioxide  is an important precursor  (1) because its
photolysis in ambient air  leads to  the formation of oxygen atoms that combine
with molecular oxygen to form ozone; and  (2) because it reacts with acetyl-
peroxy  radicals to form  peroxyacetyl nitrate, a phytotoxicant and  a  lachryma-
tor.  Although ubiquitous, nitrous oxide (N,,0) is unimportant in the production
of oxidants  in  ambient  air because it is virtually inert in the troposphere.
                                   3-126

-------
3.7.3 ' Atmospheric Reactions of Ozone and Other Oxidants  Including Their  Role
       in Aerosol Formation
     The chemistry of  the  polluted atmosphere is exceedingly complex, but an
understanding of the basic  phenomena  is not difficult  to acquire.   Three
processes occur:   the  emission of precursors to  ozone  from predominantly
manmade sources;  photochemical  reactions that take .place  during the  disper-
sion and transport of  these precursors; and scavenging processes that reduce
the concentrations of both 03 and precursors along the trajectory.
     The specific atmospheric  reactions of ozone and  of  other  photochemical
oxidants such  as peroxyacetyl  nitrate  and  hydrogen  peroxide are  becoming
increasingly well-character!zed.   The reactions of these species  result in
products and processes that may have significant environmental and health- and
weIfare-related  implications, including effects on biological systems, nonbio-
logical materials, and such phenomena as visibility degradation and acidifica-
tion of cloud and rain water.
3.7.3.1   Formation  and  Transformationof Ozone and Other Photochemical  Oxi-
dants.  In the troposphere, ozone  is formed through the dissociation  of N0? by
sunlight to yield an oxygen atom, which then reacts with molecular oxygen (0_)
to produce  an 0- molecule.   If  it  is present,  NO  can  react rapidly with  0,  to
form NOp and an  0^ molecule.  In the absence of competing  reactions,  a steady-
state or equilibrium concentration of 0- is  soon  established between  GO,  NOp,
and NO (National  Research Council, 1977a).  The injection  of  organic  compounds
into  the atmosphere  upsets  the  equilibrium  and allows the ozone to accumulate
at  higher  than  steady-state  concentrations.   The  length of the induction
period before the accumulation of 03 begins  depends  heavily on the  initial
NO/NO,, and NMOC/NO   ratios  (National Research Council, 1977a).
     i—            /\
     The major role  played  by organic compounds in smog reactions is  attribut-
able to the hydroxyl radical (OH), since it reacts with essentially all organic
compounds (e.g.,  Atkinson,  1985; Herron  and Huie, 1977, 1978; Dodge and Arnts,
1979;  Niki  et a!.,   1981).  Aldehydes,  which  are  constituents of automobile
exhaust  as  well  as  decomposition  products  of  most atmospheric  photochemical
reactions  involving  hydrocarbons,  and  nitrous acid  (MONO), are important
sources  of  OH radicals,  as is  0~  itself.  Other  free  radicals,  such  as hydro-
and alkylperoxy  radicals and the nitrate (NO,) radical play  important roles in
photochemical air pollution.
                                    3-127

-------
     The presence  of  organic  compounds,  oxides  of  nitrogen, and  sunlight does
not mean that the  photochemical  reactions will  continue  indefinitely.  Termi-
nation reactions gradually  remove NCL from the  reaction-mixtures, such that
the photochemical  cycles  slowly come to an end unless fresh NO and N0? emis-
sions are  injected into the atmosphere.  Compounds containing nitrogen,  such
as PAN, nitric  acid (HNO-), and peroxynitric acid (HNO»), as well as organic
and inorganic nitrates, are formed in these termination reactions.
     Recent studies on the photooxidation of organic compounds under simulated
atmospheric conditions  have been reasonably successful.   The  rate constants
for the reaction of OH  radicals  with  a  large number of organic compounds have
been measured  (e.g.,  Atkinson  et a!.,  1979; Atkinson et al., 1985).   The
mechanisms of the  reactions of paraffinic compounds are  fairly  well under-
stood, as are those of olefinic compounds, at least for the smaller compounds.
Photooxidation reactions of the aromatic compounds, however, are poorly under-
stood.
     In the presence  of NO , natural hydrocarbons (i.e.,  those  organic com-
                           s\
pounds emitted  from vegetation) can also undergo photooxidation reactions to
yield 0,,  although most naturally emitted hydrocarbons  are  olefins  and  are
scavengers as well as producers  of 0-  (e.g.,  Lloyd et a!., 1983; Atkinson
                    (          .        w
et al.» 1979;  Kamens  et al., 1982; Killus  and  Whitten,  1984; Atkinson and
Carter, 1984).
3.7.3.2  Atmospheric Chemical Processes InvolvingOzone.   Ozone can react with
organic compounds  in  the  boundary layer of the troposphere  (Atkinson and
Carter, 1984).   It is important to recognize, however, that organics undergo
competing  reactions with  OH radicals in the daytime (Atkinson et al., 1979;
Atkinson,  1985)  and,  in certain  cases,  with  N03 radicals during the  night
(Japar and Niki,  1975;  Carter et al., 1981a; Atkinson et al., 1984a,b,c,d,e;
Winer et al., 1984), as well as photolysis, in the case of aldehydes and other
oxygenated organics.  Only  for organics whose  ozone reaction  rate constants
                     -21    3         ~1    -1
are  greater  than  ~10     cm  molecule   sec    can consumption by ozone  be
considered to be atmospherically  important (Atkinson and Carter, 1984).
     Ozone reacts  rapidly with the acyclic mono-, di-,  and tri-alkenes  and
                                                                           — 1 O
with cyclic alkenes.  The rate constants for these reactions range from ~10
to ~10     cm  molecule"  sec"   (Atkinson and Carter, 1984), corresponding to
atmospheric lifetimes  ranging from a  few minutes for the more  reactive cyclic
alkenes, such as the  monoterpenes,  to several days.  In  polluted atmospheres,
                                   3-128

-------
a significant portion  of the consumption Of  the  more reactive alkenes will
occur via reaction with  ozone  rather  than with  OH radicals,  especially in the
afternoons during photochemical oxidant episodes.   Reactions between ozone and
alkenes can  result  in aerosol formation  (National  Research Council,  1977a;
Schuetzle and  Rasmussen, 1978), with  alkenes of  higher carbon numbers  the
chief contributors.
     Because of their respective rate constants,  neither alkanes (Atkinson and
Carter, 1984) nor alkynes  (Atkinson and  Aschmann, 1984)  are expected to react
with ozone in the atmosphere,  since competing reactions  with OH radicals have
higher rate constants (Atkinson et a!., 1979'; Atkinson, 1985).
     The aromatics react with ozone,  but quite slowly (Atkinson  and Carter,
1984), such  that their reactions with ozone  are expected to be unimportant in
the  atmosphere.   Cresols are  more reactive  toward ozone than  the  aromatic
hydrocarbons (Atkinson and Carter, 1984), but their reactions with OH  radicals
(Atkinson, 1985)  or NQ3 radicals  (Carter et a!.,  1981a;  Atkinson et a!.,
1984d) predominate.
     For oxygen-containing organic compounds, especially those without carbon-
carbon double bonds,  reactions with ozone are slow.   For carbonyls and ethers
(other than  ketene) that  contain  unsaturated carbon-carbon bonds,  however,
much faster reactions are observed (Atkinson  and  Carter, 1984).
     Certain reactions of ozone other than  its reactions with organic  com-
pounds are important  in  the  atmosphere.   Ozone reacts rapidly with NO to form
N02, and  subsequently with NO,, to produce  the nitrate (N03)  radical  and an
oxygen molecule.  Photolysis of ozone can be a significant pathway  for the
formation  of OH radicals, particularly  in  polluted  atmospheres when ozone
concentrations are at  their  peak.
     Ozone may  play  a role  in the oxidation  of S02 to H^SO,,  both  indirectly
in  the gas  phase (via formation of OH radicals  and  Criegee biradicals) and
directly in aqueous droplets.          -
3.7.3.3   Atmospheric  Reactions of  PAN, H00^,' and  HCOQH.   Because  PAN is  in
            ULU"'" ^   \   "   """'"	""""" ~""r""vrr  L ' ™«™ '«>«	•'•"•"•"      £"-"jT -"••-      iuu.uuui-.ui
equilibrium with acetyl  peroxy radicals  and  NOg,  any  process that  leads  to the
removal of either of these species will  lead  to the decomposition  of PAN.  One
such  process  is the reaction  of NO with acetyl  peroxy  radicals.   This can
lead, however, to the  formation of OH radicals.   Thus, PAN  remaining overnight
from  an  episode on the  previous day  can react with  NO  emitted from morning
traffic  to  produce  OH radicals (Cox  and Roffey,  1977; Carter et al., 1981c)
                                    3-129

-------
that will enhance smog formation on that day (e.g., Tuazon et al,, 1981a).   In
the absence of  significant  NO concentrations, and in regions of moderate to
lower temperatures,  PAN  will  persist in the  atmosphere  (Wellington et a!.,
1984; Aikin et  a!.,  1983)  and contribute to the long-range transport of NO .
                                                                           r\
     Although hydrogen peroxide formed  in the gas  phase from the reactions  of
hydroperoxyl radicals plays a role in HO   chemistry in the troposphere, and
                                         j\
especially in the  stratosphere (Crutzen and Fishman, 1977; Cox and Burrows,
1979), its major  importance arises from its  high  solubility  in  water.  The
latter ensures  that a large  fraction  of gaseous H-Op will be taken  up in
aqueous droplets.   Over  the past decade, evidence has accumulated that tiJ^o
dissolved in cloud,  fog,  and  rainwater may play an important, and, in acidic
droplets (i.e., pH  <5), even  a dominant  role  in  the oxidation of SQ2 to HLSO,
(e.g., Hoffman  and  Edwards, 1975;  Martin and Damschen, 1981;  Chameides and
Davis, 1982; Calvert and Stockwell, 1983,  1984;  Schwartz,  1984).   Hydrogen
peroxide may also  play  a role in  the  oxidation  of N0? dissolved in aqueous
droplets, although relevant data are limited and additional research is required
(see, e.g., Gertler et a!., 1984).   Substantial  uncertainties remain concerning
the quantitative role of HpOp  in acidification of aqueous particles and droplets
(Richards et a!., 1983).
     Because it can be scavenged rapidly into water droplets, formic acid can
                                                           t
potentially function as an oxidant in cloud water and rain water.   Thus, HCOOH
is an example  of  a compound that is a non-oxidant or weak oxidant in the gas
phase but that  is  transformed upon incorporation in aqueous solutions  into  an
effective oxidizer of S(IV).   Although much uncertainty remains concerning the
quantitative role of HCOOH and the higher organic acids,  they potentially play
a minor but still significant role in the acidification of rain.

3.7.4  Meteorological and Climatological Processes
     Meteorological  and Climatological  processes are important in  determining
the extent  to  which precursors to ozone and other photochemical  oxidants can
accumulate, and thereby  the  concentrations of ozone and  other oxidants that
can result.  The  meteorological  factors most important in  the formation and
transport of ozone and  other  photochemical oxidants in the lower troposphere
are:   (1)  degree  of atmospheric  stability; (2) wind speed  and  direction;
(3) intensity and wavelength of sunlight; and (4) synoptic weather conditions.
These  factors  are  in turn dependent  upon  or interrelated with geographic,
seasonal, and other Climatological factors.
                                   3-130

-------
     Incursions of ozone from the stratosphere are an additional  source of the
ozone found in the lower troposphere.   The physical and meteorological  mechanisms
by which  ozone is  brought  into the troposphere from  the  stratosphere are
important in determining the resulting ground-level concentrations, ground-level
locations impacted, and the  seasonality of incursions  of stratospheric ozone.
3.7.4.1  Atmospheric Mixing.  The concentration of a pollutant in  ambient air
depends significantly on the degree of atmospheric mixing that occurs from the
time the pollutant or its precursors are emitted and the arrival  of the pollu-
tant at the  receptor.   The  rate at which atmospheric mixing proceeds and the
extent of the  final  dilution depends on  the  amount of turbulent mixing that
occurs and on  wind speed and direction.   Atmospheric stability is one of the
chief determinants  of turbulent mixing since  pollutants do  not spread  rapidly
within stable  layers  nor  do they mix upward  through stable layers to  higher
altitudes.
     Temperature inversions, in which the temperature increases with increasing
altitude, represent the  most stable atmospheric conditions.  Surface  inver-
sions  (base  at ground level) and elevated  inversions  (the entire layer  is
above the surface)  are  both common (Hosier,  1961; Holzworth, 1964)  and both
can  occur  simultaneously at the same  location.   Surface inversions show a
diurnal pattern, forming at night in the absence of solar radiation but break-
ing up by about mid-morning as the result of surface heating by the sun (Hosier,
1961; Slade,  1968).   Elevated  inversions can persist  throughout the day  and
pollutants can be  trapped  between  the  ground surface and  the base  of the
inversion.   The persistence of elevated  inversions is a major meteorological
factor contributing to  high pollutant concentrations  and  photochemical smog
conditions along the California coast (Hosier, 1961; Holzworth, 1964; Robinson,
1952).  In coastal  areas generally, such as  the  New England coast  (Hosier,
1961) and along the Great Lakes (Lyons and Olsson, 1972), increased atmospheric
stability (and diminished mixing)  occurs in summer and fall as the result of
the temperature differential between the water and the land mass.
     The  depth of the layer in  which turbulent  mixing can occur  (i.e., the
"mixing height") shows geographical dependence.  Summer morning mixing heights
are  usually  >300  m in the  United States  except  for  the Great Basin  (part of
Oregon,  Idaho, Utah,  Arizona,  and most  of Nevada), where the mixing  height  is
~200 m  (Holzworth,  1972).   By mid-morning, mixing heights  increase  markedly
such that only a few coastal areas have mixing heights <1000 m.
                                   3-131

-------
     Summer afternoon mixing heights are generally an indication of the poten-
tial for recurring photochemical oxidant problems.  Photochemical smog problems
in the United States are somewhat unexpected since the lowest afternoon mixing
height is ~600 m  (Holzworth,  1972).   Elevated inversions having bases <500 m
(i.e.j low-level  inversions)  occur in the United States,  however,  with the
following frequencies:  90 percent on the California coast; >20 percent on the
Atlantic coast (New Jersey to Maine); >5 percent  along the Great Lakes; and 5
to 10 percent from Louisiana to Arkansas and eastward to about Atlanta, Georgia.
For  most  areas of  the  United States, though, the  persistence  through the
afternoon of  low-level  stable layers is a rare event, occurring on <1 day in
20 (Holzworth and Fisher, 1979).
3.7.4.2  Wind Speed and Direction.   For areas  in  which mixing heights are not
restrictive, wind speed and,  in  some cases, wind  direction are major determi-
nants of pollution  potential.   Since strong winds dilute precursors to ozone
and other photochemical  oxidants, a location may have good ventilation despite
the  occurrence of persistent  inversions (e.g., San  Francisco).  Conversely,
light winds can  result  in high oxidant levels even if the mixing  layer  is
deep.
     The frequency of weak winds, then, is important in oxidant formation.  In
industrialized, inland areas east of the Mississippi River, surface inversions
in the morning  coupled  with wind speeds £2.5  m/sec (£6  mi/hr)  occur with a
frequency >50 percent (Holzworth  and Fisher, 1979).   These surface  inversions
break up by afternoon, however, permitting dispersion.
     The effects  of wind speed and  direction  include the  amount of dilution
occurring in  the  source  areas,  as well  as along the trajectory followed by an
urban or source-area plume.  Regions having steady prevailing winds, such that
a given air parcel  can  pass over a  number  of significant source areas, can
develop significant levels of pollutants even in the absence of weather patterns
that  lead to  the stagnation type of air  pollution  episodes.   The Northeast
states are  highly susceptible to pollutant plume transport effects, although
some  notable  stagnation episodes  have  also  affected this area  (e.g.,  Lynn
et a!., 1964).   Along the  Pacific  Coast, especially along the coast of
California, coastal  winds  and a  persistent low inversion layer contribute to
major pollutant  buildups  in urban source areas and downwind along  the  urban
plume trajectory  (Robinson, 1952; Neiburger et a!., 1961).
                                   3-132

-------
3.7.4.3  Effects of Sunlight and Temperature.   The effects  of  sunlight on
photochemical oxidant formation, aside  from the role of  solar  radiation in
meteorological processes, are  related to its intensity and its spectral dis-
tribution.   Intensity varies diurnally, seasonally, and with latitude, but the
effect of  latitude  is  strong only in the winter.   Experimental studies  have
verified the  effects on  oxidant formation of  light intensity (Peterson,  1976;
Demerjian  et  al.,  1980)  and its diurnal variations  (Jeffries  et  al., 1975;
1976), as  well  as on the overall photooxidation process (Jaffee et al.,  1974;
Winer et al., 1979).
     A correlation  between  high oxidant  concentrations and warm,  above-normal
temperatures  has  been  demonstrated  generally (Bach,  1975;  Wolff and  Lioy,
1978) and- for specific locations,  e.g., St.  Louis  (Shreffler and Evans, 1982).
Coincident meteorology appears  to  be the cause of the observed correlation.
Certain synoptic weather conditions  are favorable both for the occurrence of
higher temperatures and  for the formation of ozone and other oxidants, so that
temperature is  often  used to forecast the potential for high oxidant concen-
trations (e.g.,  Wolff  and Lioy, 1978; Shreffler and Evans, 1982).  Data from
smog chamber  studies  show an effect of temperature on ozone formation (e.g.,
Carter et  al.,  1979b;  Countess et al.,  1981),  but the effect is thought to
result from  the volatilization and reaction  of chamber wall contaminants as
the temperature is increased.
3.7.4.4    Transport of OzoneandOther Oxidants and Their Precursors.   -The
levels of  ozone and other oxidants  that will  occur at a  given  receptor site
downwind of  a precursor source area  depend  upon  many interrelated factors,
which include but are not restricted to:  (1) the  concentrations of respective
precursors leaving  the  source  area;  (2)  induction  time;  (3)  turbulent mixing;
(4) wind speed  and wind direction;  (5) scavenging during transport;  (6) in-
jection of new emissions from  source areas  in the  trajectory of the  air mass;
and (7) local and synoptic weather conditions.
     Ozone  and other photochemical oxidants  can  be transported hundreds of
miles  from the  place  of origin of their precursors,  as  documented by  the
numerous studies  on transport phenomena that were described in the 1978 cri-
teria  document  for ozone and other photochemical  oxidants (U.S. Environmental
Protection Agency, 1978a).   In that  document, transport phenomena were classi-
fied  into  three categories, depending upon transport distance:   urban-scale,
                                    3-133

-------
mesoscale, and  synoptic-scale.   In urban-scale transport, maximum concentra-
tions of  Qg  are produced about 20 miles or so (and about 2 to 3 hours) down-
wind from the  major pollutant source areas.   In mesoscale transport,  0~ has
been observed  up  to 200 miles downwind from  the  sources  of its precursors.
Synoptic-scale  transport is associated with  large-scale,  high-pressure air
masses that may extend over and persist for many hundreds of miles.
     Urban-scale transport  has been  identified as  a  significant,  characteris-
tic feature  of the oxidant problem  in  the  Los Angeles Basin (Tiao et al.,
1975), as well  as in San Franciso, New York, Houston, Phoenix, and St. Louis
(e.g., Altshuller,  1975; Coffey  and  Stasiuk,  1975; Shreffler and  Evans, 1982;
Wolff et  al.,  1977a).   Simple advection  of a photochemically reactive air
mass, local  wind  patterns,  and  diurnal wind cycles appear to  be the main
factors involved in urban-scale transport.
     Mesoscale  transport is in  many respects an  extension  of urban-scale
transport and  is  characterized  by the development  of urban  plumes.   Bell
documented cases  in 1959 in which precursors  from the Los Angeles Basin and
the resultant  oxidant plume were transported over the coastal Pacific Ocean,
producing elevated oxidant concentrations in San  Diego County  the next day
(Bell, 1960).   Similar scales  of transport have been  reported by Cleveland
et al. (1976a,b) for  the New York-Connecticut area; by Wolff and coworkers and
others (Wolff  et  al., 1977a,c;  Wolff and Lioy, 1978; Clark and Clarke, 1982;
Clarke et al.,  1982;  Vaughan  et al., 1982)  for  the Washington,  DC-Boston
corridor;  and  by  Westberg  and coworkers for  the  Chicago-Great Lakes area
(Sexton and  Westberg, 1980; Westberg et al., 1981).   These and other studies
have demonstrated  that ozone-oxidant plumes from major urban areas can extend
downwind  about 100 to 200 miles  and  can have  widths  of tens of miles  (Sexton,
1982), frequently  up  to  half the length of the plume.
     Synoptic-scale transport  is characterized by the general and widespread
occurrence of elevated oxidants and precursors on a regional or air-mass scale
as the result  of certain favorable weather conditions, notably, slow-moving,
well-developed  high-pressure,  or anti-cyclonic, systems characterized by weak
winds and limited vertical mixing (Korshover, 1967;  1975),  The  size of the
region that  can be affected has been described  by Wolff and coworkers, who
reported  the occurrence  of  haze and elevated ozone levels in an area extending
from the  Midwest  to the Gulf Coast (Wolff et al.,  1982) and the occurrence of
elevated  ozone concentrations  extending in a  virtual  "ozone  river"  from the
                                   3-134

-------
Gulf Coast to  New England that affected anywhere  from  a few hundred square
miles to a thousand  square miles during a 1-week  period  in July 1977 (Wolff
and Lioy, 1980).
3.7.4,5  Stratospheri'c-Tropospheric Ozone Exchange.   The  fact that ozone  is
formed in the  stratosphere,  mixed downward, and incorporated into the tropo-
sphere, where  it  forms  a more or less  uniformly mixed  background  concentra-
tion,  has  been known in  various  degrees of detail  for many  years (Junge,
1963).  It is  widely accepted that the  long-term  average tropospheric back-
ground concentration  of  about 30 ppb to 50 ppb results primarily, though not
exclusively,  from  the transfer of stratospheric ozone  into the upper tropo-
sphere, followed  by  subsequent dispersion throughout the troposphere (e.g.,
Kelly et a!., 1982).
     The exchange of ozone between the stratosphere and the troposphere in the
middle latitudes  occurs  to a major extent in events  called "tropopause folds"
(TF)  (Reiter,  1963;  Reiter and Mahlman, 1965; Danielsen, 1968; Reiter, 1975;
Danielsen and  Mohnen,  1977;  Danielsen, 1980), in  which the polar  jet stream
plays  a  major  role.   From recent studies, Johnson and Viezee (1981) proposed
four types or mechanisms of TF injection and concluded that two of these, both
of which are consistent  with theory,  could  cause substantial  effects in  terms
of high  ozone  concentrations at ground  level.  They concluded, in addition,
that all low-pressure trough  systems may induce TF events and cause the trans-
tropopause movement of ozone-rich air into the troposphere (Johnson and Viezee,
1981).
3.7.4.6   Stratospheric Ozone atGround Level.   From a detailed  review  of
studies  on background tropospheric  ozone, Viezee  and Singh  (1982) concluded
that  the  stratosphere is a major but not the sole source of background ozone
in the unpolluted troposphere, a conclusion reached  by  other  investigators as
well  (e.g.,  Kelly et al., 1982).  The  stratospheric ozone reservoir shows a
strong seasonal  cycle that is reflected at  ground-level. At some stations
that  monitor background  ozone levels, average  spring background  levels may be
as high  as  80 ppb, with average fall levels ranging  from 20 to 40 ppb (e.g.,
Singh  et al.,  1977; Mohnen,  1977; U.S.  Environmental  Protection Agency, 1978a).
Viezee and  Singh (1982) and  Viezee  et  al.  (1983)  concluded that  relatively
high  ozone  concentrations can occur  for  short  periods  of time (minutes,  to a
few  hours) over local areas  as a  result of stratospheric  intrusions.
                                    3-135

-------
     A number  of  investigators have attempted to  quantify the amount of the
surface ozone  that  can be attributed to  stratospheric  sources.   The method
most commonly  used  is based on the  assumption  that beryl!ium-7  ( Be) is a
unique tracer for air parcels of stratospheric origin.  Calculated correlations
between surface ozone and  Be show, however, that their relationship is highly
variable  (e.g.,  Kelly et al.,  1982;  Ferman and Monson,  1978;  Johnson and
Viezee, 1981;  Husain  et  al., 1977).  Singh  et al.  (1980)  and Viezee  and Singh
(1982) have pointed out  problems with using this  technique  to quantify the
contribution of  stratospheric  ozone to surface  ozone.   Singh  et al. (1980)
concluded that "the experimental  technique involving a  Be/0- ratio to esti-
mate the  daily stratospheric component of  ground-level 0, is  unverified and
considered to  be  inadequate for air quality  applications"  (p.  1009).   This
group of  investigators  have suggested,  however, that   Be may  be  used, under
appropriate meteorological  conditions, as a qualitative tracer for air masses
of  stratospheric  origin (Johnson  and  Viezee, 1981;  Viezee  et al.,  1979).
     Other methods  used  to  attempt to quantify  the stratospheric  component of
surface ozone include aircraft observations of TF events coupled with calcula-
tions of  downward  ozone  flux,  and examination of surface ozone data records.
From such data, Viezee et al. (1983) concluded that direct ground-level contri-
butions from  stratospheric ozone  are infrequent (<1 percent of the time),
short-lived, and associated with ozone concentrations £0.1 ppm.
     Notwithstanding difficulties with quantifying its contribution to surface
ozone, however, stratospheric  ozone is clearly  present  in atmospheric surface
layers, and the meteorological  mechanisms responsible have been described by a
number of investigators  (e.g., Danielsen,  1968; Wolff  et al.s 1979; Johnson
and Viezee, 1981).
3.7.4.7   Background Ozone from Photochemical Reactions.  Whereas stratospheric
ozone  is  thought by  many investigators   to be  the dominant  contributor to
background  levels  of  ozone, as discussed  above,  other investigators have
concluded that as much as  two-thirds of the annual  average background concen-
trations  may  result from photochemical reactions.   For example, Altshuller
(1986), in a recent review article, has concluded that photochemically generated
ozone should equal  or exceed the stratospheric contribution at lower-elevation
remote locations; and that photochemically  generated ozone from manmade emissions
probably  constitutes most of the ozone measured at more polluted rural locations
during the  warmest  months of the year.   His conclusions were based, in part,
                                   3-136

-------
on an analysis of global circulation (e.g., Levy et al., 1985) and photochemical
modeling .studies  (e.g.,  Fishman  and Seller, 1983; Fishman and Carney, 1984;
Fishman et al.,  1985;  Dignon and Hameed,  1985).   In these modeling studies,
the photochemical  contribution  to background ozone levels was  estimated to
range from ~15 ppb (long-term) to ~80 ppb (summertime), depending on the level
of NO  emissions assumed.
     /v
     Studies on  the  role of NO   in  nonurban ozone photochemistry  have shown
                               •r\
that ozone formation at many  of  the  locations is  not NO -limited,  but depends
                                                        P\
on VOC reactions, as well (e.g., Martinez and Singh,  1979; Kelly et al., 1984;
Liu et al., 1984).  Background NO concentrations  at most remote,  clean  loca-
                                  y\
tions range from <0.05 ppb upward.   Mean  concentrations  of  NO  at nonurban
                                                               /\
locations in the United States east of the Rocky Mountains range from ~1 ppb
to 10 ppb (Altshuller,  1986; see also Sections 3.5 and 3.7.5).  These background
concentrations of  NO   are  higher than previously  thought  (see, e.g., Singh  et
                    )\
al., 1980; Kelly et al., 1984, regarding global models  and assumed reservoirs
of NOX).
     The contributions  of  biogenic VOC  to  background ozone,  although a matter
of controversy in recent years, appear not to be significant under most atmos-
pheric conditions, since ambient air concentrations of biogenic  VOC are  quite
low, even at rural sites (Altshuller, 1983).
     Thus, photochemistry  and stratospheric intrusions are both  regarded as
contributing to  background ozone concentrations, but the apportionment of
background to respective sources  remains a matter  of investigation.

3.7.5  Sources,  Emissions,and Concentrations of Precursors toOzone and Other
       Photochemical Oxidants
     As  noted  earlier,  photochemical production of ozone depends  both on the
presence of precursors, volatile organic compounds (VOCs) and nitrogen  oxides
(NO ), emitted  by manmade  and by natural  sources; and on suitable conditions
of  sunlight, temperature,  and other meteorological factors.   Because of the
intervening requirement for meteorological conditions  conducive  to the  photo-
chemical generation  of ozone, emission inventories are not as direct predic-
tors of  ambient  concentrations of secondary pollutants such as ozone and other
oxidants as they are for primary  pollutants.
3.7.5.1   Sources and Emissions of Precursors.   Emissions  of  manmade  VOCs
(excluding  several relatively unreactive  compounds such  as  methane)  in the
                                   3-137

-------
United States  have  been estimated at 19.9 Tg/yr for 1983 (U.S. Environmental
Protection Agency,  1984a).   Retrospective  estimates  show that manmade VOC
emissions rose from about 18.5 Tg/yr in 1940 to about 27.1 Tg/yr in 1970 (U.S.
Environmental Protection  Agency,  1986).   An examination of trends in manmade
VOC emissions  for  1970 through 1983 shows  that the annual emission  rate for
manmade VOCs decreased some 26 percent during this period.  The main sources
nationwide are  industrial processes,  which emit a wide variety of VOCs, such
as chemical  solvents;  and transportation,  which includes  the emission of VOCs
in gasoline  vapor  as  well as  in  gasoline  combustion  products.  Estimates  of
biogenic  emissions  of  organic  compounds  in the  United States are  highly
inferential but  data  suggest that the yearly rate  is  the  same  order  of  magni-
tude as  manmade emissions.   Most  of  the  biogenic emissions actually occur
during the growing  season,  however, and the kinds  of compounds emitted are
different from those arising from manmade sources.
     Emissions of manmade NO  in the United States were estimated at 19.4 Tg/yr
                            /\
for 1983.  Retrospective  estimates show that manmade NO  emissions rose from
                                                        f\
about 6.8 Tg/yr in 1940 to about 18.1 Tg/yr' in 1970 (U.S.  Environmental  Protec-
tion Agency,  1986).   Annual emissions of  manmade  NO   were some 12  percent
                                                    jfX
higher in  1983  than in 1970,  but  the rate leveled  off  in the  late 1970s and
exhibited  a  small  decline from about 1980  through  1982 (U.S.  Environmental
Protection Agency, 1984a).  The increase over the period 1970 through 1983 had
two main  causes:   (1)  increased fuel combustion in  stationary  sources such as
power plants;  and  (2)  increased fuel combustion in  highway motor vehicles, as
the result of the  Increase in vehicle miles  driven.   Total  vehicle miles
driven increased by 42 percent over the 14 years in question.
     Estimated biogenic NO   emissions  are based on uncertain  extrapolations
                          yx
from very  limited  studies,  but appear to be about an order of magnitude less
than manmade NO  emissions.
               /\
3.7.5.2  Representative Concentrations in Ambient Air.
3.7.5.2.1  Hydrocarbons in  urban areas.   Most of  the available ambient air
data on  the  concentrations  of nonmethane  hydrocarbons  (NMHC)  in urban  areas
have been  obtained  during the 6:00 to 9:00 a.m.  period.   Since hydrocarbon
emissions are at their peak during that period of the day, and since atmospheric
dispersion is  limited that  early in the morning,  NMHC concentrations measured
then generally  reflect maximum diurnal levels.  Representative data  for urban
areas show mean NMHC concentrations between 0.4 and 0.9 ppm.'
                                   3-138

-------
     The hydrocarbon composition  of  urban atmospheres  is  dominated by species
in the C2 to C-,Q molecular-weight range.  The paraffinic hydrocarbons (alkanes)
are most prominent,  followed by aromatics and  alkenes.   Based  on speciation
data obtained  in  a number of urban areas, alkanes generally constitute 50 to
60 percent of the hydrocarbon burden in ambient air, aromatics 20 to 30 percent,
with alkenes and acetylene making up the remaining 5 to 15 percent (Sexton and
Westberg, 1984).
3.7,5.2.2   Hydrocarbons in nonurbanareas.    Rural   nonmethane  hydrocarbon
concentrations  are  usually one to two  orders  of  magnitude lower than those
measured in urban areas (Ferman, 1981;  Sexton and Westberg, 1984).  In samples
from sites  carefully selected to guarantee their rural character, total NMHC
concentrations  ranged  from 0.006 to 0.150 ppm  C  (e.g., Cronn,  1982; Seila,
1981;  Holdren  et  al.,  1979).   Concentrations of individual species  seldom
exceeded 0.010 ppm C.  The bulk of species present in  rural areas are alkanes;
ethane,  propane,  ri-butane,  iso-pentane,  and ri-pentane  are  most abundant.
Ethylene and propene are sometimes present at  <0,001  ppm C, and toluene  is
usually  present at  ~0.001 ppm C.   Monoterpene concentrations  are usually
£0,020 ppm  C.   During the summer months,  isoprene  concentrations as  high  as
0.150  ppm  C have  been  measured (Ferman, 1981).  The maximum  concentrations of
isoprene  usually  encountered, however, are  in  the  range of 0.030 to 0.040
ppm C.
3.7.5.2.3  Nitrogen  oxides in  urban  areas.  Concentrations of NO  . like hydro-
                                                                f\
carbon concentrations,  tend  to peak in urban areas  during the  early morning,
when atmospheric  dispersion  is limited and automobile  traffic  is  dense.   Most
NO   is emitted as nitric  oxide (NO),  but the NO  is  rapidly converted to N0?,
  x,                                                                        *~
initially  by thermal oxidation and  subsequently  by  ozone and peroxy radicals
produced  in atmospheric photochemical  reactions.   The  relative  concentrations
of  NO  versus  N0? fluctuate  day-to-day,  depending on diurnal and day-to-day
fluctuations in ozone  levels  and  photochemical  activity.
     Urban  NO   concentrations  during the 6:00 to  9:00  a.m. period in 10 cities
             /\
ranged from 0.05 to 0.15  ppm in  studies done  in  the last 5 to  7 years  (e.g.,
Westberg  and  Lamb, 1983; Richter, 1983; Eaton et  al.,  1979),  although concen-
trations  two  to  three times  higher occur in  cities  such as  Los Angeles. ,
Concurrent  NMHC measurements for these 10 cities showed  that NMHC/NO   ratios
                                                                     J\
ranged from 5  to  16.
                                    3-139

-------
3.7.5.2.4  Nitrogen oxides  In  nonurban areas.   Concentrations  of  NO   in  clean
                                                                   f\
remote  environments  are usually  <0.5 ppb (Logan, 1983).   In  exceptionally
clean air, NO  concentrations  as low as 0.015 ppb have been recorded (Bellinger
             f\
et a!,, 1982).   Concentrations of NO  at nonurban  sites1.;in the northeastern
s                                     f\.
United  States  appear  to be higher than  NO   concentrations in the west by a
                                          A,
factor  of ten (Mueller  and Hidy,  1983).  From the  limited amount of data
available, NO  concentrations  in unpopulated nonurban areas in the west average
             *\
<1 ppb; but  in nonurban northeastern  areas  average NO  can exceed 10 ppb.
^~                                                   "   Px

3.7.6  Source-Receptor (Oxidant-Precursor) Models
     In order  to apply knowledge of the atmospheric chemistry of precursors,
and of  ozone and other photochemical oxidants, during  their  dispersion and
transport, models  describing  these phenomena have been  developed  in a variety
of forms  over the past 15 years.  Most  of  these  models relate the rates of
precursor emissions from mobile and stationary sources,  or precursor atmos-
pheric  concentrations,  to  the resulting ambient  concentrations of secondary
pollutants that impact receptors at .downwind sites.  For this reason they have
been described as source-receptor models.
     Current  air quality,  or  source-receptor,  models can  be  classified  as
either  statistical  or  computational-dynamic.    Statistical   models  are
generally based  on a  statistical analysis  of historical  air  quality data,
and are not  explicitly concerned with atmospheric  chemistry  or meteorology.
An example of empirical models is the linear rollback concept.
     Computational, or dynamic,  models attempt  to describe mathematically the
atmospheric  chemical  and physical  processes influencing air pollution forma-
tion  and  impacts.   Examples of  computational models  include  trajectory and
fixed-grid airshed models.   Two phenomenologically different approaches have
been employed in dynamic models with respect to the coordinate systems chosen.
A coordinate system fixed  with respect to the earth is  termed Eulerian,  while
in  Lagrangian models  the  reference frame moves  with  the air parcel  whose
behavior  is  being simulated.
3.7.6.1   Trajectory Models.   In trajectory  models,  a moving-coordinate  system
describes pollutant transport  as influenced  by  local meteorological conditions.
Trajectory models  provide  dynamic descriptions  of atmospheric source-receptor
relationships that are simpler and less  expensive to derive than  those obtained
from fixed-cell models.
                                   3-140

-------
     The simplest form  of  trajectory model is the empirical kinetic modeling
approach  (EKMA).    This approach  was  developed  from earlier  efforts
(Dimitriades, 1972)  to use smog chamber data to develop graphical relationships
between- morning NMOC  and  NO  levels and afternoon maximum  concentrations of
                            /\    .               •     i
ozone.   In applying EKMA,  the Ozone Isopleth Plotting Package (OZIPP) (Whitten
and Hogo,  1978) is used to generate ozone isopleths  at  various levels of
sophistication corresponding to  "standard"  EKMA,  "city-specific"  EKMA,  or the
simplified trajectory model (F.R., 1979).   The EKMA isopleths generated are used
to determine the  relative  degree of control of precursor emissions needed to
achieve a given percentage reduction in ozone.
     The use of EKMA in ozone abatement programs is relatively widespread.   It
is therefore worth  noting  the  general  control  implications  of  EKMA  isopleths.
For areas  with high  levels  of morning precursor emissions and  meteorology
conducive to oxidant formation, such as Los Angeles, for example, EKMA  isopleths
predict that (1) at high NMQC/NQ  concentration ratios, reductions in NO  will
                                y\                                       J\
decrease ozone  formation;  (2)  at moderate  NMOC/NO  ratios,  reductions  in NMOC
                                                  s\.
and NO  will  decrease ozone formation; and  (3)  at  very low NMOC/NO  ratios,
      /\                                                            y\
increases in NO   will inhibit ozone formation.   These predictions cannot be
                /\
assumed to  apply  to all urban  areas, or even  to  all  high-oxidant urban  areas,
since the shape of  the  EKMA isopleths  is a function  of numerous  factors, many
of which are location-specific.   For discussions of the specific assumptions
employed in  EKMA  and  the underlying chemistry and  meteorology,  the primary
literature should be consulted (e.g., Dimitriades, 1970, 1972, 1977a,b; Dodge,
1977a,b; Whitten  and  Hogo, 1977; U.S.   Environmental Protection Agency, 1977,
1978a; Whitten, 1983).   Likewise,  the  primary literature should be consulted
for additional  data and discussions on the respective  effects  on ozone  forma-
tion of controlling NMHC and NO  (e.g., Liu and Grisinger,  1981; Chock  et al.,
1981; Kelly, 1985;  Kelly et al., 1986;  Glasson and Tuesday, 1970; Dimitriades,
1970, 1972, 1977a,b).
3.7.6.2  Fixed-Grid Models.   Fixed-grid models,  also  called regional airshed
models, are  based on  two- or  three-dimensional  arrays of grid cells and are
the  most sophisticated  source-receptor models  presently  available.    Such
models are computationally complex  and  require the most extensive set of input
data; but they also provide the most realistic treatment of the  various processes
involved in photochemical air  pollution formation.
                                   3-141

-------
3.7.6.3   Box Models.   Box models (Hanna, 1973;  Demerjian  and Schere, 1979;
Derwent and Hov,  1980) are the  simplest  of  dynamic models.   They treat the
atmosphere as  a  single cell, bounded by  the mixing layer, having  an  area on
the order of 100 square miles.
3.7.6.4  Validation and Sensitivity Analyses for Dynamic Models.  All present-
ly available source-receptor models require a degree  of simplifying  assump-
tions to deal with practical limitations imposed by existing computer capabil-
ities, time and  cost  constraints, or  lack of knowledge concerning  inputs such
as boundary  conditions,  emissions,  or detailed  reaction  mechanisms.  The
reliability and  applicability  of any particular model therefore depends upon
its specific limitations, data  requirements, and degree of validation against
experimental  data from ambient air measurements or environmental chamber runs.
Reliability and  applicability also  depend   on the  quality of the chemical
kinetics mechanisms used  to define the 0,-HC-NO  relationship.
                                        O      X
     Attempts  are made to validate model predictions  by comparing them with
real  observations; and operating parameters  are  often  varied  to determine the
sensitivity of the model  to respective parameter changes  (Gelinas and Vajk,
1979).   In addition,  the extent of agreement  between the results from two
simulations can  be tested.   In this way, completely different  models may be
compared, or an  internal  component, such  as  the  chemical kinetics  mechanisms,
may be substituted and the model run  again  to ascertain the effect of such
substitutions.
3.8 REFERENCES
Aikin, A. C.; Herman, J. R.; Maier, E. J. R.; McQuillan, C. J. (1983) Influence
     of peroxyacetyl nitrate (PAN) on odd nitrogen in the troposphere and lower
     stratosphere. Planet.  Space Sci. 31: 1075-1082.
Akimoto, H.; Bandow, H.; Sakamaki, F.; Inoue, G.; Hoshino, M.; Okuda, M. (1980)
     Photooxidation of  the  propylene-NQ -air system: studies  by long-path Fourier
     transform  infrared spectrometry. tnviron.  Sci.  Techno!.  14:  172-179.
Allwine, G.;  Lamb,  B.; Westberg, H.  (1983) Application of atmospheric tracer
     techniques for determining biogenic hydrocarbon fluxes from an oak tracer
     In: Proceedings of forest environmental measurements conference; October;
     Oak Ridge, TN.
                                   3-142

-------
Altshuller, A.  P.  (1975) Evaluation of  oxidant  results at CAMP sites in the
     United States. J. Air Pollut. Control Assoc. 25: 19-24.

Altshuller, A. P. (1983) Review: natural volatile organic  substances  and  their
     effect on air quality in the United States. Atmos.  Environ.  17:  2131-2165.

Altshuller, A. P. (1986) Review paper: 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. Environment. 20:  245-268.

Altshuller, A.  P.;  Bufalini,  J.  J.  (1971)  Photochemical aspects of air pollu-
     tion: a review.  Environ. Sci. Techno!. 5: 39-64.

Altshuller, A. P.; McPherson, S. P.  (1963) Spectrpphotometric  analysis of alde-
     hydes in the  Los  Angeles atmosphere.  J.   Air  Pollut.  Control Assoc.
     13: 109-111.

Ames, J.;  Myers, T. C.; Reid,  L.  E.; Whitney, D.  C.; Golding, S. H.; Hayes,
     S. R.; Reynolds, S. D. (1978) The user's  manual  for the SAI  airshed  model.
     San Raphael, CA: Systems Applications, Inc. No.  EM78-89.

Anbar, M.; Taube, H.  (1954) Interaction  of nitrous  acid with hydrogen peroxide
     and with water.  J. Am. Chem. Soc. 76: 6243-6247.

Anderson,  G.  E.; Hayes,  S. R.; Hillyer, M, J.;  Killus, J.  P.; Mundkur, P.  V.
     (1977) ,Air  quality  in the Denver metropolitan region: .1974-2000. Denver,
     CO:  Region  VIII, U.S.  Environmental Protection Agency. EPA-908/1-77-002.

Andrews,  A. W.;  Thibault,  L.  H.;  Lijinsky,  W.  (1978) The relationship between
     mutagenicity  and carcinogenicity of some nitrosamines. Mutat.  Res.  51:
     319-326.

Appel,  B.  R. ; Kothny, E.  L.;  Hoffer,  E. M.; Wesolowski, J.  J. (1980) Sulfate
     and  nitrate data from the California Aerosol  Characterization Experiment
     (ACHEX).  In:  Hidy,  G.M.; Mueller,  P.  R.;  Grosjean,  D.;  Appel,  B. R.;
     Wesolowski,  J.  J.,  eds.  The character  and  origins of  smog  aerosols:  a
     digest of results from the California Aerosol  Characterization Experiment
     (ACHEX). New York,  NY: John Wiley and Sons; pp.  315-335.  (Advances in en-
     vironmental  science and  technology: v. 9).

Arnts,  R.  R.; Gay, B. W.,  Jr. (1979) Photochemistry of  some naturally emitted
     hydrocarbons.  Research  Triangle Park, NC:  U.S.  Environmental  Protection
     Agency,  Office of Research and  Development; EPA report no.  EPA-600/3-79-081.
     Available from:  NTIS, Springfield,  VA; PB80-131980.

Arnts,  R.  R.; Meeks, S, A.  (1981) Biogenic hydrocarbons:  contribution to the
     ambient  air in selected  areas.  Atmos.  Environ.  15: 1643-1651.

Arnts,  R.  R.; Seila,  R.  L.;  Kuntz, R.  L.; Mowry, F.  L.; Knoerr, K.  R.; Dudgeon,
     A.  C. (1978)  Measurement of  alpha-pinene  fluxes from a loblolly pine forest.
     In:  Proceedings  of the  4th  joint conference  on sensing of environmental
     pollutants,  American  Chemical  Society.  Washington, DC: American Chemical
     Society.
                                    3-143

-------
Arnts, R.  R.;  Gay,  B.  W.;  Bufalini,  J.  J.  (1981) Photochemical  oxidant poten-
     tial  of the biogenic hydrocarbons. In:  Bufalini,  J.  J.; Arnts,  R.  R.,
     eds.  Atmospheric  biogenic hydrocarbons, v. II. Ann Arbor, MI: Ann Arbor
     Science Publishers; p. 117.

Ashby, H.  A.;  Stahman, R. C.;  Eccleston,  B. H.; Hum, R. W. (1974)  Vehicle
     emissions—summer to winter.  In: Society of Automotive  Engineers meeting;
     October; Toronto, Canada. Warrendale,  PA:  Society  of Automotive  Engineers;
     SAE technical paper 741053.

Association  of Bay Area  Governments. (1979a)  Application  of photochemical
     models, Vol.  I:  The use  of photochemical  models in urban ozone  studies.
     Research Triangle  Park,  NC:  U.   S.  Environmental  Protection Agency;  EPA
     report no.  EPA-450/4-79-025.

Association  of Bay Area  Governments. (1979b)  Application  of photochemical
     models, Vol.  Ill:  Recent sensitivity tests and other applications of the
     LIRAQ model. U. S. Environmental Protection Agency, Research  Triangle  Park,
     NC; EPA report no. EPA-450/4-79-027.

Atkinson,  R.  (1985) Kinetics  and  mechanisms of the gas phase  reactions  of
     the hydroxyl radical with organic compounds under  atmospheric conditions.
     Chem. Rev.: in press.

Atkinson,  R.;  Aschmann,  S.  M. (1984) Rate constants for the reactions of 03
     and OH radicals with a series of alkynes.  Int.  J.  Chem.  Kinet. 16: 259-268.

Atkinson,  R.; Carter, W. P. L. (1984) Kinetics  and  mechanisms of the  reactions
     of ozone with organic compounds  in  the gas phase.  Chem.  Rev.  84:  437-470.

Atkinson,  R.;  Lloyd,  A.  C. (1981) Smog  chemistry and urban  airshed modeling.
     In:  Rodgers,  M.  A.  J.;  Powers,  E.  L., eds. Oxygen and oxy-radicals in
     chemistry and biology. New York; Academic  Press, pp. 559-592.

Atkinson,  R.;  Lloyd,  A.  C. (1984) Evaluation of kinetic and mechanistic  data
     for modeling of photochemical smog. J.  Phys. Chem. Ref.  Data  13:  315-444.

Atkinson,  R.;  Darnall, K.  R.; Pitts, J.  N., Jr. (1978)  Rate constants for the
     reactions of  OH  radicals and ozone with cresols at 300 ± 1 K. J. Phys.
     Chem. 82: 2759-2761.

Atkinson,  R.;  Darnall, K. R.; Lloyd, A. C.  ; Winer, A. M.;  Pitts, J.  N., Jr.
     (1979)  Kinetics  and mechanisms of  the reactions of the hydroxyl radical
     with  organic  compounds  in the gas  phase. In:  Pitts,  J.  N.,  Jr.;  Hammond,
     G.  S.;  Gollnick,  K.; Grosjean,  D.,  eds. Advances in photochemistry, v.
     11. New York, NY: John Wiley  and Sons.  pp.  375-488.

Atkinson,  R.;  Carter,  W.  P.  L.;  Darnall,  K. R.; Winer, A.  M.;  Pitts, J.  N.,
     Jr.  (1980)  A smog chamber and  modeling study of  the  gas phase  NO -air
     photooxidation of toluene and the  cresols.  Int. J. Chem. Kinet.  12:  779-836.

Atkinson,  R.;  Aschmann,  S. M.; Winer,  A.  M.;  Pitts, J. N., Jr.  (1981) Rate
     constants for the gas-phase  reactions of 03 with a series of  carbonyls at
     296 K.  Int. J. Chem.  Kinet.  13:  1133-1142.


                                   3-144

-------
Atkinson, R.;  Aschmann,  S.  M.; Carter, W. P.  L.; Winer, A.  M. ;  Pitts,  J.  H. ,
     Jr. (1982a)  Alky!  nitrate formation from the NO  -air photooxidations of
     C2-C8 n-alkanes. J. Phys. Chem. 86: 4563-4569.

Atkinson, R.;  Aschmann,  S.  M. ; Fitz,  D.  R.;  Winer,  A, M.;  Pitts, J. N. , Jr.
     (1982b) Rate  constants  for the gas phase reactions of 03 with selected
     organics at 296 K. Int. J. Chem.  Kinet.  14: 13-18.

Atkinson, R.;  Lloyd,  A.  C.; Winges, L. (1982c)  An updated chemical  mechanism
     for hydrocarbon/NO /SO  photooxidation suitable for  inclusion  in  atmospheric
     simulation models. Atmos. Environ. 16: 1341-1355.

Atkinson, R.;  Aschmann,  S.  M.; Carter, W. P.  L.; Winer,  A.  M.;  Pitts,  J.  N.,
     Jr. (1982d) Kinetics of the gas-phase reactions of OH radicals  with alkyl
     nitrates at 299 ± 2 K. Int. J. Chem. Kinet. 14: 919-926.

Atkinson, R.;  Carter,  W.  P.  L.; Winer,  A.  M. (1983)  Effects  of temperature
     and pressure on  alkyl  nitrate yields  in  the NO  photooxidations of
     n-pentane and n-heptane. J. Phys. Chem.  87: 2012-2018.

Atkinson, R.;  Plum,  C. N.; Carter, W. P. L.;  Winer, A. M.;  Pitts,  J.  N.,  Jr.
     (1984a) Rate constants for the gas-phase reactions of N03 radicals with a
     series  of organics in air at  298 ± 1 K. J. Phys. Chem.  88:  1210-1215.

Atkinson,  R.;  Pitts, J. N.,  Jr.;  Aschmann,  S.  M. (1984b)  Tropospheric re-
     actions of  dimethyl  sulfide with N03 and OH radicals.  J.  Phys. Chem. 88:
     1584-1587.

Atkinson, R,; Aschmann, S. M.; Winer,  A. M.;  Pitts, J. N., Jr.  (1984c)
     Kinetics  of the gas-phase  reactions  of N03 radicals with  a series  of
     dialkenes,  cycloalkenes,  and monoterpenes  at 295 ± 1  K.  Environ.  Sci.
     Technol.  18: 370-375.

Atkinson, R.;  Carter,  W.  P.  L.; Plum, C. N.;  Winer, A.  M.;  Pitts,  J.  N.,  Jr.
     (1984d) Kinetics  of the  gas-phase reactions of N03 radicals with  a series
     of aromatics at 296 ±  2  K.  Int. J. Chem.  Kinet. 16:  887-898.

Atkinson, R.;  Plum,  C. N. ; Carter, W. P.  L. ;  Winer, A.  M.;  Pitts,  J.  N.,  Jr.
     (1984e) Kinetics  of the  gas phase reactions of N03 radicals with  a series
     of alkanes  at 296 ± 2  K.  J. Phys. Chem.  88: 2361-2364.

Atkinson, R.;  Aschmann, S.  M.;  Carter, W.  P.  L.; Winer,  A.  M.;  Pitts, J.  N.,
     Jr. (1984f)  Formation  of alkyl  nitrates  from the  reaction of branched and
     cyclic  alky peroxy radicals with NO.  Int. J.  Chem. Kinet.  16:  1085-1101.

Atkinson, R.;  Aschmann, S.  M.;  Winer,  A.  M.;  Pitts,  J. N.,  Jr.  (1985)  Kinetics
     and atmospheric implications of  the  gas-phase  reactions  of N03 radicals
     with a  series of  monoterpenes  and related organics at 294 ± 2 K.  Environ.
     Sci. Technol. 19:  159-163.        .

Attmannspacher,  W.;  Hartmannsgruber,  R.  (1973)  On extremely  high  values of
     ozone  near  the  ground.  Pure Appl. Geophys.  106-108:  1091-1096.
                                    3-145

-------
Automotive News.  (1982a)  Sales of diesel-powered  cars  in the U.S. In: 1982
     market data book issue. April 28: 74.

Automotive News.  (1982b)  Automotive news analysis  of  new-car registrations.
     In: 1982 market data book issue. April 28: 8.

Automotive News.  (1983a)  Sales of  new  diesel-powered  cars in the  U.S.  In:
     1983 market data book  issue. April 27: 70.

Automotive News.  (1983b)  Domestic  car  sales.  1982 market data  book  issue.

Bach, W.  D.,  Jr.  (1975) Investigation of ozone and ozone precursor concentra-
     tions in nonurban locations in the eastern United  States; phase  II. meteo-
     rological analyses. Research Triangle Park, NC: U.S.  Environmental Protec-
     tion Agency, Office of Air Quality Planning and Standards;  EPA report no.
     EPA-450/3-74-034a. Available  from  NTIS,  Springfield,  VA;   PB-246899.

Baldwin,  A.  C.;  Barker,  J.  R.; Golden,  D.  M.; Hendry,  D. G.  (1977)  Photo-
     chemical  smog.  Rate parameter  estimates  and computer simulations.  J.
     Phys. Chem. 81: 2483-2492.

Bandow, H.; Okuda, M.; Akimoto, H.  (1980) Mechanism of  the gas-phase  reactions
     of C3H6 and N03 radicals. J. Phys. Chem. 84:  3604-3608.

Barth,  D. S.  (1970) Federal motor  vehicle  emission goals for CO, HC and NO
     based on  desired  air  quality  levels. J.  Air Pollut. Contr.  Assoc. 20:
     519-523.

Bartok, W.;  Crawford,  A.  R.;  Piegari, J.  (1971) Systematic field study of NO
     emission control methods  for utility boilers.  Research Triangle  Park, N(T:
     U.S. Environmental Protection  Agency; APTD-1163.   Available from:  NTIS,
     Springfield, VA; PB-210739.

Bass, A.  M.;  Glasgow,  L.  C.;  Miller, C.; Jesson,  J. P.;  Filkin, D. L.  (1980)
     Temperature  dependent  absorption cross  sections  for formaldehyde (CHgO):
     the  effect  of formaldehyde on  stratospheric  chlorine chemistry.  Planet.
     Space Sci. 28: 675-679.

Batt,  L.  (1979) Gas phase  decomposition  of alkoxy radicals.  Int.  J.  Chem.
     Kinet. 11: 977-993.

Bird, R.  B.;  Stewart,  W.  E.;  Lightfoot,  E.  N. (1960)  Concentration  distri-
     bution  in  turbulent  flow. In:   Transport phenomena.  New  York:  John Wiley
     and  Sons, pp. 626-635.

Bell, G.  B. (1960) Meteorological conditions during oxidant episodes  in
     coastal  San  Diego County in October and  November, 1959.  Sacramento, CA:
     Department of Public Health.

Bjerknes, J.. (1951) Extratropical  cyclones.  In:  Compendium  of  meteorology.
     Boston, MA: American Meteorological  Society;  pp.  557-598.
                                    3-146

-------
Black, F. M.  (1977)  The impact of  emissions  control  technology on passenger
     car  hydrocarbon  emisson rates  and patterns.  In: Dimitriades,  B.,  ed.
     International  conference on photochemical  oxidant  pollution and  its
     control;  September 1976; > Raleigh, NC.  Proceedings:  v.  II.  Research
     Triangle  Park, NC:  U.S.  Environmental Protection Agency;  pp.  1053-1068;
     EPA  report  no.  EPA-600/3-77-QQlb.  Available from:  NTIS,  Springfield,
     VA; PB-264233.

Black, F. M.; Bradow, R. L. (1975) Patterns of hydrocarbon emissions from  1975
     production cars.  Presented at:  Society of Automotive Engineers'  national
     fuels  and lubricants  meeting;  June;  Houston,  TX.  Warrendale, PA:  Society
     of Automotive Engineers; SAE paper no. 750681.

Black, F. M.; High, L. E.  (1979) Diesel hydrocarbon emissions,  particulate and
     gas phase. Presented  at: Society  of  Automotive  Engineers' symposium on
     diesel particulate emissions measurement characterization;  February;  Detroit,
     MI. Warrendale, PA: Society of Automotive Engineers.

Black, F. M.; High, L. E.; Lang, J. M.  (1980) Composition of  automobile  evapo-
     rative  and tailpipe hydrocarbon emissions.  J.  Air Pollut.  Control Assoc.
     30: 1216-1221.

Bluraenthal,  D. L.;  Richards, L. W.; Macias,  E.  S.; Bergstrom,  R.  W.;  Wilson,
     W.  E.;  Bhardwaja,  P.  S. (1981)  Effects  of  a coal-fired power plant and
     other  sources on southwestern visibility  (interim summary  of EPA's  project
     VISTTA). Atmos. Environ. 15: 1955-1969.

Bellinger,  M.  J.;  Parrish, D.  D.; Hahn, C.;  Albritton,  D. L.;  Fehsenfeld, F.
     C.  (1982) NO  measurements in  clean  continental  air. In:  Proceedings of
     2nd  symposium  on composition of the nonurban troposphere;  May; Williams-
     burg,  VA.  (see also,  Bellinger, M.,  et al. (1984)  NO   measurements  in
     clean  continental  air and analysis of the  contributing meteorology.  J.
     Geophys.  Res. 89: 9623-9631).

Box,  E.  0.  (1981) Foliar  biomass:  data base of  the International Biological
     Program and other sources. In:  Bufalini, J. J.;  Arnts,  R.  R., eds.  Atmos-
     pheric biogenic  hydrocarbons, v.  1.  Ann Arbor,  MI: Ann Arbor Science;
     pp.  107-119.

Braddock, J.  N.; Bradow, R.  L.  (1976) Emission patterns  of diesel-powered  pas-
     senger cars.  Presented  at meeting of Society of Automotive Engineers,
     Houston,  TX, June 3-5,  1975. SAE Paper No.  750682.

Braverman,  T.  N.;  Layland, D.  E.  (1982) Application  of  urban emission inven-
     tories to photochemical modeling.  Proceedings of Air Pollution Control
     Association  Specialty Conference on Emission Inventories and Air Quality
     Management,  Kansas  City, MO, April.

Breeding,  R.  J.;  Lodge,; J.  P., Jr.;  Pate,  J.  B.;  Sheesley,  D.  C.; Klonis, H.
     B.;  Fogle, B.; Anderson, J.  A.;  Englert,  T.  R.;  Haagenson, P. L.; McBeth,
     R.  B.; Morris, A.  L.; Pogue, R.; Wartburg,  A. F. (1973) Background trace
     gas  concentrations  in the  central  United States.  JGR J.  Geophys.  Res. 78:
     7057-7064.
                                    3-147

-------
Brock, J. R.; Durham, J. L. (1984) Aqueous aerosol as a reactant  i  the atmos-
     phere. In: Calvert, J. G., ed. S02, NO and N02 oxidation mechanisms:  atmos-
     pheric considerations, Boston,  MA:  Butterworth Publishers;  pp. 209-250.
     (Teasley, J. I., ed. Acid precipitation series, v. 3).

Bruckmann,  P. W.;  Winner, H. (1983) Infrared spectroscopic study of peroxy-
     acetyl nitrate (PAN) and its decomposition products. Environ. Sci, Techno!.
     17: 352-357.

Bucon, H. W.; Macko, J. F.; Taback, H. J. (1978) Volatile organic compound (VOC)
     species  data  manual.   Prepared  by KVB  Engineering,  Inc.,  Tustin, CA.
     Research Triangle Park, NC: U.S. Environmental Protection Agency; EPA report
     no. EPA-450/3-78-119. Available from: NTIS, Springfield, VA; PB80-113822.

Bufalini, J.  J.; Arnts,  R.  R.,  eds.  (1981)  Atmospheric  biogenic  hydrocarbons,
     Vols. 1 and 2. Ann Arbor, MI; Ann Arbor Science,

Bufalini, J. J.; Lonneman, W.  A., eds. (1977) Proceedings of symposium on  1975
     northeast oxidant transport study. Research Triangle Park, NC: U.S. Environ-
     mental Protection Agency. Report no. EPA-600/3-77-017.

Burns,  R.  C.;  Hardy,  R.  W.  F.  (1975)  Nitrogen  fixation  in bacteria and higher
     plants. New York: Springer-Verlag.

California  Bureau  of Air Sanitation.  (1966) The  oxides of nitrogen  in  air
     pollution. Sacramento, CA: California Department of  Public Health.

Calvert, J. G.;  Mohnen,  V. (1983) The  chemistry  of acid formation.  In:  Acid
     deposition: atmospheric  processes in eastern North America. Washington,
     DC: National Academy Press.

Calvert, J. G.;  Stockwell,  W.  R.  (1983)  Acid generation in the  troposphere by
     gas-phase chemistry. Environ. Sci. Technol. 17: 428A-443A.

Calvert, J. G,; Stockwell, W.  R. (1984) The mechanism and rates of the gas-phase
     oxidation of  sulfur dioxide and nitrogen oxides  in the atmosphere.  In:
     Calvert, J. G., ed., Acid precipitation: S02, NO and N02 oxidation mechan-
     isms:  atmospheric considerations.  Boston,  MA:  Butterworth  Publishers.
     (Teasley, J. I., ed., Acid precipitation series v. 3)

Calvert, J. G.;  Su,  F.;, Bottenheim,  J.  W.;  Strausz,  0.  P.  (1978)  Mechanism of
     the homogeneous  oxidation of sulfur dioxide  in  the troposphere.  Atmos.
     Environ. 12: 197-226.

Carter, W.  P.  L;  Carnal 1, K.  R.;  Lloyd,  A.  C.;  Winer, A.  M.;  Pitts,  J.  N.,
     Jr.  (1976)  Evidence for  alkoxyl  radical isomerization  in photooxidation
     of C4~C6 alkanes under simulated atmospheric  conditions. Chem. Phys,  Lett.
     42: 22-27.                                 ,

Carter, W.  P. L.; Lloyd, A. C.; Sprung, J.  L.; Pitts, J.  N., Jr.  (1979) Computer
     modeling of smog chamber data:  progress in validation of detailed mecha-
     nism for the photooxidation of propene and _n-butane  in photochemical  smog.
     Int. J. Chem.  Kinet. 11:  45-101.
                                    3-148

-------
Carter, W. P. L.; Winer, A. M.; Darnall, K. R.;
     chamber studies  of temperature effects  in
     Sci, Techno!. 13: 1094-1100.
                Pitts, J. N.,
                photochemical
Jr. (1979b) Smog
 smog.  Environ.
Carter, W.  P.  L.;  Winer, A. M.;  Pitts,
     sink  for  phenol and  the cresols:
     Environ. Sci.  Techno!. 15:. 829-831.
         J.  N.,  Jr.  (1981a)  Major atmospheric
         reaction  with the nitrate radical.
Carter, W.  P.  L.;  Tuazon, E. C.; Winer, A. M.;  Pitts, J.  N.,  Jr.  (1981b)  Gas
     phase  reactions  of N,N-dimethylhydrazine with ozone and NO  in simulated '
     atmospheres.  In:  Scanlan,  R,  A.; Tannenbaum,  S.  R.,  eds.  tr-nitroso com-
     pounds. Washington, DC: American Chemical Society;  pp.  118-131.  (American
     Chemical Society symposium  series no.  174).
Carter, W. P. L.; Winer, A. M.;
     nitrate on  the initiation
     15: 831-834.
Pitts, J. N., Jr. (1981c) Effect of peroxyacetyl
of photochemical smog. Environ. Sci. Techno!.
Carter, W.  P.  L.;  Winer, A. M.;  Pitts,  J.  N.,  Jr.  (1982) Effects of kinetic
     mechanisms and hydrocarbon composition  on  oxidant-precursor relationships
     predicted by  the EKMA isopleth technique.  Atmos.  Environ. 16: 113-120.

Cato, G. A.; Buening, H. J.; DeVivo, C.  C.;  Morton,  B.  G.;  Robinson,  J.  M.  (1974)
     Field testing: Application of combustion modifications to  control  pollutant
     emissions from  industrial  boilers—phase  I. Research  Triangle Park,  NC:
     U.S. Environmental  Protection Agency;  EPA report  no.  EPA-650/2-74-078a.
     Available from:  NTIS,  Springfield,  VA;  PB-238920.

Cato, G.  A.;  Muzio,  L.  J.;  Shore, D.  E.  (1976)  Field testing:  Application of
     combustion modifications  to  control pollutant  emissions from industrial
     boilers—phase II.  Research  Triangle  Park, NC:  U.S.  Environmental  Protec-
     tion Agency;  EPA report no.  600/2-76-086a.

ChameideSj W. L.;  Davis, D. D.  (1982)  The  free  radical  chemistry of cloud  drop-
     lets and  its  impact upon  the  composition  of rain. JGR J. Geophys. Res.
     87: 4863-4877.

Chameides, W. L.;  Walker, J. C. G.  (1976)  A  photochemical  theory of tropospheric
     ozone. J. Geophys.  Res. 78:  8751-8760.

Chameides,  W.  L:;  Stedman, D.  H.;  Dickerson,  R. R.; Rusch, D. W.; Cicerone,
     R.  J.  (1977)  NO  production  in  lightning. J.   Atmos.  Sci.  34: 143-149.
                     "
Chameides,  W.L.;   Tan,  A.  (1981). The two-demensional  diagnostic model for
     tropospheric  OH: An uncertainty analysis.  J. Geophys.  Res. 86: 5209-5223.

Chatfield,  R.  B.;  Rasmussen,  R,  A.  (1977)  Hydrocarbons,  f1uorocarbons, and
     oxidants  observed  at a rural  site  near St. Louis. In: Proceedings from
     national AIChE meeting; March; Houston, TX.

Chock,  D.  P.;  Dunker, A. M.;  Kumar,  S.; Sloane, C.  S. (1981)  Effect of NO
     emission  rates on  smog formation  in the California South Coast Air Basin.
     Environ. Sci. Techno!.  15: 933-939.
                                    3-149

-------
Clark, T. L.; Clarke, J. F. (1982) Boundary layer transport of NO  and 03 from
     Baltimore, Maryland—a case  study.  Presented at: 75th annual meeting of
     the Air Pollution Control Association; June; New Orleans, LA. Pittsburgh,
     PA: Air Pollution Control Association; paper no. 82-24.3.

Clarke, J. F.; Ching, J. K. S.; Brown, R. M.; Westberg, J.; White, J. H. (1982)
     Regional transport of ozone.  Presented at: Third conference on air pollu-
     tion meteorology;  January.  Boston,  MA: American Meteorological Society;
     paper no. 1.1.

Cleveland, W. S.; Guarino, R.; Kleiner, B.; McRae, J. E.; Warner, J. L. (1976a)
     The analysis  of the ozone problem  in  the northeast United States. In:
     Specialty conference on  ozone/oxidants—interactions with the total environ-
     ment.  Pittsburgh,  PA: Air Pollution Control  Association;  pp.  109-120.

Cleveland, W. S.; Kleiner, B.; McRae, J. E.; Warner, J. L. (1976b) Photochemical
     air pollution:  transport from New York City  area into Connecticut and
     Massachusetts. Science (Washington, DC) 191: 179-181.

Coffey, P. E. (1977) The issue of natural organic emissions.  In: International
     conference on oxidants,  1976—analysis of evidence and viewpoints: part IV.
     Research Triangle Park,  NC: U.S. Environmental  Protection Agency; EPA report
     no. EPA-600/3-77-116.  Available from:  NTIS, Springfield, VA; PB-277462.

Coffey, P. E.; Stasiuk, W. N.  (1975) Evidence of atmospheric  transport of ozone
     into urban areas. Environ. Sci. Techno!. 9: 59-62.

Coffey, P. E,; Stasiuk, W. N.; Mohnen, V. (1977) Ozone in  rural and urban areas
     of New  York  State.  In:  Dimitriades, B., ed. International conference on
     photochemical oxidant pollution and its control; September 1976; Raleigh,
     NC. Proceedings: volume  I. Research Triangle Park, NC: U.S. Environmental
     Protection Agency, Office of Research and Development; pp. 89-96; EPA report
     no. EPA-600/3-77-001a. Available  from:  NTIS,  Springfield,  VA;  PB-264232.

Cole,  H.  S.;  Newberry,  C.  F.; Cox,  W.;  Moss,  G.  K.;  Layland,  D.  (1982)  Appli-
     cation of  the airshed model  for ozone control  in St. Louis. Proceedings
     of the  75th  Annual  Meeting of the Air  Pollution Control  Association.  New
     Orleans, LA, June 20-25.

Council  for  Agricultural  Science and Technology.  (1976)  Effect of  increased
     nitrogen fixation on  stratospheric ozone. CAST  report no. 53.

Countess, R. J.; Wolff, G. T.; Whitbeck, M. R. (1981) The  effect of temperature
     on ozone formation in the propene/nitrogen dioxide/air system. J. Environ.
     Sci. Health. A16: 1-8.

Cox, R. A.; Burrows, J. P.  (1979) Kinetics  and mechanism of the disproportion-
     ation of H02  in the gas  phase.  J. Phys. Chem. 83: 2560-2568.

Cox,  R.  A.;  Penkett, S. A. (1972)  Aerosol  formation from sulfur dioxide  in
     the presence  of ozone and olefinic  hydrocarbons.  J.  Chem.  Soc.  Faraday
     Trans. I 68: 1735-1753.
                                    3-150

-------
Cox, R. A.; Roffey, M. J. (1977) Thermal decomposition of peroxyacetyl nitrate
     in the  presence of  nitric  oxide.  Environ. Sci. Technol.  11:  900-906.

Cox, R.. A.;  Sheppard, D.  (1980)  Reactions  of OH radicals with gaseous sulfur-
     compounds. Nature 284: 330-331.

Cox, R. A.;  Derwent,  R.  G. ; Kearsey,  S. V.; Batt,  L.;  Patrick, K.  G. (1980)
     Photolysis of  methyl  nitrite:  kinetics  of  the  reaction of the  methoxy
     radical with 02. J.  Photochem. 13: 149-163.

Crawford, A. R. ; Manny, E. H. ; Bartok, W.  (1974) Field testing:  application  of
     combustion modifications  to control  NO  emissions  from utility  boilers.
     Research Triangle Park, NC: U.S. Environmental  Protection  Agency; EPA report
     no.  EPA-650/2-74-066. Available  from:  NTIS, Springfield,  VA;  PB-237344.

Crawford, A. R. ; et al. (1978) Control of  utility boiler and gas turbine  pollu-
     tant emissions by combustion modification-phase I.  Research Triangle Park,
     NC:  U.S. Environmental Protection Agency;  EPA report no.  EPA-600/7-78-036a.
     Available from NTIS PB, Springfield,  VA; 281-078. Cited in Hall  and  Bowen
     (1982).

Cronn, D. R. (1982) Monitoring of ambient  halocarbons and natural hydrocarbons
     in the  Smoky  Mountains.  In: Proceedings from  2nd symposium on composi-
     tion of the nonurban troposphere; May;  Williamsburg, VA.

Crutzen,  P.  J.  (1982) The global  distribution  of hydroxyl.  In: Goldberg, E.
     D., ed. Atmospheric chemistry. New York, NY: Springer-Verlag;  pp. 313-328.

Crutzen,  P.  J.; Ehhalt,  D.  H.  (1977)  Effects  of nitrogen  fertilizers and
     combustion in  the stratospheric  ozone layer. Ambio.  6:  112-117.

Crutzen, P. J.; Fishman, J. (1977)  Average concentrations of OH in  the tropos-
     phere,  and  the budgets of  CH4,  CO, H2,  and CH3CC13. Geophys.  Res. Lett.
     4: 321-324.

Daniel sen,  E.  F.  (1968)  Stratospheric-tropospheric  exchange based on radioac-
     tivity, ozone  and potential vorticity.  J.  Atmos. Sci. 25:  502-518.

Daniel sen,  E.  F.  (1980)  Stratospheric source for unexpectedly large values of
     ozone  measured over the Pacific Ocean during  Gametag,  August 1977.  JGR
     J. Geophys. Res. 85: 401-412.

Danielsen,  E.  F. ;  Mohnen,  V.  A.  (1977) Project duststorm report: ozone trans-
     port,  jjn situ measurements,  and meteorological analyses  of tropopause
     folding. JGR J.  Geophys.  Res.  82:  5867-5877.

Darnall,  K.  R. ;  Lloyd, A. C. ;  Winer, A.  M. ; Pitts, J.  N.,  Jr.  (1976)  Reac-
     tivity  scale for atmospheric  hydrocarbons  based on  reaction with hydroxyl
     radical.  Environ. Sci. Technol.  10: 692-696.

Darnall,  K.  R.; Atkinson,  R.;  Pitts,  J.  N., Jr.  (1979) Observation  of biacetyl
     from  the  reaction of OH  radicals  with o-xylene.  Evidence for  ring clea-
     vage.  J.  Phys.  Chem.  83:  1943-1946.
                                    3-151

-------
Dasgupta, P. K. (1980a) Discussion on "The importance of atmospheric ozone and
     hydrogen peroxide in oxidizing sulphur dioxide in cloud and rainwater [on
     Penkett et a!., 1979]." Atmos. Environ. 14: 272-275.

Dasgupta, P. K.  (1980b) The  importance  of atmospheric ozone and hydrogen per-
     oxide in oxidizing sulfur dioxide  in cloud and rainwater—further discussion
     Con Penkett'et al.5 1979]. Atmos.  Environ. 14: 620-621.

Davis, D. D.;  Heaps, W.;  Philen,  D.;  McGee, T.  (1979) Boundary layer measure-
     ments of the OH radical in the vicinity of an isolated power plant plume:
     S02 and N02 chemical conversion times.  Atmos.  Environ. 13: 1197-1203.

Davis, D. R.;  Jensen,  R.  E. (1976)  Low level  ozone and  weather systems.  In:
     Specialty conference on ozone/oxidants—interactions with  the total  environ-
     ment; March; Dallas, TX. Pittsburgh, PA: Air Pollution Control Association;
     pp. 242-251.                               -

Decker, C.  E.;  Ripperton,  L. A.; Worth,  J. J.  B. ; Vukovich,  F. M.;  Bach, W.
     D.; Tommerdahl, J. B.; Smith, F.;  Wagoner, D. E.  (1978) Study of the forma-
     tion and transport of ambient oxidants in the western Gulf Coast and north-
     central and northeast regions of  the  United States.  Research Triangle
     Park, NC: Research Triangle  Institute.

Delwiche, C. C. (1970) The nitrogen cycle. Sci. Amer.  223: 137-146.

Dement, W.  A.;  Tyson,  B.  J.; Mooney, H.  A.  (1975) Mechanism of monoterpene
     volatilization in Sal via mellifera.  Phytochemistry  14: 2555-2557.

Demerjian,  K.  L.;  Schere,  K. L.  (1979) Applications  of a photochemical  box
     model  for  03  air  quality  in  Houston, Texas,  Proceedings of Air  Pollution
     Control Association  Specialty Conference on  Ozone/Oxidants:  Interactions
     with the Total Environment,  pp. 329-352.

Demerjian,  K.  L.;  Kerr,  J.  A.; Calvert,  J. G. (1974) The mechanism of photo-
     chemical  smog formation.   In:   Pitts, J.  N., Jr.;  Metcalf,  R.  L.,  eds.
     Advances  in environmental  science  and technology:  v.  4.  New York,  NY:
     John Wiley and Sons, Inc.; pp. 1-262.

Demerjian,  K.  L.;  Schere,  K.  L.;  Peterson,  J.  T.  (1980)  Theoretical  estimates
     of actinic (spherically integrated)  flux and photolytic rate constants of
     atmospheric species  in the  lower  troposphere.  In:  Pitts, J. N.,  Jr.;
     Metcalf,  R. L.; Grosjean,  D.,  eds. Advances  in environmental  science and
     technology: volume  10.  New  York,  NY: John Wiley and Sons; pp.  369-459.

Derwent, R.  G.;  Hov,  0.  (1980) Computer modeling studies of  the  impact of
     vehicle exhaust emission  controls  on photochemical air pollution forma-
     tion in the United Kingdom.  Environ. Sci. Techno!.  14:  1360-1366.

Derwent, R. G.; Eggleton, A. E. J.; Williams, M.  L.;  Bell,  C. A. (1978)  Elevated
     ozone  levels  from natural sources.  Atmos.  Environ.  12:  2173-2177.

Dietzman, H. E.; Parness, H. A.;  Bradow,  R. L.  (1980)  Emissions from  trucks by
     chassis version 1983 transient procedures.  Presented at Society of Auto-
     motive  Engineer's Meeting. SAE Paper No. 801371.


                                   3-152

-------
Dietzman, H.  E.;  Parness, M. A.;  Bradow,  R.  L.  (1981)  Emissions  from gasoline
     and diesel delivery trucks by chassis transient cycle. ASME Energy Sources
     Technology Conference and Exhibition. January; Houston, TX.

Dignon, J.; Hameed, S. (1985) A model investigation of the impact of  increases
     in anthropogenic NO  emissions between 1967 and 1980 on tropospheric ozone.
     J. Atmos. Chem. 3: 491-506.

Dimitriades,  B,  (1970)  On  the function of hydrocarbon and nitrogen oxides in
     photochemical smog formation. Washington, DG: U. S. Department of Commerce,
     Bureau of Mines. Report of Investigations, RI 7433.

Dimitriadesj  B.  (1972)  Effects of hydrocarbon and  nitrogen  oxides  on photo-
     chemical smog formation. Environ. Set. Tech, 6: 253-260.

Dimitriades,  B. (1974) Proceedings of the solvent reactivity conference. Research
     Triangle Park,  NC:  U.S.  Environmental Protection Agency; EPA report no.
     EPA-650/3-74-010.  Available  from:  NTIS, Springfield,  VA;  PB-238296.

Dimitriades,  B. (1977a) An alternative to the Appendix-J method for calculating
     oxidant- and N02-related  control  requirements.. In:  International Confer-
     ence on  Photochemical Oxidant Pollution and Its Control. Proceedings, Vol.
     II.  Research Triangle  Park, NC: U.  S.  Environmental  Protection Agency.
     EPA-600/3-77-001b.                                         .-.-....

Dimitriades,  B.  (1977b) Oxidant  control  strategies.  Part  I. Urban oxidant-
     control  strategy  derived  from existing smog chamber data. Environ. Sci.
     Technol. 11: 80-88.

Dimitriades,  B.  (1981)  The  role of natural  organics in photochemical  air pol-
     lution:  issues  and research needs.  J.  Air  Pollut.  Control  Assoc. 31:
     229-235.                                            ..-•-.-

Dimitriades,  B.;  Dodge, M. C. s eds,  (1983) Proceedings of the Empirical Kinetic
     Modeling Approach  (EKMA) validation  workshop.  Research  Triangle  Park, NC:
     U.S. Environmental Protection Agency. Report no.  EPA-600/9-83-014.

Dodge, M. C.  (1977a) Combined use of modeling techniques and smog chamber data
     to  derive  ozone-precursor relationships.  In:  International  Conference on
     Photochemical  Oxidant  Pollution and Its Control.  Proceedings,  Vol.  II.
     Research Triangle Park,  NC: U.  S>  Enviornmental Protection  Agency.
     EPA-600/3-77-001b.

Dodge,  M.  C.  (1977b) Effect of selected parameters  on predictions of a photo-
     chemical model. Research Triangle  Park, NC: U. S. Environmental  Protection
     Agency.  EPA  600/3-77-048.

Dodge,  M.  C.; Arnts, R. R.  (1979) A new mechanism for the reaction of ozone
     with olefins.  Int. J. Chem.. Kinet. 11:  399-410.

Doyle, G. J.; Tuazon, E. C.; Graham, R, A.;  Mischke, T. M.;  Winer, A.  M.;  Pitts,
     J.  N.,  Jr.  (1979) Simultaneous  concentrations  of ammonia and nitric acid
     in  a polluted  atmosphere and their equilibrium relationship  to  particulate
     ammonium nitrate.  Environ.  Sci. Technol. 13: 1416-1419.


                                   3-153

-------
Drives, P. J.  (1977) TRACE (Trajectory  atmospheric  chemistry and emissions)
     user's guide.  Santa Monica,  CA:  Pacific Environmental  Services, Inc.  PES
     Public.  No. TP-016.

Duce, R.  A.;  Mohnen, V. A.; Zimmerman,  P.  R.;  Grosjean, D.;  Cautreels, W.;
     Chatfield, R.; Jaenicke, R.; Ogren, J.  A.; Pellizzari,  E, D.; Wallace,  G.
     T. (1983) Organic material in the global troposphere.  Rev. Geophys. Space
     Phys. 21: 921-952.

Dunker, A. M.  (1980) The response of an atmospheric  reaction-transport model
     to changes in  input function. Atmos. Environ. 14: 671-679.

Dunker, A. M.  (1981) Efficient calculation  of  sensitivity  coefficients for
     complex atmospheric models. Atmos.  Environ. 15:  1155-1161.

Durbin, P. A.;  Hecht, T. A.; Whitten,  G.  I. (1975)  Mathematical  modeling of
     simulated photochemical  smog.  Research  Triangle Park,  NC: U. S. Environ-
     mental Protection Agency; EPA report no. EPA-650/4-75-026.

Dutkiewicz, V.  A.;  Husain,  L. (1979) Determination of stratospheric ozone  at
     ground  level  using  'Be/Ozone  ratios.   Geophys.   Res.  Lett.  6:  171-174.

Eaton, W.  C.; Decker, C. E.; TommerdahlI, J.  B.; Dimmock,  F.  E. (1979) Study of
     the  nature  of  ozone, oxides of  nitrogen, and nonmethane hydrocarbons  in
     Tulsa, Oklahoma;  v.  1: project  description and  data summaries.  Research
     Triangle Park,  NC:  U.S.  Environmental  Protection Agency, Office of Air
     Quality  Planning and Standards;  EPA report no.  EPA-450/4-79-008a. Avail-
     able from: NTIS, Springfield, VA; PB-300481.

Erickson, R.  E.; Yates,  L. M.; Clark, R. L.; McEwen,  D. (1977) The  reaction of
     sulfur dioxide with ozone  in water and  its possible atmospheric signifi-
     cance. Atmos.  Environ. 11: 813-817.

Evans, R.  C.; Tingey, D. T.; Gumpertz, M. L.; Burns,  W. F.  (1982) Estimates of
     isoprene and monoterpene emission  rates in  plants.  Bot. Gaz.  (Chicago)
     143:  304-310.

F.R. (1971, August  14)  36:  15486-15506.

F.R. (1979, November  14) 44: 65667-65670.

F.R. (1981, January 22)  46: 7193-7194.

Fabian, P.; Pruchniewicz, P. G. (1977) Meridional distribution of ozone  in  the
     troposphere  and its seasonal variations. J,  Geophys. Res. 82:  2063-2073.

Falls, A. H.; McRae,  G,  J.; Seinfeld, J. H.  (1979) Sensitivity and  uncertainty
     of reaction mechanisms for photochemical  air pollution.  Inter.  J.  Chem.
     Kinet. 11:  1137-1162.

Farber, R. J.;  Huang, A. A.; Bregman,  L. D.; Mahoney, R. L.; Eatough, D.  J.;
     Hansen,  L.  D.; Blumenthal, D.  L.;  Keiffer,  W.  S. ;  Allard,  D. W. (1982)
     The  third dimension  in  the  Los Angeles basin.  Sci.   Total  Environ.
     23:  345-360.
                                    3-154

-------
Federal Highway Administration  (1978)  Traffic volume trends,  tables  5A,  5B,
     and  9A.  November-December  1978,   computer  printout;  available  from
     Washington, DC: Federal Highway Administration.

Ferman, M. Z.  A. (1981) Rural nonmethane hydrocarbon concentrations and composi-
     tion. In: Bufalini, J. J.; Arnts,  R. R., eds. Atmospheric biogenic hydro-
     carbons.  Ann Arbor, MI:  Ann Arbor Science Publishers, Inc.

Ferman, M. Z.  A.; Monson, P. R. (1978) Comparison of rural and urban air  quality.
     Paper presented at 71st annual meeting of the Air Pollution Control  Asso-
     ciation;  Houston, June.

Finlayson, B.  J.;  Pitts, J. N.,  Jr.  (1976) Photochemistry of the  polluted
     troposphere. Science 192: 111-119.

Finlayson-Pitts, B. J.; Pitts, J. N., Jr. (1982) An assessment of the atmospheric
     chemistry  of  oxides  of sulfur and  nitrogen:  acid  deposition  and ozone.
     Final Report  to  Office of Technology Assessment, Congress of the United
     States.  Available from NTIS, Springfield, VA.

Finlayson, B.  J.; Pitts, J. N., Jr.; Atkinson, R. (1976)  Low pressure gas phase
     ozone-olefin  reactions:  chemiluminescence,  kinetics, and  mechanisms.  J.
     Am. Chem. Soc. 96: 5356-5367.

Fisher, T. M.  (1980) The evolution of automotive emission  controls.   Presented
     at:  73rd  annual  meeting of  the Air Pollution Control Association; June;
     Montreal, Canada. Pittsburgh, PA:  Air Pollution Control Association; paper
     no. 80-1.1.

Fishman, J. (1984) Ozone in the troposphere. In: Stratospheric ozone. Whitten,
     R. C.; Prasad, S., eds. New  York; Van Nostrand Reinhold.
                               \

Fishman, J.; Carney, T. A.  (1984) A one-dimensional photochemical model of  the
     troposphere with  planetary boundary-layer parameterization. J. Atmos.  Chem.
     1: 351-376.

Fishman,  J.;  Crutzen,  P.  J. (1978) The  origin  of ozone  in the troposphere.
     Nature (London) 274: 855-858.

Fishman, J.; Seiler, W. (1983)  Correlative nature of ozone and carbon monoxide
     in the troposphere:  Implications  for the tropospheric  ozone  budget. J.
     Geophys.  Res. 88: 3662-3670.

Fishman,  J.;  Solomon,  S.;  Crutzen,  P.  J.  (1980)  Observational  and  theoretical
     evidence in support of a significant jji situ photochemical source of tropo-
     spheric ozone. Tell us  31:  432-446.

Fishman,  J. ;  Vukovich, F. M. ;  Browell,  E.  V.  (1985) The photochemistry  of
     synoptic-scale ozone  synthesis:  Implications for the global  tropospheric
     ozone budget. J.  Atmos. Chem. 3:299-320.
                                    3-155

-------
Flyckt, D. L.; Westberg, H. H.; Holdren, M. W. (1980) Natural organic emissions
     and their  impact  on air quality.   Presented  at:  73rd annual  meeting of
     the Air Pollution Control Association; June; Montreal, Quebec. Pittsburgh,
     PA: Air Pollution Control Association; Paper 80-69.2.

Galloway, J.  N.; Likens, G. E.; Keene, W. C.; Miller, J. M. (1982) The composition
     of  precipitation  in  remote  areas of the world.  JGR J.  Geophys. Res.
     82: 8771-8786.

Gelinas,  R.  J.; Vajk,  J.  P.  (1979)  Systematic sensitivity analysis  of  air
     quality simulation models. Research Triangle Park, NC: U.S. Environmental
     Protection Agency. EPA-6QQ/4-49-035.

Gerstle, R.  W.;  Peterson,  R.  F.  (1966) Atmospheric  emissions  from  nitric acid
     manufacturing  processes. Cincinnati,  OH:  U.S. Department of Health,
     Education, and Welfare,  National  Center for Air Pollution  Control;  Public
     Health  Service  publication  no.  999-AP-27. Available  from: NTIS, Spring-
     field, VA; PB-190239.

Gertler, A. W.; Miller, D. F.; Lamb, D.; Katz, U. (1984) Studies of sulfur  dioxide
     and nitrogen  dioxide reactions  in haze and  cloud.  In: Durham, J.  L., ed.
     Chemistry of particles,  fogs and  rain. Boston, MA: Butterworth Publishers;
     pp.  131-160.  (Teasley,  J.  I.,  ed. Acid  precipitation series, v.   2).

Gibbs, R. E.; Hyde, J.  D.; Whitby, R.  A.; Choudhury, D. R.  (1983)  Characterization
     of emissions and fuel economy of  in-use diesel automobiles. Research Triangle
     Park,  NC:    U.S.  Environmental  Protection Agency;   EPA  report  no.
     EPA-600/3-83-087.   Available from:  NTIS,  Springfield, VA; PB83-262071.

Gillani, N.  V.;  Kohli,  S.; Wilson, W.  E.  (1981)  Gas-to-particle conversion of
     sulfur  in  power plant plumes—I.  Parameterization of the conversion rate
     for  dry,  moderately  polluted ambient  conditions.  Atmos.  Environ.   15:
     2293-2313.

Gipson,  G.  L.  (1984) Guidelines  for using carbon-bond mechanisms in city-
     specific EKMA.  Research Triangle Park, NC:  U.S.  Environmental Protection
     Agency.  Report no. EPA-450/4-84-005.

Glasson, W. A.; Tuesday, C. S. (1970)  Inhibition of atmospheric photooxidation
     of hydrocarbons by nitric oxide.  Environ. Sci. Techno!.  4: 37-44.

Graedel,  T.  E.  (1978)  Chemical  compounds  in  the atmosphere. New  York,  NY:
     Academic Press.

Graedel, T.  E.;  Farrow, L. A.;  Weber, T.  A.  (1976) Kinetic  studies  of the
     photochemistry  of  the urban troposphere. Atmos.  Environ.  10:  1095-1116.

Graedel, T.  E.;  Weschler, C. J.  (1981)  Chemistry  within  aqueous atmospheric
     aerosols and raindrops.  Rev. Geophys. Space Phys.  19:  505-539.

Graham,  R. A.;  Winer,  A.  M.; Pitts, J.  N., Jr.  (1977)  Temperature dependence
     of  the  unimolecular decomposition of pernitric acid  and its  atmospheric
     implications. Chem.  Phys. Lett. 51: 215-220.
                                    3-156

-------
Graham, R. A.;  Johnston,  H.  S.  (1978)  The photochemistry of N03 and the kine-
     tics of the N205-03 system. J. Chem.  Phys. 82: 254-268.

Graham, R. A.; Winer, A. M.; Pitts, J. N., Jr.  (1978)  Pressure  and  temperature
     dependence of  the  unimolecular decomposition of  H02N02.  J.  Chem.  Phys.
     68: 4505-4510.

Grosjean, D.  (1982) Formaldehyde and  other  carbonyls  in Los Angeles ambient
     air. Environ.  Sci. Techno!. 16: 254-262.

Grosjean, D.  (1984)  Photooxidation of  methyl  sulfide, ethyl  sulfide,  and
     methanethiol.  Environ.  Sci. Techno!.  18:  460-468.

Grosjean, D.;  Lewis,  R. (1982)  Atmospheric  photooxidation  of methyl  sulfide.
     Geophys. Res.  Lett. 9:  1203-1206.

Grosjean, D.;  Fung, K.  (1984)  Hydrocarbons  and carbonyls in Los  Angeles Air.
     J. Air Pollut. Control  Assoc.  34: 537-543.

Haagenson, P.  L.;  Shapiro,  M.  A.;  Middleton,  P.  (1981)  A case  study relating
     high ground  level  ozone to enhanced photochemistry and isentropic trans-
     port from  the  stratosphere. JGR J.  Geophys.  Res.  86: 5231-5237.

Halfpenny, E.;  Robinson,  P.  L.  (1952a)  Pernitrous acid: the reaction between
     hydrogen peroxide  and nitrous  acid,  and the  properties of  an intermediate
     product. J. Chem.  Soc.  928-938.

Halfpenny,  E.;  Robinson,  P.  L.  (1952b)  The  nitration  and  hydroxylation of
     aromatic compounds by pernitrous  acid.  J.  Chem. Soc. 939-946.

Hanna,  S.  R.  (1973) A  simple  dispersion model  for the analysis of chemically.
     reactive pollutants.  Atmos. Environ.  7:  803-817.

Hanst,  P. L. .(1970) Infrared spectroscopy and infrared lasers in  air pollution
     research and monitoring.  Appl. Spectrosc.  24: 161-174.

Hanst,  P.  L. ;  Gay, B.  W.  (1977) Photochemical reactions among formaldehyde,
     chlorine,  nitrogen dioxide in air.  Environ.  Sci.  Techno!.  11: 1105-1109.

Hanst,  P.  L. ;  Wilson, W.  E.; Patterson,  R. K.; Gay,  B. W.,  Jr.; Chaney,  L. W.
     (1975)  A  spectroscopic study of California smog.  Research Triangle  Park,
     NC:  U.S. Environmental  Protection Agency, National  Environmental Research
     Center;  EPA  report no.  EPA-650/4-75-006.  Available from:  NTIS, Spring-
     field, VA; PB-241022.

Harding,  L.  B.;  Goddard,  W.  A.,  III.  (1978)  Mechanisms of gas-phase  and
     liquid-phase ozonolysis.  J. Am.  Chem. Soc. 100:  7180-7188.

Harris,  G.  W.;  Carter, W.  P.  L.;  Winer, A.  M.; Pitts, J. N., Jr.;  Platt, U.;
     Perner,  D. (1982) Observations of nitrous acid in the  Los Angeles  atmos-
     phere  and  implications for predictions of ozone-precursor relationships.
     Environ. Sci.  Techno!.  16:  414-419.
                                    3-157

-------
Harrison, R. M.;  Laxen,  D.  P.  H.  (1978)  Sink processes for tetralkyllead com-
     pounds in the atmosphere. Environ. Sci. Techno!.  12:  1384-1392.

Hatakeyama, S.;  Okuda, M.; Akimoto,  H.  (1982) Formation  of  sulfur dioxide
     and methanesulfonic acid  in the photooxidation of dimethyl sulfide  in  the
     air. Geophys. Res. Lett.  9: 583-586.

Hatakeyama, S.; Akimoto, H. (1983) Reactions of OH radicals with mathanethiol,
     dimethyl sulfide,  and dimethyl  disulfide in air.  J. Phys.  Chem.  87:
     2387-2395.

Hayes, S.  R.; Hogo,  H.  (1982)  An  uncertainty analysis of EKMA-derived control
     requirements for  ozone abatement in the South Coast Air Basin. Presented
     at  75th  annual  meeting of the  Air Pollution Control  Association,  New
     Orleans, LA, June 20-25.

Hegg, D. A.; Hobbs,  P. V.  (1981) Cloud water  chemistry  and the production  of
     sulfates in clouds. Atmos. Environ. 15: 1597-1604.

Heikes,  B.  G.; Lazrus,  A.  L.; Kok, G.  L.;  Kunen, S. M.; Gandrud,  B.  W.;
     Gitlin, S. N.;  Sperry, P. D. (1982) Evidence for aqueous phase  hydrogen
     peroxide  synthesis  in the  troposphere.  JGR  J. Geophys.  Res. 87:
     3045-3051.

Heikes,  B.  G.;  Thompson, A. M. (1983)  Effects of heterogeneous  processes on
     N03, HONO, and  HN03 chemistry in the  troposphere.  J. Geophys. Res. 88:
     10883-10895.

Hendry,  D.  G.;  Baldwin, A. C.; Barker,  J.  R.; Golden, D. M.  (1978) Computer
     modeling  of  simulated photochemical  smog.  Research Triangle  Park,  NC:
     U.S. Environmental Protection Agency, Office  of  Research and Development;
     EPA report no.  EPA-600/3-78-059. Available from: NTIS,  Springfield, VA:
     PB-284025.

Herron,  J.  T.;  Huie, R.  E.  (1977)  Stopped-flow studies  of  the mechanism of
     ozone-alkene reactions in the  gas phase.  Ethylene.  J. Am.  Chem.  Soc. 99:
     5430-5435.

Herron, J. T.;  Huie, R. E.  (1978) Stopped flow studies of mechanisms  of
     ozone-alkene reactions in gas  phase propene  +  isobutene.  Int. J.  Chem.
     Kinet. 10: 1019-1040.

Herron,  0.  T.;  Martinez,  R. I.; Huie, R.  E.  (1982) Kinetics and energetics  of
     the Criegee intermediate  in the  gas phase. I. The Criegee intermediate in
     ozone-alkene reactions. Int.  J.  Chem. Kinet.  14:  201-224.

Hill,  A.  C. (1971)  Vegetation:  a sink  for  atmospheric pollutants.  J.  Air
     Pollut. Control Assoc. 21: 341-346.

Hilliard, J. C.;  Wheeler,  R.  W.  (1979)  Nitrogen  dioxide  in  engine exhaust.
     Presented  at  passenger car meeting; June; Dearborn,  MI. Warrendale, PA:
     Society of Automotive  Engineers;  SAE paper no. 790691.
                                    3-158

-------
Hoffman, M. R.;  Edwards, J. 0.  (1975)  Kinetics  of the  oxidation of sulfite by
     hydrogen  peroxide  in  acidic  solution.  J.  Phys.  Chem.  79; 2096-2098.

Hoffman, M. R.;  Jacob,  D.  J.  (1984) Kinetics and mechanisms of the catalytic
     oxidation of dissolved sulfur dioxide in aqueous solution: an application
     to nighttime fog water chemistry. In: Calvert, J. G., ed. S02, NO and N02
     oxidation mechanisms:  atmospheric considerations. Boston, MA: Butterworth
     Publishers; pp.  101-172.  (Teasley,  J.  I.,  ed.  Acid  precipitation series,
     v. 3).

Holdren, M. W.;  Westberg,  H.  H.;  Zimmerman,  P.  R.  (1979) Analysis of monoter-
     pene hydrocarbons in rural atmospheres. JGR J. Geophys. Res. 84: 5083-5088.

Holdren, M. W.;  Spicer,  C. W.;  Hales,  J.  M.  (1984) Peroxyacetyl  nitrate  solu-
     bility and  decomposition  rate  in acidic  water.  Atmos.  Environ. 18:
     1171-1173.

Holdren, M.;'Spicer, C.; Sticksel, P.; Nepsund,  K.; Ward, G.;  Smith,  R.  (1982)
     Implementation and  analysis  of hydrocarbon grab  samples  from Cleveland
     and Cincinnati 1981.  Ozone monitoring study.  Research Triangle Park, NC:
     U.S.  Environmental  Protection Agency; EPA  report no.  EPA-905/4-82-001.
     Available from: NTIS, Springfield, VA; PB83-197673.

Holzworth, G.  C. (1964) Estimates  of mean maximum  mixing depths  in the contig-
     uous United States. Mon. Weather  Rev. 92:  235-242.

Holzworth, G.  C.  (1972)  Mixing heights, wind speeds,  and potential  for  urban
     air pollution  throughout the contiguous United States.  Research Triangle
     Park, NC:  U.S.  Environmental Protection Agency;  publication  no. AP-101.
     Available from: NTIS, Springfield, VA; PB-207103.

Holzworth, G.  C.;  Fisher,  R.  W.  (1979) Climatological summaries  of  the  lower
     few kilometers of rawinsonde observations.  Research Triangle Park, NC:
     U.S.  Environmental  Protection Agency,  Environmental Sciences  Research
     Laboratory; EPA report no. EPA-600/4-79-026.

Hoshino, M.; Akimoto, H.; Okuda,  M. (1978) Photochemical  oxidation of benzene,
     toluene,  and  ethyl benzene initiated  by  OH radicals in the  gas phase.
     Bull. Chein. Soc.  Japan 51: 718-724.

Hosier,  C.  R.  (1961)  Low-level inversion  frequency in the contiguous United
     States. Mon. Weather  Rev.  89: 319-339.

Howard,  C.  J.  (1977)  Kinetics of  the  reaction of H02 with N02. J. Chem.  Phys.
     67: 5258-5263.

Hull,  L. A. (1981) Terpene ozonolysis  products.  In:  Bufalim",  J.  J.;  Arnts,  R.
     R., eds.  Atmospheric  biogenic hydrocarbons:  v.  II.   Ann  Arbor, MI: Ann
     Arbor Science Publishers;  p.  161.

Hunsaker,  D.  (1981)  Selection of biogenic  hydrocarbon  emission factors for
     land  cover classes found in the San Francisco  Bay Area.  California
     Association of Bay Area Governments; air quality  technical memo  31.
                                    3-159

-------
Hunter, S.  C.,  et  al.  (1978)  Evaluation of combustion modifications for emis-
     sions  reduction  in  industrial  processes.  ASME  Paper  No.  78-WA/APC-8.

Hunter, S.  C.;  Carter, W. A,; McElroy,  M.  W.;  Cherry, S.  S.;  Buening,  H.  J.
     (1977). Application  of combustion modifications to industrial  combustion
     equipment. Research  Triangle Park, NC:  U.  S.  Environmental  Protection
     Agency; EPA report no. EPA-60Q/7-79~15a.

Husain, L.; Coffey, P. E.; Meyers, R.  E.; Cederwall,  R. T.  (1977) Ozone
     transport  from stratosphere  to  troposphere. Geophys.  Res.  Lett.  4:
     363-365.

Jaffee, R. J.; Smith, F.  C., Jr.; West,  K. W. (1974)  Study  of  factors affecting
     reactions  in  environmental  chambers:  final report on phase III.  Research
     Triangle Park, NC: U.S. Environmental Protection Agency,  Office of  Research
     and  Development;  EPA report no.  EPA-650/3-74-004b. Available  from:  NTIS,
     Springfield, VA PB-256252.

Japar, S. M.;  Niki, H.  (1975)  Gas-phase reactions of the nitrate radical with
     olefins. J. Phys. Chem. 79:  1629-1632.

Jeffries, H.  E,; Johnson, G.  T.  (1981) Predicting  ozone  frequency distri-
     butions.  In:  Proceedings of EKMA Model Validation  Workshop,  December
     15-17, Research Triangle Park, NC.

Jeffries, H.;  Fox,  D.;  Kamens,  R. (1975) Outdoor smog chamber studies.  Effect
     of  hydrocarbon reduction on nitrogen  dioxide.  Research Triangle Park,
     NC:  U.S. Environmental Protection Agency;  EPA report  no.  EPA-650/3-75-011.
     Available from: NTIS, Springfield,  VA;  PB-245829.

Jeffries, H.;  Fox,  D.; Kamens,  R. (1976) Outdoor smog chamber studies:  light
     effects relative to  indoor  chambers. Environ. Sci. Technol. 10:. 1006-1011.

Jeffries, H.  E.;  Sexton, K.  G.;  Salmi,  C. N. (1981)  Effects of chemistry  and
     meteorology on ozone calculations using simple  trajectory models  and the
     EKMA procedure. Research Triangle Park, NC:  U.  S.  Environmental Protection
     Agency; EPA report no. EPA-450/4-81-034.

Jeffries, H.  E.;  Sexton, .K. G.;  Kamens, R.  M.;  Holkman, M. S.  (1985) Outdoor
     smog chamber  experiments to test photochemical  models. Research Triangle
     Park, NC: U. S. Environmental Protection Agency.

Johnson,  W. B.; Viezee, W. (1981) Stratospheric ozone in  the  lower  troposphere—
     I. Presentation and  interpretation  of aircraft  measurements. Atmos.  Environ.
     15:  1309-1323.

Joshi,  S. B.  (1979) Houston field study (1978):  formaldehyde  and total  alde-
     hydes  monitoring  program.  EPA contract no. 68-02-2566. Research Triangle
     Park, NC: Northrop Services, Inc.-Environmental  Sciences;  technical  report
     no.  ESC-TR-79-22.

Joshi, S. B.; Adger, R. A.; Barnes, D.  A.;  Kaiserman, M.  (1981) Aldehyde sampling
     field  study  (1980).  Research Triangle Park, NC: Northrop Services, Inc.-
     Environmental  Sciences; technical report no.  TR-81-02.


                                   3-160

-------
JSA, Inc. (1978)  Draft environmental  impact  statement  for  the  South  Tahoe  public
     utilities district,  waste water  facilities  planning  program.  Jones and
     Stokes Associates, Inc.

Junge, C.  E.  (1963)  Air  chemistry and radioactivity.   New York, NY:  Academic
     Press, (van Mieghen, J.;  Hales,  A.  L., eds.  International  geophysics  series
     v. 4).                                           .   ,

Kacmarek, A.  J. ;  Solomon, I.  J.;  Lustig, M.  (1978) Preparation and properties
     of peroxyacetyl nitrate.  J.  Inorg.  Nucl.  Chem. 40:  574-576.

Kamens,  R.  M. ;  Gery, M.   W. ;  Jeffries,  H.  E.;  Jackson,  M.;  Cole,  E.  I.  (1982)
     Ozone-isoprene  reactions:  product formation  and aerosol potential.  Int.
     J. Chem. Kinet. 14:  955-975.

Kan, C. S. ; Su, F.;  Calvert,  J. G.;  Shaw,  J.  H.  (1981)  Mechanism  of the ozone-
     ethene  reaction in  dilute N2/02  mixtures.  J.  Phys. Chem.  85:  2359-2363.

Kelly, N.  A.  (1985)  Ozone/precursor relationships in the Detroit metropolitan
     area derived  from captive-air  irradiations  and an  empirical  photochemical
     model. J. Air Pollut.  Control  Assoc.  35:  27-34.

Kelly, N. A.; Wolff, G. T. ;  Ferman,  M.  Z.  A.  (1982) Background  pollutant measure-
     ments  in air  masses  affecting  the eastern half of  the  United States--!.  Air
     masses arriving from the northwest.  Atmos.  Environ. 16:  1077-1088.

Kelly, N. A.; Wolff, G. T.;  Ferman,  M.  A.  (1984) Sources and sinks  of ozone in
     rural  areas.  Atmos.  Environ.  18:  1251-1266.

Kelly, N. A.; Ferman, M.  A.;  Wolff,  G.  T.  (1986) The  chemical  and meteorological
     conditions associated  with high  and  low ozone concentrations in south-
     eastern Michigan and nearby  areas of Ontario.  J. Air  Pollut.  Control  Assoc.
     36: 150-158.

Kenley,  R.  A.; Davenport,  J. E. ;  Hendry, D. G.  (1981) Gas-phase  hydroxyl
     radical reactions. Products  and pathways for the reaction  of OH with aro-
     matic  hydrocarbons.  J.  Phys.  Chem.  85:  2740-2746.

Killus,  J.  P.;  Whitten,  G.  Z. (1982) A mechanism describing the photochemical
     oxidation of  toluene in smog.  Atmos.  Environ.  16:  1973-1988.

Killus,  J.  P.; Whitten, G.  Z.  (1984) Isoprene:  a photochemical  kinetic
     mechanism. Environ.  Sci.  Techno!.  18: 142-148.

Kim,  C.  M.  (1973) Influence  of  vegetation type  in the intensity of ammonia
     and  nitrogen  dioxide liberation from soil.  Biochem.  5: 163-166.

Kircher,  C.  C.; Margitan, J.  J.;  Sander, S.  P.  (1984) Pressure  and temperature
     dependence  of the reaction  N02 + N03 + M --> N205 + M.  J.  Phys. Chem.  88:
     4370-4375.

Kirk^-Othmer encyclopedia of chemical technology.  (1981) 3rd ed. New  York,  NY:
     John  Wiley &  Sons,  Inc., v.  16: 701.
                                    3-161

-------
Knoerr, K.  R.;  Mowry,  F.  W. (1981)  Energy balance/Bowen ratio technique for
     estimating hydrocarbon  fluxes.  In:  Bufalini, J. J.; Arnts,  R.  R.,  eds.
     Atmospheric biogenic hydrocarbons, v. 1. Ann Arbor, MI: Ann Arbor Science.

Knoppel, H.; Versino, B.; Peil, A.; Schauenburg, H.; Vissers, H. (1982) Quanti-
     tative determination of terpenes  emitted by conifers.  In:  Versino,  D.;
     Ottj H., eds. Physicochemical behavior of atmospheric pollutants. Boston,
     MA: D. Reidel Publishing Company; pp. 89-98.

Korshover,  J.  (1967) Climatology  of  stagnating  anticyclones  east of the  Rocky
     Mountains,  1936-1965.   Cincinnati,  Ohio: U.S.  Department of  Health,
     Education  and Welfare,  National Center  for  Air  Pollution  Control;  Public
     Health Service publication no. 999-AP-34. Available from: NTIS, Springfield,
     VA; PB-174709.

Korshover,  J.  (1975) Climatology  of  stagnating  anticyclones  east of the  Rocky
     Mountains, 1936-1975. Washington, DC: U.S.  Department of Commerce, National
     Oceanic and Atmospheric Administration; NOAA technical memorandum ERL AR-55.
     Available from:  NTIS, Springfield, VA; PB-257368.

Kunen, S.  M.;  Lazrus,  A.  L. ; Kok, G. L.; Heikes, B. G. (1983) Aqueous oxida-
     tion  of  S02  by  hydrogen peroxides.  JGR  J,.  Geophys.  Res.  88:  3671-3674.

Lamb, B. K.; Westberg,  H. H.; Quarles, T.; Flyckt, D. L. (1983) Natural hydro-
     carbon emission rate measurements from selected forest  sites.  Research
     Triangle Park, NC:  U.S. Environmental Protection Agency; EPA report no.
     EPA-600/3-84-001.   Available  from: NTIS,  Springfield,  VA; PB84-124981.

Lamb, R.  G.  (1977) A case study of stratospheric ozone  affecting ground-level
     oxidant concentrations. J. Appl. Meteorol.  16: 780-794.

Larsen, R.  I.  (1969)  A new  mathematical  model of air pollutant concentration
     averaging  time and frequency. J.  Air Pollut.  Control  Assoc.  19:  24-30.

Larson, T.  V.;  Horike,  N. R.;  Harrison,  H.  (1978)  Oxidation  of sulfur dioxide
     by oxygen and ozone in  aqueous solution: a kinetic study with significance
     to atmospheric rate processes. Atmos. Environ. 12: 1597-1611.

Lee, Y.-N.;  Senum, G.  I.; Gaffney, J.  S. (1983) Peroxyacetyl nitrate (PAN)
     stability, solubility,   and reactivity—implications for tropospheric nitrogen
     cycles and precipitation chemistry.  Presented at CACGP Symposium, Oxford,
     England; September.

Leighton,  P. A. (1961) Photochemistry of  air pollution. New York, NY: Academic
     Press, Inc.

Leihkanen,  H. E.; Beckman, E. W.  (1971) The effect of leaded and unleaded gaso-
     lines  on  exhaust  emissions as influenced by combustion chamber deposits.
     Presented  at Society of Automotive Engineers' national combined  fuels and
     lubricants,  powerplant, and  truck  meetings; October;  St.  Louis, MO.
     Warrendale,  PA:  Society of  Automotive  Engineers;  SAE  paper no. 71083.
                                   3-162

-------
Leone, J. A.;  Seinfeld,  J.  H. (1984)  Updated  chemical  mechanism for atmos-
     pheric photooxidation  of toluene.  Int.  J,  Chem.  Kinet.  16: 159-193.

Leone, J. A.;  Flagan,  R.  C.; Grosjean, D.; Seinfeld, J. H. (1985) An outdoor
     smog chamber and  modeling study of  toluene-NO   photooxidation.  Int.  J.
     Chem. Kinet. 17: 177-216.                     x

Levy, H.  II; Mahlman,  J.  D.;  Moxim,  W.  J.  (1985)  Tropospheric  ozone:  The role
     of transport.  J. Geophys. Res. 90: 3753-3772.

Lindley, C. R.  C.;  Calvert, J. G.;  Shaw, J.  H.  (1979)  Rate  studies  of the
     reactions of the  (CH3)2N radical  with  02,  NO, and  N02.  Chem.  Phys.  Lett.
     67: 57-62.

Linnell,  R.  H.; Scott,  W.  E.  (1962)  Diesel  exhaust composition and odor
     studies.  J. Air Pollut. Control Assoc. 12: 510-515, 545.

Liu, C.  S.; Grisinger, J. E.  (1981) Review of SAI Airshed Model Sensitivity Study
     Conducted for the South Coast Air  Basin by California Air Resources Board.
     AQMP Technical  Paper No.  5.  South Coast Air  Quality Management District;
     El  Monte, CA.

Liu, M.-K.; Seinfeld,  J.  H.  (1975)  On  the  validity of  grid and  trajectory
     models of urban air pollution. Atmos. Environ. 9: 555-574.

Liu, S.  C.; Cicerone,  R.  J.; Donahue,  T. M.; Chameides, W. L. (1977) Sources
     and sinks of atmospheric N20 and the possible ozone reduction due  to  indus-
     trial fixed nitrogen fertilizers.  Tell us.  29: 251-263.

Liu, M.-K.; Morris, R. E.;  Killus, J. P.  (1984) Development of a  regional  oxidant
     model and application  to the northeastern United States.  Atmos.  Environ.
     18: 1145-1161.

Lloyd,  A.  C.  (1979) Tropospheric chemistry of aldehydes.  In:  Herron, J. T.;
     Huie, R.   E.; Hodgeson, J. A., eds. Chemical  kinetic data needs for modeling
     of  the lower  troposphere:  proceedings of  a  workshop;  May 1978;  Reston,
     VA.  Gaithersburg,  MD:   U.S.  Department of Commerce, National Measurement
     Laboratory; Bureau of  Standards special publication 557.

Lloyd,  A.  C.;  Lurmann, F.  W.; Godden,  D. A.;  Hutchins, J.  F.;  Eschenroeder,
     A.   Q. ; Nordsieck, R. A.  (1979) Development  of the ELSTAR photochemical
     air  quality simulation model and  its evaluation  relative to  the  LARPP data
     base.  ERT Doc. No.  P-5287-500,  Environmental  Research and  Technology,
     Inc., Westlake Village,  CA.

Lloyd,  A. C.;  Atkinson, R.; Lurmann, F. W.; Nitta, B. (1983) Modeling
     potential  ozone  impacts  from  natural  hydrocarbons. I. Development and
     testing of  a chemical  mechanism for  the NO -air  photooxidations  of isoprene
     and  orpinene  under  ambient  conditions.  Amos.  Environ. 17: 1931-1950.

Logan,  J.  A.  (1983) Nitrogen  oxides  in the troposphere: global  and regional
     budgets.  JGR J. Geophys.  Res. 88:  10785-10807.
                                    3-163

-------
Logan, J.A.;  Prather,  M.J.; Wofsy, S.C.;  McElroy,  M.B.  (1981) Tropospheric
     chemistry: A global perspective.  J. Geophys. Res. 86: 7210-7254.

Lonneman, W.  A.  (1977)  Ozone and hydrocarbon  measurements  in recent oxidant
     transport studies.  In:  Dimitriades,  B,, ed. International conference on
     photochemical oxidant pollution and its control; September 1976; Raleigh,
     NC.  Proceedings: volume I. Research Triangle Park, NC: U.S. Environmental
     Protection Agency, Environmental  Sciences Research Laboratory;  pp. 211-223;
     EPA report  no.  EPA-600/3-77-001a.  Available from: NTIS,  Springfield,  VA;
     PB-264232.

Lonneman, W. A. (1979) Ambient hydrocarbon measurements in Houston.  In: Ozone/
     oxidants: interactions  with  the  total environment II:  proceedings of a
     specialty conference;  October; Houston,  TX. Pittsburgh,  PA:  Air  Pollu-
     tion Control Association; pp. 278-287.

Los Angeles County,  California.  (1971) Profile  of air pollution control.  Los
     Angeles,  CA: Los Angeles Air Pollution Control District.

Ludwig^F.  L.;  Reiter,  E.;  Shelar, E.; Johnson, W. B. (1977)  The  relation of
     oxidant levels to precursor emissions and meteorological  features. Volume
     I: analysis  and findings.  Research Triangle Park, NC:  U.S.  Environmental
     Protection Agency,  Office of Air Quality  and Standards;  EPA final  report
     no.  EPA^450/3-77-022a.  Available from: Springfield, VA;  PB-275001.

Lurmann,  F.;  Godden,  D.; Lloyd, A. C.; Nordsieck,  R.  A.  (1979)  A Lagrangian
     photochemical air quality simulation model. Adaptation to the St.  Louis-RAPS
     Data Base.  Vol.  I.  Model Formulation. Vol.  II.  User's Manual.  Research
     Triangle  Park,  NC:  U.  S.  Environmental  Protection Agency; EPA report  no.
     EPA-600/8-79-015a,b.

Lynn, D. A.;  Steigerwald,  B. J.; Ludwig,  J, H.  (1964) The  November-December
     1962 air  pollution  episode in  the  eastern United States. Cincinnati,  OH:
     U.S. Department of  Health,  Education  and  Welfare, Division of Air Pollu-
     tion;  Public Health Service publication no.  999-AP-77.  Available from:
     NTIS, Springfield, VA; PB-168878.

Lyons, W. A.;  Olsson,  L. E.  (1972) Mesoscale  air pollution transport in the
     Chicago lake breeze. J. Air Pollut. Control Assoc. 22: 876-881.

MacCracken, M. C.; Wuebbles, D. J.; Walton, J. J.; Duewer, W.  H.;  Grant, K.  E.
     (1978) The Livermore regional air  quality model: I. Concept and development.
     J. Appl.  Meteorol. 17: 254-272.

Magnotta, F.;  Johnston, H.  S. (1980)  Photodssociation  quantum yields  for
     the N03 free radical. Geophys. Res. Lett. 7: 769-772.

Makdo, M. W.;  Troe,  J.   (1982)  Analysis of unimolecular  reaction of N205 + M
     j N02 + N03  + M. Int. J.  Chem. Kinet.  14: 399-416.

Marakov, B. N.  (1969)  [Liberation of nitrogen dioxide from  soil]. Pochovove-
     denie. 1: 49-58.
                                   3-164

-------
Marchesani, V. J.;  Towers,  T.;  Wohlers, H.  C.  (1970)  Minor sources of air
     pollutant emissions. J. Air Pollut. Control Assoc. 20: 19-22.

Martin, L. R.  (1984) Kinetic studies of sulfite oxidation in aqueous solution.
     In: Calvert, J. G., ed. S02s NO and NOg oxidation mechanisms: atmospheric
     considerations. Boston, MA: Butterworth Publishers; pp. 63-100. (Teasley,
     J. I., ed., Acid precipitation series, v. 3).

Martin, L. R.;  Damschen, D. E.  (1981) Aqueous oxidation of sulfur dioxide by
     hydrogen peroxide at low pH. Atmos. Environ. 15: 1615-1621.

Martinez, J.  R.;  Singh,  H.  B.   (1979)  Survey of the  role of NO   in nonurban
     ozone formation.  SRI  International,  Menlo Park,  CA.  Final report to
     U.S. Environmental  Protection Agency, SRI Project 6780-8. Research Triangle
     Park, NC: U.S. Environmental Protection Agency.

Martinez, R.  I.;  Herron, J. T.; Huie,  R.  E.  (1981)  The mechanism of ozone-
     alkene reactions  in the gas phase. A mass spectrometric study of  the
     reactions of  eight  linear  and branched-chain alkenes. J. Am, Chem. Soc.
     103: 3807-3820.

Maxwell,  C.;  Martinez,  J.  R.  (1982) Performance evaluation of three chemical
     models used with the Empirical Kinetic Modeling  Approach (EKMA). Presented
     at:  75th  annual  meeting of the Air  Pollution  Control Association. New
     Orleans, LA, June 20-25.

McElroy,  C. T.; Kerr, J. B. (1977) A limb  scan measurement of atmospheric nitro-
     gen  dioxide. Presented at:  fall meeting of the American Geophysical Union;
     San  Francisco, CA.

McKee,  H. C.; Clark, J.  M.; Wheeler, R. J. (1962) Measurement of diesel engine
     emissions. J. Air Pollut.  Control Assoc. 12: 516-521, 546.

Meyer,  E. L.; Gipson, G. L.; Dimitriades,  B. (1981) Evaluation of the empirical
     kinetics  modeling  approach. Paper 81-6.4, Annual  Meeting of the Air
     Pollution Control Association, Philadelphia, Pennsylvania.

Meyers, T.  C.,  et al. (1979) User's  Guide to the SAI Trajectory Model. SAI
     Report No. EP-79-3, Systems Application, Inc., San  Rafael,  CA.

Mohnen,  V.  A. (1977)  Review  and analysis. In:   International conference  on
     oxidants,, 1976—Analysis of evidence and viewpoints.  Part III.  The issue
     of stratospheric ozone  intrusion.  Research Triangle Park, NC:  U.S.
     Environmental Protection Agency.  Report no.  EPA-600/3-77-115.

Morris,  E.  D.,  Jr.,  Niki,  H.   (1974)  Reaction  of the nitrate radical with
     acetaldehyde  and propylene. J. Phys.  Chem.  78: 1337-1338.

Motor  Vehicle Manufacturers Association.  (1984)  MVMA Motor Vehicle  Facts and
     Figures  (annual  summary issues).  Detroit,  MI: Motor Vehicle Manufacturers
     Association of the  United  States,  Inc.

Mudd,  J.  B. (1966)  Reaction of  peroxyacetyl  nitrate with glutathione. J. Biol.
     Chem. 241: 4077-4080.


                                    3-165

-------
Mueller, P.  K.;  Hidy,  G. M.  (1983)  The sulfate regional experiment  (SURE):
     report  of  findings. Palo Alto,  CA:  Electric Power Research  Institute;
     report no. EA-1901.

National Research  Council.  (1977a)  Ozone and  other photochemical oxidants.
     Washington, DC: National Academy of Sciences.

National Research  Council.  (1977b) Nitrogen oxides. Washington, DC:  National
     Academy of Sciences.

Naylor, M.  H.;  Lee,  D.  C.;  Smith,  M.  J.  (1981)  Southeast Las Vegas valley air
     quality measurement project 1980-81. Final  report.  Las Vegas,  Nevada:
     Clark County Health District.

Neiburger, M.; Johnson,  D. S.; Chen, C. -W.  (1961) Studies of the  structure of
     the atmosphere over the eastern Pacific Ocean in the summer:  I.  the  inver-
     sion over  the eastern north Pacific  Ocean.  Berkeley,  CA:  University of
     California Press; publications  in meteorology,  v.  1, no. 1.

Nicksic, S.  W.;  Harkins, J.; Mueller,  P.  K. (1967)  Some analyses  for PAN and
     studies of its structure. Atmos. Environ.  1: 11-18.

Niki, H.; Daby, E. E.; Weinstock, B.  (1972)  Mechanisms  of smog  reactions.  Adv.
    • Chem. Ser. 113: 16-57.

Niki,  H.;  Maker,  P.  D.; Savage, C. M.;  Breitenbach,  L.  P.  (1977) Fourier
     transform IR spectroscopic  observation  of  pernitric acid formed  via  HOO  +
     N02 -*• HOON02. Chem. Phys. Lett. 45: 564-566.

Niki, H.; Maker, P. D.;  Savage,  C. M.;  Breitenbach,  L.  P. (1978) Mechanism for
     hydroxyl radical  initiated  oxidation of olefin-nitric  oxide  mixtures in
     parts per mi 11 on concentrations. J. Phys.  Chem. 82: 135-137.

Niki,  H.;  Maker,  P.  D.; Savage, C. M.;  Breitenbach,  L.  P.  (1979) Fourier
     transform  infrared  (FTIR)  studies  of gaseous and particulate nitrogenous
     compounds.  In:  Grosjean, D.,  ed.,  Nitrogenous Air Pollutants. Ann Arbor,
     MI: Ann Arbor Science  Publishers,  Inc.; pp.  1-16.

Niki, H.; Maker, P. D.;  Savage,  C. M.;  Breitenbach,  L.  P. (1981) An FT-IR study
     of a transitory product  in  the  gas-phase ozone-ethylene reaction.  J.  Phys.
     Chem. 85: 1024-1027.

Niki,  H.;  Maker,  P.  D.; Savage,  C.  M.; Breitenbach, L. P.  (1983) An FT-IR
     study of the mechanism for  the  reaction HO+  CH3SCH3. Int.  J.  Chem. Kinet.
     15: 647-654.

Novak,  J.  (1986)  Atmospheric Sciences Research Laboratory,  U.S. Environmental
     Protection Agency,  Research Triangle Park, NC.  Personal communication to
     B. E. Tilton, Environmental Criteria  and Assessment Office, U.S.  Environ-
     mental  Protection Agency, Research Triangle  Park,  NC.

Noxon,  J.  F.  (1976)  Atmospheric  nitrogen fixation by lightning. Geophys.
     Res. Lett. 3: 463-465.
                                    3-166

-------
Noxon, J,  F. ;  Norton,  R.  B.;  Marovich," E.  (1980)  N03 in the troposphere.
     Geophys.  Res. Lett. 7: 125-128.

NPN:  National Petroleum News  (annual  fact books).  Des Plaines,  IL:   National
     Petroleum News.

Oltmans, S. J.  (1981) Surface measurements  in clean air. J.  Geophys.  Res.
     86:  1174-1180.

Parrish, D. D.; Fahey, D.  W.; Williams, E. J.; Liu,  S. C.; Trainer, M.; Murphy,
     P.  C.; Albritton,  D.  L.; Fehsenfeld,  F.  C.  (1986)  Background ozone and
     anthropogenic ozone enhancement at Niwot  Ridge, Colorado. J.  Atmos.  Chem.
     4:  63-80.

PEDCo Environmental, Inc.  (1978) Assessment  of gasoline  toxicity.  Prepared for
     U.S.  Environmental Protection  Agency  under contract no.  68-02-2515.
     Research Triangle Park, NC: Office of Air Quality Planning  and Standards.

Penkett, S. A.  (1983) Atmospheric chemistry:  PAN  in the natural atmosphere.
     Nature (London) 302:  293-294.

Penkett, S. A.; Brice, K.  A. (1986) The spring maximum in photooxidants  in the
     Northern Hemisphere troposphere.  Nature 319: 655-657.

Penkett, S. A.;  Jones,  B.  M.  R.; Brice, K.  A.; Eggleton, A.  E.  J.  (1979) The
     importance of atmospheric ozone and hydrogen peroxide in oxidising  sulphur
     dioxide i.n cloud and rainwater. Atmos.  Environ. 13:  123-137.

Perner, D.; Schmeltekopf, A.; Winkler, R.  H.;  Johnston,  H.  S.; Calvert,  J. G.;
     Cantrell, C.  A.; Stockwell, W.  R. (1985) A laboratory and field study 'of
     the equilibrium N205 5 N03  + N02. J.  Geophys.  Res.  90:  3807-3812.

Peterson, J. T. (1976)  Calculated actinic  fluxes (290-700 nm) for air pollution
     photochemistry applications.  Research Triangle Park,  NC:  U.S.  Environ-
     mental Protection Agency, Environmental  Sciences  Research Laboratory; EPA
     report  no.   EPA-6QO/4-76-025.  Available  from:  NTIS, Springfield,  VA;
     PB-240125.

Pitts, J.  N. ,  Jr.; Winer, A.  M.;  Darnall, K. R.;  Lloyd, A. C.;  Doyle,, G. J.
     (1977) Hydrocarbon reactivity and the  role  of hydrocarbons, oxides of
     nitrogen, and aged smog in the production of photochemical  oxidants. In:
     Dimitriades,  B., ed.  International  conference  on  photochemical  oxidant
     pollution and its  control; September  1976; Raleigh, NC.  Proceedings: v.
     II.  Research Triangle Park,  NC:  U.S.  Environmental  Protection  Agency;
     EPA-600/3-77-001b.

Pitts, J.  N. ,  Jr.; Grosjean,  D. ;  Van  Cauwenberghe, K.;   Schmid,  J. P.; Fitz,
     D. R.  (1978)  Photooxidation of  aliphatic amines under simulated  atmospheric
     conditions:  formation of  nitrosamines,  nitramines,  amides,  and photochemical
     oxidant. Environ.  Sci. Techno!. 12: 946-953.
                                    3-167

-------
Pitts, J. N.,  Jr.;  Darnall, K. R.; Carter, W. P. L.; Winer, A. M.; Atkinson,
     R. (1979)  Mechanisms  of photochemical reactions  in  urban air.  Research
     Triangle  Park,  NC:  U.S.  Environmental Protection Agency;  EPA report  no.
     EPA-600/3-79-110.  Available  from:  NTIS,  Springfield, VA; PB80-132384.

Pitts, J. N.,  Jr.;  Winer,  A. M.; Harris, G. W.; Carter, W. P.  L.; Tuazon,  E.
     C. (1983)  Trace nitrogenous  species in  urban  atmospheres. EHP  Environ.
     Health Perspect. 52: 153-157.

Pitts, J. N.,  Jr.;  Sanhueza, E.; Atkinson,  R.;  Carter,  W. P.  L.; Winer,  A.
     M.; Harris, G.  W.; Plum, C. N. (1984a) An investigation of the dark forma-
     tion of nitrous acid  in environmental  chambers.  Int.  J.  Chem.  Kinet.  16:
     919-939.

Pitts, J. N., Jr.; Biermann, H. W.; Winer, A. M.; Tuazon,  E. C. (1984b)
     Spectroscopic identification  and measurement of gaseous  nitrous  acid in
     dilute auto exhaust. Atmos. Environ. 18: 847-854.

Pitts, J. N., Jr.; Biermann, H. W.; Atkinson, R.; Winer, A. M.  (1984c)
     Atmospheric implications  of  simultaneous nighttime measurements  of  NOs
     radicals and HONO. Geophys. Res.  Lett. 11:  557-560.

Pitts, J. N.,  Jr.; Winer,  A. M.;  Aschmann,  S.  M.;  Carter,  W.  P. L.;  Atkinson,
     R. (1985)  Experimental protocol  for determining  hydroxyl  radical  reaction
     rate constants.  Estimation of atmospheric  reactivity. Research Triangle
     Park: U.S. Environmental Protection Agency; EPA  report no. EPA-600/3-85-000.

Platt, U.;  Perner, D.;  Harris,  G.  W.; Winer,  A.  M.; Pitts, J.  N., Jr.  (1980a)
     Observations of nitrous acid  in  an  urban atmosphere by differential optical
     absorption. Nature 285: 312-314.

Platt, U.;  Perner, D.;  Winer,  A.  M.;  Harris,  G.  W.; Pitts, J.  N., Jr.  (1980b)
     Detection  of NOa in the polluted troposphere by differential optical  ab-
     sorption.   Geophys. Res. Lett. 7:  89-92.

Platt, U.;  Winer, A. M.; Biermann, H.  W.;  Atkinson, R.;  Pitts,  J.  N.,  Jr.
     (1984) Measurement of nitrate radical  concentrations in continental  air.
     Environ. Sci. Techno!. 18: 365-369.

Plegue, J.  (1983) Diesel-car sales plunge  48 pet.  Automotive News (May 2): 1,
     70.

Price, J. H. (1976)  A study of  factors associated with high urban ozone
     concentrations  in  Texas.   In:  Specialty conference on ozone/oxidants—
     interactions with  the total  environment. Pittsburgh,  PA: Air Pollution
     Control Association; pp. 282-292.

Prior, E. J.;  Schiess,  J.  R.;  McDougal, D. S. (1981) Approach to forecasting
     daily maximum ozone levels in St. Louis. Environ. Sci. Technol. 15: 430-436.

Rao, T.  K.;  Young,  J.  A.;  Lijinsky,  W.; Epler,  J.  L. (1979) Mutagenicity of
     aliphatic  nitrosamines  in Salmonella typhimurium. Mutat.  Res.  66:  1-7.
                                    3-168

-------
Rasmussen, R. A.  (1964)  Terpenes:  their analysis and fate in the atmosphere.
     St. Louis,  MO: Washington University.  Ph.D. thesis.

Rasmussen, R. A.  (1970)  Isoprene:  identified  as  a forest-type  emission to the
     atmosphere. Environ. Sci. Technol. 4: 667-671.

Rasmussen, R. A. (1972)  What do the hydrocarbons from trees contribute to air
     pollution?  J. Air Pollut. Control Assoc.  22: 537-543.

Rasmussen, R. A.; Chatfield, R. B.; Holdren, M. W.; Robinson, E. (1976) Hydro-
     carbon  levels  in  a  midwest open-forested area. Technical  report to Coor-
     dinating Research Council, Inc., Atlanta, GA.

Reiter, E. R. (1963)  A case study of radioactive fallout. J. Appl.  Meteorol.
     2: 691-705.

Reiter, E. R. (1975) Stratospheric-tropospheric exchange processes.  Rev.  Geophys.
     Space Phys. 13: 459-473.

Reiter, E. R. (1978) Atmospheric transport processes - 4. Radioactive  tracers.
     Available from: National Technical Information Service, U.S. Department of
     Commerce, Springfield, VA. Report no. TID-27114. (see also, Reiter,  E.  R.
     (1977) The transport of radioactive debris and ozone from  the  stratosphere
     to the ground. Arch. Met. Geophys. Biokl. 26(ser A): 323-340).

Reiter, E. R.; Mahlman, J. D. (1965) Heavy radioactive fallout  over the southern
     United States, November 1962.  JGR J. Geophys. Res. 70: 4501-4519.
                                                                        , j ,
Reynolds, S.   D.  et al. (1979) Photochemical modeling of transportation control
     strategies.  v.I.  Model  development,  performance  evaluation, and strategy
     assessment.  Prepared  for the  Federal Highway Administration,  Office of
     Research, Washington, DC.

Reynolds, S.  D.;  Hogo, J.; Oliver, W.  R.;  Reid,  L,  E.  (1982)  Application of
     the SAI  Airshed Model to the  Tulsa metropolitan area. SAI  No.  82004.  Systems
     Applications, Inc., San Rafael, CA.

Richards, L.  W.;  Anderson,  J.  A.;  Blumenthal, D.  L.;  McDonald, J.  A.; Kok, G.
     L.; Lazrus, A. L. (1983) Hydrogen peroxide and sulfur (IV) in  Los Angeles
     cloud water. Atmos. Environ.  17: 911-914.

Richter,  H.  G.  (1983) Analysis of organic compound data gathered during  1980
     in  northeast corridor cities. Research Triangle Park,  NC:  U.S. Environ-
     mental  Protection Agency,  Office of Air Quality Planning and  Standards;
     EPA  report no.  EPA-450/4-83-017.  Available  from: NTIS, Springfield,  VA;
     PB84-116052.

Richter,  H.  G.  (1986) Monitoring  and  Data  Analysis  Division,  Office  of  Air
     Quality Planning and  Standards,  U.S.  Environmental Protection Agency,
     Research Triangle  Park,  NC.  Personal  communication  to B. E.  Tilton,
     Environmental  Criteria  and  Assessment  Office,  U.S.  Environmental-
     Protection Agency,  Research Triangle Park,  NC.
                                    3-169

-------
Ripperton, L. A.; Worth, J. J. B.; Vukovich, F. M.; Decker, C. E. (1977) Research
     Triangle Institute studies of high ozone concentrations  in  nonurban areas.
     In:  Proceedings of the International Conference on Photochemical Oxidant
     Pollution and  Its  Control,  v.I. Dimitriades,  B. ,  ed.  Research Triangle
     Park,  NC:  U.S.  Environmental  Protection  Agency.  EPA  600/3-77-OOla.
     pp. 413-424.

Roberts, J.  M. ;  Fehsenfeld,  F.  C.;  Albritton,  D.  L.; Sievers,  R.  E.  (1983)
     Measurement of  monoterpene  hydrocarbons  at Niwot Ridge,  Colorado.  JGR J.
     Geophys. Res.  88: 10,667-10,678.

Robinson, E.  (1952)  Some  air  pollution  aspects  of the Los  Angeles temperature
     inversion. Bull. Am.  Meteorol.  Soc. 33: 247-250.

Robinson, E. ;  Robbins,  R.  C.  (1975)  Gaseous  atmospheric pollutants  from urban
     and natural sources.  In:  Singer, S. F., ed. The  changing  global environment.
     Boston: D. Reidel Publ.  Co.

Roth,  P.  M.; Reynolds, S. D.;  Anderson,  G.  E.;  Pollack,  R.  I.; Yooke, M.
     A.; and Killus, J. P. (1976) An evaluation of  methodologies  for assessing
     the impact of oxidant control strategies.  EF76-112R. Systems Applications
     Inc., San Rafael, CA.

Roth,  J. A.;  Sullivan,  D. E.  (1981)  Solubility of ozone in water.  Ind.  Eng.
     Chem. Fundam.  20: 137-140.

Rubino, R. A.; Bruckman,  L.; Magyar, J. (1976)  Ozone  transport.  J.  Air Pollut.
     Control Assoc. 26: 972-975.

Salop, J.; Wakelyn, N. T.; Levy, G.  F.; Middleton,  E. M.; Gervin, J. C.  (1983)
     The application of forest classification from  Landsat  data  as  a basis  for
     natural hydrocarbon  emission estimation and photochemical  oxidant model
     simulations in  Southeastern Virginia.  J.  Air  Pollut.  Control  Assoc. 33:
     17-22.

Sakamaki,  F.;  Hatakeyama, S.;  Akimoto, H.(1983) Formation of nitrous  acid
     and nitric  oxide in  the heterogeneous dark reaction of  nitrogen dioxide
     and water  vapor in a smog  chamber.  Int.  J.  Chem.  Kinet.  15:  1013-1029.

Sanadze, G. A.; Dolidze, G. M. (1962) C5H8 (isoprene) type  in  volatile emission
     from leaves of plants [transl.]. Soobshch. Akad. Nauk  Gruz.  SSR 57: 1222.

Sandberg, J. S.; Basso, M. J.; Okin, B. A. (1978) Winter rain  and summer ozone:
     a  predictive   relationship.  Science  (Washington,  DC) 200:  1051-1054.

Schere,  K.  L.;  Shreffler, J.  H.  (1982a) Evaluation of four urban-scale photo-
     chemical  air  quality simulation models.  Draft.  Office of  Research and
     Development,  Research Triangle  Park, NC:  U.S.  Environmental  Protection
     Agency.

Schere, K. L.; Shreffler, J. H.   (1982b) Final evaluation of urban-scale
     photochemical  air  quality  simulation models.  Draft.  Office of Research
     .and Development,  U.S.  Environmental  Protection Agency, Research Triangle
     Park, NC.


                                   3-170

-------
Scheutzle, D, ; Rasmussen,  R. A.  (1978)  The  molecular composition of secondary
     aerosol   particles  formed  from terpenes. J.  Air Pollut.  Control Assoc.
     28: 236-240.

Schrenk, H. H.; Heimann, H.; Clayton, G. D.; Gafafer, W. M.; Wexler, H. (1949)
     Air pollution  in Donora,  PA:  epidemiology  of the unusual  smog episode of
     October  1948:  preliminary report.  Washington,  DC:  Public  Health Service;
     Public Health Service bulletin no. 306.

Schwartz,  W.  (1974) Chemical  characterization  of model aerosols.  Research
     Triangle Park,  NC:  U.S.  Environmental Protection Agency; EPA  report  no.
     EPA-650/3-74-011.   Available  from:  NTIS,  Springfield,  VA;   PB-238557. '

Schwartz,  S.  E.  (1984)  Gas-aqueous reactions  of sulfur and nitrogen oxides in
     liquid-water clouds.  In:  Calvert, J.   G.,  ed.,  S02, NO and  N02  Oxidation
     Mechamisms:   Atmospheric  Considerations, Chapter 4.   Ann  Arbor Science
     Publishers,   Acid  Precipitation Series,  v.  3.   Boston, MA:   Butterworth
     Publishers,  pp. 173-208.

Scott, W.E.;  Stephens,  E.R.; Hanst,  P.L. ;  Doerr,  R.C. (1957)  Further develop-
     ments in the chemistry of the  atmosphere.  Proc. Amer.  Petrol. Inst.  37
     (III): 171-183.

Seila,  R.  L.  (1981) Nonurban  hydrocarbon concentrations in ambient air north
     of  Houston,  TX. In:  Bufalini,  J.  J.;  Arnts, R.  R.,  eds.   Atmospheric
     biogenic  hydrocarbons. Ann Arbor,  MI:  Ann  Arbor Science  Publishers,  Inc.

Seinfeld,  J.  H. ;  Wilson, K. R.  (1977)  The issue of  air quality simulation
     model utility.  In:  International  Conference on Oxidants,  1976 - Analysis
     of  Evidence  and Viewpoints,  Part  VI.  Research Triangle  Park, NC:   U.S.
     Environmental  Protection  Agency. EPA-600/3-77-118.

Sexton,  K.  (1982) Evidence of  an  additive  effect for small city plumes. Pre-
     sented  at:  the 75th annual meeting of the  Air Pollution  Control Associa-
     tion; June;  New Orleans,  LA.  Pittsburgh, PA: Air Pollution  Control  Asso-
     ciation;  paper no.  82-31.4.  (see  also, Sexton, K. (1983)  Evidence for an
     additive effect for ozone plumes from  small  cities. Environ.  Sci. Technol.
     17: 402-407.)

Sexton,  K.; Westberg, H. (1980)  Elevated ozone  concentrations  measured downwind
     of  the Chicago-Gary urban complex. J.  Air  Pollut.  Control Assoc.  30:  911-
     914.

Sexton,  K.;  Westberg,  H.  (1984) Nonmethane hydrocarbon composition of urban
     and rural atmospheres. Atmos.  Environ.:  in press.

Shafer,  T. B.; Seinfeld, J. H.  (1985)  Evaluation  of chemical reaction
     mechanisms  for photochemical  smog. Part  III.  Sensitivity  of the empirical
      kinetic  modeling approach (EKMA)  to chemical mechanism and  input  parameters.
      Research Triangle  Park, NC:   U.S.  Environmental  Protection  Agency.

Shreffler, J.  H.;  Evans, R. B.  (1982)   The  surface ozone record  from the regional
      air pollution  study,  1975-1976. Atmos.  Environ.  16: 1311-1321.
                                    3-171

-------
Singh, H. B.; Hanst, P. L. (1981) Reroxyacetyl nitrate  (PAN)  in  the  unpolluted
     atmosphere: an important reservoir for nitrogen oxides.  Geophys.  Res.  Lett.
     8: 941-944.

Singh, H. B.; Salas, L. J. (1983) Peroxyacetyl nitrate  in the free troposphere.
     Nature (London) 302: 326-328.

Singh, H.  B.; Ludwig,  F.  L.;  Johnson, W. B.  (1977)  Ozone  in clean  remote
     atmospheres: Concentrations  and variabilities. SRI Project 5661.  Menlo
     Park, CA: Stanford Research Institute (SRI).

Singh, H, B.; Ludwig, F.  L.; Johnson, W.  B. (1978)  Tropospheric  ozone:  concen-
     trations and variabilities in clean remote  atmospheres.  Atmos.  Environ.
     12:  2185-2196.

Singh, H.  B.;  Viezee,  W.; Johnson, W, B.; Ludwig,  F. L. (1980)  The  impact of
     stratospheric ozone  on  tropospheric air  quality.  J. Air Pollut.  Control
     Assoc. 30: 1009-1017.

Singh, H.  B.;  Salas,  L.  J.; Stiles,  R.  E.  (1982)  Distribution  of  selected
     gaseous organic mutagens and suspect carcinogens in ambient air.  Environ.
     Sei. Techno!. 16: 872-879.

Singh, H.  B.; Jaber, H.  M.;  Davenport,  J.  E.  (1984)  Reactivity/volatility
     classification of  selected organic chemicals: Existing  data.  (Prepared
     by  SRI for  U.  S.  EPA). Research Triangle Park,  NC: U.S. Environmental
     Protection Agency; EPA-600/3-84-082.

Slade, D.  H.,  ed.  (1968) Meteorology and atomic  energy 1968.  Washington,  DC:
     U.S.  Atomic Energy  Commission.  Available from:  NTIS,  Springfield, VA;
     TID-24190.

Smith, L.  R.;  Black,  F.  H. (1980) Characterization of  exhaust emissions from
     passenger cars equipped with three-way catalyst control  systems.  Presented
     at  Passenger Car  Meeting,  Society of Automotive  Engineers,  Dearborn,  MI;
     June 9-13.  SAE paper no.  800822.

Soderlund, R.; Svensson,  B. H.  (1976) The global  nitrogen cycle.  In:  Soderlund,
     R.; Svensson, B. H.., eds;  Nitrogen,  phosphorus, and sulfur—global cycles.
     SCOPE report 7. Ecol. Bull (Stockholm) 22: 23-73.

Souza, A. F.; Daley, P. S. (1978) U.S. Air Force  turbine engine  emission survey;
     v.  Ill:  engine  model summaries. Tyndall  Air Force Base', FL: Civil and
     Environmental Engineering  Development Office, Air Force Systems Command;
     CEEDO-TR-78-34.

Spicer,  C.  W.;  Holdren,  M. W.; Keigley,  6. W.  (1983) The ubiquity of peroxy-
     acetyl nitrate  in the continental  boundary  layer. Atmos.  Environ. 17:
     1055-1058.

Spicer,  C. W.; Joseph, D. W.; Sticksel,  P. R.; Ward, G.  F.  (1979) Ozone sources
     and transport  in  the northeastern  United States.  Environ.  Sci.  Techno!.
     13: 975-985.
                                    3-172

-------
Spicer, C. W.;  Joseph,  D.  W.; Sticksel, P. R. (1982) An investigation of the
     ozone plume from a small city. J. Air Pollut. Control Assoc. 32: 278-281.

Springer, K.  J.; Stahman, R. C. (1977a) Diesel car emissions—emphasis on parti-
     culate and sulfate. Presented at: the international automotive engineering
     congress and exposition; February-March; Detroit, MI. Warrendale, PA:  Society
     of Automotive Engineers; SAE paper no, 770254.

Springer, K.  J.; Stahman, R. C. (1977b) Unregulated emissions from diesels  used
     in trucks  and buses.   Presented at:  International  automotive engineering
     congress  and  exposition; February-March;  Detroit,  MI.  Warrendale,, PA:
     Society of Automotive Engineers; SAE paper no. 770258.

Stair,  R.  (1961)  The spectral radiant  energy from the sun through  varying
     degrees of smog at Los Angeles.  In: Proceedings of the third national  air
     pollution symposium; 1955; Pasadena, CA. pp. 48-54.

Stelson, A. W.; Friedlander, S. K.; Seinfeld, J.  H. (1979) A note on  the equi-
     librium  relationship  between  ammonia  and nitric acid and particulate
     ammonium nitrate. Atmos. Environ. 13: 369-371.

Stephens, E.  R. (1964) Absorptivities for infrared determination  of peroxyacyl
     nitrates. Anal. Chem. 36: 928r929.

Stephens, E.  R. (1967) The formation  of molecular oxygen by alkaline  hydrolysis
     of peroxyacetyl nitrate. Atmos.  Environ. 1:  19-20.

Stephens,  E.  R.  (1969)  The  formation, reactions,  and properties of peroxyacyl
     nitrates  (PANs)  in photochemical air pollution.  In:  Pitts,  J.  N.,  Jr.;
     Metcalf,  R.  L.,  eds.  Advances  in environmental sciences v.  1. New York,
     NY: Wiley-Interscience; pp. 119-146.

Stephens, E.  R.; Burleson, F. R.;  Holtzclaw,  K. M. (1969) A damaging  explosion
     of peroxyacetyl nitrate.  J.  Air Pollut. Control Assoc. 19:  261.

Sticksel,  P.  R.;.  Spicer, C. W.; Westberg,  H. (1979)  Ozone plumes from small
     cities and airborne observations of ozone  in a moving high pressure system.
     Presented at: the 72nd annual  meeting of the Air Pollution Control Associa-
     tion; June; Cincinnati, OH. Pittsburgh,  PA:  Air Pollution  Control Associa-
     tion; paper no. 79-58.6

Stockwell, W.  R.;  Calvert,  J.  G.  (1983)  The mechanism of the  H0-S02 reaction.
     Atmos. Environ. 17: 2231-2235.

Su, F.; Calvert, J. G.; Shaw, J. H.  (1980) An FT-IR spectroscopic study of  the
     ozone-ethene  reaction  mechanism  in  02-rich mixtures.  J.  Phys. Chem.
     84: 239-246.

Sze, N. D.; Rice, H. (1976) Nitrogen  cycle factors contributing to  N20
     production from fertilizers.  Geophys. Res. Lett. 3: 343-346.

Tiao,  G. C.; Box, G. E.  P.; Hamming,  W. J.  (1975) Analysis of  Los Angeles photo-
     chemical  smog data:  a statistical overview. J. Air Pollut.  Control Assoc.
     25: 260-268.


                                    3-173

-------
Tilden, J.  W.;  Seinfeld, J, H. (1982) Sensitivity  analysis  of  a  mathematical
     model  for  photochemical  air  pollution.  Atmos. Environ. 16:  1357-1364.

Tilton, B,  E.;  Bruce, R. M. (1981) Review of criteria  for vapor-phase  hydro-
     carbons.  Research  Triangle  Park,  NC:  U.S. Environmental  Protection
     Agency: EPA report  no.  EPA-600/8-81-022. Available from: NTIS,  Springfield,
     VA; PB82-136516.

Tingey, D. T.; Burns, W. F.  (1980) Hydrocarbons  emissions from  vegetation.  In:
     Proceedings,  symposium on effects of air pollutants on Mediterranean and
     temperate forest ecosystems, Riverside, CA, June 22-27.

Tingey, D.  T.;  Manning,  M.; Grothaus,  L. C.;  Burns, W.  F.  (1980) Influence of
     light and temperature  on  monoterpene emission  rates from slash  pine.  Plant
     Physio1, 65:  797-801.

Tingey, D. T.; Evans, R.; Gumpertz, M. (1981) Effects of environmental  conditions
     on isoprene emission from live oak. Planta  152: 565-570.

Tuazon, E.  C.;  Graham,  R.   A.;  Winer, A. M.; Easton, R.  R.;  Pitts, J. N.,  Jr.
     Hanst,  P.  L.  (1978a)  A kilometer pathlength  Fourier-transform infrared
     system  for  the study  of trace pollutants in ambient and synthetic atmos-
     pheres. Atmos.  Environ. 12: 865-875.

Tuazon, E.  C.;  Winer, A. M.;  Grahami R. A.; Schmid, J,  P.;  Pitts, J. N.,  Jr.
     (1978b) Fourier transform infrared detection of nitramines in irradiated
     amine-NQ  systems.  Environ. Sci. Techno!. 12:  954-958.
             f\

Tuazon, E. C.; Winer, A. M.; Graham, R.  A.; Pitts,  J. N., Jr. (1980) Atmospheric
     measurements  of trace  pollutants by kilometer-pathlength FT-IR  spectroscopy.
     In:  Pitts,  J. N.,  Jr.; Metcalf,  R. L.;  Grosjean,   D.,  eds.  Advances  in
     environmental  science  and technology: v. 10. New York,  NY:  John Wiley and
     Sons; pp. 259-360.

Tuazon, E. C.; Winer, A. M.; Pitts, J. N., Jr. (1981a)  Trace pollutant  concen-
     trations in  a multiday smog episode  in  the California South Coast Air
     Basin by long-path!ength  Fourier-transform  infrared spectroscopy.  Environ.
     Sci. Techno!.  15: 1232-1237.

Tuazon, E. C.; Carter, W. P. L.; Winer,  A. M.; Pitts, J. N., Jr.  (1981b) Reac-
     tions  of  hydrazines with  ozone under  simulated atmospheric conditions.
     Environ. Sci.  Techno!.  15: 823-828.

Tuazon, E. C.; Carter, W. P. L.; Brown,  R. V.; Atkinson, R.; Winer,  A.  M.; Pitts,
     J. N.,  Jr.  (1982)  Atmospheric reaction  mechanisms of amine fuels. U.S.
     Air  Force final  report, ESL-TR-82-17.

Tuizon, E.  C.;  Atkinson, R.;  Plum,  C.  N.;  Winter, A.  M.;  Pitts, J.  N., Jr.
     (1983)  The reaction of gas phase  N205 with  water vapor. Geophys.  Res.  Lett.
     10:  953-956.

Tuazon, E.  C.;  Sanhueza, E.;  Atkinson,  R.;  Carter, W.   P.  L.;  Winer, A. M.;
     Pitts,  J. N.,  Jr. (1984)  Direct determination  of the equilibrium constant
     at 298  K for  the  N02 + N03 -*  N205 reactions. J. Phys.  Chem.  88:  3095-3098.


                                    3-174

-------
U.S. Department of Commerce. (1968) Climatic atlas of the United States. Asheville,
     NC: National Climatic  Center.  [Reprinted (1977) by the National Oceanic
     and Atmospheric Administration.]

U.S. Department  of  Energy.  (1978) Energy data  reports. Washington,  DC:  U.S.
     Department of Energy; publication no. EIA-Q049/1.

U.S. Department  of  Health,  Education,  and Welfare (1970) Air quality criteria
     for photochemical oxidants. Washington, DC: National Air Pollution  Control
     Administration; NAPCA report no. AP-63. Available from: NTIS, Springfield,
     VA; PB82-242231.

U.S. Environmental  Protection  Agency.  (1977) Uses, limitations and  technical
     basis of procedures  for quantifying relationships between photochemical
     oxidants and precursors. Research Triangle Park, NC: Office of  Air  Quality
     Planning and Standards Report no. EPA-450/2-77-021a.

U.S. Environmental  Protection  Agency.  (1978a) Air quality criteria  for  ozone
     and  other  photochemical  oxidants.  Research  Triangle  Park,  NC:  U.S.
     Environmental  Protection  Agency,  Environmental Criteria and  Assessment
     Office; pp.  116-135; EPA  report  no.  EPA-6QO/8-78-004.  Available from:
     NTIS, Springfield, VA; PB80-124753.

U.S.  Environmental   Protection  Agency.  (1978b) Procedures  for quantifying
     relationships  between  photochemical  oxidants and precursors: supporting
     documentation.   Research Triangle Park, NC: Office of Air Quality  Planning
     and Standards;  EPA report  no. EPA-450/2-77-021b.

U.S. Environmental  Protection  Agency.  (1980a) Procedures for the preparation
     of emission  inventories  for volatile organic  compounds:  v.  1.  2nd ed.
     Research Triangle Park, NC: U.S.  Environmental  Protection Agency; EPA
     report  no.  EPA-450/2-77-Q28.  Available from:  NTIS,  Springfield,  VA;
     PB-275292.

U.S. Environmental  Protection Agency. (1980b) Final emission inventory require-
     ments for  1982 ozone state implementation  plans.  Research  Triangle Park,
     NC:  U.S. Environmental Protection Agency; EPA report no.  EPA-450/4-80-016.
     Available from: NTIS, Springfield, VA; PB81-060434.

U.  S. Environmental  Protection  Agency. (1980c) Guideline for use of  city-specific
     EKMA  in  preparing ozone SIPs.  Research Triangle Park,  NC: U.S.  Environ-
     mental Protection Agency;  EPA report no. EPA-450/4-80-027.

U.S. Environmental  Protection  Agency.  (1981a) Comparative application of the
     EKMA  in  the Los Angeles area.  Research Triangle Park,  NC: U.S.  Environ-
     mental  Protection Agency,  Office of Planning and  Standards;  EPA report
     no. EPA-450/4-81-030d. Available from: NTIS,  Springfield,  VA; PB82-234204.

U.S. Environmental  Protection  Agency.  (1981b)  Procedures for emission  inven-
     tory preparation, v. I: emission  inventory fundamentals. Research Triangle
     Park, NC:  Office of Air Quality  Planning  and Standards;  EPA report no.
     EPA-450/4-81-026a.  Available from:  NTIS,  Springfield,  VA; PB82-240102.
                                    3-175

-------
U.S. Environmental  Protection  Agency.  (1981c) Procedures for emission inven-
     tory preparation, v. II: point sources. Research Triangle Park, NC: Office
     of Air Quality Planning and Standards; EPA report no. EPA-450/ 4-81-026b.
     Available from: NTIS, Springfield, VA; PB82-240110.

U.S. Environmental  Protection  Agency.  (198ld) Procedures for emission inven-
     tory preparation, v. Ill:  area sources. Research Triangle Park, NC: Office
     of Air Quality Planning and Standards; EPA report no. EPA-450/ 4-81-026c.
     Available from: NTIS, Springfield, VA; PB82-240128.

U.S. Environmental  Protection  Agency.  (1981e) Procedures for emission inven-
     tory preparation,  v. IV: mobile  sources.  Research Triangle Park,  NC:
     Office  of  Air  Quality Planning  and  Standards;   EPA  report  no.
     EPA-450/4-81-026d.  Available  from:  NTIS, Springfield, VA;  PB82-240136.

U.S. Environmental  Protection  Agency.  (1981f) Procedures for emission inven-
     tory preparation, v. V: bibliography.  Research Triangle  Park,  NC:  Office
     of Air Quality Planning and Standards; EPA report no. EPA-450/ 4-81-026e.
     Available from: NTIS, Springfield, VA; PB82-240144.

U.S. Environmental  Protection Agency.  (1981g)  The  sensitivity  of complex
     photochemical model  estimates to detail in  input  information.  Research
     Triangle  Park,  NC:  U. S.  Environmental  Protection  Agency, EPA  report no.
     EPA-450/4-81-031a.

U.S. Environmental  Protection  Agency.  (1982a) Air quality criteria  for  oxides
     of nitrogen.  Research Triangle Park, NC: U.S.  Environmental Protection
     Agency; EPA-600/8-82-026. Available  from:  NTIS, Springfield,  VA;  PB83-
     131011.

U.S. Environmental Protection Agency. (1982b) Air quality criteria  for particu-
     late matter  and sulfur  oxides.  Research Triangle Park, NC:  U.S.  Environ-
     mental Protection Agency; EPA-600/8-82-029. Available from:  NTIS, Springfield,
     VA; PB84-156777.

U.S. Environmental Protection Agency (1984a).  National  air pollutant emission
     estimates,  1940-1983.  Research Triangle  Park,  NC:  U.S.  Environmental
     Protection Agency.  EPA-450/4-84-028.

U.S. Environmental  Protection  Agency (1984b) Guideline  for using the carbon
     bond mechanism  in city-specific EKMA. Research  Triangle  Park, NC:  U.S.
     Environmental Protection Agency; EPA report no.  EPA-450/4-84-005.

U.S. Environmental  Protection  Agency (1984c) User's  manual for OZIPM2:  Ozone
     isopleth  plotting with  optional mechanisms/version 2. Research Triangle
     Park,  NC:  U.S.   Environmental  Protection  Agency;  EPA  report  no.
     EPA-450/4-84-024.

U.S. Environmental  Protection  Agency (1986) National air pollutant emission
     estimates,  1940-1984.  Research Triangle  Park,  NC:  U.S.  Environmental
     Protection Agency.  Report no. EPA-450/4-85-014.
                                   3-176

-------
Vaughan, W. M.; Chan, M.; Cantrell, B.; Pooler, F. (1982) A study of persistent
     elevated pollution episodes  in the Northeastern United States.  Bull.  Am.
     Meteorol. Soc. 63: 258-266.

Viezee, W.;  Singh,  H.  B.  (1982) Contribution of stratospheric ozone  to ground
     level ozone  concentrations—a scientific  review of existing evidence.
     Final report  (revised  draft).  EPA grant CR  809330010.  Menlo Park, CA:
     SRI International.

Viezee, W.; Johnson, W. B.; Singh, H.   B. (1979) Airborne measurements  of strato-
     spheric ozone  intrusions into the  troposphere over the United States:  final
     report.   Atlanta,  GA: Coordinating Research  Council;  SRI project  6690.

Viezee, W.; Singh,  H. B.; Shigeishi, H. (1982) The impact of  stratospheric  ozone
     on tropospheric air  quality—implications from  an analysis  of  existing
     field data:  final report. Atlanta,  GA:  Coordinating Research Council.
     Available from: NTIS, Springfield, VA;  PB82-256231.

Viezee, W.; Johnson, W. B.; Singh, H.   B. (1983) Stratospheric ozone  in the  lower
     troposphere-II. Assessment of downward  flux  and ground-level impact. Atmos.
     Environ. 17: 1979-1993.

Vukovieh,  F.  (1977) Review  and analysis.  In: International  conference on oxi-
     dants,  1976~-Analysis  of evidence and  viewpoints.  Part V.  The  issue of
     oxidant transport. Research  Triangle  Park, NC.  U.S. Environmental  Protec-
     tion  Agency;   EPA  report no. EPA-600/3-77-117.  Available from:  NTIS,
     Springfield, VA; PB-278750.

Wellington,  T.  J.;  Atkinson,  R.; Winer,  A.  M.  (1984) Rate constants for the
     gas  phase  reaction of  OH radicals with peroxyacetyl  nitrate (PAN) at 273
     and 296 K. Geophys.  Res.  Lett, 11: 861-864.

Weast,  R.  C.,  ed.  (1977)  CRC  handbook of  chemistry and physics.  57th ed.  Boca
     Raton,  FL: CRC Press; p.  D-194.
                              *
Wells,  C.  G.  (1981).   Methods of  estimating foliar biomass. In:   Bufalini, J.
     J.; Arnts, R.  R.,  eds. Atmospheric biogenic  hydrocarbons, v. 1.   Ann  Arbor,
     MI: Ann Arbor  Science; pp. 107-114.

West,  E.  S.; Todd, W.  R.; Mason,  H.  S.; Van Bruggen, J.  T.  (1966) Textbook of
     Biochemistry,  4th  Ed. New York,  NY; The Macmillan Co.

Westberg,  H.  (1977) Analysis of  evidence  and viewpoints.  Part IV.  The issue
     of natural organic emissions. International  Conference on Oxidants,  1976.
     Research Triangle  Park,  NC:  U.S.  Environmental  Protection Agency.  EPA report
     no.  EPA  600/3-77-116.

Westberg,  H.;  Lamb, B. (1983) Ozone  production and  transport in  the Atlanta,
     GAj  region.  Final report; EPA. grant no.  CR 809221.  Research  Triangle
     Park,  NC:  U.S. Environmental  Protection  Agency, -Atmospheric  Sciences
     Research  Laboratory.
                                    3-177

-------
Westberg, H.; Allwine, K. J.; Elias, D.  (1976) Vertical ozone distribution above
     several  urban and adjacent  rural  areas across the United  States.  In:
     Specialty  conference  on  ozone/oxidants—interactions  with the  total
     environment.  Pittsburgh, PA: Air Pollution Control Association; pp. 84-95.

Westberg, H.; Allwine,  K. J.; Robinson,  E.  (1978a) The transport of oxidant
     beyond  urban  areas:  light hydrocarbon and oxidant  data, New England Study,
     1975.  Research  Triangle Park,  NC:   U.S. Environmental Protection  Agency,
     Office  of  Research and Development;  EPA  report no.  EPA-6007 3-78-006.
     Available from: NTIS, Springfield,  VA;  PB-277043.

Westberg, H.; Allwine, K.; Robinson, E.  (1978b) Measurement  of light hydrocarbons
     and  studies  of oxidant transport—Houston study 1976.  Research Triangle
     Park,  NC:   U.S.  Environmental  Protection Agency,  Environmental Sciences
     Research Laboratory; EPA report no.  EPA-600/3-78-062. Available from: NTIS,
     Springfield,  VA; PB-285891.

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.

Whitby, R. A.; Coffey, P. E. (1977) Measurement of  terpenes  and  other  organics
     in an Adirondack mountain pine forest.  JGR J.  Geophys.  Res.  82: 5928-5934.

Whitten,  G.  Z.  (1983) The chemistry of smog formation: a review of current
     knowlege. Environ.  Int. 9: 447-463.

Whitten,  G.  Z.;  Gery, M. W.(1986) Development  of CBM-X mechanisms for urban
     and  regional  AQSM's.  Prepared under  EPA contract  no.  68-02-3738 (M.  C.
     Dodge,  Project  Officer) by Systems  Applications,  Inc., San Raphael, CA.
     Research Triangle Park, NC:  U.S. Environmental  Protection Agency.

Whitten,  G.  Z,; Hogo, H.  (1977) Mathematical modeling of simulated photochemical
     smog.  Research  Triangle Park,  NC:  U.S. Environmental Protection  Agency.
     Report  no. EPA-600/3-77-011.

Whitten,  G.  Z.;  Hogo, H. (1978)  User's manual' for  kinetics model and ozone
     isopleth plotting package. Research Triangle Park, NC:  U.S.  Environmental
     Protection Agency;  EPA  report no.  EPA-600/8-78-014a.

Whitten,  G.  Z.;  Hogo, H.  (1981) Application of the  empirical kinetic modeling
     approach to  urban areas,  v. II: Tulsa. Research Triangle Park, NC: U.S.
     Environmental  Protection Agency. EPA report no. EPA-450/4-81-005b.

Wight, G. D.; Wolff,  G. T.;  Lloy, P. J.;  Meyers, R.  E.; Cederwall, R.  T. (1978)
     Formation and transport of ozone  in the northeast quadrant of the United
     States.  In:  Morris, A. L.;  Barras,  R.  C.;  eds. Air quality meteorology
     and  atmospheric  ozone.  Philadelphia,  PA: American  Society for Testing and
     Materials; ASTM  STP  653.

Wilson, W.  E.  (1978) Sulfates  in the atmosphere:  a progress report on project
     MISTT.  Atmos.  Environ.  12: 537-548.
                                    3-178

-------
Wine, P. H.; Kreutter, N. M.; Gump, C. A.; Ravishankara,  A.  R.  (1981)  Kinetics
     of OH reactions with the atmospheric sulfur compounds  H2S,  CH3SH,  CH3SCH3
     and CH3SSCH3. J. Phys.  Chem. 85: 2660-2665,

Winer, A. M. (1983) Investigation of the role of natural  hydrocarbons  in photo-
     chemical smog formation in California:  final  report. Contract  no.  AO-056-32.
     Sacramento, CA: California Air Resources Board.

Winer,  A.  M.;  Lloyd,  A.  C.;  Darnall,  K.  R.;  Pitts, J.  N., Jr.  (1976) Relative
     rate  constants  for  the reaction of the hydroxyl  radical with  selected
     ketone, chloroethenes  and monoterpene  hydrocarbons. J.  Phys.  Chem. 80:
     1635-1639.

Winer,  A.  M.;  Breuer, G. M.; Carter, W. P.  L.; Darnall,  K.  R.;  Pitts,  J.  N.,
     Jr.  (1979) Effects  of  ultraviolet  spectral  distribution on the  photo-
     chemistry  of simulated  polluted atmospheres.  Atmos.  Environ. 13:  989-998.

Winer,  A.  M.;  Atkinson,  R.;  Pitts,  J.  N.,  Jr.  (1984) Gaseous nitrate radical:
     possible  nighttime  atmospheric  sink for  biogenic organic  compounds.
     Science (Washington, DC) 224: 156-159.

Wolff,  G.  T.;  Lioy, P.  J.  (1978)  An empirical  model for forecasting maximum
     daily ozone levels  in the northeastern  U.S. J.  Air Pollut.  Control Assoc.
     28: 1034-1038.

Wolff,  G. T.; Lioy, P. J. (1980) Development of an ozone  river associated with
     synoptic-scale episodes in the eastern  United States.  Environ.  Sci. Techno!.
     14: 1257-1260.

Wolff,  G.  T.; Lioy, P. J.; Meyers, R. E.; Cederwall, R. T.; Wight,  G.  D.; Pasceri
     R;  E.;  Taylor, R.  S. (1977a) Anatomy of two  ozone transport episodes in
     the Washington,  D.C.,  to  Boston,  Mass., corridor. Environ. Sci. Techno!.
     11: 506-510.

Wolff,  G.  T.; Lioy, P. J.; Wight, G. D.; Meyers,  R.  E.; Cederwall,  R.  T. (1977b)
     An investigation of long-range transport  of  ozone across the  midwestern
     and eastern United  States. Atmos.  Environ. 11:  797-802.

Wolff,  G.  T.;  Lioy, P. J.;  Wight,  G.  D.; Pasceri, R. E. (1977c) Aerial  inves-
     tigation  of  the ozone  plume phenomenon.  J.  Air  Pollut.  Control  Assoc.
     27: 460-463.

Wolff,  G.  T.;  Lioy,  P. J.; Wight, G. D.;  Pasceri,  R. E. (1977d) Aerial  investi-
     gation  of  photochemical oxidants over the  northeast.  In:  Bufalini, J. J.;
     Lonneman,  W.  A., eds.   Proceedings of symposium on 1975 northeast  oxidant
     transport  study.  Research Triangle  Park,  NC:  U.S.  Environmental Protection
     Agency. Report no:  EPA-600/3-77-017.

Wolff,  G.  T. ;  Ferman,  M. A.;  Monson,   P.  R.   (1979)  The distribution  of
     beryl!ium-7  within  high-pressure systems  in  the eastern United States.
     Geophys.  Res.  Lett. 6: 637-639.  (See  also "Correction,"  Geophys. Res.
      Lett.  6:  816,  1979.)
                                    3-179

-------
Wolff, G.  T.;  Lioy,  P. J.; Wight, G. D. (1980) Transport of ozone associated
     with an air mass. J.  Environ. Sci. Health. A15: 183-199.

Wolff, G.  T.;  Kelly,  N.  A.;  Ferman,  M.  A.  (1981)  On the sources  of summertime
     haze  in the  eastern United States.  Science (Washington,  DC) 211:703-706.

Wolff, G.  T.;  Kelly,  N.  A.;  Ferman,  M.  A.  (1982)  Source regions  of summertime
     ozone and  haze  episodes in the  eastern  United States.  Water, Air,  Soil
     Pollut. 18: 65-81.

Zika, R.;  Saltzman,  E.;  Chameides, W.  L.; Davis, D. D.  (1982) H202  levels  in
     rainwater collected in South Florida and the Bahama Islands.  JGR J. Geophys.
     Res: 87: 5015-5017.

Zimmerman, P.  R.  (1977)  Procedures  for conducting hydrocarbon emission  inven-
     tories  of biogenic sources and  some  results of:recent  investigations.
     In:  Emission inventory/factor workshop,  v.  2.  Research Triangle Park,
     NC: U.S.  Environmental  Protection Agency,  Office  of Air Quality Planning
     and  Standards;   EPA  report.  Available  from:  NTIS, Springfield, VA;
     PB-289123.

Zimmerman, P.  R.  (1978) Testing of Hydrocarbon Emissions from Vegetation and
     Development  of  a Methodology for Estimating Emission Rates  from Foliage.
     Final  Report,  Draft.  Prepared by  Washington State University  for  U.S.
     Environmental  Protection  Agency under•Contract  No.  DU-77-C063. U.S.
     Environmental Protection Agency, Research Triangle  Park, NC.

Zimmerman, P.  R.  (1979) Determination  of emission  rates of hydrocarbons from
     indigenous species  of vegetation  in the  Tampa/St.  Petersburg,  Florida,
     area: final  report; appendix C.  Research Triangle  Park,  NC:  U.S. Environ-
     mental Protection Agency;  EPA report no. EPA-904/9-77-028.  Available from:
     NTIS, Springfield,  VA; PB-297057.

Zimmerman, P. R.  (1980)  Natural  sources of ozone  in  Houston:  natural  organics.
     In:   Proceedings of specialty conference on ozone/oxidants—interactions
     with  the  total  environment; October 1979;  Houston, TX.  Pittsburgh, PA:
     Air Pollution Control Association; pp. 299-310.
                                    3-180

-------
                4.   SAMPLING AND MEASUREMENT OF OZONE AND OTHER
                    PHOTOCHEMICAL OXIDANTS AND THEIR PRECURSORS
4.1  INTRODUCTION
     Detailed information is presented in this chapter on methods for sampling
and measuring  ozone,  "total oxidants,"  hydrogen  peroxide,  and peroxyacetyl
nitrate and  its  higher homologues.   Because of their utility  in atmospheric
research and in  the  application of oxidant abatement programs,  methods for
sampling and measuring the  organic and  inorganic precursors to oxidants are
described as well.  The information presented here should  prove  helpful  to
state and local air pollution agencies and to researchers investigating health
and welfare effects.   The chief reason for presenting such information, however,
is to provide relevant information: (1) for assessing the accuracy of aerometric
data on these  pollutants;  and (2) for determining  the  impact of respective
measurement and calibration methods on existing data on the health and welfare
effects of  ozonej  total  oxidants,  and  individual  other oxidants.  Primary
emphasis is  placed in this  chapter on techniques considered satisfactory for
routine monitoring, on the  effects of changes in calibration  procedures for
ozone  measurements,  on the  relationship between  ozone  and "total oxidant"
measurements, and on  the  accuracy and reliability  of methods  for measuring
oxidants not routinely monitored in ambient air.
     Since the publication  of the 1978  criteria  document on ozone and other
photochemical oxidants  (U.S.  Environmental  Protection Agency, 1978a),  a new
procedure for  calibrating ozone measurements has been promulgated by  EPA as
the Federal  Reference Method for  calibration.  In addition, EPA has  continued
efforts to institute and codify a formal  nationwide program of quality assurance
for the routine  monitoring  of pollutants in  ambient  air.   Some examples of
these procedures are documented in this chapter as they apply to actual opera-
tion of the  analytical instrumentation.   Detailed descriptions of analytical
procedures,  quality  assurance  procedures,  and reporting  requirements are
contained in Quality  Assurance  Handbook  for Air Pollution Measurement  Systems
(U.S. Environmental Protection Agency, 1977b).  Pertinent rules and regulations
are contained in the Federal Register (1979a,b,c,d,e).
                                    4-1

-------
     Brief summaries are provided below of requirements pertaining to quality
assurance and sampling for ozone monitoring.
4.2   QUALITY ASSURANCE  AND OTHER SAMPLING FACTORS  IN  MONITORING  FOR
4.2.1  Quality Assurance in Ambient Air Monitoring for Ozone
     Quality assurance as defined by EPA rules and regulations consists of two
distinct functions.  One  is the assessment of the quality of monitoring data
by estimating their precision and accuracy.  The  other is the control and
possible improvement of the quality of the  ambient air  data by implementation
of quality control policies, procedures,  and corrective actions.
     Each quality control   program, developed by the individual  States and
approved by the EPA Regional Administrator, must include operational procedures
for each of the following activities:

      1.  Selection of methods, analyzers, or samplers (prescribed refer-
          ence and  equivalent methods for ambient air monitoring  are
          described elsewhere in this chapter);
      2.  Installation of equipment;
      3.  Calibration (test concentrations  for ozone must  be  obtained by
          means of the ultraviolet (UV) photometric calibration procedure
          described elsewhere  in  this chapter or by  means  of  a certified
          ozone transfer standard);
      4.  Zero/span checks and adjustments of automated analyzers;
      5.  Control checks and their frequency;
      6.  Control limits for zero,  span, and other  control checks, and
          respective corrective actions  when such limits are surpassed;
      7.  Calibration and  zero/span checks  for multiple range analyzers;
      8.  Preventive and remedial  maintenance;
      9.  Quality control procedures for air pollution episode monitoring;
     10.  Recording and validation of data;
     11.  Documentation of  quality control  information.

     A one-point precision  check must be carried out at least every 2 weeks on
each automated analyzer  used for ozone,  using a precision test gas of  known
concentration.   Each calendar quarter,  at  least 25  percent of the  analyzers
                                    4-2

-------
used by the  State  and Local Air Monitoring  Stations  (SLAMS) for monitoring
ozone must be formally audited by an independent operator by challenging with
at least one audit gas of known concentration in each of the four concentration
ranges.   Similar requirements  are  set  forth for monitoring networks designed
to assess Prevention of Significant Deterioration (PSD) requirements.
     In addition  to requirements and  recommendations associated  with the
selection, installation,  and maintenance of monitoring equipment, the above-cited
Federal  Register publications discuss  certain  design criteria for monitoring
networks (SLAMS and the National  Aerometric Monitoring Stations, NAMS).  Included
are requirements on siting of monitors  in order to obtain ozone concentrations
that are  representative  of regions of varying  dimensions.   For example,  a
"middle scale" monitor would represent  conditions close to sources of NO  such
                                                                        /v
that local ozone scavenging effects might be of significance.  A "neighborhood
scale" monitor, on  the other hand, would be  located somewhere  in a  reasonably
homogeneous  urban  subregion having dimensions  of  a few kilometers.   Other
"scales" applicable to siting  of ozone monitors include  urban scale,  which
would be  used to estimate  concentrations  characteristic  of an area having
dimensions between several and 50 kilometers or to measure high concentrations
downwind of an area with high precursor emissions;  and regional scale,  used to
typify concentrations  over portions of a  major metropolitan complex  up to
dimensions of hundreds of kilometers.   For  ozone SLAMS stations, applicable
scales are middle,  neighborhood,  urban, and regional.  Requirements for NAMS
stations for ozone  are neighborhood and urban scale.   Two ozone  NAMS stations
are expected to be sufficient for each urban area:   one for specific transport
conditions leading to high ozone; and the other for monitoring peak concentra-
tions relative to population exposure.

4.2.2  SamplingFactors in Ambient Air Monitoring for Ozone
     Sampling factors may have a crucial effect on the quality and  utility of
measurements both  in ambient  air  and  in controlled  laboratory  situations.
Sampling techniques and strategies must preserve the  integrity of a represen-
tative fraction of ambient air and must be consistent with the specific purpose
of the measurement.  In this section, the significance of some sampling factors
will be discussed briefly.  For more detailed discussions of this subject,  the
reader is referred to Ott (1977) and to reports prepared for EPA by the National
Research Council (U.S. Environmental Protection Agency, 1977a; National Research
Council, 1977).
                                    4-3

-------
4.2.2.1   Sampling  Strategies  and Air Monitoring  Needs.   Air monitoring data
relevant to assessing ambient Og or oxidant levels are collected for a variety
of specific needs, including:

     1.   Data to be used in trend analysis as indicators of the state of
          attainment of ambient air quality standards.
     2.   Data to  be  used in development of Og  control  strategies and
          evaluation of their effectiveness.
     3.   Data to be used in the development and validation of air quality
          simulation models  capable of application to  the Qg problem.
     4.   Data to be used in investigation of causes of the ozone problem
          both in general and in specific localities.
     5.   Data to be used in special research studies such as the effects
          of ambient air pollution on human health and welfare.

     Each specific purpose or need requires special considerations with regard
to air sampling  strategy.   For example, 5  or more years of On data might be
required for the adequate assessment of trends that resulted from the applica-
tion of  a particular  control strategy rather than trends  that resulted from
chance local meteorological conditions.  In contrast,  the validation of an air
quality  simulation model  might require only a few carefully  chosen days of
very detailed measurements  of 0,,  hydrocarbons,  and NO , as well as detailed
meteorological data and time-varying emissions along the trajectory of the air
parcel in question.
4.2.2.2   Air Monitoring Site Selection.  Ozone in the lower troposphere is a
product  of  photochemical  reactions  that involve  sunlight, hydrocarbons, and
oxides of nitrogen.   In  typical urban atmospheres, ozone precursors react to
produce  ozone at such  a rate that the 0, reaches its daily peak level in the
middle of the day  at locations  downwind from the source-intensive  center-city
area.  Thus, if peak Og concentrations are to be measured, monitoring stations
should,  in general,  be located downwind  from  city centers.   This downwind
distance may be on the order of 15 to 30 kilometers (9 to 19 miles), depending
on predominant wind patterns in the area (U.S.  Environmental Protection Agency,
1977a).   This distance may be  highly  area-specific,  however.   For example,
ozone maxima in the Los Angeles plume have been observed as far downwind as 50
to 70 km.
                                    4-4

-------
     Once a  station is  located,  additional sampling  considerations, arise
because of the chemical  reactivity and instability of the (k molecule.  Ozone
reacts extremely rapidly with NO and with some hydrocarbon compounds, including
most of those emitted  by vegetation.  Also, 0, decomposes readily on contact
with the surface of many materials.   Consideration of these effects led to the
development of specific  criteria  for locating an Qg monitoring station (U.S.
Environmental Protection Agency,  1977a;  National  Research Council, 1977).
Briefly, the  inlet of the  sampling  probe of the ozone  analyzer  should be
positioned 3 to  15 meters (10 to 49 feet)  above ground, at least 4 meters
(13 feet) from large trees,  and 120 meters (349 feet) from heavy automobile
traffic.  Sampling  probes should be  designed so  as to minimize 0, destruction
by surface reaction or by reaction with NO.
     Another consideration that  has  significance for the selection of  sites
for air monitoring stations  is  the fact that ambient .monitoring  data, as
routinely obtained, have  limited validity as absolute measures of air quality.
This limitation arises from the fact that, at ground level, the ambient atmos-
phere is inhomogeneous as a  result of a  continuous influx of fresh emissions,
incomplete mixing,  and destruction of 0^ by fresh and unreacted emissions  and
destruction on surfaces.  In view of such inhomogeneity, monitoring data from
a fixed  network provide  measures of  air  quality  at a discrete number of loca-
tions but may  not detect temporal and spatial variations in ozone concentra-
tions of  a localized nature.  This  problem can be alleviated by  use  of a
greater density  of monitoring stations or  by  use of a validated air quality
model.  Such models are  capable of helping quantify the  emission, dispersion,
and chemical reaction  processes.   Their  outputs can provide data on the dis-
tribution of air quality  concentrations between widely spaced ambient monitors.
     The emphasis  in  this section has been on a brief discussion of sampling
strategies.  The  word  sampling is also widely considered to mean those tech-
niques  that  are  required to obtain a parcel of air that is representative of
the polluted atmosphere, and to maintain its  integrity until a measurement of
concentration has  been  carried, out.   Considerations relating to this, meaning
of sampling are discussed as appropriate  in the following sections on measure-
ment techniques.
                                    .4-5

-------
4.2.3  Measurement Methods forTotal Oxidants and Ozone
4.2.3.1  Total Oxidants.  Although ozone was first unambiguously identified in
polluted atmospheres  by infrared spectroscopy (Stephens et a!., 1956b; Scott
et al.» 1957), the  earliest procedures for routinely monitoring 0, and other
oxidizing  species  in the  atmosphere were  based  on iodometry.   lodometric
techniques are inherently  non-specific  in  that a  variety of oxidizing  species
in addition to Og  may be positive interferences, whereas reducing agents are
negative interferences.  Thus,  the name "total oxidants" was coined because
the technique responded to 0, and other oxidants such as peroxides, peroxyacetyl
nitrate (PAN), and  nitrogen dioxide (NO*).  Total oxidants are then actually
defined by the particular iodometric procedure used, since the response to the
various oxidizing species present will depend on the details of the procedure.
This will  be  more evident when  interferences  are  discussed below.  The use of
the word "total" is  in  itself a misnomer.   The measurement does not reflect a
sum of the oxidizing species present because the various oxidants present in
the atmosphere react to produce iodine at different stoichiometries and differ-
ent rates.  In spite  of these difficulties, the measurement of total oxidants
was a  useful  method for characterization  of  the  atmosphere because of  its
correlation with the principal  oxidant, 0^;  and, consequently,  there is a
large  oxidant data base  available.   For  these reasons, the  two  principal
methods used for monitoring total oxidants are discussed briefly below.
     The bulk of the total oxidants data base was obtained by the use of two
types of continuous  monitoring  instruments.  In both types, an air sample is
continuously  scrubbed by an aqueous  reagent containing  potassium  iodide  (KI).
Ozone  and  other  oxidants produce  iodine  (tri-iodide ion)  according to  the
reaction:

                    03 + 31" + H20 -»• I3" + 02 + 20H                    (4-1)

In colorimetric oxidant instruments, the iodine is measured photometrically by
ultraviolet absorption  (Littman and Benoliel, 1953).  In the other common type
analyzer the  iodine  produced is measured by electrochemical means  (Brewer and
Mil ford, 1960; Mast  and Saunders, 1962).   Many other chemical techniques for
oxidants have been proposed and in some cases applied, but for these reference
is made to the original literature (Hodgeson, 1972; Katz, 1976).
                                    4-6

-------
     The interferences for both colorimetric and amperometric 0^ analyzers are
other oxidizing and reducing species  in the atmosphere.  The major oxidant in
ambient air by  far is 0^ (Chapter 5); and the other oxidants present, except
NOp, are considered part of the total oxidants  measured rather than inter-
ferences.   The only significant reducing interference known is SOp.
     The magnitude of the NC^  interference is variable (Tokiwa et a!., 1972;
Intersociety Committee,  1970).   For the  Brewer amperometric cell, the inter-
ference from NOp is only 6 percent of an equivalent concentration of 03 (Tokiwa
et al.s 1972).   For colorimetric oxidant analyzers, NO, interference equivalents
vary from  20 to 32 percent depending on 03  concentration  (Tokiwa  et al.,
1972),   A "corrected"  oxidant measurement is obtained by simultaneous measure-
ment of NOp  and correction of the  corresponding total  oxidant measurement.
The  interference  from SOp  is  quantitative  for both colorimetric and  elec-
trochemical oxidant measurements,  with one mole of SQp consuming one mole of
tri-iodide ion.  If the SOp concentration is less than that of total oxidant
and  SOp is  simultaneously measured, the total oxidant  may  also be corrected
for  SOp.  This  was the procedure previously  applied  in the older aerometric
data for California.   For many areas  of the  East Coast and Midwest, such a
correction was not possible and preferential SOp scrubbers were used (Saltzman
and Wartburg, 1965; Mueller et al., 1973).   These scrubbers could be effective
in the  hands of skilled operators  but their  use was not without problems.
Among these problems  were partial  oxidation  of  NO  to NOp and  of  HpS to SOp,
and partial removal of 03 when the scrubber was wet or contaminated  (Hodgeson,
1972).
4.2.3.2  Ozone
4.2.3.2.1  Gas-phase  chemi1uminescence.  Many of the 0, oxidation reactions
are  sufficiently energetic that  they  produce  electronically excited  products,
intermediates,  or  reactants,   which in turn may chemiluminesce (Zocher and
Kautsky, 1923;  Bowman and  Alexander,  1966).   Although  well known for many
years,  such  reactions were not applied to chemical analysis until the 1960s.
In 1965, Nederbragt reported  a detector that employed chemiluminescence from
the  reaction of 0-, with ethylene for  measurement of 0, in the vicinity  of
large  accelerators (Nederbragt  et  al.,  1965; Warren  and  Babcock,  1970).
Applications to  atmospheric analysis  were a  natural consequence  (Stevens and
Hodgeson, 1973).   The reference method for 0, originally promulgated by  EPA
for  compliance  monitoring was  the  Og-ethylene chemiluminescence method (F.R.,

                                    4-7

-------
1971).  Appendix D  of 40  CFR5  Part 50,  describes  the principle of the method,
including a method of calibration (F.R., 1971; C.F.R.,  1971).  Since then, the
measurement principle has  remained  the same but  calibration procedures have
undergone extensive  revision  as  discussed below (Section 4.2.4).   It is also
noteworthy that the reference method is specific for 03, whereas the data used
for establishing the  standard were  based on measurement  of  total  oxidants.
This issue is addressed in Section 4.2.5.
     A flow of sample air (1 to 5 liter/min) containing 03 and a small  flow of
pure ethylene are  mixed at atmospheric pressure  in a  small  reaction chamber
closely coupled to  the  photocathode of a photomultiplier tube.   The reaction
between Qg and  ethylene produces a small fraction of  electronically excited
formaldehyde.  Chemiluminescence  from  this  excited state results in a broad
emission band  centered  at  430 nm (Finlayson et  al.,  1974).  The  emission
intensity that  is  monitored is a linear  function of  Q3 concentration from
0.001 to greater than 1 ppm.  The relation between intensity and concentration,
i.e.,  instrument  calibration, must be  determined for  each  instrument with
standard concentrations of  0- in air.   The minimum detection  limit and the
response time are functions of detector design.   Detection limits of 0.005 ppm
and response times of less than 30 seconds are readily attained, however,  with
modest design features.   For example,  cooling the photomultiplier improves the
sensitivity but is  not  normally required.  There are  no  known interferences
among the common atmospheric pollutants.  There have been reports of a positive
interference when  0, is  measured in  the presence of  water  vapor,  i.e.,  a
signal enhancement  of 3 to  12 percent  in  high humidity as opposed to measure-
ment of the  same  concentration of 03  in  dry  air (California Air Resources
Board, 1976).  Where  this may be a real problem, it can be minimized by per-
forming calibrations with humidified air.  Finally, in order to obtain accept-
able measurement precision and constant span,  analyzers must contain means for
maintaining constant air and ethylene flow rates.
     Ambient air monitoring reference  and equivalent  methods have  been pub-
lished by EPA (F.R., 1975a).  This regulation prescribes methods of testing and
performance specifications  that  commercial  analyzers must meet in order to be
designated as a reference method or as  an equivalent method.  An analyzer may
be designated as  a reference method if  it  is based on the same principle as
the reference chemiluminescence  method and meets performance specifications.
An automated equivalent method must meet the prescribed performance specifi-
cations and  show  a consistent relationship with  a  reference method.   These
                                    4-8

-------
specifications for 03 analyzers are listed in Table 4-1.  Commercial analyzers
that have been  designated  as reference or equivalent  methods  are listed in
Table 4-2.   Information concerning the applications supporting the designation
of analyzers as reference or equivalent methods may be obtained by writing the
U.S.  Environmental   Protection Agency,  Environmental   Monitoring  Systems
Laboratory,  Research Triangle Park, NC 27711.
4.2.3.2.2  Gas-sol id chemi1uminescence.  The first chemiluminescence technique
for (k was developed by Regener for stratospheric measurements (Regener, 1960)
and later for  measurements in the troposphere (Regener, 1964).  The reaction
of On with Rhodamine-B adsorbed on activated silica gel produces chemilumines-
cence in the  red region of the spectrum  characteristic of the fluorescence
spectrum of Rhodamine-B.  The intensity is a linear function of CL concentration,
the minimum detection  limit can be lower than 0.001 ppm,  and  no atmospheric

         TABLE 4-1.   PERFORMANCE SPECIFICATIONS FOR AUTOMATED
                               OF OZONE ANALYSIS
Performance parameter
Range
Noise
Lower detectable limit
Interference equivalent
Each interference
Total interference
Units
ppm
ppm
ppm
ppm
ppm
Specification
0 to 0.
0.005
0.01
±0.02
0.06
5



Zero drift, 12 and 24 hour
Span drift, 24 hour
ppm
±0.02
20% of upper range limit
80% of upper range limit
Lag time
Rise time
Fall time
Precision
20% of upper range limit
80% of upper range limit
percent
percent
minutes
minutes
minutes
ppm
ppm
±20.0
±5.0
20
15
15
0.01
0.01
Source:  F.R. (1975a); C.F.R. (1975).
                                    4-9

-------
        TABLE 4-2.  LIST OF DESIGNATED REFERENCE AND EQUIVALENT METHODS
                               OF OZONE ANALYSIS
Identification and source
Beckman Model 950A Ozone
Fed.
Vol.
42
Register notice
Page Date
28571 6/03/77
Designa-
tion
(E=equiv.
Ref . R=ref . )
(F.R. , 1977a) R
 Analyzer
  Beckman Instruments
  2500 Harbor Boulevard
  Fullerton, CA 92634

Bendix Model 8002 Ozone            41
 Analyzer                          45
  The Bendix Corporation
  Post Office Drawer 831
  Lewisburg, WV 24901

Columbia Scientific Industries     44
 Model 2000 Ozone Meter
  11950 Jollyvilie Road
  Austin, TX 78759
        5145
       18474
2/04/76
3/21/80
(F.R.,  1976a)
(F.R.,  1980a)
       10429   2/20/79   (F.R.,1979a)
Dasibi Model 10Q8-AH Ozone
 Analyzer
Dasibi Model 1003-AH
 1003-PC or 1003-RS Ozone
 Analyzers
  Dasibi Environmental Corp.
  616 E. Colorado Street
  Glendale, CA 91205
MEC Model 1100-1 Ozone Meter,      41
 MEC Model 1100-2 Ozone Meter,     42
 or MEC Model 1100-3 Ozone Meter
  Columbia Scientific Industries
  11950 Jollyville Road
  P.O. Box 9908
  Austin, TX 78766

Meloy Model OA 325-2R Ozone        40
 Analyzer
Meloy Model OA 350-2R Ozone        40
 Analyzer
  Columbia Scientific Industries
  11950 Jollyville Road
  Austin, TX 78759
48     10126   3/10/83   (F.R.,  1983)       E
42     28571   6/03/77   (F.R.,  1977a)      E
       46647   10/22/76  (F.R.,  1976b)
       30235    6/13/77  (F.R.,  1977b)
       54856   11/26/75  (F.R., 1975b)

       54856   11/26/75  (F.R., 1975b)
                             R

                             R
                                      4-10

-------
   TABLE 4-2 (continued).
LIST OF DESIGNATED REFERENCE AND EQUIVALENT METHODS
   OF OZONE ANALYSIS
Identification and source
Fed.
Vol.
Register notice
Page Date
Designa-
tion
(E=equiv.
Ref . R=ref . )
Monitor Labs Model 8810            46
 Photometric Ozone Analyzer
  Monitor Labs, Incorporated
  10180 Scripps Ranch Boulevard
  San Diego, CA 92131

Monitor Labs Model 8410E           41
 Ozone Analyzer
  Monitor Labs, Incorporated
  10180 Scripps Ranch
  San Diego, CA 92121

PCI Ozone Corporation Model        47
 LC-12 Ozone Analyzer
  PCI Ozone Corporation
  One Fairfield Crescent
  West Caldwell, NJ 07006

Philips PW9771 03 Analyzer         42
  Philips Electronic Instruments,  42
  Incorporated
  85 McKee Drive
  Mahwah, NJ 07430

Thermo Electron Model 49
  UV Photometric Ambient 03        45
 Analyzer
  Thermo Electron Corporation
  Environmental Instruments Division
  108 South Street
  Hopkinton, MA 01748
               52224   10/26/81  (F.R., 1981)
               53684   12/08/76  (F.R.,  1976d)
               13572   03/31/82  (F.R., 1982)       E
               38931   08/01/77  (F.R,, 1977c)
               57156   11/01/77  (F.R., 1977d)
               57168   08/27/80  (F.R., 1980b)
  interferences  have  been observed (Hodgeson et al., 1970).  The technique is,

  in  fact,  more sensitive than the  gas-phase Nederbragt method and does  not

  require critical control of flow rate. It had the disadvantage in the original

  analyzer  built,  however,  that  frequent and  periodic internal calibration

  cycles were required to compensate for changes and decaying sensitivity of the

  surface of the detector (Regener, 1964; Hodgeson et al., 1970).

       Improvement was  made  in  the stability  of the  surface response in a  modi-

  fication  added by  Bersis and Vassiliou (1966),  in which  gallic  acid is  also

                                      4-11

-------
adsorbed on the surface in excess.  The Q3 apparently reacts with and consumes
the gallic acid rather than Rhodamine-B.  An energy transfer step to Rhodamine-B
subsequent to the  initial  reaction  results  in  the  same  chemiluminescence  from
the dye compound,  which  is now no  longer  consumed.   A commercial analyzer,
Phillips Model  PW9771,  is  based on this principle and has been designated as
an equivalent method under EPA regulations.
4.2.3.2.3  Ultraviolet photometry.  Ozone  has  a moderately strong absorption
band in the  ultraviolet  (UV), with a  maximum  very near the mercury 254  nm
emission line.  This  band is essentially a  continuum  near  250  nm.  The molar
absorption coefficient at the mercury line has been measured by several inves-
                                                                     —i   ~i
tigators with good agreement and has  an accepted  value of 134 ± 2 M   cm
(base 10) at 0°C and 1 atm (Hampson et a!., 1973).   Ultraviolet absorption has
long been  used  as  a method of measuring gas-phase 0, in fundamental  chemical
and physical  studies.  Some  of the  first atmospheric  0, measurements were,  in
fact, made by UV photometry; e.g., the Kruger Photometer.   These early instru-
ments and  the problems  with their  use  are  described  more completely in the
first criteria  document for photochemical  oxidants  (National  Air Pollution
Control  Administration,  1970).   The major problem with the older photometric
instruments was the large imprecision involved in measuring the very small ab-
sorbance values obtained.
     This problem  of  adequate sensitivity with moderate pathlengths has been
overcome by modern digital  techniques for measuring small absorbancies.   The
first instrument of this new generation of  photometers was marketed by Dasibi
of Glendale,  California,  in the early 1970s.  The details of this instrument
have been  described by Bowman and Horak (1972).  Other  commercial  instruments
have since been marketed and, along with the Dasibi, have been designated as
equivalent methods by EPA  (Table 4-2).  All  of these  instruments operate
effectively as  double-beam  digital  photometers.   A transmission signal  is
averaged over a finite  period of time  with 0, present and is compared to a
similar transmission  signal  obtained  through an otherwise  identical reference
air stream from which the 0~ has been preferentially scrubbed; e.g.,  using a
manganese dioxide  scrubber.   The  electronic comparison  of  the  two  signals can
be converted directly into a digital display of-Og concentration.
     The UV photometric  technique has  the  advantages, like gas-solid chemilu-
minescence, that a reagent gas flow is not required and that sample air flow
control  is  not  critical.  In addition, the measurement is in principle  an
                                    4-12

-------
absolute  one,  in that the concentration  can  be computed directly from  the
measured  absorbance  since  the absorption coefficient and the  pathlength are
known.  This capability  is used extensively for the  purpose  of 03  calibration
as discussed in  Section  5.5.5.   Commercial  UV photometers for Oo can serve a
dual  function  as a standard for 0^  calibration and  as  a means for measuring
ambient ozone  concentrations.  In  practice, UV  photometric analyzers  that are
used  for  monitoring  Og concentrations in the atmosphere are calibrated  with
standard  Og  samples  in  order to  compensate for possible 03 losses  in the
sampling  and  inlet  systems.   A  UV  photometric  analyzer has the potential
disadvantage that  any molecular  species that absorbs at 254 nm (e.g., SO,,,
benzene,  mercury vapor)  and  that may also be  removed along with  03 during the
reference  cycle  can  interfere.  Documentation  of such  interference  during
atmospheric monitoring is lacking at present.

4.2.4  Generation and Calibration Methods for Ozone
     Unlike the  other criteria pollutants, 03 is a thermally unstable species
that  must be generated in situ during the calibration  of analyzers used for
atmospheric monitoring.   This creates  special  requirements  not encountered
with  other pollutants and thus this  section  deals with means  for  generating
dynamic air  streams  containing  stable 0^ concentrations  and  chemical and
physical  means for absolute measurement of these concentrations.
4.2.4.1   Generation.   Ozonized  samples of air can be produced by a number of
means,  including photolysis  (Brewer and Mil ford, 1960), electrical discharge
(Toyama and Kobayashi, 1966), and  radiochemical  methods (Steinberg and Dietz,
1969).  Electrical  discharges are  useful  for  producing  high  concentrations of
Oo  in air for  other  applications; e.g., 0, chemistry.   Radiochemical methods
would be  ideal  except for their cost and  required safety features.   By far the
most  common method,  however,  for generating low  concentrations of 0,  in air in-
volves the photolysis of molecular oxygen.

                     02 + hY(X<200  nm) +20                            (4-2)

                     0 + 02 + M (M =  N2 or 02) •*  03 + M                (4-3)

One of the most  common photolytic  generators uses a mercury  vapor 6-  or 8-inch
PenRay  photolysis  lamp positioned  parallel to a quartz  tube  through which air
flows  at  a controlled rate.   The  0, concentration  is simply varied by means
                                     4-13

-------
of an adjustable and calibrated mechanical sleeve placed over the lamp envelope
(Hodgeson et  al.,  1972)  or by varying the voltage or current supplied to the
lamp.

4.2.4.2  Calibration
4.2.4.2.1  KI procedures:  original EPA reference method.   The  output  of
photolytic CL generators  can  provide  air  samples  containing  stable CL  concen-
trations over a considerable period of time with careful control of flow rate,
lamp voltage, temperature,  and pressure.   It is necessary to calibrate these
generators periodically  with  an absolute reference  method.   Prior  to 1975,
there were as many as seven different calibration methods for CL employed to
varying extents  in this  country  (National  Research  Council, 1977).    In an
attempt to standardize the  methodology, EPA published a reference calibration
procedure with the reference method in 1971 when the oxidant (as CL) standards
were promulgated (F.R., 1971).  This method was the 1 percent neutral buffered
potassium iodide  (NBKI)  procedure, a technique that had been used by EPA and
other agencies for some time.
     During the early 1970s,  it became evident that there were serious defi-
ciencies with the  NBKI reference  method.  Several problems with  the  NBKI pro-
cedure, summarized by a  joint EPA-NBS workshop in 1974 (Clements, 1975), in-
cluded the gradual continued release of iodine after sampling, variable results
obtained with different types of impingers, reagent impurities, and a positive
bias when compared to other 0, measurement methods.   An interagency collabora-
tive study was  undertaken to  intercompare iodometric methods used by the Los
Angeles Air  Pollution Control  District  (LAAPCD),  the California  Air  Resources
Board (CARB), and  EPA, using  UV photometric 0, measurements  as the reference.
The results of this study (DeMore et al.,  1976) demonstrated the positive bias
of the NBKI methods.   Concurrent with and after these earlier reports, a large
number of  individual  studies  ensued.   The history of these  studies  has  been
reviewed  in  the previous  criteria document (U.S. Environmental  Protection
Agency, 1978a)  and by Burton  et al.  (1976).  The  major  conclusions from  these
studies are presented below.

     1.   Results obtained by NBKI procedures are higher than those obtained
          by  UV photometry or gas-phase  titration by  5 to  25 percent,
          depending on details of the procedure.
                                    4-14

-------
     2.    When Og  is measured in the  presence  of humidified air, NBK!
          results tend to be even higher by another 5 to 10 percent (e.g.,
          California Air  Resources  Board,  1976).  The  reason  for this
          apparent moisture effect is not known.
     3.    In general,  NBKI techniques are  subject  to large imprecision
          because of procedural variation.

The EPA then evaluated four alternative calibration procedures  (Rehme  et  a!,,
1981) and selected  UV  photometry as the  reference  procedure because of its
superior accuracy and precision and its simplicity of use.
     Although NBKI  methods are no longer  used in  this country for  the  purpose
of calibration, there is a considerable data base available on health and wel-
fare effects, as  well  as atmospheric chemistry and monitoring, that is based
on these methods  as standards.   Therefore, it  is important to consider how
these data  may be evaluated and  compared  to newer effects and aerometric  data
based on the new UV calibration standard.   Since a systematic bias is known to
exist between calibrations by KI methods and UV photometric methods, it should
be possible, in  principle,  to apply correction factors to convert from a KI
reference to a UV photometric reference.  There are several problems inherent
in attempting such corrections, however. A fairly wide  range of variations has
been reported  in  the literature on the comparison  of KI  and UV photometric
measurements.  As discussed previously, the presence  of moisture  in  the cali-
bration air increases  the magnitude "of the bias.   Fortunately, both the CARB
and the LAAPCD procedures called for  the  consistent  use  of humidified air,
whereas the EPA  reference method prescribed the use of  dry air.   In  addition,
the elapsed time  between sample collection arid color measurement will also
affect the  magnitude of the bias because  of  the  slow  liberation  of iodine
after sampling (Clements, 1975;  Beard et  al.,  1977;  Hodgeson,  1976).  Other
unknown experimental  factors may  also influence the bias, e.g., impinger
design (Clements, 1975; Beard et al., 1977).
     An assessment  has been made of the previous  KI versus  UV intercomparisons,
and recommendations are  given in Table 4-3 for  correction factors  to apply to
calibration data  for conversion from UV to a KI  reference  or vice versa.  It
should be emphasized that these factors  could  validly  be applied to correct
for a calibration bias only and can  not  be applied for  comparison  of data
where other effects are present, e.g., the comparison  of oxidants versus O
                                    4-15

-------
         TABLE 4-3.  FACTORS FOR INTERCOMPARISON OF DATA CALIBRATED BY
                      UV PHOTOMETRY VERSUS KI COLORIMETRY

Calibration method                                     Ratio, KI/UV
EPA, 1% NBKI                                           1.12 ±0.05
CARB, 2% NBKI                                          1.20 ± 0.05
LAAPCD, 1% UKI                                         0.96a

 Correction for this method not recommended; only one intercomparison has been -
 reported.

data where the effects  of  oxidizing or  reducing  interferences must  be consid-
ered.  In this assessment, consideration was given only to  those studies  in
which the KI procedure was compared directly to UV photometry.   The recommended
value for data based on the CARB  method assumes the use of humidified air.
The value recommended  for  the EPA method  assumes  that  dry air was used and
that color measurement was made immediately after sample collection.
     The uncertainties  assigned  reflect the fact that a range  of values has
been reported  for the  ratios  in  previous  studies.   Finally,  whenever any
attempt is made to convert from  one data base to another,  these uncertainties
must be added by  conventional  error propagation  techniques to the uncertainty
inherent with the original measurement.
4.2.4.2.2  Ultravi olet  photometry  method.   A  major reason UV photometry was
designated as the calibration  procedure was the excellent precision of the
photometric measurement.   In  the collaborative study by Rehme et al. (1981),
measurement with  ten individual  UV photometers gave only a 3.4  percent  varia-
bility when  compared to a  reference  measurement system.   Other significant
factors in the  selection of UV photometry were the inherent simplicity of UV
photometric measurements and the ready  availability of commercial instruments
that can  also serve  well as transfer  standards between laboratory photometers
and field Og  analyzers.  (See National  Research Council,  1977,  and McElroy,
1979, for a discussion  of transfer standards.)
     It was  also presumed  that  UV photometry gives more  accurate  results,
since the accuracy is determined primarily by the 03 absorption coefficient,
which is well known  (Hampson et al.,  1973; DeMore and Patapoff, 1976).  Although
                                    4-16

-------
there is little  doubt that the accuracy of 0^ measurements has been signifi-
cantly  improved  by  conversion to the UV  basis,  some question still exists
regarding the absolute  relation  between Og measurements by UV photometry and
Og measurements  by GPT  measurements based  on  either  an  NBS  standard reference
material (SRM) nitric oxide (NO) gas cylinder or an N02 SRM permeation tube.
These intercomparisons  have been made by several investigators over the past
10 years and have been summarized by Burton et al. (1976) and Paur and McElroy
(1979).   The agreement  between GPT and  UV  measurements  was  generally close to
unity, although in some cases GO measurements by GPT have shown a small positive
bias with respect to UV measurements (e.g., Rehme et al., 1981; DeMore et al.,
1976).  DeMore and Patapoff (1976) reported a ratio  of  unity  between simulta-
neous measurements of (K  by GPT and  UV with  a 5 percent uncertainty on the
ratio of these  measurements.  In a recent detailed  study conducted at the
National Bureau of Standards (NBS) (Fried and Hodgeson, 1982), 0-j measurements
made with an NBS standard  photometer  (Bass et al., 1977) were compared to GPT
measurements of  0,, that were standardized  against both  NO cylinders (NBS SRM)
and NQp permeation tubes (NBS SRM).   Since the measurement  of flow rates is  a
critical GPT variable and  has been considered as a  major source  of error in
GPT measurements  (DeMore and Patapoff,  1976), NBS facilities were  used for
making absolute  flow  measurements  by both gravimetric  and  volumetric  means.
The results of this study  were that values of 0~  measured by  GPT  based on NOg
or NO SRMs agreed to within  less than 1 percent, but that values of 0, measured
by UV were lower by 3 percent.  For a consideration of possible error sources,
reference is made to  the  original article (Fried and Hodgeson,  1982).  In
summary, the UV  photometric 03 standard agrees quite closely with the NO and
N02 measurement  standards  by GPT, as it should in principle.   The resolution
of any  small biases that remain  seems an  appropriate matter for consideration
by EPA and NBS.
     The measurement  principle for the  absolute measurement of Og by UV photom-
etry  is the same as that used by instruments  for  monitoring atmospheric 0, as
described in Section  4.2.3.2.3  (Bowman  and Horak, 1972; DeMore et al., 1976;
Bass  et al.,  1977).   Ozone  is measured in a  dynamic flow system by measuring
the transmission,  I/Io, of ozonized clean air in an absorption cell of path-
length H.  When  the concentration is to  be expressed in  units of ppm,  meas-
urement of temperature and pressure  is also  required.  The 0, concentration
may then be calculated  directly from the Beer-Lambert equation:
                                    4-17

-------
                   rn-1     =     xo                   (4-4)
                   LU3Jppm    «£  x 273     ^
where or = 0- absorption coefficient at 254 nm, 1 atm, and 0°C,
                     -1  -1
        = 308 ± 4 atm  cm   (log base e),

and

      T = temperature, °K;
      P = pressure, torr.

Laboratory photometers used for primary  Og calibrations  have  pathlengths of 1
to 5 meters  and sophisticated digital electronic  means  for measuring small
absorbancies (Bass et a!., 1977; Bowman and Horak, 1972).
     The conversion of a commercial 0~ photometric monitor to a photometer for
use as  a  transfer standard for calibration  has  been described by Paur and
McElroy (1979).  Definitions are in order here. A primary standard UV photometer
is one that meets the requirements and specifications given in the 1979 revision
of the  Oo  measurement and calibration procedures  (F.R. ,  1979e).  A transfer
standard as used by EPA is a device or a method that can be calibrated against
a primary photometer and transferred to another location for calibration of 03
analyzers.  Commercial Q~  photometers have served well  in  this  regard,  but
other devices  have been  used as well;  e.g., calibrated generators and GPT
apparatus.  Guidelines on transfer standards for 0, have been published by EPA
(McElroy, 1979),  and  reference has already been  made to  the NAS  discussion on
transfer standards (National Research Council, 1977).
     The use of UV photometry is unique  in air pollution measurements  in that
it is based  on a physical measurement principle rather than a chemical stan-
dard.    It  is then worthwhile to trace how the measurement chain works from a
primary standard to field measurements.  The primary standard is referenced to
the accepted 0, absorption coefficient.  Transfer standards are then calibrated
with primary photometers  maintained  at EPA, NBS,  and elsewhere.  The  use of
commercial photometers in this regard has been described by several investiga-
tors (DeMore et al.,  1976; Hodgeson et al., 1977).  These and other kinds of
transfer  standards are then used to  calibrate 03 analyzers used for  field
measurements.
                                    4-18

-------
4.2.4.2.3  Other procedures.   Although UV photometry has been specified as the
reference calibration procedure, other procedures are available that can give
equivalent results.   These include a modified KI method, which was allowed as an
interim alternative  method for the calibration of  Og  monitors when the UV
method was designated in 1979.  Other KI methods that have been used success-
fully in  Europe  are  also  briefly discussed here.   Finally, the GPT method is
reviewed  since it has been used extensively  in  this country  and was discussed
above with regard to the cross-check of method accuracies.
     A major problem with  NBKI techniques is the slow  release of  iodine and
continued color  development  after  sampling.   Flamm (1977) evaluated the rate
of this iodine production  and found that  it was  the  same as  the  rate with
which hydrogen peroxide (H^O^) releases iodine from the same solution.  Based
on this observation  and  a consideration of other possible species that might
be responsible, Flamm concluded that certain buffer anions, including phosphate,
catalyze the formation of hLCU and yield stoichiometries for iodine production
greater than 1.  Measurements  made with a  1  percent KI  reagent containing 0.1
M boric acid (BAKI), pH=5, did not exhibit this phenomenon, nor did the original
EPA  method  when phosphate was omitted from  the  reagent (Hodgeson,  1976).
     The  BAKI method was  evaluated as one of four alternative techniques ,in
the collaborative study conducted by EPA (Rehme et a!,, 1981).  No significant
bias was  observed between BAKI and the reference technique based on UV photom-
etry.  An analysis,  however, of BAKI measurements by ten volunteers revealed
a  large  system-dependent variability, and thus the BAKI technique was not
recommended as an  independent calibration method.  It is noteworthy that the
system variability attributable  to calibration was reduced  somewhat if each
operator  assumed a  molar  absorption coefficient for iodine (as I0) of 2.56 x
      -1   -1
104  M cm   at  352  nm  rather than  independently  measuring  the absorption
coefficient with standard  I2 solutions  as  these procedures usually prescribe.
Measurement  systems  based on the BAKI  procedure  may  still be certified as
transfer  standards provided  the  guidelines for  certification given in the EPA
technical assistance document for such standards are followed (McElroy, 1979).
     Methods based on iodometry have been used  in Europe for some  time for the
calibration of 03  analyzers.   Bergshoeff  (1970) described a  method for use  in
the  Netherlands, in  which  thiosulfate  is  added  to  the  KI reagent  (KIT method)
along with 0.1 M phosphate buffer.  The iodine  released  is immediately reduced
by the thiosulfate and  the  amount of  iodine consumed  is determined by back-
titration of  the thiosulfate.   This method  has the advantage that problems
                                    4-19

-------
associated with iodine instability in solution are eliminated.  In the Federal
Republic of Germany,  the standard is based  on a  2 percent  KI  reagent with 2
percent KBr (KIBr Method) and a low concentration (0.02 to 0.03 M) of phosphate
buffer (Van de Wiel et al.5  1979).  These  techniques  have been compared to UV
and GPT measurement procedures  by Van de  Wiel, et al. (1979).  Measurements
made with the KIBr method were  in essential  agreement with  measurements by UV
or GPT, while measurements by the KIT method were too high by 15 to 25 percent,
depending on the  relative  humidity of the samples.  Modifications have since
been made in the  KIT  method  by  the addition  of KBr and reduction  of the phos-
phate concentration.
     The gas-phase titration (GPT) method employs the moderately rapid bimolec-
ular reaction between Og and NO to produce NG« (Rehme et al., 1974):

                         NO + 03 -» N02 + 02                           (4-5)

This approach was,  in fact, one  of  the early methods used to measure the
absorption coefficient of 0^ (Clyne  and Coxon, 1968) and yielded excellent
agreement with other absolute techniques (DeMore and  Patapoff, 1976).   When NO
is present in excess, no side  reactions occur and  the  stoichiometry is as
given above.   This method has the distinct advantage that it gives an absolute
relation among three  common  pollutants.   A measurement of the quantity of NO
or Og  consumed  or NOg produced provides a simultaneous  measurement of the
other two species and the  GPT  procedure has been used  in  all  three modes.
This calibration  technique  is  often  used in the calibration of chemilumines-
cence NO  (NO +  N09)  analyzers.  In order to obtain accurate concentration
        f\          Lm
measurements in the procedure as normally employed, accurate flow measurements
are required; and this  is  the  principal complexity  and difficulty with this
procedure (DeMore and Patapoff, 1976).  Stedman et al. (1976)  has employed an
appropriate NO detector  to  make flow ratio  measurements and thus avoid the
requirement for  absolute flow  measurements.  Because of  unexplained biases
between GPT measurement  systems and the UV reference  in the EPA collaborative
study, the GPT method was not recommended  as an independent calibration tech-
nique  (Rehme  et al., 1981).   It  is  still  allowed,  however,  as a transfer
standard in accordance with  the EPA guidelines for these standards (McElroy,
1979).
                                    4-20

-------
4.2.5  Relationship between Methods for Total Oxldants and Ozone
     When the  ambient air  quality standards  for  criteria pollutants were
originally established, a numerical standard was set for photochemical oxidants
as defined by measurements based on iodometric techniques.  Much of the health-
related and welfare-related  evidence  used as the basis for the standards was
obtained using the total oxidant instrumentation discussed above.  The reference
method specified  in 1971,  however,  was  the chemiluminescence measurement
of Og.  Instrumental  methods for the specific measurement of  atmospheric  0,
became commercially available  in 1970.  These had several  practical  advantages
over total oxidant Kl-based instruments.  These advantages were greater sensi-
tivity, precision, specificity—no interferences from ambient SOp and N02--and
improved reliability in routine monitoring.  Second, the data available showed
that 03 was  the  major contributor to total oxidant measurements, that 03 was
the major contributor to observed health  and welfare effects, and that 0,
could probably serve,  then,  as the best  surrogate for measurements of total
oxidants and for  controlling effects  of oxidants  in ambient air (see reviews
in Burton et al., 1976; U.S. Environmental Protection Agency, 1978a).       :
     Notwithstanding  the  promulgation  of  standards for  ozone rather than
photochemical  oxidants by EPA in  1979, an examination  of the temporal .and
quantitative relationships  between total  oxidant and Q*3 data remains of con-
siderable interest, largely because earlier data and many  newer data on health
and welfare  effects were obtained by means  of total  oxidant methods.  Aside
from the  relative paucity of data  on  simultaneous  measurements,  there are  two
distinct problems  in  making such  comparisons. The first is the difficulty in
estimating  the contributions  to the  total oxidant measurements from other
oxidizing species  such as N02 and from reducing  species such as SQp. 'The
presence  of such species could cause  the total  oxidant measurements to be
either higher  or lower than 0., concentrations.  The  second difficulty is  in
estimating  the bias  created between past  and present data as a result of the
change from the NBKI to the UV photometry  calibration procedures.  Fortunately,
these two problems can be treated separately and  the latter problem vanishes
for comparison of simultaneous 0^ and  oxidant  data obtained using the same
calibration procedure.                                         . .     •
     In the  sections  below, the relationship that  should  exist between total
oxidant and On is  considered from  an evaluation of  the response of NBKI measure-
ments to  other oxidizing and reducing species.   The predicted  relationship is
                                    4-21

-------
then compared to  data obtained in simultaneous  field  measurements  of total
oxidant and Og.
4.2.5.1  Predicted Relationship.  The predicted total oxidant measurements can
be expressed  as the  sum of the contributions from  oxidizing  and reducing
species that release or consume iodine in NBKI reagent:

               [Total Ox] = a[03] + 2. b. [Ox]. + c[N02]              (4-6)
                          - d[S02] - I. e. [Red].

In this  equation, [Ox],  and [Red], represent the concentrations of other
oxidizing and reducing  species  in the  atmosphere.  The  atmospheric concentra-
tions of other  reducing species, such  as  H2S,  are normally quite  low  compared
to S02  concentrations  (Stevens et al.,  1972b,  and  references therein) and
these species will  not be  considered  further here.   If the concentrations
above are true atmospheric concentrations, the constants a,  b,  c,  and d repre-
sent the efficiencies with which the various species  release or consume iodine.
For example, the  value  of the constant a  for  an  oxidant instrument calibrated
by the CARB 2 percent NBKI method would be approximately 1.2 (Section 4.2.4.2).
Since the instruments  are calibrated with ozonized  air, the factor a repre-
sents the bias  of the calibration method used.   If  the 03 concentration is
overestimated because of  calibration bias,  then so  are the contributions of
the other species by  the  same factor;  i.e., the  constants b, c, and d are all
higher than their true  values  by the same constant,  a.  Therefore, it should
in principle be possible to correct total oxidant data for calibration bias by
dividing both sides of the equation above by a.

          [Total Ox]1 = [Total  Ox]/a                                  (4-7)
                      = [03] + I. b1. [Ox].  + c'[N02] - d'[S02]

     Other atmospheric oxidants that have been identified and that may contrib-
ute to the  total  oxidant reading are hydrogen peroxide (H202), small organic
peroxides (e.g.,  methyl  and ethyl  hydroperoxide), peracetic acid, peroxyacyl
nitrates (Cohen et  al., 1967),  and  pernitric  acid (Niki  et al.,  1976).   An
estimation of the contribution  of these  species  to the  total oxidant  measure-
ment is  quite difficult because individual  b^'s  have  not  been measured and
there are  few data  available  on atmospheric concentrations  of individual
species.   The magnitude of  the efficiency term  will  depend  not only on the
                                    4-22

-------
stoichiometry of the  oxidation  reaction,  but also on the rate.  A summary of
the effects of various oxidants on NBKI reagent and the Mast oxidant meter is
given in Table 4-4  (Cohen et a!., 1967;  Purcell  and Cohen, 1967; Burton et
al., 1976).

   TABLE 4-4.   RESPONSE OF NBKI REAGENT AND MAST METER TO VARIOUS OXIDANTS

Ozone
Peracetic acid
Hydrogen peroxide
Acetyl peroxide
Ethyl hydroperoxide
n- Butyl hydroperoxide
tert-Butyl hydroperoxide
Nitrogen oxides (NO )
f\
Peroxyacetyl nitrate (PAN)
Peroxypropionyl nitrate (PPN)
NBKI
Aa
A
B
B
B
B
B
D (10% as N02)
D
D
Mast
E
NAb
D
N
NA
NA
NA
D (10%)
N
D
aA = immediate color development; B = slow color development; 0 = positive
 interference; E = good response; and N = no response (or negligible).
 NA = Data not available.
Source:  Cohen et al. (1967).

     By contrast,  reaction  efficiencies for N02 and  S02 are relatively  well
known.  Tokiwa  et  al.  (1972) observed reaction efficiencies of 6 percent for
the Mast  oxidant meter,  22  percent  for  a 10  percent KI  colorimetric analyzer,
and variable (20 to 32 percent) for a 20 percent KI colorimetric analyzer.  It
is well documented that SO^  is a quantitative negative  interference with a 100
percent efficiency for reducing the oxidant  reading by  an amount equivalent to
the SO, concentration present (Cherniack and Bryan, 1965; Saltzman and Wartburg,
1965).
                                    4-23

-------
     Returning to  the analytical expression  for total  oxidant,; an oxidant
value corrected for NQ2 and SOg interferences can be expressed as below, assum-
ing that no other significant reducing interferences are present:

          [Total  Ox]corr  = [Total Ox] - c[N02] + [SOg]               (4-8)
                        "  = [03] + z. b. [Ox].

Thus, a total oxidant measurement for which  legitimate  corrections or compen-
sations for  NOp  and  S02  have  been  made should  always be higher than a
simultaneous 0, measurement  by an amount that  is a  function of  the type and
concentrations of other oxidants present.  The only major qualifications to this
prediction are that  both  types of measurements must be sampling the same air
mass and  be calibrated with  respect to the same reference; that  no  other
significant reducing  interferences  are  present;  and that 03 losses within
sample inlet systems are insignificant.   On the other hand, total oxidant data
uncorrected for SOp  and  NQ2 interferences may be higher or lower than corre-
sponding 03 data,  depending on the concentrations of these pollutants.   Because
of the potential presence of these interferences, it  is  quite  difficult or
impractical to compare oxidant and  03 measurements  during evening and early
morning hours, when  03  concentrations  are quite low.  Therefore, in the com-
parison of total  oxidant and 03 simultaneous field measurements  below,  emphasis
is placed on comparison of peak hourly averages.
4.2.5.2    Empirical Relationship Determined from Simultaneous Measurements.
Several precautions  should  be taken in performing simultaneous  measurements.
Both kinds of instruments must be calibrated frequently with the same ozonized
air stream that has  been analyzed by a common reference method.   In a simul-
taneous comparison, daily calibrations should be made with an 03 generator and
the generator  output should  be analyzed weekly.  Both instruments  should
sample the same air  parcel.   Routine maintenance  should be frequent to ensure
constant gas and reagent flow rates, clean sample inlet systems,  etc.   Finally,
in any meaningful  comparison of 03 and oxidant data,  simultaneous measurements
of NOg and  S02  should be made.   If  chromium trioxide scrubbers are used to
remove SOp  in the  inlet to the oxidant  instrument,  these must be frequently
tested to  ensure that 03 is  not  also removed during continued use, particu-
larly under very humid conditions.   These scrubbers  may cause some additional
bias by oxidation of NO to NOo.
                                    4-24

-------
     The earliest comparative  study reported was by Renzetti and Romanovsky
(1956).  This study  compared  a phenolphthalein total oxidant monitor,  a KI
continuous oxidant  monitor,  a  rubber-cracking apparatus,  and an open-path
ultraviolet spectrometer, which monitored the  UV absorption at characteristic
Og absorption wavelengths.   The only meaningful measurements for consideration
here are the  KI  oxidant and UV Og  measurements,  since  these are similar to
measurement methods  used later.   The UV G~ spectrometer differed in a number
of  respects  from the 0, photometers  in  use  today.   Measurements were made
across an open optical path of 325 ft of the transmission at three wavelengths,
X-, = 265  nm,  Xg = 313  nm, and A.., = 280 nm.   Intensity  ratios at the three
wavelengths were  used  to minimize the effects  of  other  UV absorbers and of
particulate scattering.  Some  non-CU absorption may still  have been present,
and, if so, the  measured values would be higher than the true 0, values.  The
published absorption coefficients at these three wavelengths were used to com-
pute 0, concentrations (Vigroux, 1952).   Measurements were made over a 4-month
period.  Figures  4-1 and 4-2  are illustrative of the data  obtained for a
monthly average and a single day, respectively.  Peaks of total  oxidant and of
03 occurred at  the  same time, but  the  0,  maximum was usually less than the
total oxidant maximum.  The UV 03 data were usually higher in the wings at low
Oo concentrations.  Renzetti  and Romanovsky attributed the higher total oxidant
reading to the  presence of "other  oxidants" and estimated concentrations of
other  oxidants  of 0.1 to 0.4  ppm,  depending on 0- concentration.   Since the
total  oxidant instrument was  calibrated by an  NBKI method, this estimate is
almost certainly too high and a large portion of the difference between oxidant
and 0-  may  have been a result  of the 20 to 25 percent positive calibration
bias.  Since interferences may be present in the UV measurement and simultane-
ous measurements  of  N02 and SOp were not available at the  time,  no  attempt  is
made to make any more quantitative  assessment of this study.
     A later study (Cherniack and Bryan, 1965)  compared a 10 percent colorimet-
ric KI  oxidant  instrument, a  Mast oxidant  meter (Brewer  and Mil ford,  1960),  a
galvanic-cell oxidant  instrument  (Hersch and Deuringer, 1963),  and a UV 0^
photometer (Bryan and Romanovsky, 1956).   This latter instrument was similar
in  principle  to present-day photometers.  The precautions  noted above were
taken.  All  the instruments were  calibrated  with  respect to  the  2 percent UKI
calibration procedure  used by the  LAAPCD.  Simultaneous SOp and NO^ measure-
ments  were made, but no corrections were made  because the  concentrations were
reported to be quite small during the period of comparison. Atmospheric sampling
                                    4-25

-------
I
  %
o
   60
   SO
   40
   30
QC

111 nn
Q ZO
z
o
0 10
I   I  I   I   I   I  I  III

  	  OXIDANTSBYKI
  	O3 BY UV
                                            I   I   I   I   I   I   I   I   I   I   I   I
         I	I   I   I    I   I   I   I    I   I   I   I   I   I   I   I    I   I   !   I   I   I    I
     12 1   234
                     5   67  8  9  10 11 12  1  23  4  5  678  9 10 11  12

                      a.m.             AUGUST 1955          p.m.             P.S.T.
Figure 4-1. Comparison of ozone and total oxidant concentrations in the Pasadena area, August 1955.

Source: Renzetti and Romanovsky (1956).
   60

   50

   40

   30

   20


   10
         I   I   I   I   I   I   I   I   I   I   I   I   I   I   f  I   I   I   I   I    I   I   I

                 OXIDANTS BY Ki
O
Ul
O
Z
o
O
    0
     12  1   2  3  45678   9 10  11  12  1   2  3456   78  9  10 11 12

                       a.m.            AUGUST 27, 1955         p.m.             P.S.T.

     Figure 4-2. Comparison of ozone and total oxidant concentrations in the Los Angeles area,
     August 1955.

     Source; Renzetti and Romanovsky (1956).
                                      4-26

-------
was conducted over an unspecified period of time, and the data were referenced
to the  colorimetric  oxidant measurements.   The linear regression analysis of
the data  over  the  concentration range 0 to  0.6 ppm gave the  relationships
shown in  Table  4-5 after correction for calibration factors.  Thus, the data
show a  much  better absolute agreement  and  correlation between  CL measurements
and colorimetric total  oxidant  than  between  electrochemical  total  oxidant  and
colorimetric total  oxidant.  In  addition, these data do  not  indicate any
significant contribution by "other oxidants" to the total oxidant measurement.

  TABLE 4-5.   COMPARISON OF CORRECTED  INSTRUMENT READINGS TO COLORIMETRIC
                  OXIDANT READINGS DURING ATMOSPHERIC SAMPLING
Instrument
Mast meter
Galvanic cell
Ozone photometer
m
0.896
0.776
0.980
b
-0.013
+0.004
-0.005
r
0.868
0.867
0.982
(y = Instrument reading
   = mx + b; x = colorimetric oxidant measurement; m = slope; b = intercept;
      r = correlation coefficient)

Source:  Cherniack and Bryan (1965).
     During  the  1970s,  several  studies were conducted on the intercomparison
of On  and  total  oxidant instrumentation by  the Research Triangle Institute
(RTI)  of North  Carolina (Ballard et al.,  1971a,b;  Stevens  et a!.,  1972a,b).
Measurements were  made  for 0^ by chemiluminescence  and for  total  oxidant by a
colorimetric  KI  analyzer  and a  Mast  meter.   Calibrations were carried  out
Mast meter.   Calibrations  were  carried out  frequently with an  0., generator
calibrated by the 1 percent NBKI method.  Simultaneous N02 and S02 measurements
were also  made  and the oxidant data reported were corrected  for these inter-
ferences.  Clark et al. (1974) intercompared a  commercial UV  photometer, three
different  commercial  gas-phase  chemiluminescence analyzers,  and  a  gas-solid
chemiluminescence  analyzer by monitoring in  a rural  environment.   The instru-
ments  were all  calibrated  by a  common  reference procedure and hourly-averaged
field  measurements were collected over  a 1-month period  in August 1972.  Davis
and Jensen (1976)  reported intercomparisons  of  Mast meter total oxidant  measure-
ments  and  chemiluminescence 0- measurements. The instruments  were not calibrated
by the same  procedure,  however,  nor were any corrections attempted for SOg  and
                                    4-27

-------
NOp interferences.  Okita and Inugami (1971) reported an intercomparison of KI
total oxidant measurements with chemiluminescence 03 measurements in the urban
atmosphere of Musashino, Japan.  An intercomparison of total oxidant by KI and
CL by chemiluminescence  in  irradiated auto  exhaust  was  reported  by  Carroll et
al. (1972).   In  another extensive field study conducted at an air monitoring
station near  the Houston ship channel, Severs  and  coworkers  (Severs, 1975;
Neal et al.,  1976)  examined the  relationship  between  ozone  and total  oxidants
for this  area by making  simultaneous measurements with  a gas-phase  chemilumi-
nescence Oo monitor and a Beckman colorimetric total oxidant analyzer.  Primary
calibrations  of  the instruments  were performed periodically  using  the  EPA 1
percent NBKI  method.   No corrections were attempted for NOp or SOp interfer-
ences, but  during the latter part of this study a chromium trioxide scrubber
was placed in the inlet of the total oxidant analyzer.
     All of these 1970 studies were reviewed in the previous criteria document
(U.S.  Environmental Protection Agency, 1978a).  Only the major conclusions are
repeated  here.   In  general, the averaged data showed fairly good qualitative
and quantitative  agreement  between  the diurnal variations  of total oxidants
and Og.   The  usual  trend was a  slightly  higher value for the total oxidants
measurement at the  maximum,  a  not unexpected  result in  view of the  discussion
above.  Comparisons of monthly-averaged data taken from studies in Los Angeles
and St. Louis are shown in  Figures 4-3  and 4-4 (Stevens et  al., 1972a,b).
The total  oxidant data shown in Figure 4-4 are uncorrected and show distinct
morning and evening peaks  resulting from N02 interference  (see Chapter 3 for
diurnal patterns of N02).    Examination of  data taken from individual  days
shows considerably more variation among the methods, with total oxidant measure-
ments both higher and lower than 03 measurements.   Intercomparisons of only UV
photometric and  chemiluminescence 03 analyzers have  not shown these large
variations  (Clark et al.,  1974; Wendt, 1975).   In all probability,  these
variations result from the  large imprecision and interferences in total  oxidant
measurement.
     Two  of the  studies described  above  reported consistently  lower total
oxidant measurements.  In one of these (Davis and Jensen, 1976), the reference
KI  method was used for  calibrating  the chemiluminescence  analyzer while a
factory calibration was  used for the Mast meter.  As  pointed  out above,  other
studies have  found  low oxidant readings  for  the  Mast meter as compared  to
colorimetric  analyzers (Cherniack and Bryan, 1965; Tokiwa et al., 1972;  Stevens
et al., 1972a,b).  The use of the factory calibration would cause the
                                    4-28

-------
   .10
E
Q.
D.
2  .06
1           J


 OZONE-CHEM

 TOTAL OX-MAST


 TOTAL OX-TECH
        Figure 4-3. Measurements of ozone and total oxidants in Los Angeles,

        September 4 through September 30,1971.
        Source: Stevens et al. (1972b).

-------
*>,

CO
o
                         E
                         a
                         a
                          h
                         Z
                         O
cc
I-

UJ
U
Z
o
o
                           0.12
0.10
                           ,0,08
0.06
                           ,0.04
                	COLORIMETRIC

                     COULOMETRIC

              	CHEMiLUMINESCENCE
0.02
                                 r
                                       I   I   I  I  I   I
                                    2     4    6     8     10   12    14   16    18   20    22   24

                                                           TIME, hours


                                   Figure 4-4. Measurements of ozone and total oxidants in St. Louis,

                                   October 14 through December 21,1971.
                                   Source: Stevens et ai. (1972a).

-------
Mast readings to  be  even lower because of a calibration bias (Cherniack and
Bryan, 1965; Tokiwa et al., 1972).  The results reported by Severs and coworkers
are more difficult to evaluate.  Chemiluminescence CL values generally higher,
and sometimes considerably  higher,  than total  oxidant measurements were re-
ported,  although  the measurements were  referenced to the same calibration
procedure. Correlations were reported both with and without a chromium trioxide
scrubber  in  the oxidant  inlet.  These  results  are  inconsistent with the known
responses of the instruments and the results of other investigators.   The data
reported  for one  day of high  03, but  abnormally low oxidants, are shown in
Figure  4-5.   It  is   highly improbable  that   the  problem  is  with the
chemiluminescence 0-  measurement,  since this   response is typical of a normal
0- diurnal  variation and no other species are  known  to interfere.  It  is more
probable that some other species of pollutant  in the highly industrialized area
of the Houston ship channel repressed the response of the total oxidant analyzer,
which thus does not respond to Q3, much less to any other oxidant.
     The most recent comparison in the literature involved simultaneous 03 and
total  oxidant  measurements in the Los  Angeles basin by the California Air
Resources Board (1978)  in the  years  1974, 1976,  and  1978.   The maximum hourly
data  pairs  were correlated (Chock et  al.,  1982) and y-ielded the  following
regression equation for  1978 data, in which a  large number (927) of data pairs
were available:

                         Oxidant  (ppm) = 0.870 03 + 0.005              (4-9)
                          (correlation coefficient = 0.92)

Thus,  when  the 03 levels were  relatively high,  they were actually slightly
higher  than  total  oxidant.   The  total  oxidant data  were uncorrected  for  NO^
and S02  interferences.
      In  summary, specific 03 measurements agree  fairly well with total oxidant
corrected for  NO- and SO, interferences and Q~ is  the dominant contributor to
total  oxidant.   Indeed,  it is difficult to discern  the presence of  other
oxidants in most  total  oxidant data.   There  can,  however,  be  major temporal
discrepancies  between 03 and oxidant  data, which  are primarily  a result  of
oxidizing and  reducing  interferences with KI  measurements.   As  a result  of
these  interferences,  on any given day  the  total oxidant data may be  higher
                                     4-31

-------
UJ
to
                    0.150
                  a
                  a

                 Z

                 c 0.100
IU
o
z-
o
o

O  0.050
cc
o
 m
o
                                 — O3 BY CHEMILUMINESCENCE

                                 ,— OXIDANTS 1Y BECKMAN ACRALYZAR (Ki)
                                                  10           15


                                                     TIME, hours
                                                           20
                           Figure 4-5. Measurements of ozone and total oxidants, Houston Ship

                           Channsl, August 11,1973.
                           Source: Severs (1975).

-------
than or  lower  than  simultaneous 03 data.  The  quantitative  relationship be-
tween oxidant and (k data, such as that used by Chock et al.  (1982), is proba-
bly quite location-dependent.  From a methodologic standpoint, the measurement
of Og  is a more  reliable  indicator than  total oxidant measurements  of  oxidant
air quality, and such  difficulties and controversy as may be involved in the
intercomparison  of  0.,  and  oxidant measurements are eliminated  if  the air
quality standard is defined in terms of CL.

4.2.6  Methods for Sampling and Analysis of Peroxyacetyl Nitrate and Its
       HomoTbgues
4.2.6.1Introduction.   Since  the  discovery of  "compound  X"  (Stephens  et al.,
1956a,b), later unambiguously  identified as peroxyacetyl nitrate (PAN) (Stephens
et al., 1961), much effort has been directed toward its atmospheric measurement.
Peroxyacetyl nitrate and its homologues are products of photochemical reactions
involving hydrocarbons  and oxides of nitrogen (NO ) in the atmosphere (Stephens
                                                 ar\
et  al.,  1961).  The significance  of atmospheric PAN is  twofold.   It  is  a
potent lachrymator  and phytotoxicant  in  the ppb concentration range (Stephens
et al.,  1961;  Heuss and Glasson, 1968).   Because  of the reversible thermal
equilibrium (Hendry and Kenley, 1977),

                     CH30(02)N02  <	  CH3CO(02)- + N0£,               (4-10)

which  is sensitive to  the NO^/NO ratio, PAN may serve as an  important reservoir
for peroxy radicals and N02 (Singh and Salas, 1983a,b) and may play a signifi-
cant role  in  both the  atmospheric nitrogen cycle  and in tropospheric ozone
formation (Spicer et al., 1983).
     Only two  analytical  techniques  have been used to obtain significant data
on  ambient  PAN concentrations.  These are gas  chromatography with electron
capture  detection (GC-ECD)  and long-path Fourier-transform infrared (FTIR)
spectrometry.   Atmospheric  data on  PAN concentrations  have been  obtained
predominantly  by GC-ECD because of its relative simplicity and superior sensi-
tivity.  These analytical  techniques are  described  in  Section 4.2.6.2 along
with attendant methods of sampling.  Peroxyacetyl nitrate is somewhat analogous
to  0,  in that it is a thermodynaraically unstable oxidant and PAN  standards
must be  generated and  analyzed by some  absolute technique for the  purpose of
calibrating the  GC-ECD.   Generation and calibration techniques are discussed
separately  in Section  4.2.6.3.  Finally,  the analysis  of PAN homologues is
discussed briefly in Section 4.2.6.4.
                                    4-33

-------
4.2.6.2  Analytical Methods for PAN.   By  far the most widely  used technique
for the  quantitative determination of  ppb  concentrations  of PAN is GC-ECD
(Darley et  al.,  1963; Stephens, 1969).  With-Carbowax  or SE30  as a stationary
phase, PAN  can  be separated from components  such  as air,  water, and other
atmospheric compounds,  as  well  as ethyl  nitrate,  methyl nitrate,  and  other
contaminants that are present in PAN  synthetic mixtures.   Electron-capture
detection (using a nickel-63 source and a pulsed-current detector or a tritium
source and  a direct-current detector)  provides  sensitivity  for PAN in the ppb
and sub-ppb ranges.   A  typical column  for the separation of PAN would be 3 to
5 feet in length and 1/8 inch in diameter (i.d.), and would be run isothermally
at 25° to  60°C.   Under these conditions, a peak assignable to PAN appears
after 2 to 3 minutes.  Table 4-6 shows parameters used by several investigators
to determine trace levels of PAN by GC-ECD.
     Sample injection into  the  GC is accomplished by means of a gas-sampling
valve with  a  gas-sampling  loop  of  a few milliliters  volume (Stephens  and
Price, 1973).   Sample  injection  may be performed  manually  or  automatically.
Typically, manual  air  samples  are collected  in  50 to  200 ml ungreased glass
syringes, and purged  through  the gas-sampling valve.  Samples collected from
the atmosphere should  be analyzed as soon  as possible because PAN undergoes
thermal decomposition  in the gas phase and  at  the  surface of containers.
Automatic sample collection and injection may be accomplished by using a small
pump to pull ambient air continuously through the sampling loop of an automatic
sampling valve,  which periodically injects the sample onto the column (Stephens
and Price, 1973).  Recently, Singh and Sal as  (1983b) have used cryogenic trap-
ping of PAN, with liquid argon, from  relatively large air samples for the
purpose of measuring PAN concentrations in the sub-ppb range.
     Most of the atmospheric PAN measurements have been made in polluted urban
environments,  where  maximum concentrations  of 5  to 50  ppb and  average concen-
trations of a few ppb may  occur  (Stephens,  1969;  Lonneman et al., 1976).  For
the purpose of  such  measurements, chromatographic detection limits of 0.1 to
1 ppb are  sufficient.   The recent work of  Singh and Sal as (1983a,b) on the
measurement of PAN in the free (unpolluted)  troposphere  is illustrative of
current capabilities for measuring low concentrations.   A 50- to 200-ml  volume
of air was collected by preconcentration  into an unpacked 0.15 cm o.d.  stainless
tube of 1.24 ml  volume, at liquid argon  temperature,  prior to analysis  (see
Table 4-6).  For measurements in humid environments, the air sample was passed
                                    4-34

-------
                                      TABLE 4-6.          OF            USED IN DETERMINATION OF PAN BY GC-ECD
Reference
Heuss and Glasson,
1968
Grosjean, 1983
Darl ey et al . ,
1963
Stephens and Price,
1973
-P>
i
CO
en
Lonneman et al. ,
1976
Holdren and Spicer,
1984
Peake and Sandhu,
1982
Singh and Sal as,
1983b
Grosjean et al . ,
1984
Nielson et al. ,
1982
Column dimensions
and materials
4 ft x 1/8 in
Glass
6 ft x 1/8 in
Teflon
3 ft x 1/9 in
Glass
1.5 ft x 1/8 in
Teflon
3 to 4 ft x 1/8 in
Glass
5 ft x 1/8 in
Teflon
3.3 ft x 1/8 in
Glass
1.2 ft x 1/8 in
Teflon
1.7 ft x 1/8 in
Teflon
3.9 ft x 1/12 in
Glass
Stationary phase
SE30
(3.8%)
10% Carbowax 400
5% Carbowax 400
51 Carbowax E 400
10% Carbowax 600
Carbowax 600
5% Carbowax 600
10% Carbowax 600
10% Carbowax 400
5% Carbowax 400
Solid support
80/100 Mesh
Diatoporte S ,
60/80 Mesh
Chrontosorb P
100/200 Mesh
Chromosorb W
Chromosorb G
80/100 Mesh
treated with
dimethyl
dichlorosilane
Gas Chrom Z
60/80 Mesh
Gas Chrom I
Chroinosorb W
80/100 Mesh
Supelcoport
60/80 Mesh
Chromosorb G
Chromosorb
W - AW - DCMS
Column
temperature ,
°C
25
30
35
25
25
35
33
33
30
25
Flow rate,
(ml/min)
NAa
40
25
60
70
70
50
30
40
40
Carrier
gas
NA
N2
Na
N2
95% Ar
5% CH4
90% Ar
10% CH4
N2
95% Ar
5% CH4
N2
N2
Elution
time,
rnin
1 NA
. NA
2.17
1.75
2.7
3.00
NA
NA
5.0
6.0
Concentration
range
ppb range
ppb range
3 to 5 ppb
37 ppb
0.1 to 100 ppb
ppb range
0.2 to 20 ppb
0.02 to 0.10 ppb
2 to 400 ppb
11 ppb
aNot available.

-------
through a Nation  drier (Foulger and Simmonds,  1979)  to reduce the humidity
prior to preconcentration.  A minimum detection limit of 0.01 ppb was obtained.
Free tropospheric  concentrations in the  0.01  to  0.1 ppb range were always
observed and indicated that PAN is a natural constituent of the atmosphere and
may constitute a significant fraction of the reactive nitrogen.
     There are conflicting reports in the literature on the effects of variable
relative humidity  on  PAN  measurements by  GC-ECD.   In  1973, Stephens and Price
stated that  in preparing  PAN calibration  samples  "the diluent  (gas) should be
of normal humidity so that the  chromatogram will  be a realistic one."  Subse-
quently, Holdren  and  Rasmussen  (1976)  observed  a reduced response  to PAN
calibration samples when the relative humidity of the sample was 30 percent or
lower and a  tenfold  decrease in PAN response  when the  relative humidity  ap-
proached 0  percent.   This effect was attributed  to  an  interaction between
sample and GC  column.  Nieboer and van Ham (1976) reported that "the elution
gas stream was previously humidified .  .  . because it appeared that the height
of the PAN peak depends on the relative humidity of ambient air if dry elution
gas was  used."   In contrast to  these studies, Lonneman (1977) observed  no
effect on peak height in PAN calibration samples  in which the relative humidity
varied from 10 to 50 percent.
     In 1978, Watanabe and Stephens reported on a reexamination of the moisture
anomaly and investigated the effects of humidity on PAN storage flasks, columns,
and detectors.  A  consistent PAN loss to the  walls of  dry acid-washed glass
storage flasks was observed, but the PAN could be recovered by the addition of
moisture to  the flasks.   A tritium direct-current detector showed  no humidity
effect except for  a  small (5 to 10 percent) decrease in peak height in a few
cases at very low humidities (2 to 3 percent).   With a different GC instrument
employing a Ni-63 detector, erratic responses were observed at low humidities,
with responses  reduced 30  to  95 percent from  that  obtained  at 53 percent
relative humidity.  No conclusion was drawn on whether this difference reflec-
ted a humidity effect on the detector, the column,  or  the sampling value.
Finally, the moisture anomaly  did not appear  to  depend on column  history or
loading even after a bake-out treatment.
     In the  most  recent  study  (Grosjean et  al., 1984),  the humidity effect on
PAN calibration  samples prepared by the  dynamic  and static irradiation  of
CHoCHO-Clp-NOp mixtures  was examined.  A small  decrease  (<3 percent) was
observed in  PAN peak  height when the dry  air stream was passed either  over an
impinger containing  water or  directed  through the  water  impinger.   These
                                    4-36

-------
results are  in  contradiction  to all of the  above,  in which the response to
humidified PAN  samples  is either greater than  or  the  same  as dry  PAN  samples.
     It is noteworthy  that  the chromatographic systems employed by different
investigators often employ different materials (e.g., glass, Teflon, stainless
steel), column loadings, and detectors.  The resolution of all  the differences
noted above in regard to a' suspected humidity effect might require considerable
effort.  For the present, if a moisture effect is suspected in a PAN analysis,
the bulk  of this evidence  suggests  that  humidification of PAN calibration
samples (to  a range approximating the  humidity of the samples  being analyzed)
would be advisable.
     Conventional long-path infrared spectroscopy and Fourier-transform infrared
spectroscopy  (FTIR)  have been  used  to detect and measure atmospheric. PAN.
Sensitivity is enhanced by the use of  FTIR over conventional long-path infrared
spectroscopy.   Accurate knowledge of  the  absorptivities  of many IR bands  ,
assignable to PAN makes possible the quantitative analysis of  PAN without the
use of  calibration  standards.   The most frequently  used  IR bands have been
assigned and  the absorptivities shown in Table 4-7 have been reported.  Only
the key bands  are  shown,  but all  27  fundamentals  are infrared active and
Bruckmann and Wiliner  (1983)  have assigned  most  of them.  The assignment by
Adamson and  Guenthard  (1980)  of the bands at  1435,  1300,  and  990 cm'1 to an
impurity, CHgQNQp,  is  apparently incorrect.   Bruckmann and Willner (1983) ob-
served these same bands in a 99 percent pure PAN  sample.
     The initial  discovery  of PAN in  simulated photochemical smog was accom-
plished by long-^path infrared absorption spectrometry  (Stephens et al.,;1956b).
Some recent  simultaneous measurements  of PAN and  other atmospheric pollutants
such as 03,  HN03,  HCOOH, and HCHO have been made by  long-path FTIR spectrom-
etry during smog episodes in the Los Angeles Basin.   Tuazon et al. (1978) have
described an FTIR system operable at pathlengths  up to 2 km for ambient measure-
ments of PAN and their  trace constituents.   This  system employed an eight-mirror
multiple reflection cell with a 22.5-m base path.  The spectral windows avail-
able at pathlengths of 1 km  were  760-1300,  2000-2230, and 2390-3000  cm"1.
Thus,  PAN  could be detected by the bands at 793  and 1162  cm"1.  The 793 cm"1
                                                           »-"l
band is characteristic  of peroxynitrates, while the 1162 cm   band is reportedly
caused  by PAN only (Stephens, 1969; Hanst et al., 1982).   Tuazon et al. (1981a,b)
reported  ambient measurements  with  this  system  during a smog episode  in
Claremont, CA,  in 1978.  Maximum PAN  concentrations  ranged  from  6 to 37 ppb
over a  5-day episode; the report presented diurnal patterns for PAN and several
019QQ/A   '                          4-37

-------
                            TABLE 4-7.   INFRARED ABSORPTIVITIES OF PEROXYACETYL NITRATE (BASE 10)
CO
CO
Vapor ppm""1™"1 x 104
Mode
v(c=0)
vas(N02)
vs(N02)
v(c-o)
v(o-o)
6(N02)
Frequency,
cm"1
1842
1741
1302
1162.5
930
791.5
0.1 m, no diluent
(Bruckmann and
Willner, 1983)a
12.4
32. 6d
13.6
15.8
NA
13.4
0.1 m path;
HZ at ~1 atrn h
(Stephens, 1964)°
10.0
23.6
11.2
14.3
1.8
10.1
120 m path;
air at ~1 atm h
(Stephens, 1964)°
NAC
NA
11.4
13.9
NA
10.3
Solution
pm""1 i
Frequency,
cm"1
1830
1728
1294
1153
NA
787,
in n- octane,
(Holdren and
Spicer, 1984)
0.00041
0.00115
0.00041
0. 00042
NA
0.00044
  TTIR spectra 1-1.2
                           resolution.
   Prism spectra -10 cnr* resolutioni
   NA = Data not available.
   Resolved Q branch.

-------
other pollutants for  the 2 most severe days.  The  detection  limit given for
PAN at a 900-m pathlength was 3 ppb.
     Hanst et al. (1982) modified the FTIR system used by Tuazon et al. (1978)
by changing  it  from an eight-mirror to a three-mirror cell configuration and
by considerably  reducing the  cell  volume.  Measurements  were  made  over a
1260-m optical path folded along a 23-m base  path  at 0.25 cm   resolution.
Measurements were reported  for PAN and a  variety of other pollutants for a
2-day smog episode  at California"State University,  Los Angeles,  in 1980,  The
maximum PAN concentration observed was 15 ppb for this period of only moderate
smog intensity.  An upper limit of 1 ppb of peroxybenzoyl nitrate (PbzN) was
placed based on  observations in the vicinity  of  the PBzN band at 990 cm  .
The reports  by  Tuazon et al. (1978) and Hanst et al.  (1982) both refer to
earlier FTIR ambient air studies.
     Sampling may constitute one of the major problems  in the  analysis of
trace reactive  species, such as PAN,  by  long-path  FTIR spectrometry.  The
folded-path White cells have a large internal volume (15 m  for Tuazon et al«,
1978; 3 m  for Hanst et al., 1982).  The large internal surface area may serve
to promote  the decomposition  or  irreversible  adsorption of  reactive  trace
species.   To  minimize these  effects,  both Tuazon  et  al.  and Hanst et al.
employed  high-speed blowers to pull ambient air  through the cells at  high
velocities.   For interior cell linings, Hanst et al. employed 0.5 mm polyvinyl
chloride sheeting and Tuazon et al. used Plexiglas and FEP Teflon.
     Pitts et al. (1973) proposed a chemiluminescence technique for continuous
monitoring of ambient concentrations of PAN.  The reactions of both PAN and 0,
with triethyl amine  in the gas phase produce chemiluminescence.   The  spectra
reported  overlap somewhat with a h  x of 520 nm for the 0, reaction and \_ax
of 650 nm for  the PAN reaction.  Pitts proposed a technique  that included the
measurement of the  emission  intensity in the two regions by the use of optical
cut-off filters.  Thus,  the PAN  concentration  can be determined,  provided the
absolute 0, concentration is simultaneously measured.  Concentrations of 6 ppb
PAN were  detected  and a  lower limit of detection of 1 ppb was estimated.  No
interfering  emissions were observed from methyl nitrate,  ethyl  nitrate, ethyl
nitrite, or N02-  No  further work has been reported on the development of this
technique, and there  have been no atmospheric applications.
4.2.6.3  Generation and Calibration of PAN.  Because of the thermal instability
of dilute PAN samples and  the  explosive  nature of liquid PAN,  calibration
samples  are  not commercially available; each  laboratory involved in making
                                    4-39

-------
such measurements must prepare its own standards.  The PAN samples are prepared
by various means at concentrations in the ppm range and these must be analyzed
by some  absolute technique.   The analyzed  samples  must  then be diluted to
obtain gas-phase  samples  in  the low ppb  range  for  direct calibration of GC
instruments.  Thus, this section includes descriptions of various means of PAN
generation, methods of  analysis,  and the procedures  for  sample  handling and
storage where applicable.
     The earlier  methods  used  for the preparation of  PAN  have been summarized
by Stephens (1969).  These included (1) the photolysis of mixtures of nitrogen
oxides with organic compounds  in  air or oxygen (Stephens et a!., 1956b;
Stephens et a!.,  1961); (2)  the photolysis of alkyl  nitrite vapor in oxygen
(Stephens et a!., 1965); (3) the dark reaction of aldehyde vapor with nitrogen
pentoxide (Tuesday, 1961); and (4) the nitration of peracetic acid.  Of these
methods, the photolysis of alkyl nitrites was favored and used extensively by
Stephens and other investigators.   As described by Stephens et al.  (1965),  the
liquefied crude  mixture obtained  at  the  outlet  of  the photolysis chamber is
purified by preparation-scale  GC.   [CAUTION:   Both the liquid crude mixture
and the  purified PAN  samples are violently explosive and should be  handled
behind explosion shields  using plastic full-face protection, gloves,  and  a
leather  coat at  all times.   These PAN samples should  be handled  in the frozen
or gaseous  state whenever possible.]   The pure PAN is usually diluted to about
1000 ppm  in cylinders pressurized with nitrogen to approximately 100 psig.
When  refrigerated at  <15°C,  PAN  losses  are less than 5  percent per month
(Stephens et al.,  1965).  Lonneman et al.  (1976) used the photolysis products
without  purification  for  the calibration of GC  instruments  in the field and
discussed the  use of  fedlar bags for the preparation and transport of cali-
bration  samples.
     Gay et al. (1976) have used the photolysis of CU:  aldehyde: NOp mixtures
in air or  oxygen for  the  preparation of PAN and  a number  of  its  homologues at
high yields:

d2 + hv	>  2 Cl                                                  (4-11)

      00
d + RC-H  	»  R-C' + HC1                                          (4-12)
                                    4-40

-------
 00
RC-  + 02 + M  	>  RC-02'  + M                                       (4-13)
 0                    0
RC-02*  + N02  	>  RC-02-N02                                        (4-14)

This procedure  has been  utilized  in a portable  PAN generator that can be used
for the  calibration  of  GC-ECD  instruments  in  the field (Grosjean, 1983;
Grosjean et al., 1984).  The output of this generator is a dynamic flow of PAN
in air at  a concentration of about 2 to  450 ppb.   Dilute concentrations of
reactant gases for the photolysis chamber are obtained by passing a controlled
flow of air over C12, N02, and acetaldehyde permeation tubes.
     The other  technique for PAN preparation  in current  use  involves  the
nitration of peracetic acid.  In the 1969 review (Stephens, 1969), this approach
was considered not useful for synthesis.  Several investigators, however, have
recently reported on  a condensed-phase  synthesis of  PAN with peracetic  acid
that produces  high yields  of a pure  product free of other alkyl  nitrates
(Hendry and  Kenley,  1977; Kravetz et al.,  1980;  Nielsen et  al.,  1982; Holdren
and Spicer, 1984).  Most of these procedures call for the addition of peracetic
acid (40 percent  in  acetic  acid) to a hydrocarbon solvent (pentane, heptane,
octane) maintained at -80°C in a dry-ice acetone bath, followed by acidification
with sulfuric acid.   Nitric acid is formed in situ with stirring by the slow
addition of  sodium nitrate.  After the  nitration is  complete, the hydrocarbon
fraction,  containing PAN concentrations  of 2  to  4  mg/ml  (Nielsen  et al.,
1982), can  be  stored at -20°C  for periods longer  than a year  (Holdren  and
Spicer,  1984).   After analysis,  the  PAN-hydrocarbon solutions can be used
directly for  calibration by the evaporation of  measured microliter  volumes of
solution into Tedlar bags containing known volumes of clean air.
     The most direct method for absolute  analysis is by  infrared absorption
using absorptivities given  in Table 4-7.  This  is the technique used by Stephens
(1969; analysis  of  PAN in  N2 cylinders),  Lonneman et al.  (1976;  analysis  of
gas-phase  products from  photolysis  of ethyl nitrite); and Holdren and Spicer
(1984; analysis  of PAN in octane solutions).  Whereas long, folded-path  cells
and FTIR spectrometry are required for the analysis of atmospheric PAN, conven-
tional  IR  instruments and  10-cm gas cells  can analyze  gas standards with
concentrations  greater than 35 ppm (Stephens,  1969)  and  Holdren and Spicer
                                    4-41

-------
(1984) used 50-|jm liquid microcells for the analysis of PAN in octane solutions.
Another candidate technique for absolute PAN analysis was gas-phase coulometry
using a tandem  electron-capture detector (Lovelock et al., 1971),  Singh and
Salas (1983b) have shown, however, that this technique is unsuitable for abso-
lute PAN analysis because a significant fraction of the PAN is destroyed prior
to coulometric detection.
     The alkaline hydrolysis  of PAN to  acetate  ion  and nitrite  ion  in quanti-
tative yield  (Nicksic  et al., 1967) provides a means independent of infrared
for the quantitative  analysis of PAN.  Molecular oxygen  is also  produced  in
quantitative yield by the reaction (Stephens, 1967):

                  00
               CH3COON02 + 20H" -* CHgCO" + 0£ + N02" + HgO            (4-15)

The colorimetric determination  of nitrite  ion with  Saltzman reagent was  first
used to  measure PAN quantitatively (Nicksic  et al., 1967; Stephens, 1969;
Kravetz et al., 1980).  Nielsen et al. (1982) analyzed the hydrolyzed products
of pure PAN  samples  by ion chromatography  for  nitrite and nitrate  and found
that 4 percent  of the nitrite had been oxidized to nitrate.  Some  gas-phase
PAN calibration samples  (e.g.,  photolysis of Cl«:  acetaldehyde: N02) contain
impurities such as N02 that will yield nitrite and nitrate in aqueous solution.
Thus, Grosjean (1983) and Grosjean et al.  (1984) performed ion chromatographic
analysis of the acetate ion to determine the PAN output of a portable generator.
     An alternate calibration procedure has been proposed based on the thermal
decomposition of PAN in the presence of excess NO, and measurement by chemilumi-
nescence of the NO consumed (Lonneman et al., 1982).  The acetylperoxy radical,
CH3C(0)02, and  its  decomposition products rapidly oxidize NO to NO,,.   In the
presence of  a small  amount of  benzaldehyde, which  is  used to scavenge  the
hydroxyl radical and control the stoichiometry, simulation models predict that
5 molecules of  NO will be  removed per PAN  molecule  present.   By the use  of NO
and  PAN  standard mixtures  and the chemiluminescent measurement  of the  NO
consumed, the experimental  value was determined to  be ANQ/APAN = 4.7 ±  0.2.
This measurement could be  performed in field stations where chemiluminescent
NO analyzers are usually available.
4.2.6.4  Methods of Analysis  of Higher  Homologues.   The  GC-ECD  analyzer  is
likewise used for the higher homologues of PAN (Darley et al., 1963; Stephens,
1969; Heuss  and Glasson,  1968).   The  higher  homologues  elute  with longer
                                    4-42

-------
retention times.  The  first  observation of peroxypropionyl nitrate (PPN) in
heavily polluted air was by Darley et al.  (1963), who also measured peroxybu-
tyryl (PBN) nitrate  in synthetic mixtures by GC-ECD.  The concentrations of
the higher homologues in ambient air are usually below the detection limits of
the GC-ECD technique.  Heuss and Glasson (1968) measured peroxybenzoyl nitrate
(PBzN) in irradiated auto  exhaust samples by GC-ECD  and reported that this
homologue was 100 times  more potent than  PAN  as  a lachrymator.  The direct
analysis of PBzN  by GC-ECD is  reported  to be complicated by  interferences
(Appel, 1973).  Therefore, an analytical technique was developed  in which the
PBzN was quantitatively  hydrolyzed to methyl benzoate (MeOBz),  followed by GC
analysis for  MeOBz  using a flame ionization detector  (Appel, 1973).   An upper
limit of 0.07  ppb was placed on the concentration of PBzN in the San Francisco
bay area.  The  analysis  for the higher  homologues of  PAN in the atmosphere by
FTIR spectrometry is not feasible because  of inadequate  sensitivity,  although
Hanst et al.  (1982)  placed  an upper limit for PBzN in smoggy Los Angeles air
of 1 ppb based on absorption in the 990 cm   region.
     The higher homologues  of PAN may be  prepared for use in  calibration in
the same manner as  PAN by the  use of a  compound containing the parent alkyl
group.  Thus,  PPN and PBzN  have been prepared by the photolysis of alkyl
nitrates in oxygen (Stephens, 1969) and parent aldehydes plus chlorine and N02
(Gay et al.,  1976).  The study of Gay et al.  (1976) confirmed that the first
member of the series, peroxyformyl nitrate [HC(0)02N02], is too unstable to be
observed.  There have been few reports of the absolute analyses for the higher
homologues.    Infrared  absorption analysis  of purified samples  should be the
preferred technique.  Infrared absorptivities of homologues have been reported
by Stephens (1969) and Gay et al. (1976).

4.2.7  Methods  for Sampling and Analysis of Hydrogen Peroxide
4.2.7.1  Introduction.   Hydrogen peroxide  (HpOp) is mechanistically significant
in photochemical smog as a chain terminator and as an index of the hydroperoxyl
radical  (H02) concentration  (Bufalini  and Brubaker,  1969; Demerjian  et al.,
1974).  The major  reaction leading to the formation  of H202 is  the  recom-
bination of the hydroperoxyl radical (Graedel et al., 1976):

                          H02 + M -> H202 + 02 •+' M                     (4-16)
                                   4-43

-------
Recent studies have implicated atmospheric H000 in the aqueous-phase oxidation
              _n                            £. C.
of SOy  to SO*    and  in the acidification of  rain  (Penkett et al., 1979;
Dasgupta, 1980a,b;  Martin and  Damschen,  1981; Overton  and Durham,  1982).
     One of the  major problems in assessing the role of atmospheric HgO^ has
been a  lack  of adequate measurement methodology.   Earlier  measurements for
atmospheric hLCL  were by the titanium colorimetric method (Gay and Bufalini,
1972a,b; Bufalini  et  al.,  1972) and by a chemiluminescence technique (Kok et
al., 1978a,b).   The reported HptL concentrations (0.01  to  0.18 ppm) are now
believed to be far too  high, primarily as a  result  of artifact H202 formation
from reactions of absorbed 0,  (Zika and Saltzman, 1982; Heikes et al., 19,82;
Heikes,   1984).   Furthermore, maximum tropospheric  HJ3,, concentrations pre-
dicted by modeling calculations (Chameides and Tan,  1981; Logan et al., 1981)
and observed  in  recent  field studies (Das et al., 1983)  are  on the order of  I
ppb.   In a recent study, a chemiluminescence technique  was  employed with an
argon sample purge to remove 0~ interference, and a maximum H?0? concentration
of 1.2 ppb was  observed in a  polluted urban environment (Das et al.,  1983).
Promising techniques that have been used or proposed for aqueous- and gas-phase
HpOg are discussed below,  as well as methods  for sampling, generation, and
standardization of H202 samples.
4.2.7.2  Sampling.  Almost all of the methods  used for the measurement of
atmospheric H^O^  have used aqueous traps  for sampling.   Midget impingers (Gay
and Bufalini, 1972a,b; Kok et al., 1978b; Das et al., 1983), continuous extrac-
tors (Kok et al., 1978a), and gas washing traps (Zika and Saltzman, 1982) have
been used.  Aqueous traps have been found to be highly efficient in removing
trace concentrations  of H202  from gas streams  (Zika and Saltzman,  1982).
Atmospheric 03,  however,  which is  also absorbed  at  concentrations much higher
than HgOp, reacts in  the bulk aqueous phase and  at surfaces to produce H^Op
and thus  is  a serious interference (Zika and  Saltzman,  1982; Heikes et al.,
1982; Heikes, 1984).  Details of the aqueous chemistry of QS can also be found
in other  sources (Hoigne and Bader, 1976; Kilpatrick et al., 1955; Taube and
Bray, 1940).   Because the rate of  FLO^ production is relatively slow (Heikes,
1984), the removal  of absorbed 03  by purging immediately after sample  collec-
tion may remove  or  substantially reduce this  interference.   This  was the
approach  employed by Zika  and Saltzman  (1982)  and by  Das  et al,  (1983).
Another  problem  identified with aqueous  sampling is that other atmospheric
species  (in  particular, S02) may  interfere  with  the generation of H202 in

                                   4-44

-------
aqueous traps and  also  react with collected Hg02 to reduce the apparent H^Op
concentration measured (Heikes et al., 1982).
4.2.7.3  Measurement.  A  number of methods  for measuring  low  levels of H^CU
have been reported, including the following:

     1.   Titanium colorimetric methods.
     2.   Chemiluminescence methods.
     3.   Enzyme-catalyzed methods.
     4.   Laser diode infrared method.
     5.   Fourier-transform infrared method  (Niki et al., 1980;
             Hanst et al., 1975, 1982).
     6.   Electrochemical methods (Pisarevskii et al.,, 1980).
     7.   HpOp-olefi.n reaction (Hauser and Kolar, 1968).
     8.   Mixed-ligand complexes (Csanyi, 1981; Meloan et al., 1961).
     9.   lodometry.

Of these techniques,  only the chemiluminescence and enzyme-catalyzed methods
are  summarized  below.   A summary of  methods reported in the  literature  is
given  in Table  4-8.   Although  titanium colorimetric methods  have  been applied
in atmospheric  measurements,  these techniques are now thought to have inade-
quate  sensitivity  for the actual atmospheric concentrations  that  are present.
In addition, these techniques have questionable specificity.   The same comments
apply  to methods  6 through 8 and only the primary references are given above
for  those methods.   At  the present state-of-art, the  inevitable  presence of
water vapor absorption limits the use of Fourier-transform infrared methods to
concentrations  above  about 0.040  ppm (Hanst et al.,  1982).    The use  of a
tunable diode infrared  laser source should  eliminate  the  problem associated
with nearby water  bands,  and  this  method  is  currently  under  investigation for
atmospheric measurements  (unpublished work in progress, Schiff, 1985).   lodomet-
ric  techniques  are useful only for calibration of 1^2 standards and will be
discussed in that  section.
4.2.7.3.1   Chemi1uminescence.   Hydrogen  peroxide  in  the atmosphere may  be
detected at  low concentrations by  the chemi luminescence obtained  from (^(ID-
catalyzed oxidation  of  luminol (5-amino-2,3-dihydro-l,4-phthalazinedione) by
H«00 (Armstrong and Humphreys, 1965;  Kok et al., 1978a,b).  The  reagent is a
                                                    -5
solution containing NaOH  (pH = 12.8) and luminol, 10   M Cu(II).  The products
of the reaction with  HpOg are  3-amino-phthalic  acid,  a nitrogen molecule, and
                                   4-45

-------
                                                 TABLE 4-8.                       FOR          PfROXIDE
,, Method
Titanium
colorimetry
Chewi luminescence
Enzyme-cata 1 yzed
Enzyme-catalyzed
Enzyme-catalyzed
-P>
i Fourier-transform
5J infrared absorption
Electrochemical
H202-o1efin
reactions
Mixed-ligand
complex reagents
Reagent(s)
(1) Titanium Sulfate
-8-Quinolinol
(2) Titaniui Tetrachloride
Luminol, Cu(II)
basic solution
Scopoletin, horseradish-
peroxidase (HRP)
Leuco crystal violet,
HRP
3- (p-hydroxyphenyl )
propionic acid
None
Aqueous solutions
l,2-di-(4-pyridyl)
ethyl ene
Vanadium and
uranium hydroxamic
acid chelates
Limits of
detection
(1) 1.6 x 10-6 H
(2) ca 10-° H
0.001 to 1 ppa
1.5 x 10-11 M
10-s H
10-8 to 10-4 M6
0,040 ppm (est.)
5 x 10-6 to 1 M
10-6 to 5 x 10-4 M
10-6 M
Interferences
Positive
Alkyl hydro-
peroxides
PANd
NA
NA
NA
NAf
NA
OB
NA
Negative
S02C
S02
NA
NA
NA
None
NA
NA
NA
Applications Primary reference
Atmospheric (1) Gay and Bufalini (1972a,b)
(2) Pilz and Johann (1974);
Kok et al. (1978a)
Atmospheric, Armstrong and Humphreys (1965);
rainwater Kok et al. (1978a,b)
Atmospheric, Andreae (1955); Perschke
rainwater and Broda (1961); Zika and
Saltzman (1982)
Mottola et al. (1970)
Zaitsu and Okhura (1980)
Atmospheric3 Hanst et al. (1982)
Pisarevskii et al. (1980)
Hauser and Kolar (1968)
Csanyi (1981);
Heloan et al. (1961)
 Except where noted, detection limits are in moles/1iter(M) in aqueous solution.
 QS is an interference in all these procedures using aqueous sampling.  See Text.   NA = not available.
cThe S02 interference is reported to be small at high S02 concentrations (Gay and Bufalini, 1972b).   Studies of potential positive and
 negative interferences are incomplete for these methods.
 The reported PAN interference is small (Kok et al., 1978b).
eThe lower limit could presumably be reduced by the use of larger samples.
 With sufficient resolution,  there should be no interferences.   IR absorption by atmospheric water vapor is the major analytical
 limitation.
^Oz bands have not been observed in any long-path FTIR studies.   The estimated,lower limit of detection in these studies is
 approximately 0.04 ppm.

-------
a photon of  light at 450 nm.  The detection  limit for atmospheric samples was
given as 0.001  ppm,  and the linear dynamic  range  is 0.001 to 1 ppm.  This
technique as  initially  employed suffered the interferences from 0^  and  SC^
discussed above for aqueous traps.   Das et al. (1982) employed a static version
of the method of Kok et al. (1978a) to measure H?0? concentrations  in the 0.01
to 1 ppb range.   In addition, samples were purged with argon immediately after
collection to eliminate, reportedly, the 0^ interference.
     Recently, a modified chemiluminescence method has been reported that used
hemin, a blood  component,  as a catalyst for the lumino!-based hLOp oxidation
(Yoshizumi et al., 1984).  This method was applied to the measurement of Op
in rainwater.  There was no interference for SOp but a significant positive 0,
interference was reported.
4.2.7.3.2  Enzyme-Catalyzed Methods (Peroxidase).  This general method involves
three components:  a substrate that is  oxidizable;  the enzyme,  horseradish
peroxidase (HRP); and  hLOg.   The production or decay of the fluorescence in-
tensity of the  substrate or  reaction product is  measured as it is oxidized by
HpOg, catalyzed by HRP.   Some of the more widely used chromogenic substrates
have  been  scopoletin (6-methoxy-7-hydroxy-l,2-benzopyrone) (Andreae,  1955;
Perschke and  Broda,  1961); 3-(p_-hydroxyphenyl)propionic acid  (HPPA) (Zaitsu
and  Okhura,  1980);  and leuco crystal violet (LCV)  (Mottola et al., 1970).
     In the  scopoletin  method,  the reagent  solution is  mixed with a second
solution containing the 1^2, the concentration of which must not be less than
0.33  nor more than  0.84 times the  concentration of  scopoletin (Perschke and
Broda, 1961).   The disappearance of scopoletin fluorescence is monitored and
the  fluorescence  intensity can be used  to obtain  the  concentration of H^O^
from  a  calibration curve.   The most significant advantages of this method of
analysis are  the  specificity for H^Op, the sensitivity,  and the stoichiometry
of the  scopoletinrH^Op  reaction (1:1 mole per mole  as  long as scopoletin  is
present in a 20-fold excess over HRP).  The chief disadvantage of the scopoletin
method  is  that  the  concentration of H^Oo must  be  within a narrow  range in
order to  obtain an  accurately measurable decrease  in fluorescence.  This
limits the  usefulness  of the technique in determining  unknown HgOg  concentra-
tions over several orders  of magnitude (Armstrong and Humphreys, 1965; Andreae,
1955).  Detection limits for this technique are quite low and are in the range
of 1.5  x  10"11 M (Perschke  and  Broda, 1961) to  2  x  10"10 M (Andreae, 1955).
This  technique  has been applied to measurements of H202 in rainwater by Zika
and  Saltzman  (1982).
                                   4-47

-------
     With the  leuco  crystal  violet  (LCV)  substrate,  intensely  colored  crystal
violet is formed from the reaction of HpOp with LCV, catalyzed by HRP (Mottola
et al.,  1970).   The  absorbance is measured  at 596 nm, where the absorption
coefficient of crystal violet is 10  M  cm   , a very high value and an inherent
advantage of this  method.   The concentration of HpOp is a linear function of
the concentration of crystal violet produced.  The detection limit reported is
        _Q
about 10   M HpOp for an absorbance of 0.005 in a 5-cm cell.
     The HRP catalyzes  the  oxidation of  a wide variety  of hydrogen-donating
substrates by HpOg.  Zaitsu and Okhura (1980) have tested a number of 4-hydroxy-
phenyl compounds and found that 3-(p_-hydroxyphenyl) propionic acid (HPPA) pro-
vided the most sensitive and rapid  means for determining H202.  When HPPA
reagent  solution is  mixed  with HRP  solution and  a test solution containing
H202, a  product  is formed  that fluoresces at 404  nm following excitation at
320 nm.   The intensity of this fluorescence  is monitored as a function of H^O,,
                                                          _in               £• t-
concentration.   The  detection  limit  was  reported to  be 10   mole H000 with a
                             -8
linear range extending  to  10   mole HpOo when a test solution of only 0.1 ml
volume was  used.   Presumably the molar sensitivity  could be improved  by the
use of  larger  sample volumes.   No interference studies  were reported.  More
recently, the  acetic acid  homologue of HPPA has been employed (Kunen et al.,
1983; Dasgupta and Hwang,  1985).   The fluorescence  intensity  of the product
dimer is directly proportional to H/,0,, concentration.
     The enzymatic methods  appear to be  the most  promising aqueous, colori-
metric methods for H202 and have considerably greater sensitivity than other
colorimetric methods.  Studies of potential  atmospheric  interferences, however,
have not been reported for any of these three substrates.
4.2.7.4  Generation and Calibration Methods
4.2.7.4.1  Generation.   Standard  samples containing trace concentrations of
H202 are required for testing and calibration of various measurement methodol-
ogies. As with DO, such standards are not available and are usually prepared
at the  time of use.   A  number  of  techniques  have been  employed for generating
aqueous  standards,  but  convenient methods  for the generation  of gas-phase
standards are noticeably lacking.
     Techniques for  the  generation  of high  concentrations of  H202 have been
discussed by Shanley (1951).   Commercial  solutions of  30 percent aqueous H202
are readily available.   Trace  levels of  H009 in water  may be generated by the
                           fiO
irradiation of water with    CO ^-radiation  (Hochanadel,  1952;  Armstrong and
Humphreys,  1965)  and by enzymatic means  (Andreae,  1955).   By far the most
                                   4-48

-------
convenient  method  for generating  aqueous standards  containing  micromolar
concentrations of HpOp is  simply the serial dilution of commercial-grade.30
percent Og  (Fisher  Analytical  Reagent).  Samples prepared  in  this manner
must be standardized  and the method usually employed is the iodometric tech-
nique discussed below.   Stock  standard solutions of H^00 as low  in concentra-
          _4                                          ^  £
tion as 10   M have been found to be stable for many weeks if kept in the dark
(Armstrong and Humphreys, 1965),
     Techniques for the  convenient generation  of gas-phase standards are not
available.  With  the  increased  interest in atmospheric H«Q,,, there is,an
obvious need  for an  hLO^ generator comparable to the photolytic 0^ generator
discussed in Section 4.2.4.1.  A technique that has often been used for gener-
ating ppm concentration  levels of  H^O^ in air  has been  described by  Cohen and
Purcell (1967).  Micro!iter quantities of 30 percent \\J^^ solution are injected
into a metered stream of air that  flows  into a Teflon bag.  The  concentration
of  HgQp  in  the bag is  then determined by the iodometric  method discussed
below.
4.2.7.4.2   Calibration.   By  far  the most common method for standardizing low
concentrations of H^O* is based  on iodometry (Allen et  al., 1952; Hochanadel,
1952; Cohen et al., 1967).   Hydrogen peroxide  liberates iodine from  an iodide
solution quite slowly, but in the presence of a molybdate catalyst the reaction
is  quite  rapid.   The  iodine liberated may  be  determined by titration with
standard thiosulfate at  higher concentrations or by photometric measurement of
the tri-iodide ion at low concentrations.  The molar absorption coefficient of
the tri-iodide ion at 350 nm has been determined to be 2.44 x 10  by measuring
  mmfc      _jfl
10   to 10   M H202 solutions prepared from 0.2 M stock H202 solution standard-
ized against  primary  grade  arsenious oxide (Armstrong  and Humphreys, 1965).
The  stoichiometry  is  apparently  1 mole  of  iodine released per mole  of HpOo-
Definitive  studies of the  stoichiometry, however, have not been performed to
the same extent as those of  the  stoichiometry  for the iodometric determination
of03.
4.3   SAMPLING,  MEASUREMENT,  AND CALIBRATION METHODS  FOR  PRECURSORS  TO OZONE
     AND OTHER  PHOTOCHEMICAL OXIDANTS
     During the last decade, a number of advances have been made in the method-
ology for determining nonmethane organic compounds and oxides of nitrogen.  An
overview of these advances is discussed in the following sections. In the case
of measurement  methods  for nonmethane organic compounds (NMOCs), early ozone
                                   4-49

-------
control models did  not require speciation of the complex mixture of organics
in ambient air.   As a result, non-speciation methods  were employed for the
purpose of providing a data base for early photochemical modeling studies.  As
the air quality  models grew more sophisticated,  however,  the  need  arose  for
more specific information concerning the organic composition of the atmosphere.
Consequently, methodology was  developed  to provide  for detailed  speciation  of
NMOCs.  In addition to improving the data base for photochemical modeling, the
NMOC  speciation  techniques  have also been utilized to characterize various
sources of pollution (e.g.3 mobile versus stationary).
     The development  of  methodology for oxides of  nitrogen  has  likewise  ad-
vanced since  the original EPA  Federal  Reference  Method for measurement  of
nitrogen dioxide, the Jacobs-Hochheiser technique,  was withdrawn in 1973.  A
number of methods for nitric oxide and nitrogen dioxide have been proposed and
evaluated since then and are described here.

4.3.1  Nonmethane Organic Compounds
     Numerous sampling, analytical, and calibration methods have been employed
to determine  vapor-phase  NMOCs  in ambient air.  Some of the analytical methods
utilize detection techniques that are highly selective and sensitive to specific
functional groups or  atoms  of a compound (e.g.,  formyl group of aldehydes,
halogen), while  others respond in a more universal  manner;  that is, to the
number of carbon atoms present in the organic molecule.   In this overview of
the most pertinent measurement methods, NMOC have been arranged  in three major
classifications:  nonmethane  hydrocarbons,  aldehydes,  and other oxygenated
compounds.  Sampling, analytical, and calibration procedures are discussed for
each  classification.   Reference is  also made  to those analytical  methods
utilized in more than one of the above classifications.
4.3.1.1  Nonmethane Hydrocarbons.   Nonmethane  hydrocarbons (NMHC) constitute
the major portion of NMOC in  ambient  air  (Chapter 3).  Traditionally,  NMHC
have been measured by methods that employ a flame ionization detector (FID) as
the sensing element.  This detector was originally developed for gas chromato-
graphy and  employs  a  sensitive electrometer that measures a  change in  ion
intensity resulting from the combustion of air containing organic compounds.
Ion formation is essentially proportional to the  number of carbon atoms present
in the organic  molecule (Sevcik, 1975).  Thus, aliphatic, aromatic, alkenic,
and acetylenic  compounds all  respond similarly to  give relative responses of
1.00 ± 0.10 for  each  carbon atom present in the molecule (e.g.,  1 ppm hexane =
                                   4-50

-------
6 ppm C;  1 ppm benzene = 6  ppm C; 1 ppm propane =  3 ppm  C).  Carbon atoms
bound to oxygen, nitrogen, or halogens give reduced relative responses (Dietz,
1967).   Consequently, the FID, which is primarily used as a hydrocarbon measur-
ing method,  should  more correctly be viewed  as  an  organic carbon analyzer.
     In the  following  sections,  discussion  focuses  on the various methods
utilizing  this  detector to  measure  total  nonmethane organics.  Methods  in
which no  compound  speciation  is obtained are covered  first.   Methods  for
determining individual organic compounds are then discussed.
4.3.1.1.1  Non-speciation methods.   The  EPA  reference method  for  nonmethane
organic compounds,  which  was promulgated in 1971, involves the gas chromato-
graphic separation  of methane (CH,) from the remaining organics in an ,air
sample (F.R., 1971).   A second sample is injected  directly to the detector
without methane separation.   Subtraction of  the first  value from  the second
produces a nonmethane organic concentration.
     A number of  studies  of  commercial analyzers employing the Federal Refer-
ence Method  have  been reported (Reckner, 1974; McElroy and Thompson,  1975;
Harrison et  al.,  1977;  Sexton et  a!., 1982).   In  one of the  first studies,
analyses  of  known synthetic mixtures of hydrocarbons  were conducted by  16
users of  the reference  method  (Reckner,  1974).  The  nonmethane concentrations
tested in this study were 0.23 and 2.90 ppm C.  The results shown in Table 4-9
indicate the percentage error  from the two known concentrations.   At the  0.23
ppm level, the  majority of the measurements were in error by  amounts greater
than 50 percent.  At 2.90 ppm,  most  of the measurements were in  error by  only
20 percent or less.
     In general, all of the above  studies indicated an overall  poor performance
of  the  commercial  instruments  when  either  calibration or  ambient  mixtures
containing NMOC concentrations less than 1 ppm C were used.  The major problems
associated with using these NMOC  instruments have  been reported in a recent
technical  assistance document  (U.S.  Environmental  Protection  Agency, 1981).
The technical  assistance  document also suggests ways to reduce the effects of
existing problems.   Table 4-10 summarizes the major problem areas and lists
recommendations for  reducing these effects.
     Other approaches to the measurement of nonmethane  organics have also been
investigated.  One  such method,  developed in  1973,  utilizes the  fact that CH*
requires  more  heat  for combustion than other organics  (Poll and Zinn, 1973).
One portion  of the  air sample passes through a catalyst bed where all hydro-
carbons except CH^  are  combusted.  This  sample stream then enters an FID where
                                   4-51

-------
          TABLE 4-9.  PERCENTAGE DIFFERENCE FROM KNOWN CONCENTRATIONS
             OF NONMETHANE HYDROCARBONS OBTAINED BY SIXTEEN USERS

    Known                        % difference from given concentration
ppm
0.23
2.90
>100
6
2
50 to 100
4
0
20 to 50
3
3
10 to 20
2
2
0 to 10
1
9
Source:  Reckner (1974).

the CH* concentration  alone  is  recorded.  The  remaining portion of the sample
passes directly to  a second FID for a total organic carbon measurement.   The
simultaneous processing  of  both signals yields an  NMOC  value.   Although it
provides a  continuous  measurement  of NMOC levels, this method is  also subject
to many  of the  same shortcomings  attributed  to the  EPA reference method.
     Recently, a prototype instrument that measures NMOCs by optical  absorption
has been developed  (Manos et a!.,  1982).   The unit oxidizes NMOCs to carbon
dioxide (CO*)  and uses  a non-dispersive infrared  absorption  technique  to
measure the organic burden indirectly.   Ascarite serves to remove C02 initially
present in  air  and  a hopcalite catalyst selectively  oxidizes organics other
than methane to COg  and H20.  Since carbon  monoxide (CO)  will also oxidize to
COg during  this process,  a  dual-channel  system is utilized to correct for the
contribution from ambient CO concentrations.   This  unit performed well during
a  brief  laboratory evaluation  using  calibration standards;  however,  more
extensive laboratory and field  tests  are needed before the unit  can be  con-
sidered suitable for NMOC measurement.
     Other methods under development and evaluation include oxidation-reduction
schemes in  which  nonmethane organics  are chromatographically separated  from
methane and non-organic species and then oxidized to  C02, reduced to CH«, and
detected by FID  (U.S.  Environmental  Protection Agency, 1978b).   When organic
carbon concentrations  are greater than  100  ppb, the  reduction  step can be
eliminated and a non-dispersive infrared analyzer can be used to detect the C02
formed during the oxidation  step (Salo et a!., 1975).
     A method  for  measuring NMOC directly  has been reported  by Cox et  al.
(1982).  This  approach involves the cryogenic preconcentration of nonmethane
organic compounds and  the measurement of the  revolatilized NMOCs using flame
                                   4-52

-------
           TABLE 4-10.  PROBLEMS ASSOCIATED WITH GATHERING NMOC DATA
   WITH AUTOMATED ANALYZERS AND RECOMMENDATIONS FOR REDUCING THESE EFFECTS


Problem Areas:

1.   Contaminants may be present in compressed-gas cylinders containing
    calibration gases.
2.   Compressed-gas cylinders of calibration gases sometimes contain the
    standard in a nitrogen or argon background.  When no oxygen is blended
    with these gases, FID sensitivity is altered.
3.   The assay of calibration gases contained in compressed-gas cylinders (as
    received from the supplier) is sometimes incorrect.
4.   There are wide differences in the per-carbon response to different NMOC
    species.
5.   FID analyzers require hydrogen, which presents a potential operational
    hazard.
6.   The NMOC concentration is obtained by subtraction of two relatively large
    and nearly equal numbers (TOC-CH«=NMOC) and thus is subject to large,5 rela-
    tive errors.
7.   NMOC analyzers may exhibit excessive zero and span drift during unattended
    operation.
8.   The complex design of some NMOC analyzers creates unique problems that are
    generally not experienced in other pollutant analyzers.  Meticulous set-up,
    calibration, and operation procedures (which are analyzer-specific) are
    difficult to understand and follow.
                                                                       i i
Recommendations:                                                        . ,

1.   Calibration gases should be checked to determine the concentration of
    contaminants.                                                       i
2.   Calibration concentrations should be obtained by dynamic dilution of a
    pollutant standard with zero-grade air containing oxygen.  The dilution
    ratio should be sufficiently high (vLOO:l) to ensure that the calibration
    sample contains 20.9% ± 0,3% oxygen.
3.   All calibration standards contained in compressed-gas cylinders should be
    traceable to Standard Reference Materials from the National Bureau of
    Standards.
4.   The NMOC response should be calibrated to a propane standard.
5.   The operator should use documented procedures for hydrogen safety.
6.   All channels should be properly calibrated.
7.   The FIDs should be operated in accordance with instructions supplied by
    the manufacturer and this document.
8.   The training of qualified operators should be augmented with a Technical
    Assistance Document, which provides detailed calibration and operation
    procedures for NMOC analyzers.


Source:  U.S. Environmental Protection Agency (1981).
                                   4-53

-------
ionization detection.  The  procedure involves the following  steps.  A fixed
volume of  sample  is drawn through a  trap  cooled  to  liquid  argon temperature
(liquid Hy cannot be used since it will also condense methane and oxygen).  At
this temperature  all  NMOCs  are condensed  into the trap.  After the  residual
CH* and oxygen  are cleared from the trap by the helium carrier gas, the trap
temperature is  raised  to revolatilize the NMOC.  Oxygen  removal reduces  the
variation in response to different organic compounds.
     Jayanty et al.  (1982)  used a similar system to study the responses of a
variety of aliphatic and aromatic compounds.   Good linearity was observed with
the cryogenic trapping procedure (500 ml of air) over a concentration range of
50 to 5,000 ppb C.  Humidity did not generally interfere with the analysis.
Sample precisions of ±5 percent for single- and multiple-component gas standards
and ±10 percent for ambient samples were consistently achieved. Responses  for
aromatic compounds, however, were less than expected.  The  researchers recom-
mended additional  testing and  instrument refinement in order to resolve this
problem.
4.3.1.1.2  Speciation methods.   The primary separation technique utilized  for
NMOC speciation  is gas chromatography  (GC).  Coupled with  flame  ionization
detection, this analytical method permits the separation and identification of
many of the organic species present in ambient air.
     Compound separation is accomplished by means of both packed and capillary
GC columns.  If  high  resolution is not required and large sample volumes are
to be injected,  packed columns are employed.  The traditional packed column
may contain either (1) a solid  polymeric adsorbent (gas-solid chromatography)
or (2) an  inert support, coated with  a liquid  (gas-liquid chromatography).
Packed columns  containing an adsorbent substrate are required to separate
C^-Co compounds.   The  second  type of column can be a support-coated or wall-
coated open tubular capillary column.   The latter column  has  been widely  used
for environmental  analysis  because of its superior resolution and  broader
applicability.   The wall-coated capillary column consists of a liquid station-
ary phase  coated  or bonded (cross-linked) to the specially treated  glass  or
fused-silica tubing.  Fused-silica tubing is most commonly used because of its .
physical durability and  flexibility.   When a complex mixture is  introduced
into a GC  column,  the  carrier gas (mobile  phase)  moves the  sample through  the
packed or  coated capillary column (stationary  phase).   The chromatographic
process occurs  as  a  result of repeated sorption-desorption  of  the sample
                                   4-54

-------
components (solute)  as they  move  along the  stationary  phase.   Separation
results from the different affinities that the solute components have for the
stationary phase.
     The GC-FID  technique has been used by  numerous researchers to obtain
ambient NMOC data.   Singh (1980) drew on the cumulative experience of these
researchers to prepare  a  guidance  document for state and local  air pollution
agencies interested in obtaining speciation data.  In general, most researchers
have employed  two  gas chromatographic units to  carry  out analyses of NMOC
species in ambient  air.   The more volatile organic compounds (Cp through Cg)
are generally  measured  on one unit using packed-column technology, while the
other GC separates  the less  volatile organics using a capillary column.  In
typical chromatograms  of  urban  air,  all major peaks are identified and, on a
mass basis,  represent from  65  to  90 percent of the measurable nonmethane
organic burden.  Identification  of GC peaks is based upon matching retention
times of unknown compounds with  those of standard mixtures.   The use of  dedi-
cated computer systems  facilitates this task, but close scrutiny of the data
is still necessary  to  correct periodic misidentification of  unknown compounds
resulting from variations in retention time.  Subsequent verification of the
individual species  is normally  accomplished with gas chromatographie/mass
spectrometric  (GC/MS)  techniques.   Compound-specific detection  systems,, such
as electron capture, flame photometry, and spectroscopic techniques, have also
been employed for peak identifications.   A discussion of these systems, however,
is beyond the scope of this report.  Several documents covering these detection
systems are available (Lamb et al., 1980; Riggin, 1983).
     Because the organic  components  of the ambient atmosphere are present at
ppb levels or  lower, some means of sample  preconcentration  is  necessary to
provide sufficient material for the GC-FID system.  The two primary techniques
utilized for this purpose are the use of solid adsorbents and cryogenic collec-
tion.   The more commonly  used sorbent materials are divided into three catego-
ries:    (1)  organic polymeric adsorbents, (2) inorganic  adsorbents,  and (3)
carbon adsorbents.   Primary organic polymeric adsorbents used for NMOC analyses
include the materials Tenax GC and XAD-2.  These materials have a low retention
of water  vapor and,  hence,  large  volumes of air can  be collected.   These
materials do not,  however, efficiently  capture highly volatile  compounds such
as C«  to  Cg  hydrocarbons, nor certain  polar compounds  such as methanol and
acetone.  Primary  inorganic  adsorbents  are  silica gel, alumina, and molecular
sieves.  These materials  are more  polar than the organic polymeric adsorbents
                                   4-55

-------
and are  thus more  efficient  for  the  collection  of the  more  volatile and polar
compounds.  Unfortunately, water is  also  efficiently collected, which  in many
instances leads to rapid deactivation of the adsorbent.  Carbon adsorbents are
less polar than the inorganic adsorbents and, as a result, water adsorption by
carbon adsorbents  is  a less  significant problem.  The carbon-based materials
also tend to exhibit much stronger adsorption properties than organic polymeric
adsorbents; thus,  lighter-molecular-weight  species are more easily retained.
These same adsorption  effects  result, however, in irreversible adsorption of
many compounds.  Furthermore,  the very high thermal desorption temperatures
required  (350  to 400  °C) limit their use and also may lead to degradation of
labile compounds.   The commonly available classes of carbon adsorbents include:
(1) various conventional activated carbons; (2) carbon molecular sieves (Sphere-
carb, Carbosphere,  Carbosieve);   and  (3)  carbonaceous  polymeric adsorbents
(Ambersorb XE-340,  XE-347, XE-348).
     Although  a number of researchers have employed solid adsorbents for the
characterization of selected organic species  in air, only a few attempts have
been made to  identify and quantitate the range of organic  compounds from C~
and above.  Westberg  et  al.  (1980) evaluated  several carbon and organic poly-
meric adsorbents and found that Tenax-GC exhibited good collection and recovery
efficiencies for >Cg organics;  the remaining adsorbents tested (XAD-4,  XE-340)
were found unacceptable  for  the  lighter organic fraction.   The XAD-4 retained
>C« organic gases,  but it was impossible to  completely desorb these species
without partially decomposing the XAD-4.   Good collection and recovery efficien-
cies were provided by XE-340 only for organics of C* and above.
     Ogle et al. (1982) used a combination of adsorbents in series and designed
an automated GC-FID  system for analyzing C^ through C^g hydrocarbons.   Tenax
GC was utilized for Cg and  above, while  Carbosieve  S trapped C, through Cg
organics.  Silica  gel  followed these adsorbents and effectively removed water
vapor while passing  the C2  hydrocarbons onto a molecular-sieve 5A adsorbent.
The combined sorbents  have been  laboratory-tested with a SB-component hydro-
carbon mixture.   Good collection and recovery efficiencies were obtained.
Preliminary field  tests  have also been successful, but a very limited data
base  exists.   Futhermore, the current chromatographic procedures  utilize
packed-column  technology.  The addition  of capillary  columns to this  system
would permit  better resolution  of the  complex mixtures typically found in
ambient air.
                                   4-56

-------
     The preferred method  for obtaining NMOC data is cryogenic preconcentra-
tion (Singh, 1980).  Sample preconcentration  is  accomplished by directing air
through a packed trap immersed in either liquid oxygen (b.p. -183°C) or liquid
argon (b.p. -186°C).   For  the detection of about  1 ppb C of  an  individual
compound, a 250-cc air sample is normally processed.  The collection trap is
generally filled with  deactivated 60/80 mesh glass  beads  (Westberg et a!.,
1974), although coated chromatographic  supports, have also  been used (Lonneman
et al.,  1974).  Both of the above cryogens are  sufficiently warm  to allow air
to pass completely through the trap, yet cold enough to collect trace organies
efficiently.  This collection procedure also condenses water  vapor.   An air
volume of 250  cc  at  50 percent  relative humidity  and 25°C contains approxi-
mately 2.5  mg  of  water,  which appears  as  ice in the  collection  trap.  The
collected ice at times will plug the trap and stop the sample flow; furthermore,
water transferred to the  capillary column during the thermal des.orption step
occasionally causes plugging and other deleterious column effects.  To circum-
vent water condensation problems, McClenny et al. (1984) have employed a Nafion
tube drier  to remove water vapor selectively during the  sample  collection
step.  These researchers have constructed an automated cryogenic preconcentra-
tion sampling and analysis GC system using this drier and are currently conduc-
ting field evaluaton studies on their system.
     In  addition  to  direct sampling via preconcentration  with sorbents and
cryogenic techniques,  collection of whole air samples  is  frequently used to
obtain NMOC data.   Rigid  devices such as syringes,  glass bulbs, or metal
containers  and  non-rigid devices such as Tedlar and Teflon plastic bags are
often utilized  during  sampling.   The primary purpose of whole-air collection
is to store an  air sample  temporarily  until subsequent laboratory analysis  is
performed.  The major  problem with  this approach is  assuring the  integrity  of
the sample contents prior to analysis.  Of concern is whether sample components
of interest are adsorbed or decomposed  through  interaction with the container
walls or reaction with co-collected gases  such  as  ozone and nitrogen dioxide.
Sample condensation may also  occur  at  elevated  concentrations  or  when samples
are  stored  under  high  pressures  (i.e., in metal containers).  Contamination
from sampling  containers  is  likewise a frequent occurrence (Lonneman et al.,
1981; Seila et al., 1976).   Table 4-11 summarizes the advantages and disadvan-
tages of the primary collection media for NMOC analysis,,
4.3.1.1.3   Calibration.    Calibration  procedures  for NMOC instrumentation
require  the generation of  dilute mixtures at concentrations  expected to be
                                   4-57

-------
          TABLE 4-11.          OF ADVANTAGES AMD DISADVANTAGES OF PRIMARY COLLECTION MEDIA FOR NMOC ANALYSIS
       Sampling technique
            Advantages
            Disadvantages
    1.  Solid adsorbents
en
00
    2.  Cryogenic pre-
        concentration
   Many sorbents do not retain H20
   vapor; thus, large volumes of air
   can be processed.

   Integrated samples from a period
   of minutes to days are easily
   obtained.

   Sample cartridges are convenient
   for field use.
o  A wide range of organic material
   can be collected,

o  Artifact problems are avoided.
                               o  Collected organics are immediately
                                  available for analysis, without
                                  solvent removal or use of high
                                  thermal desorption temperatures.

                               o  Collected species are stable; good
                                  recovery efficiencies are obtained.
   No single adsorbent can be used to
   collect and recover organics of G£
   and above.

   Contamination and artifact peaks
   plague many sorbent systems.
o  Many adsorbents require high
   (>300°C) thermal desorption tem-
   peratures, which may lead to
   degradation of labile compounds,
   artifact peak formation, etc.

o  Breakthrough volume and collection
   and recovery efficiencies must be
   determined for each compound of
   interest.

o  Volume of air collected is limited
   by amount of moisture condensing.

o  Liquid argon or oxygen is necessary
   for preconcentration.

o  Field collection apparatus is bulky
   compared to adsorbent cartridges.

-------
              TABLE 4-11 (continued).  SUMMARY OF ADVANTAGES OF PRIMARY COLLECTION MEDIA FOR NMOC ANALYSIS
         Sampling technique
            Advantages
           Disadvantages
•p*
en
      3.  Rigid containers
          (Metal canisters)
      4.  Non-rigid containers
          (Teflon and Tedlar
          Bags)
o  Can be treated to make them
   chemically unreactive.

o  Durable; easy to clean, transport,
   and use.

o  Can be pressurized; leakage and
   permeation minimized.

o  Excellent stability for many trace
   species; long-term storage is
   possible.

o  Readily available.

o  Convenient for collecting integrated
   samples.

o  Good short-term stability of trace
   species.
o  High initial cost.
                                                                            o  Limited collection volume.
                                                                               Difficult to collect integrated
                                                                               samples.

                                                                               Sample may be lost to walls by
                                                                               condensation.
o  Subject to breakage (at seams)
   during handling.

o  Admits sunlight.

o  Slow permeation of chemicals out
   of and into plastic bags during
   storage.

o  Outgassing contamination from bag
   material.

o  Short storage life.
      Source:  Derived from Singh (1980); Jayanty et al.  (1982); Sexton et al
       (1976); Lonneman et al. (1981); Holdren et al.  (1982).
                                              (1982); National Research Council

-------
found in ambient  air.   Methods  for  generating  such mixtures are classified as
static or dynamic systems.
     Static systems  are generally preferred  for  quantitating NMOCs,  The most
commonly used static system is a compressed gas cylinder containing the appro-
priate concentration of the compound of interest.  These cylinder gases may
also be diluted with hydrocarbon-free air to provide multi-point calibrations.
Calibration and  hydrocarbon-free air cylinders  are  available  commercially.
Additionally,  some  standard gases  such  as propane and benzene are available
from the National  Bureau of Standards (NBS) as  certified standard reference
materials (SRM).   Commercial  mixtures  are generally referenced against these
NBS standards.  In  its  recent technical assistance document for operating and
calibrating continuous  NMOC  analyzers,  the U.S.  Environmental  Protection
Agency (1981)  recommended  propane-in-air  standards  for calibration.  Some
commercially available  propane cylinders  have been  found to contain  other
hydrocarbons (Cox et a!., 1982), so that all calibration data should be refer-
enced to NBS  standards.   Because of the  uniform carbon  response of  a  GC-FIO
system (±10 percent) to hydrocarbons (Dietz, 1967),  a common response factor
is assigned to both identified and unknown compounds obtained from the specia-
tion systems.   If these compounds  are oxygenated species, an underestimation
of the actual  concentrations will  be reported.   Dynamic calibration systems
are employed when better accuracy  is needed for these oxygenated hydrocarbon
species.   Dynamic systems are normally employed to generate jji situ concentra-
tions of the individual  compound of concern and include devices such as permea-
tion and diffusion tubes and syringe delivery systems.
4.3.1.1.4  Comparison of non-speciationversus speciation methods.   Speciation
methods involving cryogenic preconcentration have been compared with non-speci-
ation NMOC analyzers in the following studies.
     Jayanty et al.  (1982)  conducted a laboratory comparison between the pro-
totype non-speciation method described  earlier (Section 4.3.1.1.1) and their
gas chromatographic  separation  method.   Comparison of the two methods for 12
ambient air samples  collected in stainless  steel  canisters showed agreement
within ±15 percent.  Ambient air concentrations ranged from 100 to 1000 ppb C.
     Lonneman (1979) compared  total  NMOC and speciation methods during field
studies in Houston  in  1978.  Samples were collected during 3-hour integrated
time periods (6 to 9 a.m., 1 to 4 p.m.) in Tedlar bags for subsequent analysis.
The correlation coefficients for 150 measurement pairs from five sites averaged
0.74.  For data pairs of 500 ppb C and less, an average correlation coefficient
                                   4-60

-------
(r) of 0.55 was  calculated,  with a low value of 0.12 at one site.  Lonneman
attributed the low correlations to maintenance and calibration problems in the
continuous analyzers and  concluded  that the results from the, non-speciation
method are "at best  marginal  for use  in  photochemical  model  applications."
     Holdren et al.  (1982) made a similar comparison during a 2^month study at
urban sites in Cincinnati  arid Cleveland, Ohio.   They utilized a GC/cryogenic
trapping technique and  compared their results with data from state-operated
NMOC analyzers (SAROAD  data).   Ambient air samples were collected  in Teflon
bags (6 to  9  a.m.  integrated collection) and were transferred immediately to
pretreated aluminum cylinders for shipment and analysis at the central labora-
tory.  Concentrations of  NMOC ranged from 200 to 2100 ppb C.   Linear regres-
sion analyses of the non~speciated  versus  speciated data resulted in  correla-
tion coefficients  that ranged  from 0.75  to 0.92 for the  four  urban sites
(total of 67  comparisons).  Limiting  the  comparisons to concentrations of  500
ppb C and lower resulted in an average correlation coefficient of 0,10.
     Richter  (1983) compared  continuous total NMOC with GC  speciation  results
obtained  at  seven fixed ground-level  sites  used in the Northeast  Corridor
Regional Modeling Project (NECRMP).   The NMOC data were obtained in real  time,
while Teflon  bags were  used to  collected  integrated samples (6 to 9 a.m.)  for
the GC/cryogenic analyses.  Over 60 comparisons were available from each site.
Table 4-12  summarizes  statistical  information  obtained from least-squares
analysis  of the  data  (Richter,  1983).   As the table indicates, only data  from
the  East  Boston site exhibited  a high correlation coefficient.  This  study
represents the most extensive effort made yet to compare the two NMOC measuring
methods.  Emphasis was  placed on correct  instrument operation,  calibration,
etc.; and only verified data were compared.  Yet the above results  indicate that
much more work  is needed to  resolve the differences between the two methods.
4.3.1.2  Aldehydes.  Historically,  the major problem in measuring concentrations
of aldehydes  in ambient air has  been to find an appropriate monitoring technique
that is sensitive to low concentrations and specific for the various homologues.
Early techniques  for measuring formaldehyde,  the most  abundant aldehyde,  were
subject to many interferences and lacked  sensitivity at low ppb concentrations.
The  more  recently developed techniques can be utilized to accurately measure
the  various  types and amounts  of aldehydes  at  ppb  levels.  This section  de-
scribes those methods currently used  for  measuring aldehydes in ambient air.
                                   4-61

-------
             TABLE 4-12.  GC/CONTINUOUS NMOC ANALYZER COMPARISONS,
                           LEAST-SQUARES REGRESSIONS

                                                      Standard          ,
   Location            Slope    Intercept, ppm C        error          r
West End Library,      0.552         -0.552             0.672         0.169
  Washington, DC
Read Street,           0.113         -0.283             0.713         0.0077
  Baltimore, MD
Essex, MD
Linden, NJ
Newark, NJ
East Boston, MA
Watertown, MA
0.835
0.531
0.987
1.108
0.750
-0.101
NAa
-0.277
+0.095
-0.568
0.599
0.865
0.574
0.327
0.574
0.354
0.141
0.467
0.887
0.475
aData not available.
Source:  Richter (1983).

These include the chromotropic acid (CA) method for formaldehyde, the 3-methyl-
2-benzothiazolone hydrazone (MBTH) technique for total aldehydes, Fourier-trans-
form infrared  (FTIR)  spectroscopy,  and the high-performance liquid ehromato-
graphy (HPLC) method employing 2,4-dinitrophenylhydrazine (DNPH) derivatization.
4.3.1.2.1  Chromotropic acid method.  The chromotropic acid method (CA) involves
the  collection of formaldehyde in  a  midget impinger containing an aqueous
mixture of chromotropic and sulfuric acids, followed by the spectrophotometric
measurement of absorbance  of the resulting color  (Altshuller  and McPherson,
1963;  U.S.  Dept. of Health,  Education  and Welfare, 1965).  A  modification
described by Johnson  et al.  (1981) improved the accuracy  and  sensitivity of
the CA method by reducing the concentration of sulfuric acid and by eliminating
a heating cycle,  relying solely on  the  heat of  solution generated by  sulfuric
acid (Altshuller  et al., 1961; Olansky  and Deming,  1976).  Trapping formalde-
hyde in a 1 percent bisulfite solution before adding the CA solution increased
collection efficiency  from 84 percent to  92 percent with no sulfite interfer-
ences.
                                   4-62

-------
     The CA method  for measuring formaldehyde has been widely studied (Salas
and Singh,  1982;  Grosjean and  Kok,  1981;  National  Research Council, 1981;
Tuazon et al.,  1980;  Lloyd,  1979).  Originally  developed  as  a spot-test by
Eegriwe (1937), it  was adopted to quantitate formaldehyde spectrophotometri-
cally (Bricker  and  Johnson,  1945; West and  Sen,  1956)  and was modified for
ambient air measurements in the early 1960s (Altshuller et al., 1961; Altshuller
and McPherson,  1963;  U.S. Dept.  of  Health,  Education,  and Welfare,  1965).
While used widely today for both occupational and ambient air environments,
its specificity for formaldehyde, which accounts for  approximately half of
total  ambient  air  aldehydes  (see Section 3.5),  limits  its  usefulness  for
characterizing aldehyde concentrations in ambient air.
     The CA method  has been reported to be sensitive to acrolein, acetaldehyde,
phenol, nitrogen dioxide,  and  nitrates  (National  Research  Council,  1981;  Krug
and Hirt, 1977; U.S.  Dept. of  Health,  Education, and  Welfare, 1973; Sleva,
1965; Altshuller et al.,  1961).   Recent work,  however,  indicates  that neither
nitrates, nitrites, NO^,  nor  acetaldehyde at elevated  ambient  air  levels
interfere with the  CA analysis  (Johnson et al., 1981; Grosjean and Kok,  1981).
Relevant data on other interfering agents were not found.
4.3.1.2.2  MBTH method.   A spectrophotometric technique for total  aldehydes
was developed  in the  early 1960s  by  Sawicki  et al. (1961).  Known as  the  MBTH
method, it  involves the reaction of aldehyde with 3-methyl-2-benzothiazolone
hydrazone to form an  azine that is oxidized  by a ferric  chloric-sulfamic  acid
solution to form a  blue cationic  dye (Altshuller, 1983;  U.S.  Dept.  of Health,
Education, and Welfare, 1965; Hauser and Cummins, 1964; Altshuller and McPherson,
1963; Altshuller and  Leng, 1963; Altshuller et al., 1961).
     The MBTH  method  has  a reported sensitivity of  15 ppb for,  primarily,
low-molecular-weight  aldehydes  (National  Research Council, 1981).   The  method
is  subject  to  interferences  by N02 and gives an inconsistent response to
higher-molecular-weight  aldehydes  (Sawicki  et al.,  1961;  Altshuller  etal.,
1961).  Nonetheless,  the Intersociety Committee  of the  American  Public  Health
Association recommends the MBTH  colorimetric method for  determining total
aldehydes in air  (Intersociety Committee, 1977a).  Miksch and Anthon (1982)
devised a sampling  and analysis scheme that permits a single MBTH sample to be
used for both formaldehyde and  total aliphatic aldehyde determinations.
4.3.1.2.3   Fourier-transform infrared spectroscopy.    Infrared  absorption
spectroscopy has  been used to  identify and  measure  aldehydes  in  ambient  air
(Hanst  et al.,  1975,  1982; Tuazon et al., 1978, 1980, 1981a).  These studies
                                   4-63

-------
employ Fourier-transform  infrared  (FTIR)  spectrometers  interfaced to multiple
reflection cells  operating  at total optical paths of  up  to 1 km.  At such
pathlengths, a detection limit of a few ppb is achieved for formaldehyde.  The
advantages of the long-path! ength FTIR method for ambient air aldehyde measure-
ments (i.e., good specificity and direct jhn situ analysis) are offset by the
relatively high cost and lack of portability of such systems.
4.3.1.2.4  High-performance 1 iquid chromatography (HPLC) 2,4-dim'trophenylhydra-
zlne (DNPH) method.  This method  takes advantage of the reaction of carbonyl
compounds with 2,4-dinitrophenylhydrazine to form a 2,4-dinitrophenylhydrazone:
          RR'C=0 + NH2NHCgH3(N02)2 — »•  H20 + RR'C=NNHCgH3(N02)2 (4-17)

Since DNPH  is  a weak nucleophile,  the  reaction  is  carried  out  in the  presence
of acid in order to  increase protonation of the carbonyl.
     The  HPLC-DNPH  method  is  the preferred way of  measuring  aldehydes in
ambient air.   Atmospheric sampling is usually conducted with micro-irnpingers
containing an  organic solvent and aqueous, acidified  DNPH reagent  (Papa and
Turner, 1972;  Katz,  1976;  Smith  and  Drummond, 1979;  Fung and Grosjean,  1981).
After sampling is  completed,  the hydrazone derivatives are extracted and the
extract is  washed with deionized water to  remove the remaining acid and unre-
acted DNPH  reagent.   The organic layer is then evaporated to dryness, subse-
quently dissolved in a small volume of solvent, and  analyzed by reversed-phase
liquid  chromatographic techniques employing  an ultraviolet (UV) detection
system.  Analysis by a flame ionization detection  (FID) system has proved less
successful than UV  because the  derivatives are not  really amenable to GC-FID
analysis.
     An improved procedure has been reported that is  much simpler than the
above aqueous  impinger method (Lipari and Swarin,  1982; Kuntz et a!,,  1980;
Tanner  and  Meng, 1984).   This  scheme utilizes a midget impinger containing an
acetonitrile solution  of DNPH and an acid catalyst.  After sampling, an aliquot
of  the  original  collection solution  is directly  injected into the liquid
chromatograph.  This approach eliminates  the extraction  step and several
sample-handling procedures associated with  the DNPH-aqueous  solution; and
provides  much  better recovery efficiencies.  Lipari and Swarin (1982) have
reported  detection  limits of 20, 10, 5,  and  4 ppb  for formaldehyde, acetalde-
hyde,  acrolein,  and  benzaldehyde,  respectively,   in 20-liter  air  samples.
These  researchers  have also  developed a newer technique  employing a  solid
                                   4-64

-------
adsorbent cartridge containing  Florisil  coated with 2,4-dinitrophenyldrazine
(Lipari and Swarin, 1985).   Collection efficiencies greater than 90 percent
were obtained at  low  ppb levels,  A  detection  limit of 1 ppb of HCHO for a
100-liter air sample was achieved.
4.3.1.2.5   Calibrationof aldehyde measurements.   Since  they  are  reactive
compounds, it is  extremely difficult to make  stable calibration mixtures of
aldehydes in pressurized gas cylinders.  Although gas-phase aldehyde standards
are available commercially, the vendors do not guarantee any level  of accuracy.
     Formaldehyde standards  are generally  prepared by  one  of several methods.
The first method utilizes dilute commercial formalin (37 percent formaldehyde,
w/w).  Calibration  is  accomplished by the  direct  spiking  into  sampling  impin-
gers of the diluted mixture or by evaporation into known test volumes, followed
by impinger collection.   Formaldehyde can also be prepared by heating  known
amounts of paraformaldehyde,  passing the effluent gases  through a  methanpl-
liquid nitrogen slush  trap to remove  impurities,  and collecting the remaining
formaldehyde.   Paraformaldehyde  permeation tubes  have also been used (Tanner
and Meng, 1984).
     For  the  higher-molecular-weight  aldehydes, liquid solutions can be evap-
orated or pure aldehyde vapor can  be generated  in  dynamic gas-flow systems
(permeation tubes,  diffusion tubes, syringe delivery  systems,  etc.).   These
test atmospheres  are  then passed through the appropriate aldehyde collection
system and  analyzed.   A comparison of these data, with the direct spiking of
liquid aldehydes  into  the particular  collection system, provides a  measure of
the overall collection efficiency.
4.3.1.2.6  Comparison of measurement  methods.  Several side-by-side comparisons
of the chromotropic acid method  (CA)  with  other  methods  have  been  reported.
Grosjean  and  Kok  (1981) compared the CA  method  (Johnson  et al., 1981) with
HPLC-DNPH  (Fung  and Grosjean,  1981)  and  FTIR spectroscopy (Tuazon et  al.,
1978).  They found  fairly  close agreement between the  CA and HPLC-DNPH methods,
but noted consistently higher results with FTIR.  Corse (1981)  sampled ambient
air with  CA (U.S. Dept. of Health, Education, and Welfare, 1965), MB'TH  (U.S.
Dept.  of  Health,  Education, and Welfare, 1965), and HPLC-DNPH  methods  (Kuntz
et al.,  1980).  An  examination of tabulated data  from  the Corse study  shows  a
consistent  and considerable difference between CA and  HPLC measurements.  For
25  CA measurements, formaldehyde  averaged  8.8  ppb;  while HPLC measurements
from  the  same sampling train averaged 14.2 ppb.   Overall,  formaldehyde, levels
were  approximately  60 percent  higher with HPLC than  with CA  measurements.
                                    4-65

-------
Because blanks were not utilized for the HPLC analyses, however, the HPLC data
are subject to uncertainty, since blank corrections can affect results substan-
tially (Altshuller, 1983a).   During laboratory studies, Kuntz  et al.  (1980)
reported reasonable agreement (±7 percent}  among  the  HPLC-DNPH, CA, FTIR, and
GC-FID methods when  low ppb  levels of formaldehyde, acetal dehyde , propional-
dehyde, hexanal , and benzal dehyde were generated.
4.3.1.3  Other Oxygenated  Organic Species.   As mentioned earlier, unsubsti-
tuted hydrocarbons constitute the  major fraction of vapor-phase organic com-
pounds occurring in  ambient air.   Aldehydes as a class of volatile organics
appear second  in abundance.   With the exception of formic acid  (Hanst  et al.,
1982; Tuazon et  al.,  1981a, 1980, 1978), other oxygenated species are seldom
reported.  The lack of data for oxygenated hydrocarbons is somewhat surprising
since significant quantities  of these species are  emitted directly into the
atmosphere by  solvent-related industries (methanol, ethanol, acetone, etc.;
see Chapter 3).  Along with manmade emissions, natural sources of oxygenated
hydrocarbons also  contribute  to  this  total.  In addition to these direct
emissions, it  is also expected that photochemical  reactions  of hydrocarbons
with oxides of nitrogen, ozone, and hydroxyl radicals will produce significant
quantities of oxygenated products.
     Difficulties in  sample collection  and  analysis may account for this lack
of data.  The  adsorptive nature of  the  surfaces that  contact  these oxygenated
species often  complicates  the process of compound quantisation.  The approach
used for analysis of oxygenated and other polar organic compounds is to decrease
adsorption by  deactivating the interior surfaces  of  analytical hardware.  A
novel method has been reported, however, in which the reactive compounds of
interest are  modified rather  than  the  surfaces  with which these compounds
interact (Osman  et  al.,  1979; Westberg et al., 1980).  In these studies, the
laboratory derivatization  of  vapor-phase alcohols  and acids  (silylation) was
investigated to  evaluate the  potential  of  such a procedure  for stabilizing
these  polar  compounds prior  to analysis.   Results indicate  that silylation
procedures greatly reduce  adsorption of alcohols  and  acids and  that, qualita-
tively,  the  silylated derivatives  can  be  detected via the  GC-FID system.

4.3.2  Nitrogen Oxides
     In highly polluted  urban air,  the  two  most  abundant oxides of nitrogen
compounds are  nitric  oxide (NO) and nitrogen dioxide  (NO).   Both compounds,
together designated as NO  , participate in the photooxidation reactions in the
                         /s,
                                   4-66

-------
atmosphere that  lead to  the  formation of ozone  (Chapter  3).   Other minor
reactive oxides of nitrogen in ambient air include peroxyacyl nitrates, nitrogen
trioxide, dinitrogen pentoxide, and peroxynitric acid.  In rural areas, and in
urban areas  late in the  day  and at night, these  species  may  constitute a
significant fraction of the total oxides of nitrogen in ambient air (Singh and
Hanst, 1981;  Kill us and Whitten, 1985).
     Analytical methods for NOp  and NO are briefly described in this  section.
Older methods, including the former Federal Reference Method (Jacobs-Hochheiser
method), are described in the criteria document on nitrogen oxides prepared by
the U.S. Environmental Protection Agency (1982).
4.3.2.1  Measurement Methods forNQg and NO.   In  1976,  the continuous chemi-
luminescence  method  was  promulgated  as the  new  Federal  Reference Method
(F.R., 1976c).  This method measures atmospheric concentrations of N02 indirectly
by  first reducing  or  thermally decomposing the gas quantitatively  to  NO  (with
a converter),  reacting  NO with 0.,, and measuring  the  light intensity  from the
chemiluminescent reaction (Fontijn et a!., 1970).   Two types of converters have
been  employed for converting NOg to NO.  In carbon or ferrous sulfate converters,
NO/,   is  chemically reduced  to NO with  the concurrent oxidation of converter
material.  Thermal converters, such as stainless or molybdenum  steel  converters,
spontaneously  decompose N02 to  NO at  elevated temperatures.   The type  and
severity of  interfering species  will  depend  upon  the converter used  and the
temperature of operation.   The NO in the air  stream is measured separately and
subtracted from  the  previous  NO  (NO  plus NQ«) measurement to  yield  the NQ2
concentration.   Typical  commercial  chemiluminescence instruments  can  detect
levels as low as 2.5 ng/m3 (0.002 ppm) (Katz,  1976).
      While all oxides of  nitrogen and organic  nitrogen compounds are thermody-
namically unstable with respect to the formation of NO, the  rate of the conver-
  \
sion  to NO is infinitesmally slow under normal  conditions.   In  the presence of
reducing agents  or at  elevated temperatures,  however,  the conversion may
become  quite  rapid.   A number of studies  have shown for  example, that some
species found in ambient  air can undergo reduction by converters,  resulting in
positive interferences.   Winer et al.  (1974), for example, found that peroxy-
acetyl  nitrate (PAN) and  various nitrogen compounds are reduced by the converter
to  NO and that nitroethane and nitric  acid are  partially reduced in the system
when  a  carbon (reducing) converter or a molybdenum (thermal decomposition)
converter is  used.   Joshi and Bufalini (1978) reported  positive interferences
from  halocarbons  when a heated carbon  converter was used;  they  also  suggested
                                   4-67

-------
that stainless steel converters may be subject to interferences from chlorinated
hydrocarbons.  Other evidence  suggests that  ammonia  (NH3) may  be converted to
NO  in  high-temperature thermal  converters  (U.S.  Environmental  Protection
Agency, 1982).   Positive  interferences resulting  from  the presence of PAN and
HNO« on afternoons of high oxidant concentrations can exceed 30 percent of the
NOg concentrations  (Spicer et  a!., 1979).  Grosjean  (1982)  also reported that
positive interferences  from  nitric acid  and  PAN during NOg  analysis by chemi-
luminescence  can cause a  50 to 60 percent  NO,  overestimation during smog
conditions in Los Angeles.  In less severe smog, the overestimation should not
be this high.
     Other acceptable  methods  for measuring  ambient  NQ2 levels, including two
methods designated  as  equivalent methods, are  the Lyshkow-modified Griess-
Saltzman method,  the  instrumental colorimetric  Griess-Saltzman method,  the
triethanolamine method, the  sodium  arsenite method, and the TGS-ANSA method
[TGS-ANSA = triethanolamine, guaiacol (crmethoxyphenol), sodium metabisulfite,
and 8-anilino-l-naphthalene  sulfonic  acid].   The sodium arsenite method and
the TGS-ANSA method were designated as equivalent methods in 1977.   While some
of  these  methods measure the  species  of interest directly, others require
oxidation, reduction,  or  thermal  decomposition of the  sample, or separation
from interferences,  before measurement.  These methods have been described in
the 1982 criteria document on nitrogen oxides (U.S. Environmental  Protection
Agency, 1982).
     In addition  to  the wet chemical methods  for measuring N02, other tech-
niques have been  investigated.  Maeda et al. (1980) have reported a new chemi-
luminescence  method  based on the reaction of  NO, with luminol (5-amirio-2,3~
dihydro-l,4-phthalazine dione), with  a detection limit of about 50 parts per
trillion (ppt) and  linearity over a range  of  0.5 ppb  to 100 ppm.   Work is
under way  by Maeda and coworkers to  remove  the  interferences  of 0^ and SOp.
Wendel et al. (1983) have reported the development of a lumino!-based instrument
for the continuous measurement of N02 in ambient air.   In the early instrument
developed by  Maeda et al. (1980), response time was slow and the luminol  solution
pool in the detector cell showed chemiluminescence long after NO^ was removed.
The Wendel et al. (1983) instrument has rapid response time and luminol contacts
the air sample  via a  filter paper strip rather than a pool of liquid.   The
method has a  detection  limit of 30 parts per trillion and has been freed of 03
interference  through modifications  to the inlet  system and the  addition of
                                   4-68

-------
Na/jSQo to  the luminol solution.  Other workers  have endeavored to improve
chemiluminescence analyzers through  physical- modifications (e.g., Ridley and
Hewlett,  1974; Schiff et al.,  1979).
     Lipari (1984) has  developed  a solid-sorbent method for measuring N0» in
ambient air.  The NOj, is quantitatively collected by means of a commeridally
available  air  sampling  cartridge  containing Florisil  (magnesium  silicate)
coated with diphenylamine  (DPA).   The N02 reacts with DPA to form 2-nitro-4-
nitro-, and  N-nitroso-DPA  derivatives.   The products are  eluted from the
cartridge with methanol, acidified with a 1  N HC1 catalyst to ensure  recovery
of the DPA derivatives,  and analyzed by  HPLC-UV.  No  interferences from  NO, '
0,5 HNO,,  S02  and  humidity were found, but  PAN  was  found to produce a 50%
positive  interference.   Storage stability  tests indicated  that  cartridge
blanks are  stable  for at least 3 months and that samples can be transported,
stored, and analyzed for at least a period of 4 weeks after collection without
significant sample loss  or degradation.   The reported limit of detection is
0.1 ppb NO,, for a 2000-liter air sample, which corresponds to an 8-hr sampling
period at a flow rate of 4 liters/min.
     Molecular correlation  spectrometry,  in which an  absorption  band of a
sample is  compared with a  corresponding band stored in the spectrometer, has
been  applied  in analysis of  N02  (Williams  and Kolitz, 1968).  Instruments
processing the second derivative of  sample transmissivity  have also been  used
(Hagar and Anderson, 1970), as have infrared lasers and infrared spectrometers
(Hanst, 1970;  Hinkley and  Kelley,  1971; and  Kreuzer  and Patel, 1971).  Tucker
et al. (1973,  1975)  reported  on instruments based on the principle of laser-
induced fluorescence  at optical  frequencies.   Fincher  et al.  (1977) described
detection  of  1 ppb  N0?  with a technique  based  on fluorescence by a  pulsed
xenon flashlamp.  Long-pathlength differential optical absorption spectroscopy
has  also  been employed  to  monitor  N0? in  the troposphere  (Platt et al., 1980,
1984).
     The preferred approach for measuring NO is also the continuous chemilumi-
nescence  method.   Other methods  for measuring NO directly  include ferrous
sulfate  absorption  and spectrophotometric measurement of  the resulting  ion
(Norwitz,  1966),  ultraviolet  spectroscopy  (Sweeny et al.,  1964), infrared
spectroscopy  (Lord  et al., 1975),  and ultraviolet fluorescence  (Okabe and
Schwartz,  1975).  Mass  spectrometry and gas  chromatography may also be employed.
                                   4-69

-------
4.3.2.2  Sampling Requirements.  When  sampling for NO , long residence times
in sampling lines should be avoided.  In the ambient air, the rate of photoly-
sis of N09  (forming NO and 0, and thus 0.,)  is  almost equal to the rate of the
         £»                              o
reaction of the  NO  and 03 to form  N0«,  In sampling lines, photolysis stops
but NO continues  to  react with Q3,  producing NOp.  The magnitude of the dark
reaction of NO with 03 depends, of  course,  on  the concentrations of NO and 03
in the sample  being  analyzed,  as well as on the residence time of the sample
in the line.  This dark reaction has practical  consequences only under certain
conditions.   In moderately polluted urban  areas, steady-state concentrations
of NO are almost certainly too low at the period of maximal 03 to cause signi-
ficant errors in obtaining NOX or 03 measurements. Conversely, when 03 is at a
minimum, as in the  early morning or possibly  even in the  late afternoon, NO
(and NOp) may be  at maximal  levels, and no  significant errors would be intro-
duced.   If  the  concentrations of NO and 03 are both low  as  in  some  rural
areas, or during  those brief periods in polluted areas when NO and 0, diurnal
patterns cross, then measurement errors could  be introduced.  Such errors are
not likely to be significant, however.
     Techniques for  limiting sampling errors  sampling  to given  levels  of
tolerance are reviewed by Butcher and  Ruff  (1971).  In general, only glass or
Teflon materials should be used in sampling trains.  Among absorbents, granules
impregnated with  triethanolamine are reported  to be the  best, converting only
2  to  4  percent of the  incoming NOp to NO  (Intersociety Committee,  1977b;
Huygen,  1970).  The most frequently used oxidizer is chromic  oxide on a fire-
brick granule support  (Intersociety Committee, 1977b; Levaggi et al., 1974).
4.3.2.3  Calibration.   Calibration procedures for NO  measurements methods are
critical for obtaining accurate analyses.   Measurement methods for NO and NOp
are calibrated (1) by  sampling a gas  stream of  known concentration prepared
from standard reference materials (SRMs); and (2) in the case of chemilumines-
cence analyzers,  by determining converter efficiency.   The SRM  for  NO is a
cylinder of compressed NO  in N2 (50 and 100 ppm).  The  initial accuracy and
the stability with time of this mixture were found to be quite good in a study
conducted at the  National Bureau of Standards  (Hughes, 1975).  An accuracy of
1 percent of the labeled concentrations was obtained as determined from pressure
or gravimetric measurements.   Concentration changes of  less  than  1 percent
were observed over a  7-month period.   The  SRM for NOp  is  the N0« permeation
tube (O'Keeffe and Ortman,  1966;  Scaringelli et al., 1970).   These'tubes are
                                   4-70

-------
calibrated by determining the permeation rate by means of weight-Toss measure-
ments at a constant temperature; or, in some cases, by means of micromanometric
measurements.  Both the NO and NOg SRMs are commercially available from sources
traceable to SRMs maintained at the National Bureau of Standards.
     Procedures for preparing  calibrated gas streams call  for an  accurately
measured dilution  of  the output of the  NO  and  NOp SRMs.   Likewise,  the flow
from the NO-in^ cylinder must be accurately measured, as well as the diluent
air flow.  The flow of diluent air over the NO, permeation tube, maintained at
constant temperature, must also be accurately measured.  An alternative proce-
dure commonly employed for the calibration of NO,  N09, and NO  measurements is
                                                    f—        s\
the use of gas-phase titration (Rehme et a!., 1974).  A constant flow of 0., is
added to a diluted gas stream of NO at a higher concentration and the reaction
mixture  is monitored  with  an NO monitor,  e.g.,  chemiluminescence.   Because  of
the straightforward bimolecular reaction,

          NO + 03 •* N02 + 02                 "                         (4-18)

the NO  consumed  is equivalent  to  the 0-  added and  the  N00  produced.  Thus,  if
                                       
-------
large  existing  data base  that employed measurements for "total  oxidants,"
non-specific  iodometric  techniques are  discussed and  compared  to current
specific (U measurements.
4.4.1.1   Quality Assurance and Sampling.   A  quality assurance  program is
employed by the U.S. Environmental Protection Agency for assessing the accuracy
and precision of monitoring data and for maintaining and improving the quality
of ambient  air  data.  Procedures and operational details have been prescribed
in each of the following areas: selection of analytical methods and instrumen-
tation  (i.e.,  reference and equivalent  methods);  method specifications for
gaseous standards;  methods for  primary and secondary  (transfer  standards)
calibration; instrumental zero and span check requirements, including frequency
of checks,  multiple-point  calibration  procedures,  and preventive  and  remedial
maintenance  requirements;  procedures  for air  pollution episode monitoring;
methods for recording and  validating  data;  and information on  documenting
quality control (U.S. Environmental Protection Agency, 1977b).  Design criteria
for DO  monitoring stations, to help ensure the  quality of aerometric data,
have been established  (U.S. Environmental  Protection Agency, 1977a; National
Research Council, 1977).
4.4.1.2  Measurement Methods for Total Oxidantsand Ozone.  Techniques for the
continuous  monitoring  of total oxidants and 0^ in ambient air are summarized
in Table 4-13.  The earliest methods used for routinely monitoring oxidants in
the atmosphere were based on iodometry.  When atmospheric oxidants are absorbed
in potassium iodide (KI) reagent, the  iodine produced,

                         03 +  31" + H20 -»• I3~ + 02 + 20H"             (4-18)

is measured by ultraviolet absorption in colorimetric instruments and by
amperometric means  in  electrochemical  instruments.   The term "total oxidants"
is of  historical  significance  only and  should  not be construed  to mean that
such measurements yield  a sum of the  concentrations of the oxidants in the
atmosphere.   The  various  oxidants in the atmosphere  react to yield iodine  at
different rates and with different stoichiometries.  Only ozone reacts immedi-
ately to  give  a quantitative yield of iodine.  As discussed below, the total
oxidants measurement correlates  fairly well  with the specific measurement of
ozone,  except  during periods  when significant  nitrogen dioxide (NO^) and
sulfur  dioxide  (S02)  interferences are present.  The major problem with the
total  oxidants  measurement was  the effect -of  these interferences.  Total
                                   4-72

-------
                                               TABLE 4-13.   SUMMARY  OF  OZONE  MONITORING TECHNIQUES
Principle
Continuous
colorimetric
Continuous
electrochemical
-p> Chemi luminescence
i
Chemi luminescence
Ultraviolet
photometry
Reagent
10(20)% KI
buffered at
pH = 6.8
2% KI
buffered at
pH = 6.8
Ethyl ene,
gas-phase
Rhodamine-B
None
Response
Total
oxidants
Total
oxidants
03-specific
03-specific
03-specific
Minimum
detection limit
0.010 ppm
0.010 ppm
0.005 ppm
0.001 ppm
0.005 ppm
Response ,
time, 90% FSa
3 to 5 minutes
1 minute
< 30 seconds
< 1 minute
30 seconds
Major
interferences
N02(+20%, 10%l
S02(-100%)
N02(+6%)
S02(-100%)
None"
None
Species that
absorb at 254
References
(I) Littman and Benoliel (1953)
Tokiwa et al. (1972)
Brewer and Mil ford (1960)
Mast and Saunders (1962)
Tokiwa et al. (1972)
Nederbragt et al. (1965)
Stevens and Hodgeson (1973)
Regener (1960, 1964)
Hodgeson et al. (1970)
Bowman and Horak (1972)
nm
aFS = full  response.
 A signal  enhancement of 3 to 12% has  been reported for measurement of 03  in humid versus dry air (California Air Resources Board,  1976).
cNo significant interferences have been reported in routine ambient air monitoring.   If abnormally high concentrations of species that
 absorb at 254 nm (e.g., aromatic hydrocarbons  and mercury vapor)  are present,  some positive response may be expected.  If high aerosol
 concentrations are sampled,  inlet filters must be used to avoid a positive  response.

-------
oxidants instruments  have  now  been  replaced by  specific ozone monitors in all
aerometric  networks  and in  most research laboratories.   Biases  among and
between total oxidants  and ozone methods are still  important,  however,  for
evaluating existing data on health and welfare effects levels where concentra-
tions were measured by total oxidants methods.
     The reference  method  promulgated by EPA for  compliance monitoring  for
ozone is the chemiluminescence technique based on the gas-phase ozone-ethylene
reaction (F.R.,  1971).  The technique is specific for ozone, the response is a
linear function of  concentration,  detection  limits of 0.001 to 0.005 ppm are
readily obtained,  and response  times are 30 seconds  or  less.   Prescribed
methods of testing and prescribed performance specifications that a commercial
analyzer must meet in order to  be  designated as a  reference method or  an
equivalent method have been published by EPA (F.R., 1975b).  An analyzer may be
designated as a  reference  method if it is based on the same principle as the
reference method  and  meets performance  specifications.  An acceptable equiva-,
lent method  must  meet the  prescribed  performance specifications and  also show
a consistent relation with the reference method.
     The designated  equivalent  methods  are  based on  either the gas-solid
chemiluminescence procedure  or  the ultraviolet absorption procedure (Table
4-13).  The first designated equivalent method was based on ultraviolet absorp-
tion by ozone of the mercury  254 nm  emission line.   The  measurement is in
principle an  absolute one, in that the  ozone concentration can be computed
from  the measured transmission  signal  since  the absorption coefficient  and
pathlength are accurately  known.  In the gas-solid chemiluminescence analyzer,
the reaction  between  ozone and  Rhodamine-B adsorbed on activated silica pro-
duces chemiluminescence, the intensity  of which is  directly proportional to
ozone concentration.
4.4.1.3  Calibration  Methods.   All  the  analyzers  discussed above  must  be
calibrated periodically with ozonized air streams, in  which the ozone concen-
tration has  been  determined by  some  absolute technique.   This  includes  the
ultraviolet  (UV)  absorption  analyzer, which,  when  used for continuous ambient
monitoring, may experience ozone losses in the  inlet system because of contami-
nation.
     A primary ozone  calibration system consists of  a  clean air source, ozone
generator, sampling manifold,  and means for measuring absolute ozone concentra-
tion.  The  ozone  generator most often used is a photolytic source employing  a
mercury lamp  that irradiates a quartz tube through which  clean  air  flows at  a
                                   4-74

-------
controlled rate (Hodgeson et al., 1972).   Once the output of" the generator has
been calibrated by a primary reference method, it may be used to calibrate 0~
transfer  standards,  which are  portable  generators, instruments,  or other
devices used  to calibrate  analyzers  in the  field.   Reference  calibration
procedures that have been used for total  oxidants and ozone-specific analyzers
in the United States are summarized in Table 4-14.
     The original reference calibration  procedure promulgated by EPA was the
1 percent neutral buffered potassium  iodide (NBKI)  method  (F.R., 1971).  This
technique was  employed  in most of the United  States,  with the exception of
California.    The  California Air Resources Board  (CARB)  (1976)  and the  Los
Angeles Air Pollution Control  District (LAAPCD) employed different versions of
iodometric techniques.   Procedural details of the calibration methods  are
summarized in Table 4-14.  A number of studies conducted between 1974 and 1978
revealed  several  deficiencies with KI methods, the  most  notable of which were
poor precision  or inter!aboratory comparability and a positive bias  of  NBKI
measurements  relative  to simultaneous absolute  UV absorption measurements.
The positive  bias observed is peculiar to the use of phosphate buffer in the
NBKI techniques.  The  bias was not observed  in  the unbuffered  LAAPCD method
(which nevertheless suffered from poor precision), nor in the 1 percent  EPA KI
method without phosphate buffer (Hodgeson et al., 1977), nor in a KI procedure
that used boric acid as buffer (Flamm, 1977).  A summary of results of these
prior studies  was presented in the previous criteria document  (U.S. Environ-
mental Protection Agency,  1978a) and in a  review  by Burton et al.  (1976).
Correction factors  for  converting' NBKI calibration  data to a UV photometry
basis are given in Table 4-14 and discussed in Section 4.2.4.2.1,
     Subsequently, EPA evaluated four alternative reference calibration  proce-
dures and selected UV photometry -on the basis of superior accuracy and precision
and simplicity of use  (Rehme  et al.,  1981).   In  1979 regulations were amended
to specify UV photometry as the reference calibration procedure (F.R., 1979e).
Laboratory photometers  used as  reference systems  for absolute 03 measurements
have been described by DeMore and Patapoff (1976)  and Bass et al.  (1977).
     These  laboratory  photometers contain  long path cells (1  to  5 m)  and
employ  sophisticated digital  techniques  for  making effective  double beam
measurements  of small  absorbancies at low  ozone concentrations.   A primary
standard  UV  photometer is one that meets the requirements  and  specifications
given  in  the revised ozone calibration procedures  (F.R., 1979e).   Since  these
are currently  available  in only a  few laboratories,  EPA has allowed the  use of
                                   4-75

-------
                                                  TABLE 4-14.  OZONE CALIBRATION TECHNIQUES
Method
1% NBKI
2% NBKIC
1% Unbuffered
KI
UV photometry
Gas-phase
titration (GPT)
1% BAKI
Reagent
1¥ KT
X»fc M ,
phosphate buffer
pH = 6.8
2% KI
phosphate buffer
pH = 6.8
1% KI
pH = 7
None
Nitric oxide
standard reference
gas
•w KT
-Lfc M,
boric acid buffer
pH = 5
Primary standard3
Reagent grade
arsenious oxide
Reagent grade
potassium biiodate

03 absorptivity at
Hg 254 nm emission
line
Nitric oxide SRM
(50 ppm in N2)
from NBS .
Standard KI039
solutions
Organization
and dates
EPA
1971-1976
CARB
until 1975
LAAPCD
until 1975
All
1979-present
EPA, States
1973-present
EPA
1975-1979
Bias,
Purpose Ws] -/[03]
Primary reference 1.12 ± 0.05
procedure -
Primary reference 1.20 ± 0.01
procedure
Primary reference 0.96
procedure
Primary reference
procedure
Alternative reference 1.030 ± 0.015f
procedure (1973-1979)
Transfer standard (1979-present)
Alternative reference 1.00 ± 0.05
procedure • .
 In the case of the iodometric methods, the primary standard is the reagent used to prepare or standardize  iodine  solutions.
 The uncertainty limits represent the range of values obtained in several independent studies.
cPre-humidified air used for the ozone source.                               .
 Only one study available (DeMore et al., 1976).                                                                       ,    -
eUV photometry used as reference method by CARB since 1975.  This technique used as an interim, alternative  reference procedure  by
 EPA from 1976 to 1979.
 This is the value reported in the latest definitive study (Fried and Hodgeson, 1982).  Previous-studies reported  biases fpng,ing::Trom
 0 to 10 percent (Burton et al., 1976; Paur and McElroy, 1979).                          ,       .';   ,:      :    '"  ;,,  :• f:   ;t;   ^   -k
      procedure also recommended a standard I3  solution absorptivity to be used instead of the preparation  of standard iodine solutions.

-------
transfer standards, which are devices or methods that can be calibrated against
a primary standard  and  transferred to another location for calibration of 0,
analyzers.   Examples of transfer standards that  have been  used are commercial
DO photometers, calibrated generators, and gas-phase titration (GPT) apparatus.
Guidelines on  transfer  standards  have been published by EPA (McElroy, 1979).
4.4.1.4;  Relationships of Total Oxidants and OzoneMeasurements.   The temporal
and quantitative relationship  between simultaneous total oxidants and ozone
measurements has been examined in  this chapter because of the existence of a
data base obtained  by total oxidants  measurements.  Such a comparison is com-
plicated by the  relative  scarcity  of simultaneous data, the presence of both
positive (NOg) and negative ($02) interferences in total oxidants measurements
of ambient air,  and the change in the  basis  of  calibration.  In particular,
the presence  of NOp  and  SOp interferences prevent the  establishment of a
definite quantitative relationship between ozone and oxidants data.   Neverthe-
less,   some  interesting  conclusions can be drawn and are summarized below.
     Studies concluded in the early to mid-1970s were reviewed.,in the previous
criteria document (U.S.  Environmental Protection Agency, 1978a).  Averaged data
showed  fairly  good qualitative  and  quantitative agreement between diurnal
variations of  total oxidants  and  ozone.   For example, uncorrected monthly
averaged data  from Los Angeles  and  St. Louis showed  distinct  morning and
evening peaks  resulting  from NOp interference (Stevens et al.,  1972a,b). The
most recent comparison in the literature involved simultaneous ozone and total
oxidant measurements  in the  Los Angeles basin by the California Air Resources
Board  (1978)  in  1974,  1976,  and 1978.  The  maximum hourly data pairs were
correlated (Chock  et al.,  1982) and yielded  the  following  regression  equation
for 1978, in which a large number (927) of data pairs were available:

                         Oxidant (ppm) = 0.870 03 + 0.005
                         (Correlation coefficient = 0.92)             (4-19)

The oxidant data were uncorrected for NOg and SO, interferences, and on individ-
ual days maximum oxidant  averages  were both  higher than and lower than ozone
averages.
     In summary, specific ozone measurements  agree fairly well with total oxi-
dants  corrected  for N0? and S02 interferences,  and in such corrected total
oxidants measurements ozone is the dominant contributor.  Indeed, it is diffi-
cult  to discern the presence  of other  oxidants  in corrected .total oxidant
                                   4-77

-------
 data.   Without corrections there  can  be major temporal  discrepancies  between
 ozone  and oxidants data,  which are primarily a result of oxidizing jmd reducing
 interferences with KI measurements.   As a result of these interferences, on
 any given day the total  oxidant values may be higher than or  lower than simul-
 taneous ozone data.  The  measurement  of ozone is a  more reliable Indicator
 than total  oxidant measurements of oxidant air quality.
 4.4.1.5    Methods for  Sampling and Analysis of Peroxyacetyl  Nitrate andIts
 Homologues.   Only two analytical techniques have been used to obtain signifi-
•cant data on ambient peroxyacetyl  nitrate (PAN) concentrations.   These are  gas
 chromatography with electron capture detection (GC-ECD) and long-path Fourier
 transform infrared (FTIR) spectrometry.   Atmospheric data on  PAN  concentrations
 have been obtained predominantly by GC-ECD because of its relative simplicity
 and superior sensitivity.   These techniques have been described in this chapter
 along  with attendant methods  of sampling, PAN generation, absolute  analysis,
 and calibration.
     By far the most widely used technique for the quantitative determination
 of ppb concentrations of PAN  and its homologues  is  GC-ECD (Darley  et al.,
 1963;  Stephens, 1969).   With  carbowax or SE30 as a stationary phase, PAN,
 peroxypropionyl  nitrate (PPN),  peroxybenzoyl  nitrate (PBzN),  and other homo-
 logues (e.g., peroxybutyryl nitrate) are readily separated from components  such
 as air, water, and other atmospheric compounds, as  well  as  ethyl nitrate,
 methyl nitrate, and other contaminants that are present in synthetic mixtures.
 Electron-capture  detection provides sensitivities  in  the ppb and  sub-ppb
 ranges.   Typically, manual air samples are collected in 50- to 200-ml  ungreased
 glass  syringes and purged through the gas-sampling valve.   Continuous  analyses
 are performed by pumping  ambient  air through a gas  sampling  loop of an auto-
 matic  valve, which periodically injects the sample  onto the  column.   Samples
 collected from the atmosphere  should  be analyzed as soon as  possible  because
 PAN and its  homologues  undergo thermal  decomposition in the  gas  phase and  at
 the surface of containers.  The recent work of Singh and Salas (1983a,b) on
 the measurement of PAN  in the  free (unpolluted) troposphere (see  Chapter 5) is
 illustrative of  current  capabilities for measuring low concentrations.  A
 minimum detection limit of 0.010 ppb was obtained.
     The literature contains  conflicting  reports  on the effects of variable
 relative humidity  on PAN measurements by GC-ECD.   If  a moisture, effect is
                                    4-78

-------
suspected in a PAN analysis, the bulk of this evidence suggests that humidiff-
cation of PAN  calibration  samples (to a range approximating the humidity of
the samples being analyzed) would be advisable.
     Conventional long-path  infrared spectroscopy and Fourier-transform in-
frared spectroscopy  (FTIR)  have been used to detect  and measure atmospheric
PAN.  Sensitivity is enhanced by the use of FTIR.  The most frequently used IR
bands have been  assigned  and the absorptivities  reported  in  the literature
(Stephens, 1964; Bruckmann and Willner, 1983; Holdren and Spicer, 1984) permit
the quantitative analysis of PAN without calibration standards.  The absorptiv-
ity of the 990 cm   band of PBzN, a higher homologue of PAN, has been reported
by Stephens (1969).   Tuazon et al. (1978) describes an FTIR system operable at
v
pathlengths up to 2 km for ambient measurements of PAN and other trace constit-
uents.  This system  employed an eight-mirror multiple reflection cell with a
22.5-m base path.   Tuazon  et al. (1981b) reported maximum PAN concentrations
ranging from 6 to 37 ppb over a 5-day smog episode in Claremont, CA.  Hanst et
al. (1982) made measurements with a 1260-m folded optical path system during a
2-day smog episode  in Los Angeles in  1980.  An  upper limit of 1 ppb  of PBzN
was placed, and  the  maximum  PAN concentration observed was 15  ppb.  The Jarge
internal surface area of the White cells may serve to promote the decomposition
or irreversible adsorption of reactive trace species such as PAN.  High volume
sampling rates and inert internal surface materials are used to minimize these
effects.
     Because of the thermal instability of dilute PAN samples and the explosive
nature of liquefied  PAN,  calibration samples are  not commercially  available
and must be prepared.  Earlier methods used to synthesize PAN have been summa-
rized by  Stephens (1969).  The  photolysis  of alkyl  nitrites in oxygen was the
most  commonly  used procedure and may still be used  by some  investigators.  As
described by Stephens et  al. (1965), the liquefied crude mixture obtained at
the  outlet  of the photolysis chamber is purified by preparative-scale GC.
[CAUTION:  Both  the liquid crude mixture  and  the purified PAN samples are
violently explosive  and should  be  handled  behind  explosion  shields using
plastic full-face protection,  gloves, and a leather coat at all times.]  The
pure  PAN  is  usually diluted to  about  1000 ppm in cylinders pressurized with
nitrogen to 100 psig and stored at reduced temperatures, <15°C.
     Gay et al.  (1976) have used the photolysis of chlorine: aldehyde: nitrogen
dioxide mixtures  in  air or oxygen for the preparation of PAN and a number of
its homologues at high yields.  This procedure has been utilized in a portable
                                   4-79

-------
PAN generator that  can be used for  the calibration of GC-ECD instruments in
the field (Grosjean, 1983; Grosjean et a!., 1984).
     Several investigators have recently reported on a condensed-phase synthesis
of PAN by  nitration of peracetic acid in a hydrocarbon solvent.  High yields
are produced of a pure product free of other alkyl nitrates (Hendry ,and, Kenley,
1977; Kravetz et  al.,  1980;  Nielsen et a!., 1982; Holdren and Spicer, 1984).
After the  nitration is  complete,  the hydrocarbon  fraction  containing PAN
concentrations of 2 to 4 mg/ml can be  stored at -20°C for  periods longer than
a year (Holdren and Spicer, 1984).
     The most direct method  for absolute analysis of these PAN samples is by
infrared absorption, using the  IR  absorptivities mentioned earlier.  Cooyen-
                                                                        W
tional IR instruments and 10-cm gas cells can analyze gas standards of concen-
trations >35 ppm.   Liquid microcells can be used for the  analysis  of  PAN in
isooctane solutions.  The alkaline hydrolysis of PAN to acetate ion and nitrite
ion in quantitative yield (Nicksic et  al.,  1967)  provides  a means independent
of infrared for the quantitative  analysis of  PAN.   Following hydrolysis,
nitrite ion may be quantitatively analyzed by the Saltzman colorimetric proce-
dure  (Stephens, 1969).   The  favored procedures now use ion chromatography to
analyze for nitrite (Nielsen et al., 1982) or acetate (Grosjean, 1983; Grosjean
et al., 1984)  ions.   Another calibration procedure has been proposed that is
based on the thermal decomposition of PAN in the presence of excess and measured
NO concentrations (Lonneman et al., 1982).  The acetylperoxy radical, CHgC^Op,
and its decomposition products rapidly oxidize nitric oxide (NO) to NO^ with a
stoichiometry that has been experimentally determined.
4.4.1.6   Methods for Sampling and Analysis of Hydrogen Peroxide.    Hydrogen
peroxide (H^Op) is significant in photochemical smog as a chain terminator;  as
an index of the hydroperoxyl  radical (H02) concentration (Bufalini and Brubaker,
1969; Demerjian et al., 1974); and as a reactant  in the aqueous-phase oxidation
             «. O
of SQy to SO^   and in the acidification of rain  (Penkett et al.,  1979; Dasgupta,
1980a,b; Martin and Damschen, 1981; Overton and Durham, 1982).
     One of the major  problems, however,  in assessing the  role  of atmospheric
HgOp has been a lack of adequate measurement methodology.  Earlier measurements
(Gay and Bufalini, 1972a,b; Bufalini et al., 1972; Kok et al., 1978a,b) reporting
Hp02  concentrations  of 0.01  to 0.18 ppm are now believed to be far too high,
and to be  the result of  artifact H^Op  formation from reactions  of absorbed  Og
(Zika and  Saltzman,  1982; Heikes et  al.,  1982; Heikes, 1984).   Maximum tropo-
spheric H^Oz concentrations  predicted  by  modeling calculations  (Chameides and
                                   4-80

-------
Tan ,: 1981;  Logan  et  al.,  1981) and observed  in recent field studies (Das et
al., 1983) are on the order of 1 ppb.
     Almost all of the  methods used for the  measurement of atmospheric H202
have used aqueous traps for sampling.  Atmospheric 0,, however, which is also
absorbed at concentrations  much higher than H2Q2, reacts in the bulk aqueous
phase and at surfaces to produce H202 and thus is a serious interference (Zika
and Saltzman,  1982;  Heikes et al,, 1982; Heikes, 1984).   The removal of absorbed
On by purging  immediately after sample  collection may remove or substantially
reduce this interference (Zika and Saltzman, 1982; Das et al.,  1983).  Another
problem identified with aqueous sampling is that other atmospheric species (in
particular, SOp)  may  interfere with the generation of H^O/j in aqueous .traps
and also react with collected HpO^ to reduce the apparent concentration measured
(Heikes et al., 1982).
     Of the techniques  that have  been used  for the measurement of aqueous and
gas-phase HJ)^, only  the  chemi luminescence and enzyme-catalyzed methods are
summarized below.  The  other techniques are  now believed to have inadequate
sensitivity and specificity for the atmospheric concentrations actually present,
In addition, the use of a tunable diode infrared laser source should eliminate
the problem associated  with nearby water bands, and this method is currently
under investigation for atmospheric measurements (unpublished work in progress,
Schiff, 1985).
     Hydrogen peroxide  in the atmosphere may be detected at low concentrations
by the chemi luminescence obtained from copper(II)-catalyzed oxidaton of luminol
(5-amino-2,3-dihydro-l,4-phthalazinedtone)  by H2Q2  (Armstrong  and Humphreys,
1965; Kok et al., 1978a,b).  This technique as initially employed suffered the
interferences  from Og and SQ^ discussed above for aqueous traps.   Das et al.
(1982) employed a static version of the method of Kok et al. (1978a) to measure
HpO/j concentrations  in  the 0.01 to 1 ppb  range.   In  addition, samples were
purged with argon immediately after collection to eliminate, reportedly, the
Qg  interference.   Recently,  a modified chemi luminescence  method  has been
reported which  used  hemin,  a  blood component, as a catalyst for the luminol-
based HgQ/? oxl'dat1on  (Yoshizumi et al., 1984).
     The most  promising chemical  approach  employs the catalytic activity of
the enzyme horseradish  peroxidase (HRP) on  the oxidation of organic substrates
by H2Q2-  The production or decay of the fluorescence intensity of the substrate
or  reaction product  is measured as it  is oxidized by HOo* catalyzed by HRP.
                                   4-81

-------
Some of the  more widely used substrates  have  been scopoletin (6-methoxy~7-
hydroxyl,2-benzopyrone) (Andreae, 1955; Perschke and Broda, 1961); 3-(g-hydroxy-
phenyl)propionic  acid  (HPPA) (Zaltsu  and Okhura,  1980); and leuco crystal
violet (LCV) (Mottola et al., 1970).
     The decrease in the fluorescence intensity of scopoletin is measured as a
function of  HgO, concentration.   Detection limits have  been  reported to be
quite low  (10     M),   The chief disadvantage  of  this  approach is that  the
concentration of HpOg  must be within a narrow range to  obtain an accurately
measureable  decrease  in fluorescence.  Oxidation  of  LCV produces intensely
                                                                       5 -1
colored crystal  violet, which has a molar  absorption  coefficient of  10  M
cm   at the  analytical  wavelength, 596 nm.  The detection limit reported was
  ~8
10   M in  5  cm  cells.   Two quite similar hydrogen donor substrates have been
used.  Zaitsu and Okhura  (1980) employed 3-(p_-hydroxyphenyl) propionic  acid
and  more  recently the j>-hydroxyphenyl acetic  acid homologue is being used
(Kunen et al., 1983; Dasgupta and Hwang, 1985).  The measurement of the fluores-
cence intensity  of the  product  dimer  provides  a quite  sensitive means for the
assay of HgOg-
     As with Q-, HgOp calibration standards are not commercially available and
are  usually  prepared at the  time of use.  The  most convenient method  for pre-
paring aqueous  samples  containing  micromolar concentrations of t-LOo is simply
the  serial dilution of commercial  grade  30 percent  HgOg (Fisher Analytical
Reagent).   Techniques for the convenient generation of gas-phase standards are
not  available.   A technique  often used for generating ppm concentrations of
HgOg in air  involves the injection of microliter quantities of 30 percent Hy®?
solution into a metered stream of air that flows  into a Teflon bag.   Aqueous
and  gas-phase samples  are  then  standardized by conventional  iodometric proce-
dures (Allen et  al., 1952; Cohen et al., 1967).

4.4.2    Measurement of Precursors to Ozone and Other Photochemical Oxidants
4.4.2.1  Nontnethane Organic  Compounds.  Numerous  analytical methods have been
employed to  determine  nonmethane  organic compounds  (NMOC) in ambient air.
Measurement  methods  for the  organic species may be grouped according  to  three
major classifications: nonmethane hydrocarbons, aldehydes, and other oxygenated
compounds.
     Nonmethane  hydrocarbons  have  been determined primarily by methods  that
employ a flame ionization detector (FID) as the sensing element.   Early methods
for  the  measurement of total nonmethane hydrocarbons did not provide for
                                   4-82

-------
speeiation of the complex mixture of organics in ambient air.  These methods,
still in use for analysis of total  nonmethane organic compounds,  are essentially
organic carbon analyzers, since the response of the FID detector is essentially
proportional to the  number  of carbon atoms present  in the organic molecule
(Sevcik, 1975). Carbon atoms  bound, however, to oxygen, nitrogen, or halogens
give reduced  relative responses (Dietz, 1967).  The  FID  detector has been
utilized both  as  a stand-alone continuous  detection  system (non-speciation
analyzers have indicated an overall poor performance of the commercial  instru-
ments when  calibration or  ambient mixtures containing  nonmethane organic
compounds (NMOC) concentrations less than 1 ppm C were analyzed (e.g.,  Reckner,
1974; McElroy and Thompson,  1975;  Sexton et a!,, 1982).   The major problems
associated with the  non-speciation analyzers have been summarized  in a recent
technical assistance  document published by the U.S.  Environmental  Protection
Agency (1981).  The document also presents ways to reduce some of the existing
problems.
     Because of the above deficiencies, other approaches to the measurement of
nonmethane  hydrocarbons  are  currently  under development.   The  use  of gas
chromatography  coupled  to an  FID  system circumvents  many of the  problems
associated with continuous  non-speciation  analyzers.   This method, however,
requires sample preconcentration because the organic components are present at
part-per-billion (ppb) levels or lower in ambient air.  The two main preconcen-
tration  techniques  in present use are  cryogenic collection and the use of
solid adsorbents  (McClenny et  a!.,  1984;  Jayanty  et a!., 1982; Westberg.
et al.,  1980;  Ogle  et al., 1982).   The preferred preconcentration method for
obtaining speciated data is cryogenic collection.   Speciation methods involving
cryogenic preconcentration have also been compared with continuous nonspeciation
analyzers  (e.g.,  Richter, 1983).  Results  indicate  poor correlation between
methods  at ambient  concentrations  below 1  part-per-million carbon (ppm C).
     Aldehydes, which are both primary and secondary pollutants in ambient
air, are detected by total  NMOC and NMHC  speciation methods but can not be
quantitatively  determined by  those methods.  Primary  measurement  techniques
for  aldehydes include the  chromotropic acid  (CA)  method  for formaldehyde
(Altshuller and McPherson,  1963;  Johnson et al., 1981), the 3-methyl-2-benz-
othiazolene (MBTH)  technique  for total  aldehydes (e.g., Sawicki  et a1», 1961;
Hauser  and Cummins,  1964),  Fourier-transform  infrared (FTIR) spectroscopy
                                   4-83

-------
(e.g., Hanst et al.} 1982; Tuazon et a!., 1978, 1980, 1981a), and high-perfor-
mance liquid chromatography  employing  2»4-dinitrophenyl~hydrazine derivatiza-
tion (HPLC-DNPH) for aldehyde speciation (e.g., Lipari and Swarin, 1982; Kuntz
et a!., 1980).   The CA and MBTH methods utilize wet chemical procedures and
spectrophotometric  detection.   Interferences  from  other compounds have been
reported for both  techniques.   The FTIR method offers  good  specificity and
direct jn situ analysis of ambient air.  These advantages are offset, however,
by the relatively  high cost and lack  of portability of the  instrumentation.
On the other hand,  the HPLC-DNPH method not only offers good specificity but
can also be easily  transported  to  field sites.  A  few intercomparison studies
of the  above  methods  have been conducted  and considerable  differences in
measured concentrations were  found.   The data base  is still  quite limited at
present,  however, and further intercomparisons are needed.
     Literature  reports  describing the  vapor-phase  organic composition of
ambient air indicate that the major fraction of material consists of unsubsti-
tuted hydrocarbons  and aldehydes.   With the exception  of formic acid,  other
oxygenated species  are seldom reported.   The  lack of oxygenated hydrocarbon
data is somewhat surprising  since  significant  quantities of  these species are
emitted into the atmosphere  by solvent-related industries and since at least
some oxygenated  species  appear to be  emitted  by vegetation.  In addition to
direct emissions,  it  is  also expected that photochemical reactions of hydro-
carbons with oxides of nitrogen, 0,, and hydroxyl  radicals will  produce signi-
ficant quantities of oxygenated species.  Difficulties during sample collection
and analysis may account for the apparent  lack of data.  Attempts have been
made to decrease adsorption by deactivating the reactive surface or by modifying
the compound of interest (Osman et a!., 1979; Westberg et al., 1980).  Additional
research efforts should focus on this area.
4.4.2.2  Nitrogen Oxides.  Aside from the essentially unreactive nitrous oxide
(NgO), only two oxides of nitrogen occur in ambient  air at appreciable concen-
trations:   nitric oxide (NO)  and nitrogen  dioxide (N02).   Both compounds,
together designated as NO , participate in the cyclic reactions in the atmosphere
                         r\
that lead  to the formation of ozone.   Other minor  reactive oxides of nitrogen
in ambient air  include  peroxyacyl  nitrates,  nitrogen  trioxide, dinitrogen
pentoxide, and peroxynitric acid.
     The preferred means (Federal Reference Method)  of measuring NO and NOp is
the chemiluminescence  method (F.R., 1976c).  The measurement principle  is the
gas-phase  chemiluminescent  reaction of  0, and NO (Fontijn  et  al.,  1970).
                                   4-84

-------
While;NO is determined directly in this fashion, N02 is detected indirectly by
first reducing or  thermally decomposing the gas quantitatively  to  NO with a
converter.   The reaction of NO and 0, forms excited N02 molecules that release
light energy  that  is  proportional to the  NO  concentration.  Although the NO
chemiluminescence  is  interference-free, other  nitrogen  compounds do interfere
when ^directed through the N02 converter.  The magnitude of these interferences
is dependent  upon  the type of converter  used  (Winer et al.,  1974;  Joshi  and
Bufalini, 1978).   The detection  limit of  commercial chemiluminescence instru-
ments for N02 measurement is 2.5 MO/m3 (0.002 ppm) (Katz, 1976).
     Development of an instrument based on the  chemiluminescent reaction of
N02 with luminol (5-amino-2, 3-dihydro-l, 4-phthalazine dione) has  been reported
by Maeda et  al.  (1980).   Wendel et al.  (1983) have reported modifications of
this  luminol-based method  in which  better response time and  less interference
from Oo have been  achieved.
     Other acceptable methods  for measuring ambient N02 levels,  including two
methods  designated as equivalent methods, are the  Lyshkow-modified Griess-
Saltzman method,  the  instrumental  colorimetric  Griess-Saltzman method, the
triethanolamine method,  the sodium  arsenite method, and  the TGS-ANSA method
[TGS-ANSA = triethanolamine, guaiacol (o-methoxyphenol), sodium  metabisulfite,
and 8-anilino-l-naphthalene sulfonic  acid].   The sodium arsenite method, and
the TGS-ANSA method were designated as equivalent methods in 1977.   For descrip-
tions of these  methods,  the reader is  referred  to the  EPA criteria document
for nitrogen  oxides  (U.S.   Environmental  Protection Agency, 1982).   While  some
of  these methods  measure the  species  of interest directly,  others require
oxidation, reduction, or thermal decomposition of the  sample,  or separation
from  interferences,   before measurement.   None of these other  techniques,
however, is widely used  to  monitor air quality.
     Careful  adherence to  specified calibration procedures  is  essential  for
obtaining accurate NO measurements.  The U.S. Environmental Protection Agency
                      /\                                                  '
(1975) has issued  a technical assistance  document that  describes in detail the
two acceptable calibration  methods for NO  :  (1) the use of  standard reference
materials (SRMs) and  (2) gas-phase titration (GPT) of NO with 03.   The  SRM for
NO  is  a cylinder  of  compressed  NO  in  N2; the mixture  is  both  accurate and
stable  (Hughes,  1975).   The SRM  for N02  is the N02 permeation tube (O'Keeffe
and Ortman, 1966;  Scaringelli et  al., 1970).   The gas-phase  titration,  described
by  Rehme et  al.  (1974), is based upon the bimolecular  reaction  of  NO with 03
to  form  N02 plus 02.   The U.S. Environmental Protection Agency  (1975) recommends
                                   4-85

-------
the combined  use of GPT plus SRM procedures, using one technique to check the
other.
4.5  REFERENCES
Adamson, P.; Guenthard, Hs, H. (1980) Reinvestigation of the matrix  i.r.  spectrum
     of  peroxyacetylnitrate (PAN).  Spectrochim.  Acta  Part  A 36: 473-475.

Allen, A. p.;  Hochanadel,  C.  J.; Ghormley, J. A.;  Davis, T. W.  (1952)  Decom-
     position  of water and aqueous  solutions under mixed fast  neutron and
     gamma-radiation. J. Phys. Chem. 56: 575-586.

Altshuller, A.  P.  (1983)  Measurements of  the products  of atmospheric photo-
     chemical reactions in laboratory studies  and in ambient air—relationships
     between ozone and other products. Atmos.  Environ.  17: 2383-2427.

Altshuller, A, P.; Leng, L J. (1963) Application of the 3-methyl-2-benzothia-
     zolone hydrazone method  for atmospheric  analysis of aliphatic  aldehydes.
     Anal. Chem. 35: 1541-1542.

Altshuller, A.  P.; McPherson,  S.  P. (1963)  Spectrophotometric  analysis of
     aldehydes  in  the  Los  Angeles atmosphere. J.  Air Pollut.  Control Assoc,
     13: 109-111.

Altshuller, A.  P.;  Cohen,  I.  R.;  Meyer,  M.  E.;  Wartburg, A.  F., Jr.  (1961)
     Analysis of aliphatic aldehydes in source effluents and in  the  atmosphere.
     Anal. Chim. Acta 25: 101-117.

Andreae, W. A. (1955) A sensitive method for  the estimation of hydrogen peroxide
     in biological materials.  Nature (London) 175: 859-860.

Appel, B. R. (1973) A new and more sensitive  procedure  for analysis  of  peroxy-
     benzoyl nitrate. J. Air Pollut. Control  Assoc. 23: 1042-1044.

Armstrong, W. A.; Humphreys, W.  G. (1965)  A L.E.T.  independent dosimeter based
     on the chemiluminescent determination of H202- Can. J. Chem. 43: 2576-2584,

Ballard,  L.  F.;  Tommerdahl, J.  B.;  Decker, C. E.;  Royal,  T.  M,;  Nifong, D.  R.
     (1971a) Field evaluation of new air pollution  monitoring  systems:  the  Los
     Angeles Study:  interim report.  Research  Triangle Park,  NC:  U.S. Environ-
     mental  Protection  Agency;  report  no. APTD-0775.  Available from:  NTIS,
     Springfield, VA; PB-204444.

Ballard,  L.  F.;  Tommerdahl, J.  B.;  Decker, C. E.;  Royal, T. M.; Matus,  L.  K.
     (1971b) Field  evaluation of new air  pollution monitoring systems:  St.
     Louis Study, phase 1.  Contract  no. CPA 70-101, U.S. Environmental  Protec-
     tion Agency, Research Triangle  Park,  NC.  Available from:  NTIS,  Springfield,
     VA; PB-264232.
                                   4-86

-------
Bass,: A. M.; Ledford, A. E., Jr.; Whittaker, J. K. (1977) Ultraviolet photometer
     for ozone calibration.  In: Dimitriades, B., ed.  International conference
     on photochemical  oxidant pollution  and its  control;  September 1976;
     Raleigh, NC. Proceedings: v. I. Research Triangle Park, NC: U.S. Environ-
     mental Protection  Agency;  pp.  13-17; EPA  report  no.  EPA-600/3-77-001a.
     Available from: NTIS, Springfield, VA; PB-264232.

Beard, M.  E.;  Margeson, J. H.;  Ellis,  E.  C.  (1977) Evaluation of 1 percent
     neutral buffered  potassium iodide  procedure for calibration of  ozone
     monitors.   Research Triangle  Park,  NC:  U.S.  Environmental Protection
     Agency; EPA report no. EPA-6Q074-77-QQ5, Available from: NTIS, Springfield,
     VA; PB-267985.

Bergshoeff, G.  (1970)  Ozone determination in air. In: Pittsburgh conference:
     analytical  chemistry and  applied spectroscopy;  publication  no.  342.
     Delft, The Netherlands: IG-TNO.

Bersis, D.; Vassiliou,  E.  (1966) A chemiluminescence  method for determining
     ozone. Analyst 91: 499-505.

Bowman, L D,;  Horak,  R.  F. (1972) A continuous ultraviolet absorption ozone
     photometer. Anal.  Instrum. 10: 103-108.

Bowman, R. L.; Alexander, N. (1966) Ozone-induced chemiluminescence of organic
     compounds. Science (Washington, DC) 154: 1454-1456.

Brewer, A. W.;  Milford, J. R.  (1960)  The Oxford-Kew ozone sonde. Proc.  R.
     Soc.  London Ser. A 256: 470-495.

Bricker, C.  E.;  Johnson,  H.  R.  (1945)  Spectrophotometric method for deter^
     mining formaldehyde.  Ind. Eng. Chem. Anal. Ed. 17: 400-402.

Bruckmann, P.  W.;  Winner, H. (1983) Infrared  spectroscopic study of peroxy-
     acetyl  nitrate (PAN)  and  its decomposition  products.  Environ.  Sci.
     Technol. 17: 352-357.

Bryan,  R.  J.;  Romanovsky, J.  C,  (1956) Instrumentation for  air  pollution.
     Instrum. Autom. 29: 2432-2438.

Bufalini, J. J.;  Brubaker, K. L.  (1969)  Symposium on chemical reactions in
     urban atmospheres;  October.  Warren,  MI: General  Motors Research Labora-
     tories.

Bufalini, J. J.;  Gay,  B. W., Jr.;  Brubaker,  K. L.  (1972) Hydrogen peroxide
     formation  from formaldehyde  photooxidation  and  its presence in  urban
     atmospheres. Environ. Sci. Technol. 6: 816-821.

Burton, C.  S.;  Bekowies, P. J.; Pollack,  R.  I.;  Connell,  P.  (1976)  Oxidant/
     ozone  ambient measurement methods:  an  assessment and evaluation.  San
     Rafael, CA: Systems Applications,  Inc.; EF76-111R.

Butcher,  S.  S.;  Ruff,  R.  E.  (1971) Effect of inlet residence time on analysis
     of  atmospheric nitrogen oxides and  ozone.  Anal. Chem. 43: 1890-1892.
                                   4-87

-------
California Air  Resources  Board.  (1976) A  study  of the effect of atmospheric
     humidity on  analytical  oxidant measurement methods.  Presented ;;at:  15th
     conference [on]  methods in air pollution studies;  January; Long; Beach,
     CA.  Berkeley,  CA; California  Air Resources Board,, Air  and Industrial
     Hygiene Laboratory.
     •                      '         '               '                 •"!.•''
California Air  Resources  Board. (1978) California air quality data;; July-
     August-September.  10(13):  2-3.  Available from: Technical  Service Div.,
     California Air Resources Board, Sacramento, CA.

Carroll,  H.; Dimitriades,  B.;  Ray, R. M.  (1972) A  chemiluminescence.detector
     for  ozone  measurement.  Washington,  DC: U.S. Department  of the Interior,
     Bureau of Mines; report of investigations no.  RI7650.

C.F.R. (1971) 40:§50. Reference methods.                          .   ;

C.F.R. (1975) 40:§53.  Specifications for  automated methods.

Chameides, W.L.;  Tan, A.  (1981).  The  two-demensional diagnostic  model for
     tropospheric OH:  An uncertainty analysis.  J. Geophys.  Res. 86:  5209-5223.

Cherniack, I.;  Bryan, R.. J.  (1965) A  comparison study of various  types  of
     ozone and  oxidant detectors which are used  for atmospheric air sampling.
     J. Air Pollut. Control Assoc.  15: 351-354.

Chock,  D.  P.;  Kumar,  S.;  Herrmann,  R.  W.  (1982)  An  analysis of trends in
     oxidant air  quality  in  the South Coast  Air Basin of California.  Atmos.
     Environ. 16:  2615-2624.

Clark, T.  A.; Baumgardner, R.  E.;  Stevens, R.  K.;  Krost, K.  J.  (1974) Evalua-
     tion  of new  ozone monitoring  instruments  by measuring in nonurban atmos-
     pheres.   In:  Instrumentation for monitoring  air quality:  a symposium;
     August 1973;  Boulder, CO. . Philadelphia, PA: American Society  for Testing
     and Materials; pp. 101-111.

Clements,  J. B. (1975) Workshop on ozone  measurement  by the  potassium iodide
     method summary.  Research Triangle Park, NC: U.S.  Environmental Protection
     Agency; EPA  summary  report no. EPA-650/4-75-007.  Available from: NTIS,
     Springfield,  VA; PB-240939.

Clyne, M.  A. A.;  Coxon, J.  A.  (1968)  Kinetic studies of oxy-halogen radical
     systems.  Proc.  R. Soc.  London Ser. A 303: 207-231.

Cohen, I.  R.; Purcell, T. C.  (1967) Spectrophotometric determination of hydro-
     gen peroxide with 8-quinolinol. Anal.  Chem. 39: 131-132.

Cohen, I.  R.; Purcell, T. C.; Altshuller,  A. P.  (1967) Analysis of  the oxidant
     in photooxidation reactions. Environ.  Sci. Techno!. 1: 247-252.

Corse, E. W.  (1981) Aldehyde analyses for  the 1981  Atlanta study. EPA  contract
     no.  68-02-2566.  Research Triangle  Park, NG:  Northrop  Services,  Inc.-
     Environmental Sciences.
                                   4-88

-------
Cox, ;R±i D,; Balfour, W. D.;  Langley,  G.  J.  (1982)  Quality control  for ambient
     "level hydrocarbon  sampling and  analyses.   Presented  at:  75th annual
     meeting of the Air Pollution  Control  Association;  June;  New OrleanSj  LA.
     Pittsburgh, PA:  Air Pollution Control Association.

Csanyi, L. J.  (1981)  Analytical applications of mixed-ligand complexes. III.
     vanadium(V)-xylenol orange reagent mixture  for the spectrophotometric
     determination of traces of hydrogen peroxide.  Microchem.  J.  26:  10-16.

Darley, E. F.;  Kettner,  K. A.; Stephens, E. R. (1963) Analysis of peroxyacyl
     nitrates  by  gas chromatography  with  electron capture detection. Anal.
     "Chem. 35: 589-591.
Das, T.N.; Moorthy,  P.N.;
     determination of  low
     Chem. Soc. 59: 85-89.
Rao, K.N. (1982).
concentrations  of
Chemiluminescent method for the
hydrogen peroxide.  J.  Indian
Das, T.N.; Moorthy,  P.N.;  Rao, K.N. (1983).  Chemiluminescent measurement of
     hydrogen peroxide in the Bombay area.  Atmos. Environ. 17: 79-82.

Dasgupta, P.  K. (1980a) Discussion on "The importance of atmospheric ozone and.
     hydrogen peroxide in oxidizing sulphur dioxide in cloud and rainwater [on
     Penkett et al., 1979]." Atmos. Environ. 14: 272-275.

Dasgupta, P.  K.  (1980b) The importance  of  atmospheric  ozone and hydrogen
     peroxide  in  oxidizing  sulfur  dioxide in cloud and  rainwater—further
     discussion [on Penkett et al., 1979], Atmos. Environ. 14: 620-621.

Dasgupta, P.  K.;  Hwang,  H.  (1985).  Application  of  a  nested loop  system for
     the flow  injection  analysis  of .trace aqueous peroxide.   Anal.  Chem.  In
     press.

Davis, D. R.;  Jensen,  R.  E. (1976)  Low level  ozone  and  weather systems.  In:
     Specialty  conference   on  ozone/oxidants—interactions  with  the total
     environment.  Philadelphia, PA:   Air Pollution Control  Association;  pp.
     242-251.
Demerjian, K.  L.;  Kerr,  J. A.; Calvert, J.
     chemical  smog  formation.   In:   Pitts,
     Advances  in environmental science and
     John Wiley and Sons,  Inc.; pp. 1-262.
                 G. (1974) The mechanism of photo-
                 J. N.,  Jr.;  Metcalf, R.  L.,  eds.
                 technology:  v.  4.  New York, NY:
DeMore, W.  B.; Patapoff,  M.  (1976)  Comparison  of ozone determinations by
     ultraviolet photometry and gas-phase titration. Environ.  Sci.  Techno!.
     10: 897-899.

DeMore, W.  B.;  Romanovsky, J.  C.; Feldstein, M.; Hamming, W. J.; Mueller, P.
     K. (1976) Interagency comparison of  iodometric  methods for ozone determi-
     nation.  In:  Calibration in air monitoring: a  symposium;  August  1975;  .
     Boulder, CO.  Philadelphia,  PA: American Society for Testing and Materials;
     pp. 131-143; ASTM special  technical  publication 598.
Dietz, W. A. (1967) Response factors for gas chromatographic  analyses.
     Chromatogr. 5: 68-71.
                                             J. Gas
                                   4-89

-------
Eegriwe, E.  (1937)  Reactions and reagents  for  the detection of organic com-
     pounds.  Z. Anal. Chem. 110: 22025.

Fincher, C.  L;  Tucker,  A.  W.;  Birnbaum,  M.;  Paur, R. J.;  McClenny,  W.  A.
     (1977)  Fluorescence  ambient N02 monitor with  flashlamp pumping.   Appl
     Optics 16: 1359-1365.

Finlayson, B.  J.;  Pitts,  J. N., Jr.;  Atkinson,  R. (1974) Low  pressure gas
     phase ozone-olefin reactions: chemiluminescence,  kinetics, and mechanisms.
     J. Am. Chem. Soc. 96: 5356-5367.

Flamm, D. L. (1977) Analysis of ozone at low concentrations  with boric  acid buffered
     KI.  Environ. Sci. Techno!. 11: 978-983.

Fontijn, A.; Sabadell,"A. J.; Ronco, R. J.  (1970)  Homogeneous chemiluminescent
     measurement  of nitric  oxide  with ozone:  implications  for continuous
     selective monitoring of gaseous air pollutants.   Anal.  Chem. 42: 575-579.

Foulger, B. E.; Simmonds, P. G.  (1979) Drier for field use in the determination
     of trace atmospheric gases. Anal. Chem. 51: 1089-1090.

F. R. (1971, April 30) 36: 8186-8201.  National primary and  secondary air quality
     standards.

F.R. (1975a, February 18) 40: 7042-7070.  Ambient  air  monitoring reference and
     equivalent methods.

F.R.  (1975b,  February 18) 40:  54856.  Ambient air monitoring reference and
     equivalent methods:  notice of reference method designations.

F.R. (1976a,  February 4)  41: 5145-5146. Ambient air monitoring reference and
     equivalent methods:  reference method designation.

F.R. (1976b, October 22) 41: 46647-46648. Ambient  air  monitoring reference and
     equivalent methods:  reference method designation.

F.R. (1976c, December 1) 41: 52686-52692. National primary and  secondary ambient
     air quality standards: nitrogen dioxide measurement principle and  calibra-
     tion procedure.

F.R. (1976d, December 8) 41: 53684-53685. Ambient  air  monitoring reference and
     equivalent methods:  reference method designation.

F.R.  (1977a,  June 3) 42:28571-28572.  Ambient  air monitoring reference and
     equivalent methods:  reference and equivalent  method designation.

F.R. (1977b, June 13) 42:30235.  Ambient air monitoring reference and equivalent
     methods:  Amendment to reference methods for  ozone.

F.R. (1977c,  August 1) 42:38931-38932.  Ambient air monitoring reference and
     equivalent methods:  equivalent method  designation.

F.R.  (1977d,  November 1) 42:  57156.   Ambient air monitoring reference and
     equivalent methods:  Philips PW9771 03  analyzer; correction.


                                   4-90

-------
F.R- (19793, February  20)  44:  10429-10430.  Ambient air monitoring reference
     and equivalent method designations: Columbia Scientific  Industries Model
     2000 Ozone Meter  Thermo Electron Model 14D/E Chemiluminescent  NO-N02-N0-
     Analyser.

F.R. (1979b, May 10) 44: 27558-27604. Air programs;  ambient air quality monitor-
     ing, data reporting, and surveillance provisions.

F.R. (1979c, November  9) 44: 65069-65071.  Regulations  for  ambient air  quality
     monitoring and data reporting.

F.R. (1979d, December  14)  44:  72589-72592. Air programs; ambient, air quality
     monitoring, data reporting, and surveillance provisions  [errata].

F.R., (1979e, February 8) 44:  8221-8233.  Calibration of ozone  reference methods.

F.R. (1980s, March 21) 45:  18474. Ambient air monitoring reference  and equiva-
     lent methods; amendment to reference method for ozone.

F.R. (1980b, August 27)  45: 57168-57169. Ambient air  monitoring  reference and
     equivalent methods; equivalent method designation.

F.R. (1981, October 26) 46: 52224. Ambient air monitoring reference and equiva-
     lent methods; equivalent method designation.

F.R. (1982,  March  31)  47:  13572-13573.  Ambient air monitoring reference  and
     equivalent methods; equivalent method designation.

F.R. (1983,  March  10)  48:  10126.  Ambient air  monitoring reference and equiva-
     lent methods; equivalent method designation.

Fried, A.;  Hodgeson, J.  (1982)  Laser photoacoustic  detection  of  nitrogen  dio-
     xide  in the  gas-phase titration of nitric  oxide  with  ozone.  Anal. Chem.
     54: 278-282.

Fung, K.;  Grosjean,  D. (1981) Determination of nanogram amounts of carbonyls
     as 2,4-dinitrophenylhydrazones by  high-performance liquid chromatography.
     Anal. Chem. 53: 168-171.

Gay, B. W., Jr.; Bufalini, J. J.  (1972a) Hydrogen peroxide  in the urban atmos-
     phere.  In: Gould, R.  F., ed.,  Photochemical smog and ozone reactions.
     Washington,  DC:  American Chemical  Society;  pp.   225-263. (Advances  in
     chemistry series: 113.)

Gay, B. W., Jr.; Bufalini, J. J.  (1972b) Hydrogen peroxide  in the urban atmos-
     phere.  Environ.  Lett. 3: 21-24.

Gay, B.  W.;  Noonan,  R. C.; Bufalini, J. J.;  Hanst,  P.  L.  (1976)  Photochemical
     synthesis  of peroxyacyl nitrates  in gas  phase  via  chlorine-aldehyde
     reaction. Environ.  Sci.  Techno!. 10:  82-85.

Graedel,  T.  E.;  Farrow, L. A.;  Weber,  T.  A.  (1976)  Kinetic  studies  of the
     photochemistry of the urban troposphere. Atmos.  Environ. 10:  1095-1116.
                                   4-91

-------
Grosjean, D.  (1982)  Distribution of nitrogenous pollutants during photochemi-
     cal smog episodes.  Presented  at: 75th annual meeting  of the Air  Pollution
     Control Association; June;  New Orleans,  LA.  Pittsburgh, PA: Air Pollution
     Control Association; paper  no. 82-41.4.

Grosjean, D.  (1983)  Distribution of atmospheric  nitrogenous  pollutants at a
     Los Angeles  area smog receptor site.  Environ.  Sci.  Technol.  17:  13-19.

Grosjean, D.;  Kpk,  G. L. (1981)  Inter!aboratory  comparison study of methods
     for measuring  formaldehyde  and other aldehydes  in ambient air.   Westlake
     Village, CA:  Environmental  Research and Technology,  Inc.  Available from:
     NTIS, Springfield, VA; PB82-224486.

Grosjean, D.; Fung, K.; Collins, J.; Harrison,  J.; Breitung, E.  (1984) Portable
     generator  for on-site calibration  of peroxyacetyl  nitrate analyzers.
     Anal. Chem. 56:  569-573.

Hager, R. N.,  Jr.; Anderson, R.  C.  (1970) Theory of the  derivative spectro-
     meter.   J. Opt.  Soc. Am. 60: 1444-1449.

Hampson, R.  F., ed.;  Braun, W.;  Brown, R.  L.; Garvin,  D.; Herron, J. T.;  Huie,
     R. E.;  Kurylo,  M,  J.;  Laufer,  A.  H.; McKinley,  J. D.;  Qkabe,  H.;  Scheer,
     M. D.;  Tsang, W.; Stedman,  D.  H.  (1973) Survey  of photochemical  and rate
     data for  twenty-eight  reactions  of  interest in  atmospheric chemistry. J.
     Phys. Chem. Ref.  Data 2: 267-311.

Hanst, P. L. (1970) Infrared spectroscopy and infrared lasers  in air pollution
     research and monitoring.  Appl. Spectrosc. 24:  161-174.

Hanst, P. L.;  Wilson, W.  E.;  Patterson,  R. K.;  Gay,  B. W.,  Jr.; Chaney, L. W.
     (1975) A  spectroscopic study of California smog.  Research Triangle Park,
     NC: U.S. Environmental Protection Agency,  National Environmental  Research
     Center; EPA report no. EPA-650/4-75-006. Available from: NTIS, Springfield,
     VA; PB-241022.

Hanst, P. L.; Wong, N. W.; Bragin,  J.  (1982)  A  long-path  infrared study of Los
     Angeles smog. Atmos. Environ.  16: 969-981.

Harrison, J. W.; Timmons, M. L.; Denyszyn,  R. B.;  Decker, C. E.  (1977) Evalua-
     tion of the EPA  Reference Method  for measurement  of  non-methane hydrocar-
     bons.  Research  Triangle Park, NC:   U.S. Environmental Protection Agency;
     EPA report no.  EPA-600/4-77-033.  Available  from:  NTIS, Springfield,  VA;
     PB-278296.
                                                            >/
Hauser, T. R.;  Cummins, R. L. (1964) Increasing sensitivity of  3-methyl-2-benzo-
     thiazolone hydrazone  test for analysis  of aliphatic aldehydes in  air.
     Anal. Chem. 36:  679-681.

Hauser, T. R.;  Kolar,  A. M. (1968)  Spectrophotometric  determination of hydrogen
     peroxide  in  aqueous media with l,2-Di-(4-pyriclyl) ethylene.  Anal. Chem.
     40: 231-232.

Heikes,  B.G.  (1984)  Aqueous  H202  production from  03  in  glass impingers.
     Atmos. Environ.  18: 1433-1445.


                                    4-92

-------
Heikes; B. G. ; Lazrus, A. L. ; Kok, G. L. ; Kunen, S. M. ; Gandrud,  B. W. ; Gitlin,
  -^ S. N. ;  Sperry,  P.  D.  (1982)  Evidence for aqueous phase hydrogen peroxide
  : "^synthesis in the troposphere. JGR J. Geophys.  Res. 87:  3045-3051.

Hendry, D. G. ; Kenley,  R.  A.  (1977)  Generation of peroxy radicals from peroxy
   '">m'trates (RQ2N02).  Decomposition of peroxyacyl nitrates.  Atmos.  Environ.
   •  99: 3198-3199.

Hersch, P.;  Deuringer,  R.  (1963) Galvanic  monitoring of ozone in air. Anal.
           35: 897-899.
Heuss, J. M. ; Glasson, W. A. (1968) Hydrocarbon  reactivity  and  eye  irritation.
    .Environ. Sci. Techno!. 2: 1109-1116.

Hinkley, E.  D. ;  Kelley,  P.  L.  (1971)  Detection of air pollutants with tunable
     diode lasers.  Science (Washington, DC) 171:  635-639.

Hochanadel , C. J.  (1952) Effects of  cobalt ^-radiation on water and aqueous
     solutions. J. Phys. Chem. 56: 587-594.

Hodgeson j J.  A.  (1972) Review of analytical methods  for  atmospheric  oxidants
     measurements. Int. J. Environ. Anal. Chem.  2:  113-132.

Hodgeson, J.  A.  (1976)  A  survey of calibration  techniques for atmospheric
     ozone monitors.  Washington, DC:   National  Bureau  of Standards;  publica-
     tion no. NBSIR 76-1191. Available from: NTIS,  Springfield,  VA;  PB-263534.

Hodgeson, J.  A.;  Krost, K. J. ;  Q'Keeffe, A. E. ;  Stevens, R.  K.  (1970)  Chemi-
     luminescent  measurement  of atmospheric ozone:  response characteristics
     and operating variables.  Anal.  Chem.  42: 1795-1802.

Hodgeson, J.  A.;  Stevens, R.   K. ; Martin,  B.  E.  (1972) A stable ozone source
     applicable as a  secondary standard  for calibration of  atmospheric  monitors.
     ISA Trans. 11: 161-167.

Hodgeson, J. A.; Hughes, E. E. ;  Schmidt, W. P.;  Bass, A.  M.  (1977)  Methodology
     for standardization of atmospheric ozone measurements.  In:  Dimitriades,
     B. , ed.   International  conference on photochemical oxidant pollution and
     its control;  September 1976; Raleigh,  NC.  Proceedings: v.  I.  Research
     Triangle  Park,  NC:   U.S.  Environmental Protection Agency;  pp.  3-12;  EPA
     report  no.  EPA-600/3-77-001a. Available  from:  NTIS,  Springfield, VA;
     PB-264232.

Hoigne,  J. ;  Bader,  H. (1976)  The role of hydroxyl radical reactions in ozona-
     tion processes in aqueous solutions. Water  Res.  10:  377-386.

Holdren, M.  W. ;  Rasmussen,  R.  A.  (1976)  Moisture anomaly  of  peroxyacetyl
     nitrate  (PAN). Environ. Sci. Techno!.  10: 185-187.

Holdren, M.  W. ;  Spicer, C. W.  (1984) Field compatible calibration procedure
     for peroxyacetyl  nitrate. Environ.  Sci. Techno!.  18: 113-116.
                                    4-93

-------
Holdren, M,; Spicer, C.; Sticksel, P.; Nepsund, K.; Ward, G.; Smith, R. (1982)
     Implementation and  analysis  of hydrocarbon grab  samples  from Cleveland
     and Cincinnati 1981. Ozone monitoring study.  Research Triangle Rark, NC:
     U.S.  Environmental  Protection Agency;  EPA report  no.  EPA-905/4-82-001.
     Available from: NTIS, Springfield, VA; PB83-197673.               :

Hughes, E. E. (1975) Development of standard reference materials for afr quality
     measurement. ISA Trans. 14: 281-291.

Huygen,  I. C.  (1970)  Reaction of  nitrogen dioxide with Griess type reagents.
     Anal. Chem. 42: 407-409.

Intersociety Committee,  American  Public  Health Association.  (1970) Manual  and
     methods of air sampling and  analysis:  tentative  method for  continuous
     monitoring  of  atmospheric  oxidants  with amperometric instruments.  Health
     Lab. Sci.  7: 1.

Intersociety Committee,  American  Public  Health Association.  (1977a) Tentative
     method of analysis for formaldehyde content of the atmosphere  (MBTH-colori-
     metric method—application to other aldehydes). In:  Katz, M., ed.  Methods
     of air  sampling  and analysis. 2nd  ed.  Washington  DC:  American Public
     Health Association; pp. 308-313.

Intersociety Committee,  American  Public  Health Association,  (1977b) Tentative
     method of  analysis  for nitrogen dioxide  content  of the atmosphere.  In:
     Katz, M.,  ed.  Methods  of air sampling and analysis. 2nd ed. Washington,
     DC: American Public Health Association.

Jayanty,  R.  K.   M.; Blackard,  A.;  McElroy,  F.  F.;  McClenny,  W.  A.  (1982)
     Laboratory  evaluation  of  nonmethane  organic carbon determination in
     ambient air by cryogenic preconcentration and flame  ionization detection.
     Research Triangle Park,  NC:   U.S.  Environmental  Protection Agency;  EPA
     report  no.  EPA-6QO/4-82-019.  Available from: NTIS,  Springfield, VA;
     PB82-224965.

Johnson, D. F.;  Kok, G. L.; Sonner, R. J. (1981) Improved chromatographic acid
     technique for the determination of formaldehyde.  In:  Grosjean, D.; Kok,
     G.  L.,  eds.   Inter!aboratory comparison  study of methods  for measuring
     formaldehyde and  other aldehydes in ambient air: final report. Atlanta,
     GA: Coordinating Research Council.

Joshi, S. B.; Bufalini, J. J. (1978) Halocarbon interferences in chemilumines-
     cent measurements of NO .  Environ. Sci.  Techno!. 12: 597-599.
                            f\

Katz, M.  (1976)  Nitrogen compounds and oxidants.  In: Stern, A. C., ed.  Air
     pollution;  v. III. measurement, monitoring and surveillance of air pollu-
     tion. 3rd ed. New York, NY: Academic Press; pp. 259-305.

Killus,  J.  P.;  Whitten,  G.  Z.  (1985)  Behavior of  trace NO   species  in
     the  nighttime  urban atmosphere.  J. Geophys.  Res.  90; pp. 2430-2432.

Kilpatrick, M.  L.;  Herrich,  C.  C.; Kilpatrick, M.  (1955) The decomposition of
     ozone in aqueous solution. J. Am. Chem. Soc. 78:  1784-1789.
                                   4-94

-------
Kok, G._L.;  Partial!,  K.  R.; Winer, A. M.; Pitts, J, N., Jr.; Gay, B. W., Jr.
     (1978a) Ambient  air  measurements of hydrogen peroxide  in the California
     South Coast Air Basin. Environ. Sci. Technol. 12:  1077-1080.
    . •;•," I; at. .
Kok, G. L.;  Holler,  T.  P.; Lopez,  M.  B.; Nachtrieb,  H. A.;  Yuan,  M.  (1978b)
     Chemiluminescent method  for determination  of  hydrogen  peroxide in the
     ambient atmosphere.  Environ. Sci. Technol.  12: 1072-1076.

Kravetz, T.  M.;  Martin,  S. W.; Mendenhall, G. D. (1980) Synthesis of peroxy-
     acetyl  and  peroxyaroyl nitrates:  complexion of peroxyacetyl nitrate with
     benzene. Environ. Sci. Technol. 14:  1262-1264.

Kreuzer, L.  B.;  Patel,  C. K.  N. (1971) Nitric oxide air pollution: detection
     by optoacoustic spectroscopy. Science (Washington, DC)  173: 45-47.

Krug, E. L. R.; Hirt, W.  E. (1977) Interference  of nitrate in the  determination
     of formaldehyde by the chromotropic  acid method.  Anal.  Chem.  49: 1865-1867.

Kunen, S.  M.;  Lazrus, A.   L.; Kok, G.  L.; Heikes, B. G.  (1983) Aqueous  oxida-
     tion  of  S02 by hydrogen peroxides.  JGR  J.  Geophys.  Res.  88:  3671-3674.

Kuntz, R.;  Lonneman,  W.;  Namie,  G.;  Hull, L.  A.  (1980) Rapid determination of
     aldehydes in air analyses. Anal.  Lett. 13:  1409-1415.

Lamb, S.  I.;  Petrowski,  C.;  Kaplan,  I.  R. ; Simoneit,  B. R.  T. (1980) Organic
     compounds in  urban atmospheres:  a review of distribution,  collection and
     analysis. J. Air Pollut. Control  Assoc.  30:  1098-1115.

Levaggi, D.  A.;  Kothny,  E. L.;  Belsky,  T.;  de  Vera,  E.  R.; Mueller,  P. K.
     (1974)  Quantitative  analysis  of nitric  oxide in  presence  of nitrogen
     dioxide at  atmospheric concentrations. Environ. Sci. Technol. 8: 348-350.

Lipari, F.  (1984) New solid-sorbent method for ambient nitrogen  dioxide monitor-
     ing. Anal.  Chem. 56:  1820-1826.

Lipari, F.;  Swarin, S. (1982) Determination of formaldehyde  and  other aldehydes
     in automobile exhaust with an  improved 2,4-dinitrophenylhydrazine  method.
     J. Chromatogr. 247:  297-306.

Lipari,  F.;  Swarin,  S.   (1985)  2,4-Dinitrophenylhydrazine-coated Florisil
     sampling cartridges  for  the determination of formaldehyde in  air.  Environ.
     Sci.  Technol. 19: 70-74.

Littman,  F.  E.; Benoliel,  R.  W.  (1953)  Continuous  oxidant  recorder.  Anal.
     Chem.  25: 1480-1483.

Lloyd,  A.  C. (1979) Tropospheric  chemistry  of aldehydes.  In:  Herron, J. T.;
     Huie,  R. E.; Hodgeson, J. A.,  eds.  Chemical kinetic data  needs  for model-
     ing of  the  lower troposphere:  proceedings of a  workshop;  May  1978; Reston,
     VA.   Gaithersburg,  MD:  U.S.  Department of Commerce, National  Measurement
     Laboratory; Bureau of Standards  special  publication 557.

Logan,  J.A.; Prather, M.J.;  Wofsy,  S.C.; McElroy,  M.B. (1981)  Tropospheric
     chemistry:  A global  perspective.  J. Geophys.  Res. 86:  7210-7254.


                                    4-95

-------
Lonneman, W.A.  (1977)  PAN  measurement in dry and humid atmospheres.   Environ.
     Sci. Techno!. 11: 194-196.                                   ;  ---,"

Lonneman, W. A. (1979) Ambient hydrocarbon measurements in Houston.  In-: Ozone/
     oxidants:  interactions  with  the total  environment II:  proceedings  of a
     specialty  conference;  October;  Houston, TX.   Pittsburgh,  PA: . J-iir Pollu-
     tion Control Association; pp. 278-287.

Lonneman, W.  A.;  Kopczynski, S.  L.;  Parley,  P.  E.; Sutterfield, F.  D, (1974)
     Hydrocarbon  composition of  urban air pollution.  Environ. Sci. Techno!.
     8: 229-236.

Lonneman, W. A.; Bufalini, J. J.; Seila,  R.  L. (1976)  PAN and oxidant  measure-
     ment in ambient atmospheres. Environ. Sci. Techno!. 10: 374-380.:

Lonneman, W. A.; Bufalini, J. J.; Kuntz,  R.  L.; Meeks, S. A. (1981)  Contamina-
     tion from fluorocarbon  films.   Environ.  Sci. Techno!. 15; 99-103.

Lonneman, W.  A.;  Bufalini, J. J.; Namie,  G. R.  (1982) Calibration  procedure
     for  PAN  based on its thermal  decomposition in the presence of nitric
     oxide. Environ. Sci, Techno!. 16: 655-660.

Lord,  H.  C.;  Egan, W. D.; Johnson,  F. L.; Mclntosh, L. C. (1975) Measurement
     of exhaust emissions  in piston  and  diese! engines by dispersive spectrum.
     J. Air Pollut. Control  Assoc. 4: 136-139.

Lovelock, J.  E.;  Maggs,  R.  J.; Adlard,  E.  R. (1971)  Gas-phase coulometry by
     thermal electron attachment. Anal.  Chem.  43: 1962-1965.

Maeda,  Y.;  Aoki, K.; Munemori,  M.  (1980) Chemiluminescence method  for  the
     determination of nitrogen dioxide.   Ana!. Chem. 52: 307-311.

Manos, D. M.; Volltrauer, H. N.; Allen,  J.;  Burch,  D.  E.; Chudy, J.  S.; Pembrook,
     J. D. (1982) Totally optical ambient NMHC monitor. EPA  contract no.  68-02-
     3292. Research Triangle Park, NC:  U.S.  Environmental Protection  Agency.

Martin,  L.  R.;  Damschen, D. E. (1981) Aqueous oxidation of  sulfur dioxide  by
     hydrogen peroxide at  low pH. Atmos.  Environ. 15:  1615-1621.

Mast,  G.  M.;  Saunders, H.  E.  (1962) Research and development of the instru-
     mentation of ozone  sensing.  ISA Trans.  1: 325-328.

McClenny, W.  A.;  Pleil,  J.  D.; Holdren,  M.  W.;  Smith, R. N. (1984)  Automated
     cryogenic  preconcentration  and gas  chromatographic  determination  of
     volatile organic compounds in air.  Ana!,  Chem. 56: 2947-2951.

McElroy,  F.  F.  (1979) Transfer standards  for the calibration of ambient air
     monitoring  analyzers  for ozone: technical  assistance document. Research
     Triangle  Park, NC:  U.S. Environmental  Protection  Agency,  Environmental
     Monitoring  and Support Laboratory; EPA report  no.  EPA-600/4-79-056.
     Available  from: NTIS, Springfield,  VA;  PB-146871.

McElroy,  F.  F.; Thompson,  V.  L.  (1975)  Hydrocarbon measurement discrepancies
     among  various analyzers  using  flame-ionization  detectors.  Research
     Triangle  Park, NC:   U.S. Environmental  Protection Agency,  Environmental
     Monitoring  and Support Laboratory; EPA report  no.  EPA-60Q/4-75-Q1Q.
                                   4-96

-------
Meloan, C.  E.; Mauck,  M.;  Huffman,  C.  (1961)  Spectrophotometric determination
     of traces of hydrogen peroxide. Anal. Chem. 33: 104-106.

Miksch, P. R.; Anthon, D. W. (1982) A recommendation for combining the standard
     analytical methods  for the  determinations of  formaldehyde and total
     aldehydes in air.  Am. Ind. Hyg. Assoc. J. 43: 362-365.

Mottola, H. A.; Simpson, B. E.; Gorin, G. (1970) Absorptiometric determination
     of hydrogen  peroxide  in  submicrogram amounts with  leuco crystal  violet
     and peroxidase as catalyst. Anal. Chem. 42: 410-411.

Mueller,  P.  K.; Tokiwa,  Y.; deVera,  E.  R.;  Wehrmeister, W.  J.; Belsky, T.;
     TWiss,  S.; Inada,  M.  (1973) A  guide  for the evaluation  of atmospheric
     analyzers. Research Triangle  Park,  NC:  U.S. Environmental  Protection
     Agency, Quality  Assurance  and Environmental Monitoring  Laboratory;  EPA
     report  no.  EPA-650/4-74-014.  Available  from:  NTIS,  Springfield, VA;
     PB-233048.

National Air Pollution Control  Administration.  (1970) Air quality criteria  for
     photochemical oxidants. Washington, DC: U.S. Department  of Health, Educa-
     tion, and Welfare; publication no. AP-63.  Available from:  NTIS, Springfield,
     VA; PB82-242231.

National  Research Council. (1976)  Vapor-phase organic pollutants:  volatile
     hydrocarbons and  oxidation products.   Washington, DC:   National Academy
     of Sciences.

National  Research Council. (1977)  Ozone and other  photochemical  oxidants.
     Washington,  DC:  National  Academy of  Sciences;  pp.  45-125 and chapters
     4 and 6.

National Research Council. (1981) Formaldehyde  and other aldehydes.  Washington,
     DC: National  Academy  of Sciences.

Neal,  R.;  Severs,  R.  K.; Wenzel,  L.; MacKenzie, K.  (1976)  Simultaneous chemi-
     luminescent  ozone  and KI  oxidant measurements  in  Houston, Texas,  1975.
     In:  Speciality  conference  on ozone/oxidants—interaction  with the total
     environment.  Philadelphia,  PA:  Air Pollution Control  Association;   pp.
     180-188.

Nederbragt,  G. W.;  van der Horst,  A.; van Duijn,  J.  (1965)  Rapid ozone deter-
     mination  near an accelerator. Nature (London) 206: 87.

Nicksic,  S.  W.;  Harkins, J.;  Mueller, P. K.  (1967)  Some analyses for PAN and
     studies of its structure.  Atmos. Environ.  1: 11-18.

Nieboer,  H.; van  Ham, J.  (1976)  Peroxyacetyl  nitrate (PAN)  in relation  to
     ozone  and some meteorological  parameters  at  Delft in The Netherlands.
     Atmos.  Environ. 10: 115-120.

Nielsen, T.; Hansen, A.  M.; Thomsen,  E.  L.  (1982) A  convenient method for pre-
     paration  of  pure  standards of peroxyacetyl nitrate  for atmospheric analyses.
     Atmos.  Environ. 16: 2447-2450.
                                   4-97

-------
Niki, H.; Maker,  P.  D.; Savage,  C.  M.;  Breitenbach, L. P.  (1976)  Infrared
     Fourier-transform  spectroscoplc studies  of atmospheric reactions. In:
     12th informal conference on  photochemistry, National Bureau of Standards;
     June-July; Gaithersburg,  MD: National  Bureau of  Standards; paper N2.

Niki, H.; Maker, P. D.; Savage, C. M.; Breitenbach,  L.  P. (1980) An FTIR study
     of  the mechanism for the gas phase reaction between H02 radicals. Chem.
     Phys. Lett. 73: 43-46.

Norwitz, 6. (1966)  A colorimetric method  for  the  determination of  oxides  of
     nitrogen.  Analyst 91: 553-558.

Ogle, L. D.; Hall, R. C.; Crow, W. L.; Jones,  A. E.; Gise, J. P. (1982) Devel-
     opment of preconcentration and  chromatographic procedures  for the  contin-
     uous and unattended monitoring  of hydrocarbons  in  ambient  air.  Presented
     at: 184th  national  meeting of  the American Chemical Society; September;
     Kansas City, MO. Austin, TX: Radian Corporation.

Okabe, H.; Schwarz, F. P. (1975)  Nitric oxide  pollution monitor. U.S. Patent No.
     3,906,226 (Cl 250-304, G01J) Sept. 16, 1975.

O'Keeffe, A. E.; Ortman, G. C. (1966) Primary  standards for  trace gas analysis.
     Anal. Chem. 38: 760-763.

Okita, T.; Inugami, M. (1971) Measurement  of ozone  by chemiluminescent  method.
     Koshu Eiseiin Kenkyu Hokoku  20:  207-211.

Olansky, A.  S.; Deminp, S. N. (1976)  Optimization and interpretation of absor-
     bance  response  in the determination  of formaldehyde with chromotropic
     acid.  Anal. Chim. Acta 83:  241-249.

Osman, M.;  Hill,  H.  H.; Holdren, M.  W.;  Westberg, H.  H. (1979) Vapor-phase
     silylation of  alcohols  for air  analysis.   In:  Zlatkis,  A., ed.   Advances
     in  chromatography, ehromatography symposium; Houston, TX.

Ott, W.  (1977)  Development of criteria for siting  air monitoring stations.  J.
     Air Pollut. Control Assoc. 27:  543-547.

Overton, J.  H., Jr.; Durham, J. L. (1982)  Acidification of rain in the  presence
     of  S02, H202, 03, and HN03.  In:  Keith,  L. H.,  ed.  Energy and environmental
     chemistry, v. 2: acid rain.  Ann Arbor, MI:  Ann Arbor Science; pp.  245-262.

Papa, L.  J.;  Turner, L. P.  (1972)  Chromatographic determination of carbonyl
     compounds as their 2, 4-dinitrophenylhydrazones. II. High  pressure liquid
     chromatography. J. Chromatogr.  Sci. 10: 747-50.

Paur, R. J.;  McElroy,  F.  F. (1979) Technical assistance document for the
     calibration  of  ambient  ozone monitors. Research Triangle  Park, NC: U.S.
     Environmental  Protection Agency,  Environmental  Monitoring and Support
     Laboratory;  EPA  final report no. EPA-600/4-79-057. Available from: NTIS,
     Springfield, VA; PB80-149552.
                                   4-98

-------
Peake, E.;  Sandhu, H. S. (1982) The formation of ozone and peroxyacetyl nitrate
     (PAN)  in the  urban atmospheres of Alberta.   Presented  at:  65th annual
     conference of the  Chemical  Institute of Canada;  May-June;  Toronto,  ON,
     Canada.  Can.  J.  Chem.  61: 927-935.

Penkett, S. A.; Jones,  B.  M. R.; Brice, K. A.; Eggleton, A. E. J. (1979) The
    • importance of atmospheric ozone and hydrogen peroxide in oxidising sulphur
     dioxide in cloud and rainwater. Atmos. Environ. 13: 123-137.

Perschke,  H.;  Broda,  E.  (1961) Determination of very small amounts of hydrogen
     peroxide.  Nature (London) 190: 257-258.

Pilz, W.;  Johann,  I.  (1974) Die Bestimmung kleinster Mengen von Wasserstoff-
     peroxid in Luft. [Measurement of  very small  quantities  of hydrogen  pero-
     xide in air.]  Int. J. Environ. Anal. Chem. 3: 257-270.

Pisarevskii, A. M.;  Polozova, I.  P. Avramenko, L. V. (1980) An electrode for
     potentiometric determination  of hydrogen  peroxide in  alkaline media.   J.
     Anal.  Chem. USSR 35: 1419-1425.

Pitts, J.N., Jr.;  Fuhr,  H.; Gaffney, J.S.; Peters J.W. (1973).  Chemiluminescent
     reactions of peroxyacetyl nitrate and ozone with triethyl amine.  Possible
     atmospheric monitor for peroxyacetyl  nitrate.   Environ.  Sci. Tech. 7:
     550-552.

Platt, U.; Perner, D.;  Winer, A. M.; Harris, G. W.; Pitts, J. N., Jr. (1980)
     Detection of  N03 in the polluted  troposphere by differential  optical  ab-
     sorption.   Geophys. Res. Lett. 7:  89-92.

Platt, U.; Winer,  A.  M.; Biermann, H.  W.;  Atkinson, R.; Pitts,  J.  N.,  Jr.
     (1984) Measurement of nitrate .radical  concentrations  in continental  air.
     Environ.  Sci. Techno!.:  in press.

Poli, A.  A.;  Zinn, T.  L.  (1973)  A new continuous  monitor for methane and
     non-methane total  hydrocarbons.  Anal. Instrum. 11: 135-142.

Purcell, T. C.; Cohen,  I. R.  (1967) Microdetermination of peroxides by kinetic
     colorimetry.  Environ.  Sci. Technol. 1: 431-433.

Reckner, L. R.  (1974) Survey of  users  of  EPA  Reference Method for measurement
     of non-methane  hydrocarbon  in ambient air.  Research Triangle Park, NC:
     U.S.  Environmental  Protection Agency, Office of Research and  Development;
     EPA report no.  EPA-650/4-75-008.  Available  from: NTIS, Springfield, VA;
     PB-247515.

Regener, V. H.  (1960) On a  sensitive method for  the recording of  atmospheric
     ozone. JGR J. Geophys.  Res.  65: 3975-3977.

Regener, V. H.  (1964) Measurement of atmospheric ozone with the chemilumine-
     scent method. JGR  J. Geophys. Res. 69: 3795-3800.

Rehme,  K.A.; Martin,  B.E.; Hodgeson, J.A.  (1974)   Tentative method for the
     calibration of nitric oxide,  nitrogen dioxide, and ozone analyzers by gas
     phase titration.   U.S.  Environmental  Protection Agency, Research Triangle
     Park, N.C.  EPA-R2-73-246.

                                   4-99

-------
Rehme, K. A.;  Puzak,  J.  C.; Beard,  M.  E.;  Smith, C. F.; Paur, R. J... (1981)
     Evaluation of  ozone  calibration procedures. Research Triangle Park, NC:
     U.S. Environmental  Protection Agency,  Environmental Monitoring  Systems
     Laboratory; EPA  final report no.  EPA-600/4-80-050.  Available from:  NTIS,
     Springfield, VA; PB81-118911.

Renzetti, N. A.; Romanovsky, J. C. (1956) Comparative study of oxidants  in the
     Los Angeles atmosphere.  J. Air  Pollut.  Control Assoc. 6: 154-159.

Richter, H. G.  (1983) Analysis of organic compound data gathered  during 1980
     in northeast corridor cities.   Research Triangle Park, NC:  U.S. Environ-
     mental Protection Agency,  Office of Air Quality Planning and Standards;
     EPA report  no.  EPA-450/4-83-017.  Available  from: NTIS, Springfield, VA;
     PB84-116052.

Ridley, B.  A.;  Hewlett,  L. C. (1974) An instrument for nitric oxide measure-
     ments in the stratosphere.   Rev. Sci. Instrum. 45: 742-746.

Riggin, R.  M.  (1983)  Technical assistance document for sampling and analysis
     of toxic  organic compounds in  ambient air.  Research Triangle Park, NC:
     U.S. Environmental Protection Agency, Office of Air Quality  Planning  and
     Standards;  EPA  report  no.  EPA-600/4-83-027.  Available  from:  NTIS,
     Springfield, VA; PB83-239020.

Salas, L. J.;  Singh, H. B. (1982) Measurements of formaldehyde and acetaldehyde
     in the ambient air.   Environ. Sci. Techno!.: submitted.

Salo, A. E.; Whitz, S.;  MacPhee, R.  D. (1975) Determination of solvent  vapor
     concentrations by total combustion analysis:  a comparison of infrared
     with flame ionization detectors. Presented  at: the 68th annual meeting of
     the Air Pollution Control  Association;  June; Boston,  MA.  Pittsburgh,  PA:
     Air Pollution Control Association; paper no. 75-33.2.

Saltzman, B. E.; Wartburg, A.  F., Jr.  (1965)  Absorption tube  for removal  of
     interfering sulfur  dioxide  in  analysis of  atmospheric oxidant.  Anal.
     Chem.  37:  779-782.

Sawicki, E.; Hauser, T. R.; Stanley, T. W.; Elbert, W. (1961).  The 3-methyl-2-
     benzothiazolone  hydrazone  test: sensitive new methods for the detection,
     rapid  estimation, and determination  of  aliphatic aldehydes.   Anal.  Chem.
     33: 93-96.

Scaringelli, F.  P.;  Rosenberg, P. E.; Rehme, K.  A. (1970) Comparison of per-
     meation devices  and  nitrite ion as  standards for the colorimetric deter-
     mination of nitrogen dioxide.   Environ. Sci. Technol. 4: 924-929.

Schiff, H.   (1985).   York University, Toronto, Canada.  Personal communication:
     work in progress.

Schiff, H.  L;  Pepper, D.; Ridley,  B. A. (1979)  Tropospheric NO measurements
     up to 7 km.  JGR J.  Geophys.  Res, 84: 7895-7897.

Scott, W.E.; Stephens, E.R.; Hanst,  P.L.; Doerr,  R.C. (1957).  Further develop-
     ments  in  the  chemistry of the  atmosphere.   Proc. Amer. Petrol. Inst, 37
     (III): 171-183.

                                   4-100

-------
Seila, R. L.;  Lonneman,  W.  A.; Meeks,  S.  A.  (1976) Evaluation of polyv.inyl
     fluoride as a container  material  for air pollution samples.   J.  Environ.
     Sei. Health Part A 11:  121-130.

Sevcfk,  J.  (1975)  Detectors  in gas chromatography.  Prague,  Czechoslovakia:
     Elsevier Scientific Publishing Company.

Severs,  R.  K.  (1975) Simultaneous  total  oxidant and chemiluminescent ozone
     measurements in  ambient  air.  J.  Air  Pollut. Control Assoc. 25:  394-396.

Sexton,  F. W. ;  McElroy,  F.  F.; Michie,  R. M.,  Jr.; Thompson, V.  L.  (1982) A
     comparative evaluation  of seven  automated ambient non-methane  organic
     compound  analyzers.   Research Triangle  .Park,  NC:  U.S.   Environmental
     Protection Agency; EPA report  no. EPA-600/4-82-046, Available from: NTIS,
     Springfield, VA; PB82-230798.

Shanley. E.  S. (1951) Hydrogen peroxide. J. Chem. Educ.  28: 260-266.

Singh, H. B. (1980).   Guidance for  the collection and use of  ambient  hydrocar-
     bon species data in development of  ozone  control  strategies.   Research
     Triangle Park,  NC:   U.S. Environmental  Protection  Agency, Office of  Air
     Quality Planning and  Standards;  EPA  final  report  no.  EPA-450/4-80-008.
     Available from:   NTIS, Springfield, VA; PB80-202120.

Singh, H. B.; Hanst,  P. L. (1981) Peroxyacetyl  nitrate  (PAN)  in the unpolluted
     atmosphere: an   important reservoir  for  nitrogen  oxides. Geqphys.  Res.
     Lett. 8: 941-944.

Singh, H. B.; Salas,  L. J. (1983a)  Peroxyacetyl  nitrate  in the free troposphere.
     Nature (London)  302: 326-328.

Singh, H. B.; Salas,  1. J. (1983b)  Methodology  for  the  analysis of peroxyacetyl
     nitrate (PAN) in the unpolluted atmosphere. Atmos.  Environ. 17:  1507-1516.

Sleva, S. F.  (1965)   Determination  of  formaldehyde-chromotropiG  acid method.
     In: National Air Pollution Control Administration. Selected methods  for
     the measurement  of air pollutants. Durham,  NC:  U.S. Department of Health,
     Education, and Welfare,  Public Health Service;  publication no. 999-AP-ll.
     Available from:   NTIS, Springfield, VA; PB^-169677.

Smith, R. A.; Drummond, L (1979) Trace determination of carbonyl compounds  in air
     by  gas chromatography of their 2, 4-dinitrophenylhydrazones.  Analyst 104:
     875-877.

Spicer,  C. W.; Joseph, D. W.;  Sticksel, R.  R.; Ward, G.  F. (1979) Ozone  sources
     and transport in the  northeastern United States.   Environ.  Sci.  Techno!.
     13: 975-985.

Spicer,  C.  W.;  Holdren,  M.  W,; Keigley, G. W.  (1983) The ubiquity of peroxy-
     acetyl  nitrate   in the continental  boundary layer. Atmos. Environ. 17:
     1055-1058.
                                   4-101

-------
Stedman, D. H.; Kok, G.; Delumyea, R.; Alvord, H. H.  (1976) Redundant calibra-
     tion  of  nitric oxide, carbon monoxide,  nitrogen dioxide, and o2one air
     pollution monitors by chemical and gravimetric techniques.  In: Calibration
     in air monitoring, pp.  337-344.  Philadelphia,  PA:  Amer.  Soc.  for Testing
     and Materials. ASTM STP 598.
                                                                   'i *; - P
Steinberg, M.; Dietz, R. N. (1969) Sr-90 ozone generator for  sub-ppiri concentra-
     tion range—instruction manual.  Upton, Long  Island, NY:  Brookhaven  National
     Laboratory, Department of Applied Science.

Stephens, E. R. (1964) Absorptivities for infrared determination of peroxyacyl
     nitrates. Anal. Chem. 36: 928-929.

Stephens, E. R. (1967) The formation  of molecular oxygen by alkaline hydrolysis
     of peroxyacetyl nitrate. Atmos.  Environ. 1:  19-20.

Stephens,  E.  R. (1969) The formation, reactions,  and properties of peroxyacyl
     nitrates (PANs)  in photochemical air pollution.  In:  Pitts, J. N.-,  Jr.;
     Metcalf,  R.  L.,  eds.  Advances in environmental  sciences  v. 1. New  York,
     NY: Wiley-Interscience; pp. 119-146.

Stephens,  E.  R.;  Price,  M. A. (1973) Analysis of an  important air pollutant:
     peroxyacetyl  nitrate. J. Chem. Educ. 50: 351-354.

Stephens, E. R.; Hanst, P. L.; Doerr, R. C.;  Scott, W.  E.  (1956a) Reactions  of
     nitrogen dioxide and organic compounds in air. Ind. Eng.  Chem. 48:  1498-1504.

Stephens,  E.  R.;  Scott, W.  E.;  Hanst,  P.  L.; Doerr, R.  C.  (1956b) Recent
     developments  in  the  study of the organic  chemistry of the atmosphere.
     Presented at:  21st midyear  meeting  of the  meeting of the American Petro-
     leum  Institute's  Division of Refining;   May; Montreal, PQ,  Canada.  Proc.
     Am. Pet.  Inst. 36 (III): 288-297.

Stephens,  E.  R.;  Darley,  E.  F.;  Taylor,  0. C.;  Scott, W.  E.  (1961)  Photochem-
     ical  reaction  products  in air pollution.  Int.  J.  Air Water Pollut. 4:
     79-100.

Stephens, E. R.; Burleson, F. R.; Cardiff, E. A.  (1965) The production of pure
     peroxyacyl nitrates.  J. Air Pollut. Control  Assoc. 15: 87-89.

Stevens, R. K.; Hodgeson,  J.  A.  (1973)  Applications of chemiluminescent reac-
     tions to  the measurement of air pollutants. Anal. Chem.  45:  443A-446A,
     449A.

Stevens, R. K.; Clark, T.  A.; Decker, C. E.;  Ballard,  L. F. (1972a) Field per-
     formance  of  advanced monitors  for oxides  of nitrogen,   ozone, sulfur
     dioxide, carbon  monoxide,  methane and nonmethane hydrocarbons.  Presented
     at: 65th  annual  meeting of the  Air Pollution Control Association;  June;
     Miami, FL. Pittsburgh, PA: Air Pollution Control Association.

Stevens, R. K.;  Hodgeson, J. A.; Ballard, L. F.; Decker,  C.  E.  (1972b)  Ratio
     of sulfur  dioxide to total gaseous sulfur  compounds and ozone to total
     oxidants  in  the  Los  Angeles atmosphere  - an instrument evaluation study.
     In: Mamantov, G.; Shults,  W.  D.,  eds.  Determination of  air quality:
     proceedings  of the  ACS symposium on determination of air quality;  April
     1971; Los Angeles, CA. New York, NY: Plenum  Press; pp. 83-108.
                                   4-102

-------
Sweeney, M.  P.; Swartz, D.  J.;  Rost, G. A.;  MacPhee,  R.;  Chad, J.  (1964)
     Continuous measurements  of  oxides of nitrogen  in  auto exhaust.   J.  Air
     ..Rollut. Control Assoc. 14: 249-254.

Tanner,  R.  L.;  Meng, Z.  (1984)  Seasonal variations  in  ambient atmospheric
     levels  of  formaldehyde  and  acetaldehyde. Environ.  Sci.  Techno!. 18:
     723-726.

Taube, H.;  Bray, W.  C.  (1940) Chain reactions in  aqueous solutions containing
     ozone,  hydrogen peroxide and  acid. J.  Am.  Chem.  Soc. 62:  3357-3373.

Tokiwa, Y.; Twiss, S.; de Vera, E.  R.; Mueller, P. K. (1972) Atmospheric  ozone
     determination by amperometry and colorimetry.  In:  Mamantov,  G.;  Shults,
     W. D., eds.  Determination of  air quality: proceedings of  the  ACS symposium
     on  determination of air quality; April  1971; Los Angeles,  CA.   New York,
     NY: Plenum Press; pp. 109-130.

Toyama,  Y.;  Kobayashi,  J.  (1966)  Notes on an ozone generator and its calibra-
     tor. Pap.  Meteorol. Geophys. 17: 65-75.

Tuazon,  E.  C.;  Graham,  R.  A.; Winer, A.  M.;  Easton,  R.  R.; Pitts,  J.  N.,  Jr.
     Hanst,  P.  L.  (1978) A  kilometer pathlength Fourier-transform  infrared
     system  for the  study  of trace  pollutants in  ambient and synthetic atmos-
     pheres. Atmos.  Environ.  12: 865-875.

Tuazon,  E.  C.;  Winer,  A. M.;  Graham,  R.  A.; Pitts,  J.  N.,  Jr.  (1980) Atmos-
     pheric  measurements of  trace  pollutants by kilometer-pathlength FTIR
     spectroscopy. Adv.   Environ. Sci. Techno!. 10: 259-300.

Tuazon,  E.  C.;  Winer,  A. M.; Graham, R.  A.;  Pitts,  J. N.,  Jr.  (1981a) Atmos-
     pheric  measurements of  trace  pollutants: long path  Fourier-transform
     infrared spectroscopy.  Research Triangle Park,  NC:   U.S.  Environmental
     Protection Agency;  EPA report  no. EPA-600/3-81-026. Available  from:  NTIS,
     Springfield, VA; PB81-179848.

Tuazon,  E.  C.; Winer, A. M.;  Pitts,  J. N., Jr. (1981b) Trace pollutant concen-
     trations in  a multiday  smog episode in the  California South  Coast  Air
     Basin  by  long  path length  Fourier transform  infrared  spectroscopy.
     Environ. Sci. Techno!: 15: 1232-1237.

Tucker, A.  W.; Petersen, A. B.; Birnbaum, M.  (1973)  Fluorescence determination
     of atmospheric  NO and N02.  Appl. Opt.  12: 2036-2038.

Tucker,  A.  W.;  Birnbaum, M.;  Fincher,  C.  L. (1975)  Atmospheric N02 determi-
     nation  by  442-nm  laser  induced  fluorescence.   Appl. Opt.  14:  1418-1422.

U.S. Department of Health, Education, and Welfare. (1965) Selected methods for
     the measurement of air  pollutants.  Durham,  NC: National  Air Pollution
     Control Administration;  publication no.  PHS  999-AP-ll.  Available from:
     NTIS, Springfield,  VA; PB-169677.

U.S. Department of Health, Education, and Welfare. (1973) Manual  of analytical
     methods, Vol.  1.  Washington,  DC: U.S.  Government Printing Office. Cited
     in: National Academy of  Sciences.  (1981)  Formaldehyde  and  Other  Aldehydes.
     Washington, DC: National  Academy Press.

                                    4-103

-------
U.S. Environmental Protection Agency, (1975) Technical assistance document for
     the chemiluminescence measurement of nitrogen dioxide.  Researchfilriangle
     Park, NC:  U.S.  Environmental  Protection Agency; EPA report no. EPA-600/
     4-75-003. Available from: NTISS Springfield, VA.
U.S.
                                                            •   .-?•&
Environmental Protection Agency, (1976a) Ambient air monitoring references
and equivalent methods.  Code of Federal Regulations, Title 40,;Part 53.
                                                               - '"•' *' '\
U.S. Environmental  Protection Agency,  (1976b)  Nitrogen dioxide measurement
     principle  and  calibration procedure.  Notice  of final rulemaking. Fed.
     Regist. 41: 52686-52694, December 1.

U.S. Environmental  Protection Agency.  (1977a)  Air monitoring  strategy,for
     state  implementation  plans.  Research  Triangle  Park,  NC:  Office -of Air
     Quality Planning and Standards, Standing Air Monitoring Group;  EPA report
     no. EPA-450/2-77-010.  Available from:  NTIS, Springfield,  VA; PB-274014.

U.S. Environmental  Protection Agency.  (1977b) Quality assurance handbook for
     air pollution measurement systems, volume II: ambient air  specific methods.
     Research Triangle  Park, NC: U.S.  Environmental  Protection Agency; EPA
     report  no.  EPA-600/4-77-027.  Available  from:  NTIS,  Springfield, VA;
     PB-273518.

U.S. Environmental  Protection Agency.  (1978a) Air quality criteria  for ozone
     and other photochemical  oxidants. Research Triangle Park,  NC: U.S. Environ-
     mental Protection  Agency,  Environmental  Criteria and Assessment  Office;
     pp. 116-135;  EPA  report no.   EPA-600/8-78-004.  Available  from:  NTIS,
     Springfield, VA; PB80-124753.

U.S. Environmental  Protection Agency.  (1978b) Measurement  of volatile organic
     compounds.    Research  Triangle  Park,  NC:  U.S.  Environmental Protection
     Agency; EPA report no. EPA-450/2-78-041.  Available from:  NTIS, Springfield,
     VA.                                               -     .

U.S. Environmental  Protection Agency.  (1981)  Technical  assistance  document
     for the calibration and  operation of automated ambient non-methane organic
     compound analyzers.   Research  Triangle  Park,  NC:  U.S.  Environmental
     Protection Agency; EPA report  no. EPA-600/4-81-015. Available from: NTIS,
     Springfield, VA; PB82-147406.

U.S. Environmental  Protection Agency.  (1982) Air quality criteria for oxides
     of  nitrogen.   Research  Triangle  Park, NC:  Environmental  Criteria  and
     Assessment Office; EPA report  no. EPA-600/8-82-026. Available from: NTIS,
     Springfield, VA; PB83-131011.

Van de  Wiel,  H.  J.; Reijnders, H.  F.  R.;  Seifert,  R.;  Gies,  H.;  Lanting,  R.
     W.; Rudolf, W.  (1979) International comparative programme  on ozone refer-
     ence methods. Atmos. Environ.  13: 555-557.
                                                                            o
Vigroux, E.  (1952)  Spectrophotometry absorption  of ozone at 18°C below 3130A.
     Compt. Rend: 2529-2530.

Warren,  G.  J.;  Babcock, G. (1970)  Portable ethylene chemiluminescence ozone
     monitor. Rev. SciVInstrum. 41: 280-282.


                                    4-104

-------
Watanabe,  L;  Stephens, E,  R.  (1978) Reexamination of moisture  anomaly in
     analysis  of peroxyacetyl nitrate.  Environ.  Sci.  Techno!. 12:  222-223.

Wendel, G.  J.;  Stedman, D.  H.; Cantre11, C.  A. (1983)  Lumino!-based nitrogen
     dioxide detector. Anal. Chem. 55: 937-940.

Wendt;; (1975)  Personal  communication.  Information  regarding  California air
     quality data files.   El Monte, CA:  California  Air  Resources  Board.      i

West,  P.  W;  Sen, B. Z.  (1956)  Spectrophotometric determination of traces of
     formaldehyde.  I. Anal. Chem. 153:  177-183.

Westberg,  H.  H.; Rasmussen, R. A.; Holdren,  M. W,  (1974)  Gas  chromatographic
     analysis of ambient air for  light hydrocarbons using a chemically bonded
     stationary  phase.  Anal. Chem. 46:  1852-1854.

Westberg,  H. H.; Holdren,  M. W.; HiVI, H. H.  (1980) Analytical  methodology for
     the  identification and quantification of vapor phase organic pollutants.
     New  York,  NY:  Coordinating Research Council; CRC-APRAC project no. CAPA-
     11-71.

Williams,  D.  T.; Kolitz,  B. L. (1968)  Molecular correlation spectrometry.
     Appl, Opt.  7:  607-616.

Winer, A.  M.;  Peters, J.  M.; Smith,  J.  P.;  Pitts,  J. N.,  Jr.  (1974) Response
     of commercial  chemiluminescent NO-N02 analyzers to other nitrogen-containing
     compounds.  Environ.  Sci. Technol.  8: 1118-1121.

Yoshizumi,  K.;  Aoki,  K.;  Nouchi,  I.; Okita,  T.;  Kobayashi, T.;  Kamakura,  S.;
     Tajima, M.  (1984).   Measurement of the concentration in rainwater and of
     the  Henry's Law constant of hydrogen peroxide.  Atmos. Environ.  18:395-401.

Zaitsu, K.; Okhura, Y.  (1980) New  fluorogenic substrates  for  horseradish pero-
     xidase:  rapid  and sensitive  assays for hydrogen peroxide and peroxidase.
     Anal.  Biochem. 109: 109-113.

Zika,  R.  G.; Saltzman,  E.  S.  (1982) Interaction of  ozone  and  hydrogen peroxide
     in water:   implications for  analysis of H202 in air. Geophys. Res. Lett.
     9: 231-234.

Zocher,  H.; Kautsky,  H.  (1923) Uber  Lumineszenz bei chemischen  Reaktionen.
     [Luminescence  during  chemical reactions.] Naturwissenschaften 11:  194-199.
                                    4-105

-------

-------
  5.   CONCENTRATIONS OF OZONE AND OTHER PHOTOCHEMICAL OXIDANTS IN AMBIENT AIR
5.1  INTRODUCTION
     The data presented  in  this chapter on the  concentrations  of ozone and
other photochemical oxidants  in ambient air are intended to support and com-
plement information presented  in subsequent chapters on the effects of these
compounds.   Thus,  this chapter  describes potential exposures of human popula-
tions, crops and other vegetation, ecosystems, and nonbiological  materials  in
general terms for  the  entire nation and in specific terms for selected areas
of the country.  Since the  health and welfare effects of  ozone have been much
more  thoroughly  documented than  those  of  other related  oxidants,  primary
emphasis in this chapter is placed on  the  concentrations of ozone found in
ambient air.  Potential exposures are described by presenting data on peak and
average concentrations  nationwide and  on  seasonal  and  diurnal  patterns in
selected urban  and nonurban areas.  The  recurrence  on  consecutive days of
selected levels of ozone has been examined for  a few urban sites to aid in
understanding the  significance  of  health  effects documented in  subsequent
chapters.   Likewise, data have been included that portray representative urban
and rural  concentrations  by season and  by  time of day.  Spatial variations  in
ozone concentrations are briefly addressed since the effects on concentrations
of latitude, altitude, intracity variations, and indoor-outdoor gradients are
pertinent  to the assessment of  potential exposures of human  populations, and,
except for the  indoor-outdoor gradients,  of crops and  other vegetation and
ecosystems.
     Ozone  is the  only photochemical oxidant other than nitrogen  dioxide that
is routinely  monitored and  for which  a comprehensive aerometric data  base
exists.  Data for  peroxyacetyl  nitrate  (PAN) and its homologues and for hydro-
gen peroxide (H202) and  formic  acid  (HCOOH) have all been obtained  as part  of
special research investigations.  Consequently, no data on nationwide patterns
of occurrence are  available for these  non-ozone  oxidants;  nor are  extensive
data  available  on  the  correlations of  levels and patterns  of these oxidants
with  those of  ozone.   Data on  these oxidants are considerably more abundant
now,  however, than in  1978, when the previous criteria  document for ozone and
other  photochemical  oxidants was  published (U.S. Environmental  Protection
Agency, 1978),   Sections 5.6 and 5.7 summarize  the  available  data on these
other oxidants.
                                    5-1

-------
     The concentrations of  ozone  and related photochemical  oxidants observed

in ambient air are the net  result, as shown  in Chapter 3, of various combina-

tions of a variety of atmospheric processes,  including:


     1.   Local   photochemical   production  from  oxides  of  nitrogen  and
          reactive volatile organic compounds.

     2.   Atmospheric mixing after sunrise, such that ozone-rich air from
          above  the  nocturnal   inversion layer  is mixed  to  the surface,
          resulting in a  steady increase in  ozone during the morning and
          early afternoon.  This increase,  from ozone preserved aloft,  is
          independent of photochemistry, occurring even in the absence of
          precursors and photochemical  processes.

     3.   Transport  of  ozone produced  photochemically but  not locally.

     4.   Intrusion into the troposphere, even to ground level, of ozone-
          rich air from the stratospheric reservoir.

     5.   Formation of ozone photochemically  in the mid-troposphere, with
          subsequent intrusion into the boundary layer.

     6.   Chemical scavenging  in  the atmosphere of ozone and  other oxi-
          dants;  e.g., the reaction of ozone  with nitric oxide (NO) or the
          reaction of H202 with sulfur dioxide (S02).

     7.   Physical scavenging  in  the atmosphere of ozone and  other oxi-
          dants;   e.g.,  the  temperature-dependent decomposition  of PAN,
          the precipitation scavenging of H^O^, and  the  photolytic dis-
          sociation of ozone.

     8.   Combined  physical and   chemical   scavenging  processes  at  the
          surface of the earth; e.g., the deposition of ozone on reactive
          biological  or  nonbiological  surfaces,  such  as  vegetation,
          soils,  or certain polymers.


     These processes  include,  obviously, both manmade and natural  processes

and driving mechanisms.   The occurrence of high ozone concentrations is most

commonly associated  with recognized meteorological  conditions that involve

intense  sunlight and elevated  temperatures, and  the  variety  of processes

involved contribute to  strong  diurnal  cycles.   Peak concentrations have been
observed to occur,  however, at almost any time of day.   Ozone may be trans-

ported after  its formation  for distances up  to 1000 km or more (e.g.,  Hov et

a!., 1978; Wolff and Lioy,  1980).  As a result, high concentrations of ozone

and related oxidants  occur  not  only  near large  sources of precursors but also

in downwind nonurban  areas, and usually later  in  the day at these downwind
receptor sites.   Ozone,  and apparently  PAN,  as well,  can be transported at

                                    5-2

-------
night above the  surface  pollutant and nocturnal inversion layer (Chapter 3),
but during daylight hours can be transported considerable distances at or near
ground level (Coffey and Stasiuk; 1975a,b).
     The emphasis of  this  chapter is on the documentation of concentrations
rather than on  explanations of possible causes  of  observed  concentrations.
Probable causes and explanations are mentioned, however, where they are perti-
nent to the discussion.
     Most of the data presented in this chapter to characterize both nationwide
and  site-specific  ozone  concentrations  in  ambient  air were obtained after
1978.  Two  factors  influenced  the use of post-1978 data.  First, the current
Federal Reference Method for ozone,  chemiluminescence, and the equivalent UV
method were almost universally employed by 1979.  Second, EPA promulgated a UV
calibration method  for ozone in 1979.   Thus,  these data form a relatively
homogeneous set  for  purposes of intercomparison.  Because of the well-recog-
nized difficulties in converting from older data sets to the current reference
method, few pre-1979  aerometric  data  for ozone  are  presented  in  this  chapter,
and  then only  to demonstrate spatial variations  in concentrations that are
pertinent to exposure assessments.
5.2  TRENDS IN NATIONWIDE OZONE CONCENTRATIONS
     Whether ozone concentrations in ambient air are static, rising, or declining
over time must be determined from statistical tests using comparable aerometric
data for  a  number of years.  The  national  trend in ozone concentrations is
shown in Figure 5-1 for the 9-year period, 1975 through 1983 (U.S. Environmental
Protection  Agency,  1984a).  The  concentrations depicted  are  the  average
second-highest  1-hour  concentrations for  selected stations for  each year.
In this  context,  the second-highest  1-hour  value for  each  station is selected
from all daily maximum 1-hour values (n < 365) recorded per year  at that station.
Subsequently, this  statistic will  be referred to as the second-highest daily
maximum  1-hour  value or  simply the second-highest 1-hour value.  The 176
monitoring  stations  included in this analysis  reported at  least  50 percent  of
the possible hourly values  for  the ozone season  in at  least 7 of  these 9 years.
(The ozone  season varies at the respective sites from  4 to 12 months, depending
upon local  conditions.   The sampling period  is specified in State  Implementation
Plans.)
                                    5-3

-------
    0.18
    0.17
 i
 n.
 a
cc
UJ
o
z
o
o
UJ
z
o
N
o
0.16
    0.15
0.14
    0.13
    0.12
                                                     CA (27 Stations) _
O NAMS STATIONS (62)

T 35% CONFIDENCE
1. INTERVALS

O ALL STATIONS (176)

T 95% CONFIDENCE

i INTERVALS

A CALIFORNIA STATIONS (27)


i? ALL STATIONS EXCEPT

  CALIFORNIA (149)

 I      i      I      I
             1975   1976  1977  1978  1979   1980  1981  1982   1983



                                      YEAR


          Figure 5-1. National trend in composite average of the second highest

          value among daily maximum 1-hour ozone concentrations at selected

          groups of sites, 1975 through 1983.


          Source: U.S. Environmental Protection Agency (1984a).
                                  5-4

-------
     Ambient air monitoring  stations  in the nation are generally operated by
state and  local agencies, but among the stations  is a  small group of National
Air Monitoring Stations  (NAMS)  whose  data are reported directly to EPA.   The
trend line  for the  subset of the 62 NAMS  ozone stations, also shown in Figure
5-1, tracks  fairly  closely  the  trend for all 176 stations.  To permit a com-
parison  of aerometric data  for California, a high-oxidant area,  with the
nationwide trends data, data for California and data for all  states other than
California are also plotted  separately in Figure 5-1 and are compared below.
     For the entire 9-year period, 1975 through 1983, all subsets of monitoring
stations show a  decline  in  the composite second-highest daily maximum 1-hour
ozone concentration.  In 1979, a new,  more accurate ozone calibration procedure
was promulgated  by  the U.S.  Environmental Protection Agency (see Chapter 4).
Between  1979 and 1982, a small  decline  of 9  to 10 percent  in nationwide  ozone
concentrations occurred.   From 1982 to 1983, however, concentrations increased
by about 10 percent in California, by about 12 percent nationwide, and by about
14 percent outside  California.  Recently published data for 1984 from a somewhat
smaller  number of  stations  (163 sites),  not  depicted  in Figure 5-1, show a
decrease in second-highest 1-hour concentrations  nationwide, with ozone levels
in  1984  approximating those  recorded in  1981, which are shown  in  Figure 5-1
(U. S. Environmental  Protection Agency, 1986).
     Note  the  influence  of  ozone data  from  Calfornia.   The  California data
heavily  influence the trends data because of the  number of California monitor-
ing sites  represented in the data and because of the actual concentrations  of
ozone in California.
     Evaluation of  national  trends,  as well  as  local  or regional  trends,  in
reported concentrations  of ozone in ambient air over the past 5 to 10 years  is
complicated  by a number  of  factors, including:   (1)  the  change  in  calibration
procedure  recommended by EPA in 1978  and  promulgated in  1979  (see  Chapter 4);
(2) the  possible effects on aerometric data  of  quality  assurance  procedures
instituted by EPA in  1979; (3) the influence of diverse  regional meteorological
conditions;  and (4) changes  in precursor  emissions.
     How much of the  observed decline in  ozone concentrations from 1975 through
1982 is  attributable  to  the  1979 promulgation of  the ultraviolet (UV) calibra-
tion method as the Federal  Reference  Method is  uncertain.  To  determine  that,
the monitoring practices at  each of the 176 sites would  have to be examined  in
detail  to  find out the  calibration methods used,  the  date when the UV method
                                     5-5

-------
 was  adopted for this  site,  and whether states applied correction factors to
 prior data (see, e.g., Walker,  1985;  Hoggan,  1986; Walker, 1986).   Such an
 examination would be  necessary  because  not all  monitoring sites switched to
 the  use of the UV method  simultaneously.   For  example,  the state of  California
 (in  EPA  Region  IX)  had already  been  using the  UV method  before  it was
 promulgated in February 1979.   In  addition,  other states, in other regions,
 may  have used the boric acid-potassium iodide (BAKI) method before,  after, or
 both before and after promulgation of the UV method, since the BAKI procedure
 was  allowed by EPA  as an  interim method for 18 months following the 1979 UV
•promulgation (see Chapter  4).   Likewise,  other states  used gas-phase titration
 prior' to 1979 but either BAKI or UV procedures following the UV promulgation.
 In addition,  random errors associated  with  individual  operator practices
 occur.   Thus, the relationships among these three methods, even if monitoring
 practices at individual sites  were  known,  are complex and would preclude the
 simple application  of a  single correction factor to  these  trend data (see
 Chapter 4).
      Hunt and Curran (1982) have noted that only  EPA Region IX showed improve-
 ment in ozone air quality between 1979 and 1981 but, in contrast to other EPA
 regions,  showed no  improvement in that period versus the 1975 to 1979 period.
 Since California, which dominates Region IX,  changed its  calibration method  in
 1975, the decrease in ozone concentrations  seen  in California  from 1979  through
 19S2 (Figure 5-1) cannot be attributed to the introduction of the UV calibra-
 tion method.  The shape of the trend  line  for California is quite  similar to
 that for the  rest of the  nation over the  4-year  period  (1980 through 1983)
 following the promulgation  of  the  UV  method.   From 1982  to 1983, an increase
 in ozone concentrations occurred  nationwide,  with the percentage increase in
 California roughly  paralleling that for the  rest of the  nation.   Increased
 precursor emissions and meteorological conditions conducive to oxidant forma-
 tion appear to be the most likely causes of  the  increase in 1983 in the com-
 posite average of the second-highest daily maximum 1-hour ozone concentrations.
 An examination of emissions  data,  however, indicates that NO  emissions did
                                                              ){
 not  change significantly from 1982 to 1983 and that VOC emissions rose only 3
 percent in that period (U;S. Environmental Protection Agency, 1984b; also see
 Chapter 3).  Therefore, precursor emissions are  not thought to account for all
 of the 12  percent  increase in 0- from 1982 to 1983 nationwide (U.S.  Environ-
 mental Protection Agency,  1984a).
                                     5-6

-------
     The influence of  meteorological  conditions on ozone concentrations  has
been explored for 10 cities using an experimental index based on meteorological
parameters conducive to ozone formation (U.S. Environmental  Protection Agency,
1984a).  The index suggests that the potential for ozone formation was greater
in 1983  than  in 1981 and 1982;  however,  the index also indicates that even
greater  potential  for  ozone formation existed  in  1980  than  in  1983  and that
conditions in 1979 might  be comparable to  those in  1981 and  1982.  Thus,  this
simple index may  indicate the direction but not the magnitude of the effects
of meteorological factors on ozone formation.
     From a study of climate in the upper Midwest during the summer of 1983,
Wendland et al.  (1984) have reported that temperatures, which correlate reason-
ably well with  insolation,  were generally higher  in  the  summer of 1983 than
the 1950~through-1980  norm  for the 12-state region;  and  that cooling-degre.e
days over about one-third of the  region were  50 percent  higher than normal.
While not quantitative or conclusive, these studies do  suggest that meteorolo-
gical conditions  in  1983 contributed to the  increase in  the second-highest
value among daily maximum 1-hour ozone concentrations.
     Trends in  the composite average second-highest daily maximum 1-hour ozone
values  in  the  10 EPA  regions are shown in  Figure 5-2 (U.S. Environmental
Protection Agency, 1984a).  The  use of data beginning with 1979 avoids some of
the potential effects of changes in calibration  and quality assurance procedures
mentioned earlier.   Nine of the 10 regions show a  7 to  15 percent decrease in
this statistic  from  the 1979-1980 period  to  the 1981-1982 period.   The same
nine regions  showed  increases of 6 to 16  percent from the 1981-1982  period to
1983,  demonstrating  the  pervasiveness  of the  trend.   Only  in  the  Pacific
Northwest, Region X, was there  an  opposite trend:  +6 percent in 1981-1982 and
-9 percent in 1983.
5.3  OVERVIEW OF OZONE CONCENTRATIONS IN URBAN AREAS
     An overview of  nationwide urban ozone concentrations for 1981 is,provided
in Figures 5-3 and 5-4, which depict graphically the average daylight concentra-
tions.  Figure 5-3 shows data for  spring and summer months, the months that make
up the smog season in most though not all  areas  of the nation;  and Figure 5-4
shows  daylight  concentrations  during the  fall  and winter months.  The daylight
period of 6:00 a.m.  to 8:00 p.m.  includes the hours of  greatest human activity
                                     5-7

-------
  0,24-1
  0.20
E
D.
O.

Z
O
Z
ui
o
z
o
u
UJ
Z
o
N
o
0.16-
0,12-
0.08-
  0.04-
   0.00

EPA REGION3


NO. OF SITES
                  1979-1980 COMPOSITE AVERAGE
                  1981 -1982 COMPOSITE AVERAGE
                  1983 COMPOSITE AVERAGE
                   II


                   19
III


26
IV


14
V


40
VI


15
VII    VIII


 8     7
IX


36
X


5
           Figure 5-2. Comparison of the 1979-1980, 1981-1982, and 1983
           composite average of the second highest daily maximum 1-hour ozone
           concentrations across EPA Regions.

           aWote:  Regions are composed of the following states:
                      I  CT, MA, ME, NH, RI.VT
                     II  NJ, NY, PR, VI
                     ill  DE, MD, PA, VA, WV
                     IV  AL, FL, GA. KY, MS, NC, SC, TN
                     V  IL, IN, Ml, MM, OH,WI
                     VI  IA, KS, MO, NE
                    VII  AR, LA, NM, OK, TX
                   VIII  CO, MT, ND, SD. UT, WY
                     IX  AZ, CA. HI. NV
                     X  AK, ID, OR, WA

           Source: U.S. Environmental Protection Agency (1984a).
                                     5-8

-------
en
10
                                                                                              .14 - .16 PPM
                                                                                              .16 - .18 PPM
              Figure 5-3. Average daylight (6:00 a.m. to 8:00 p.m.) concentrations of ozone in the second
              and third quarters (April through September), 1981.
01 Apr 1983
              Source: SAROAD (19S5d). Derived by G. Duggan, OAQPS.

-------
tn
i
                                                                                             .16 - .18 PPM
              Figure 5-4. Average daylight (6:00 a.m. to 8:00 p.m.) concentrations of ozone in the first and

              fourth quarters (January through March and October through December), 1981.              01 Apr 1983
             Source: SAROAD (1985d). Derived by G. Duggan, OAQPS,

-------
outdoors; the hours when exposure of vegetation and ecosystems would be expected
to have  the greatest consequences (stomata are open in daylight and photosyn-
thesis is taking place; see Chapter 6); and the hours of greatest local forma-
tion of  ozone  and  other oxidants via  photochemistry  in  the atmosphere (see
Chapter 3).   The average  concentrations  during the spring  and summer  months
(second  and third  quarters  of the year) are clustered mainly in the 0.04 to
0.06 ppm  range.  Averages for the winter and  fall  months (first and fourth
quarters) are clustered mainly in the 0.02 to 0.04 ppm range.
     The  stations  used  in Figures 5-3  and 5-4  reported at  least 75 percent of
the possible 1-hour values per quarter.  Some stations,  however,  monitor ozone
only during the months when the potential for photochemical ozone formation is
significant in those localities.   Also, certain areas of the United States are
not monitored  routinely for ozone because of the lack of emission sources or
transport events and  thus the low potential  for significant ozone or oxidant
concentrations.  The Great  Basin and the Great Plains,  for example, are such
areas.
     Figure 5-5  shows  the nationwide frequency distributions of the first-,
second-,  and  third-highest 1-hour 0-  concentrations  at  predominantly  urban
stations  aggregated for 1979, 1980, and 1981, as well as the highest 1-hour 03
concentration at sites  of the National Air Pollution Background Network (NAPBN)
(see  Section 5.4.1) aggregated .for  the same 3 years.  Only data collected by
the Federal  Reference  Method (chemiluminescence) or the equivalent UV method
(see Chapter 4) have been used in this analysis.  A "valid site"  is one report-
ing at  least  75 percent  of the  8760  possible  1-hour  values in a year. There
were  282 such  sites in 1979,  266 in  1980, and  358  in  1981 (U.S. Environmental
Protection Agency,  1980,  1981, 1982a).   As shown by  Figure 5-5, 50  percent of
the second-highest 1-hour values from non-NAPBN sites in this 3-year  period
were  0.12 ppm  or less  and about  10  percent were  equal to  or greater  than  0.20
ppm.  At the  NAPBN sites, the collective 3-year  distribution (1979 through
1981) is such  that about 60 percent of the values are less than 0.10 ppm and
fewer than 20 percent are higher than  0.12 ppm.
     Table 5-1  lists  the second-highest 1-hour 0-  values reported for 1981
through  1983  for the  80 most populous Standard Metropolitan Statistical Areas
(SMSAs),  grouped by population (see Table 5-1  for a definition of the  "second-
highest"  values).  Collectively  these  SMSAs account for about 53 to 54"percent
of the 1981 United States population of 229.3  million.  The significant obser-
vation to be drawn from this table of  second-highest values is that the lowest
                                    5-11

-------
             99.999.8
                      99  98 " iS   90
70  60  SO 40 30  20
10
2  1 0.5 02 0.1 0.05 0.01
   0.45
   0.40
   OJi
o.
a 030
 »

O
oc


01
o
z
O
O
IU
Z
O
N
O
   0.2S
   0.20
   0.15
   0.10
   0.05
                       I   I    i    i     I    i   I  I  I   !   I     I    I    I   I
                     	 HIGHEST

                      — — 2nd-HIGHEST

                     	3rd HIGHEST

                     ....... HIGHEST, NAPBN SITES
            I  i  I    i  i  I    i    i    _L   i  I  i   i   i   i    I   I    II
                                                                                I  I
      0.01  0.05 0.1 0.2  0.5 1  2    5   10   20  30  40  50  60  70  80    90   95   98  99    99.8 99.9


                STATIONS WITH PEAK 1-hour CONCENTRATIONS < SELECTED VALUE, percent
                                                                                        99.99
      Figure 5-5. Distributions of the three highest 1 -hour ozone concentrations at valid sites
      (906 station-years) aggregated for 3 years (1979,1980, and 1981) and the highest
      ozone concentrations at NAPBN sites aggregated for those years (24 station-years).
      Source: U.S. Environmental Protection Agency (1980, 1981, 1982).

-------
      TABLE  5-1.   SECOND-HIGHEST 1-hr OZONE CONCENTRATIONS*>b REPORTED FOR
 STANDARD  METROPOLITAN  STATISTICAL AREAS HAVING POPULATIONS >0.5 MILLION,
                            1981 THROUGH 1983
                                              Ozone concentration,  ppm
Standard Metropolitan Statistical  Area   1981
1982
1983
Population >2 million

New York, NY - NJ                        0.18
Los Angeles - Long Beach, CA             0.35
Chicago, IL                              0.14
Philadelphia, PA - NJ                    0.17
Detroit, MI                              0.15
San Francisco - Oakland, CA              0.14
Washington, DC - MD - VA                 0.15
Dallas - Fort Worth, TX                  0.15
Houston, TX                              0.23
Boston, MA                               0.13
Nassau - Suffolk, NY                     0.14
St. Louis, MO - IL                       0.15
Pittsburgh, PA                           0.16
Baltimore, MD                            0.17
Minneapolis - St. Paul, MN - WI          0.10
Atlanta, GA                              0.14

     Summary statistics:
       Minimum 1-hour value              0.10
       Median 1-hour value               0.15
       Maximum 1-hour value              0.35

Population 1 to < 2 million

Newark, NJ                               0.14
Anaheim - Santa Ana - Garden Grove, CA   0.31
Cleveland, OH         :                   0.12
San Diego, CA                            0.24
Miami,  FL                                0.14
Denver  - Boulder, CO                     0.13
Seattle - Everett, WA                    0.12
Tampa - St.  Petersburg,  FL               0.12
Riverside -  San Bernardino - Ontario, CA 0.34
Phoenix, AZ                              0.16
Cincinnati,  OH -  KY - IN                 0.13
Milwaukee, WI                            0.17
Kansas  City, MO - KS                     0.12
San Jose, CA                            0.14
Buffalo, NY                              0.12C
Portland, OR - WA                        0.15
New Orleans,  LA           '               0.12
Indianapolis,  IN                         0.13
Columbus, OH             .              0.11
0.17
0.32
0.12
0.18
0.16
0.14
0.15
0.17
0.21
0.16°
0.13
0.16
0.14
0.14
0.10
0.14
0.10
0.15
0.32
0.17
0.18C
0.13
0.21
0.14
0.14
0.09
0.12
0.32
0.12
0.13
0.13
0.10
0.14
0.11
0.12
0.17
0.12
0.13
0.19
0.37
0.17
0.10
0.17
0.17
0.17
0.16
0.28
0.18
0.17
0.18
0.14
0.19
0.13
0.17
0.10
0.17
0.37
0.25
0.28
0.15
0.20
0.12
0.14
0.10
0.14
0.34
0.16
0.15
0.18
0.13
0.16
0.12
0.12
0.12
0.14
0.12
                                  5-13

-------
TABLE 5-1 (continued).  SECOND-HIGHEST 1-hr OZONE CONCENTRATIONS3'b REPORTED
FOR STANDARD METROPOLITAN STATISTICAL AREAS HAVING POPULATIONS >0.5 MILLION,
                             1981 THROUGH 1983
Standard Metropolitan Statistical Area
San Juan, PR
San Antonio, TX
Fort Lauderdale - Hollywood, FL
Sacramento, CA
Summary statistics:
Minimum 1-hour value
Median 1-hour value
Maximum 1-hour value
Population 0.5 to < 1 million
Rochester, NY
Salt Lake City - Ogden, UT
Providence - Harwick - Pawtucket, RI - MA
Memphis, TN - AR - MS
Louisville, KY - IN
Nashville - Davidson, TN
Birmingham, AL
Oklahoma City, OK
Dayton, OH
Greensboro - Winston-Salem - High Point, NC
Norfolk - Virginia Beach - Portsmouth, VA - NC
Albany - Schenectady - Troy, NY ,
Toledo, OH - MI
Honolulu, HI
Jacksonville, FL
Hartford, CT
Orlando, FL
Tulsa, OK
Akron, OH
Gary - Hammond - East Chicago, IN
Syracuse, NY
Northeast Pennsylvania
Charlotte - Gastonia, NC
Allentown - Bethlehem - Easton, PA - NJ
Richmond, VA
Grand Rapids, MI
New Brunswick - Perth Amboy - Sayreville, NJ
West Palm Beach - Boca Raton, FL
Omaha, NE - IA
Greenville - Spartanburg, SC
Jersey City, NJ
Austin, TX
Ozone
1981
0.07
0.12
0.11
0.17

0.07
0.13
0.34

0.12
0.16
0.15
0.14
0.14
0.13
0.16
0.11
0.13
0.11
0.12
0.12
0.13
0.05
0.11
0.15
0.10
0.15
0.24e
0.14
0.11
0.10
0.12
0.12
0.12
0.11
0.13
0.09
0.08
0.11
0.14
0.12
concentration,
1982
0.04C
0.14
0.09
0.16

0.04
0.13
0.32

0.11
0.14
0.15
0.13
0.14
0.11
0.15 '
O.llc
0.16
0.11
0.11
0.12
0.13
0.07
0.12
0.17r
0.10C
0.13
0.14
0.13
0.12
0.16
0.12
0.15
0.12
0.11
0.16
0.09
0.09
0.11
0.14
0.11
ppm
1983
NDd
0.12
0.10
0.15

0.10
0.14
0.34

0.12
0.14
0.15
0.15
0.16
0.12
0.15
0.11
0.13
0.12
0.13
0.12
0.13
0.06
0.14
0.19
0.11
0.13
0.13
0.17
0.08
0.13
0.15
0.14
0.14
0.14
0.19
0.09
0.09
0.11
0.16
0.12
                                  5-14

-------
  TABLE 5-1  (continued).   SECOND-HIGHEST 1-hr OZONE CONCENTRATIONS9>b  REPORTED
  FOR STANDARD METROPOLITAN STATISTICAL AREAS HAVING  POPULATIONS >0.5  MILLION,
                                 1981 THROUGH 1983
   Standard  Metropolitan  Statistical  Area
                                                   Ozone  concentration, ppm
                                             1981
1982
1983
Youngstown ~ Warren, OH
Tucson, AZ
Raleigh - Durham, NC
Springfield - Chicopee - Holyoke, MA - CT
Oxnard - Simi Valley - Ventura, CA
Wilmington, DE - NJ - MD
Flint, MI
Fresno, CA
Long Branch - Asbury Park, NJ
0.13
0.10
0.12
0.16
0.20
0.12C
0.11
0.11
ND
0.11
0.12
0.09
0.16
0.22
0.16
0.11
0.16
ND
0.11
0.11
0.13
0.19
0.21
0.18
0.11
0.16
ND
Summary statistics:
Minimum 1-hour value
Median 1-hour value
Maximum 1-hour value

0.05
0.12
0.24

0.07
0.12
0.22

0.06
0.13
0.21
d
The maximum 1-hour value  for each day  forms the data  set  from which  the  highest
and "second-highest"  concentrations are  determined.   Thus,  the  "second-highest"
1-hour concentration  occurs on  a different day from the highest and  therefore
is not necessarily the  absolute second-highest concentration for the monitoring
period.
These values  permit a comparison of potential exposures and are not  necessarily
equivalent to the "design value" used  for control  strategy  development.

'Less than 50% of days in  ozone  season.
ND = no  data.
^Questionable  data; third-highest value for 1981 was 0.125 ppm.
Source:  U.S. Environmental Protection Agency (1984a).
                                     5-15

-------
median concentrations in 1981, 1982, and 1983, for SMSAs having populations of
0.5 to 1  million,  are 0.12, 0.12,  and  0,13 ppm, values that equal or exceed
the current national ambient air quality standard for ozone.   With the possible
exception of the concentration reported for Akron,  Ohio, in 1981, which appears
to be questionable  (see footnote, Table 5-1), the highest of all the second-
highest 1-hour  concentrations  in  these  80 SMSAs occurred  in California in  all
3 years.
5.4  OVERVIEW OF OZONE CONCENTRATIONS IN NONURBAN AREAS
     As mentioned in the preceding section, very few ozone monitoring stations
are located  in  nonurban areas.   Consequently, the aerometric  data base for
nonurban areas  is  not comparable to that for urban areas.  The nonurban data
presented in this  section were obtained from two special-purpose  monitoring
networks that were  designed  to measure  ozone  concentrations at sites specifi-
cally selected to represent a variety of pristine, rural, or suburban environ-
ments.  These sites do not all represent areas totally unaffected by manmade
ozone or its precursors, as shown by the fact that some data records contain a
significant  number  of high values that are best explained as  resulting from
the transport of ozone or its precursors  from upwind  urban areas.  The data
given  here  are  intended to  show  an  overview  of nonurban concentrations in
areas with  relatively infrequent  urban  influences.  Additional data on  speci-
fic rural areas are presented in Sections 5.5.1 and 5.5.2.

5.4.1  National Air Pollution Background Network (NAPBN)
     The NAPBN  consists of eight stations located in  eight National Forests
(NF) across  the country (Figure 5-6).  The first three stations began opera-
tion in  1976 (Green Mountain NF,  Vermont;  Kisatchie NF,  Louisiana; and  Custer
NF, Montana); the  second three  in 1978  (Chequamegon NF,  Wisconsin; Mark Twain
NF, Missouri; and Croatan NF, North Carolina); and the last two in 1979 (Apache
NF, Arizona; and Ochoco  NF, Oregon).  Yearly summaries of ozone concentrations
through  1980 are shown  in Table  5-2  for  the  three sites established first
(Evans et a!.,  1983).   The principal  points of interest  in these  summary sta-
tistics  are the range  and the  arithmetic  mean  of the ozone concentrations
measured at  these  National Forest sites.  The arithmetic mean concentrations
for the  three  sites ranged from 0.027 ± 0.015 ppm at the Kisatchie NF site  in
                                    5-16

-------
                                                   CROATAN IMF
Figure 5-6, Locations of the eight national forest (IMF) stations
constituting the National Air Pollution Background Network
(NAPBN).

Source: Evans et al. (1983),
                            5-17

-------
                  TABLE 5-2.  ANNUAL OZONE SUMMARY STATISTICS FOR THREE SITES OF THE
                              NATIONAL AIR POLLUTION BACKGROUND NETWORK
Site
Kisatchie NFS LA




en
i
co Custer NF, MT




Green Mt. NF, VT




Year
1976
1977
1978
1979
1980


1976
1977
1978
1979
1980
1976
. 1977
1978
1979
1980
No. 1-hr
meas.
3448
6793
5636
6993
4438


275
7603
7674
8488
7754
1058
6483
3671
6423
8574
% of
possible
1-hr meas.
39,4
77.5
64.3
79.8
50.7


3.1
86.8
87.6
96.9
88.5
12.1
74.0
41.9
73.3
97.9
Concn. , ppm
Concn.
M1n.
NDa
ND
ND
ND
ND


0.020
ND
ND
ND
ND
ND
ND
ND
ND
ND
, Ppm
Max.
0.125
0.135
0.125
0.100
0.105


0.060
0.080
0.075
0.070
0.070
0.060
0.145
0.105
0.105
0.115
Arith.
mean
0.032
0.033
0.034
0.027
0.028


0.039
0.040
0.030
0.032
0.037
0.029
0.038
0.029
0.032
0.032
Arith.
std. dev.
0.021
0.023
0.021
0.015
0.016


0.008
0.011
0.017
0.012
0.012
0.011
0.021
0.018
0.017
0.017
Concn. , ppm
Geom.
mean
0.024
0.025
0.027
0.023
0.023


0.038
0.039
0.023
0.029
0.035
0.026
0.031
0.024
0.027
0.027
Geom.
std. dev.
2.19
2.25
2.14
1.92
1.94


1.22
1.37
2.14
1.59 .
1.41
1.76
2.00
2.01
1.86
1.90
aND = not detectable.

Source:   Evans et al.  (1983)

-------
1979 to 0.040 ± 0.011 ppm at the Custer NF site in 1977.  The arithmetic mean
concentration across all years  and all three sites was 0.033 ppm.  Fluctua-
tions in the observed  concentrations from year-to-year and site-to-site are
demonstrated by the  range  of concentrations measured and by the size of the
standard deviations  as  well.   The  lowest concentrations seen were below the
limits of detection of the chemiluminescence monitor employed,  but the highest
concentrations observed at  the  Kisatchie NF and Green Mt.  NF sites were both
above the present  ozone standard of 0.12 ppm.  In  1979,  only one excursion
over 0.12 ppm  03  was recorded at an NAPBN site.   That excursion was recorded
at the Mark Twain  NF site  in Missouri  (Evans et al., 1983; Lefohn, 1984).   In
1980, seven 1-hour excursions over 0.12 ppm were reported for Croatan NF, North
Carolina (Lefohn,  1984).  None of the NAPBN sites  recorded an ozone concentra-
tion greater than 0.12 ppm in 1981 (Evans et al.,  1983).
     These summary statistics show somewhat higher mean concentrations,  lower
maximum concentrations, and  lower,  standard deviations in data obtained at the
Custer NF site  than  at the other two, which may indicate that meteorological
conditions are less variable at that site or that the site is much less affec-
ted, if not altogether unaffected,  by manmade ozone or its precursors.
     During a  6-day period  in  1979,  the NAPBN site in the  Mark Twain NF,
Missouri, showed  ozone concentrations  well  in excess  of typical values.
Table 5-3 shows the  peak 1-hour value  for each of the 6 days.   A 1-hour  value
of 0.125 ppm,  the  maximum  observed at  any NAPBN site in 1979, was  measured  at
that site on  July 21,  1979.  Evans et al.  (1983) calculated the trajectories
of air masses  reaching the site during  the  6-day  period of July 18 through
July 23, 1979.   They ascribed the  unusually high values, including the  peak
value on the 21st, to pollutants picked up as the trajectory passed over urban
areas in the  Ohio River Valley and the Great Lakes region.   Figure 5-7 shows
the trajectories for the air parcels reaching the Mark Twain NF site at midnight
(0000), 8 a.m.  (0800), noon (1200), and 6 p.m. (1800) on July  21, 1979.  On
July 23, clouds and rain spread over the region and the airflow trajectories
shifted to  the east and south,  reducing both  the  quantities of transported
precursors and  the potential for photochemical ozone generation.
     More recent and more comprehensive data from the NAPBN sites are presented
in Table 5-4  and  in  Figures 5-8A and 5-8B.   Table 5-4 presents  the percentile
distributions of ozone  concentrations at all eight of the sites, aggregated by
quarter across  several years.
                                    5-19

-------
        TABLE 5-3.   CONCENTRATIONS OF OZONE DURING 6-day PERIOD OF HIGH
      VALUES AT NAPBN SITE IN MARK.TWAIN NATIONAL FOREST, MISSOURI, 1979
Date
July 18
July 19
July 20
July 21
July 22
July 23
1-hr maximum
03 concentration, ppm
0.080
0.100
0.115
0.125
0.120
0.050
Source:  Evans et al. (1983)

     As the  data in Table 5-4 show,  the  arithmetic mean concentrations of
ozone, for the years of data averaged, are  generally  higher in the second
quarter of the year (April, May,  June) than at other times at the NAPBN sites.
Although a few excursions of the 1-hour  concentration  above 0.12 ppm were
recorded at  some of these  sites,  as discussed above (Evans  et  al.,  1983;
Lefohn, 1984), the  distribution  given in Table 5-4 clearly indicates that 99
percent of the  1-hour  ozone concentrations measured at these sites are well
below the present ozone standard.  This is true even at those sites thought to
be  influenced  by transport  of ozone (Green Mt. NF) or  demonstrated  to be
influenced by  transport  (Mark Twain NF, Evans  et  al.,  1983).   As shown in
Table  5-4, the  highest 99th percent!le value was  0.093 ppm,  reached at both
the Green Mt,  NF,  VT,  site  (second  quarter) and the Mark Twain  NF, MO, site
(third quarter).  The maximum 1-hour ozone concentrations at these sites ranged
from  0.050 ppm at  Custer NF, MT (in the fourth quarter) to 0.155 ppm at Mark
Twain  NF, MO (in the third quarter).  The second-highest 1-hour concentration
among  maximum  daily 1-hour  values  ranged from 0.050 ppm at  Custer NF, MT
(fourth quarter) to 0.150 ppm at Mark Twain NF, MO (third quarter).
     Five of the NAPBN monitoring stations  (Apache NF, AZ; Mark  Twain  NF, MO;
Custer NF, MT;  Croatan NF, NC;  Ochoco  NF,  OR)  reported sufficient data for
1979  through 1983  to support an examination of their  second-highest 1-hour
                                    5-20

-------
MINNEAPOLIS
                                       1200 1
               MILWAUKEE

     DES MOINES
        > KANSAS CITY
           ST. LOUIS


    MARK TWAIN N.F.
     INCINNATI
     ^
LOUISVILLE
  Figure 5-7. Trajectory analysis plots for the
  NAPBN site at Mark Twain National Forest, MO,
  July 21, 1979 (distance between bars represents
  12hr).

  Source: Evans et al. (1983).
                       5-21

-------
                          TABLE 5-4.  PERCENTILE DISTRIBUTIONS OF OZONE CONCENTRATIONS
                             AT SITES OF NATIONAL AIR POLLUTION BACKGROUND NETWORK,
                                   AGGREGATED BY QUARTER ACROSS SEVERAL YEARS
                                                      (ppb)
en
 i
Site, yr» No. 1-hr
and qtr meas.
Apache, AZ




Ki satchi e ,




Mark Twain,




Custer, MT





Croatan, NC




(1980-1983)
1
2
3
4
LA (1977-1980, 1982)
1
2
3
4
MO (1979-1983)
1
2
3
4
(1979-1983)
1
2
3
4

(1978-1983)
1 • • ••••••
2
3
4

7587
7971
8407
8537

7093
6333
5462
6872

9484
10294
9155
8624

6675
8646
8751
9956


10640
11491
8389
12036
Percentile
10%

32
40
28
27

16
14
6
8

14
29
23
11

22
30
27
18


10
12
5
4
50%

39
50
40
35

32
37
22
22

32
46
42
27

31
40
39
28


26
36
23
18
80%

45
56
46
39

43
55
42
37

41
60
55
38

37
46
47
32
i

37
51
43
29
90%

49
59
51
40

53
65
53
48

47
68
65
46

40
50
50
35


44
60
53
37
95%

50
64
54
44

60
72
60
56

52
73
75
57

43
53
53
36


50
66
60
41
2nd
highest
W%

54
68
60
46

73
85
70
77

63
85
93
72

50
59
58
40


61
79
72
54
1-hr

65
90
85
55

120
115
110
90

80
115
150
85

65
80
85
50


85
110
• 80
85
Max
1-hr

75
90
90
55

125
135
110
105

85
115
155
100

65
80
85
50


95
150
95
85
Arith.
mean

39.3
49.4
39.5
34.2

32.4
38.7
25.8
25.7

31.0
47.4
43.2
28.6

30.9
40.3
38.6
26.9


27.4
36.1
27.0
19.8

-------
                     TABLE 5-4 (continued).   PERCENTILE DISTRIBUTIONS OF OZONE CONCENTRATIONS
                              AT SITES OF NATIONAL AIR POLLUTION BACKGROUND NETWORK,
                                    AGGREGATED BY QUARTER ACROSS SEVERAL YEARS
                                                       (ppb)
(Jl
ro
CO
Site, yr,
and qtr
Ochoco, OR




Green Mt. ,




Chequamegon




(1980-1983)
1
2
3
4
VT (1977-1981)
1
2
3
4
, WI (1979-1981)
1
2
3
4
No. 1-hr
meas.

7236
7861
8041
7467

7387
7752
8636
8712

5548
6085
4577
5909
Percent! le
10%

26
28
26
22

21
17
5
11

22
32
15
15
50%

35
38
38
31

31
43
26
25

36
45
30
25
80%

37
45
46
32

41
58
44
33

45
58
45
30
90%

40
46
49
36

45
69
56
37

51
68
53
32
95%

42
51
53
39

48
81
65
41

56
75
58
35
2nd
highest
99%

43
54
59
41

61
93
86
54

65
88
75
43
1-hr

55
60
75
55

120
140
110
75

80
110
90
55
Max
1-hr

55
75
80
60

135
145
115
85

80
115
95
60
Arith.
mean

33.0
37.1
37.8
29.2

32.8
42.2
29.2
24.8

36.1
48.3
33.0
23.7
           Data are weighted by the number of 1-hr concentrations measured.  Since data records were
           inadequate for some sites for some years, the years of data presented differ from site
           to site.   The percentile distributions were derived from all 1-hr values for each quarter
           and all  years listed.   The maximum 1-hr and second-highest 1-hr concentrations represent
           the single value for each out of the entire data record for respective sites.
          b                                              •' '•"'"
           These are the two highest values in the "daily maximum 1-hour" data set.

          Source:   Derived frojji Ivans (1985).  ,,'.-  ...: •_., .-.,.', =.-'  ._'.. v       ., :1 ;;', -^.

-------
a
a
 %
Z
o

§

I

ui
o
z
o
u
ui
z
o
 a.
 a.
  *,


 O


 §



 1
 ui
 U
 Z
 o
 u
 UI
 z
 o
 M
 o
      1979
                 1380
    1981

PERIOD, years
    1982
     1983
          Figure 5-8A. Second-highest value among maximum 1-hr ozone

          concentrations at five IMAPBRI monitoring stations, 1979 through

          1983.

          Source; Derived from SAROAD (1985 b-f).
    0.14
    0.12
0,10
0,08
0.06
    0.04
                                 I           I           I



                                T   95% CONFIDENCE INTERVAL
                                                       T

                                                        I
                            I


                            -L
             1979
                     1980
    1981
1982
1983
                                PERIOD, years

         Figure 5-8B. Composite averages of the second-highest value

         among daily maximum 1 -hr ozone concentrations at five NAPBN

         stations, 1979 through 1983.

         Source: Derived from SAROAD (1985 b-f).
                             5-24

-------
ozone values for  evidence  of trends.   As seen  in  Figure 5-8B, however,  the
diversity of the NAPBN site characteristics, the consequent large variation in
recorded values,  and  the  small  sample size result in such a large confidence
interval around the average  for each year  that  no trend can be determined.
The  data  presented in Figures 5-8A and  5-8B  show the NAPBN second-highest
daily maximum 1-hour  ozone concentrations to be  about one-half those reported
for urban-suburban sites for the same period (Figure 5-1).

5.4.2  Sulfate Regional Experiment Sites (SURE)
     As part of a comprehensive air monitoring project sponsored by the Electric
Power Research  Institute  (Martinez and Singh, 1979), ozone measurements were
made by the  chemiluminescence method in the  last  6 months of 1977 at nine
"nonurban" SURE sites in  the eastern United States, shown in Figure 5-9.  On
the  basis  of diurnal  NO  patterns that  indicated the influence of traffic
                        )\
emissions, five of the sites were classed as "suburban"; the other four were
classed as "rural."  The ozone data from these nine stations are summarized in
Table 5-5.   Martinez   and  Singh (1979) noted that  the  four rural  stations
occasionally recorded high values comparable with those in urban areas, but that
the  incidence was low.   They concluded that infrequent transport of ozone or
its precursors, or both, rather than local ozone generation, was the most prob-
able cause of these high values.
5.5  VARIATIONS IN OZONE CONCENTRATIONS:  DATA FROM SELECTED URBAN AND NONURBAN
     SITES
     Variations of ozone concentrations by season  and  by time of day  have
been  long known and are well  documented.   First studied in smog  chambers,
diurnal  patterns  have since been corroborated  by field investigations, and
exceptions  to  such general  patterns  have  been examined  and  documented.   Like-
wise,  field investigations  have substantiated general  seasonal patterns and
exceptions  to  them,  and have also established a  number of spatial variations
in  concentration,  such as those that  occur with latitude or with altitude.
While  it is difficult to discuss temporal and spatial variations separately,
this section is subdivided along those  lines for  convenience.
                                    5-25

-------
MN
                                          050 150250

                                          I  ' ' '  ' I
                                             km
     Figure 5-9. Location of Sulfate Regional Experiment (SURE)

     monitoring stations.



     Source: Martinez and Singh (1979).
                            5-26

-------
                TABLE 5-5.   SUMMARY OF OZONE CONCENTRATIONS  MEASURED  AT SULFATE REGIONAL EXPERIMENT
                               (SURE)  NONURBAN  STATIONS, AUGUST THROUGH DECEMBER 1977
Total no. of
measurements
Number of measurements
with concentrations:
>0.08 ppm
>0.10 ppm
>0.12 ppm
Mean
concn,. ppm
Mean of
daily
1-hour
maxima,
ppm
1-hour
maximum,
ppm
Rural sites

01
i
•xl

#1
#4
#6
#9
Montague, MA
Duncan Falls, OH
Giles Co., TN
Lewisburg, WV
3419
3441
3632
3459
60
52
63
23
33
2
5
3
21
0
0
0
0.
0.
0.
0.
021
029
026
035
0.
0.
0.
0.
044
049
052
054
0.153
0.107
0.117
0.106
Suburban sites






#2
#3
#5
#7
#8

Scranton, PA
Indian River, DE
Rockport, IN
Ft. Wayne, IN
Research Triangle
Park, NC
3410
3017
3462
3438

3495
0
29
29
0

80
0
0
0
0

10
0
0
0
0

0
0.
0.
0.
0.

0.
023
030
025
020

025
0.
0.
0.
0.

0.
035
049
046
039

050
0.077
0.099
0.099
0.080

0.118
Source:   Martinez and Singh (1979).

-------
5.5.1  Temporal Variations In Ozone Concentrations
5.5.1.1  Diurnal Variations  in OzoneConcentrations.   By  definition,  diurnal
variations are those that occur during a 24-hour period.  Diurnal patterns of
ozone may be  expected  to vary with location, depending on the balance among
the many factors affecting ozone formation, transport, and destruction,  as de-
scribed in Chapter 3 and noted in Section 5.1.   Although they vary with  locality,
diurnal patterns for ozone typically show  a rise in concentration from low or
near-zero levels  to an  early afternoon peak.  The 1978  criteria document
ascribed the diurnal pattern of concentrations to three simultaneous processes:
(1) downward  transport  of ozone from layers aloft; (2) destruction of ozone
through contact with  surfaces and through reaction with nitric oxide (NO) at
ground level; and (3) in situ photochemical production of ozone (U.S.  Environ-
mental Protection Agency,  1978;  Coffey et al., 1977;  Mohnen,  1977;  Reiter,
1977a).  Figure  5-10 shows  the diurnal  pattern of ozone concentrations  on
July 13, 1979, in Philadelphia, Pennsylvania.   On this day a peak 1-hour average
concentration of 0.20 ppm, the highest for the month,  was reached at 2:00 p.m.,
presumably as the result of meteorological factors, such as atmospheric mixing,
and local photochemical processes.  The severe depression of concentrations to
below detection limits (less than 10 ppb) between 3:00 and 6:00 a.m.  is usually
explained as  resulting from the  scavenging of  ozone  by local nitric oxide
emissions.  In this regard,  this station is typical of most urban locations.
     Diurnal  profiles  of ozone concentrations can vary from  day  to day  at a
specific site, however, because of changes in the various factors that influence
concentrations.  Such day-to-day variations are clearly demonstrated in Figure
5-11 (SAROAD, 1985c), which  shows diurnal variations in ozone concentrations on
2  consecutive days  at the same monitoring  site  in  Detroit, Michigan.  Differ-
ences  in  timing  and magnitude occur that  are  especially  noticeable between
midnight and  about  7:00 a.m.  Transport is probably involved in these nighttime
variations.   The  afternoon peak concentrations, the  actual maxima  for the 2
days,  differ  in magnitude but  not in timing.
     Composite diurnal  data, that is,  concentrations  for  each hour  of the day
averaged over multiple days  or months,  often  differ markedly  from the diurnal
cycle  shown  by concentrations for a specific  day.   In Figures 5-12 through
5-14  (SAROAD, 1985d),  diurnal data for  2  consecutive days are compared  with
composite  diurnal   data  (1-month  averages of hour-by-hour measurements)  at
three  different  kinds of sites:   center city  (Washington,  DC), rural  but near
                                     5-28

-------
O

§
LU


I
                I  I  I  I  I  I  I  I  I
      12 1  234567 89 10 lit  1  23456789
                           NOON
       «	-a.m.	HOUR OF DAY
                                   1011
"a.m."
                                         •p.m.-
     Figure 5-10. Diurnal pattern of 1 -hr ozone
     concentrations on July 13, 1979, Philadelphia, PA.

     Source: SAROAD (1985b).
                       5-29

-------
   0.12
   0.10
i%
Q.
 *
z
o
=  0.08  —
cc


UJ
o
z
o
o
UJ
z
o
N
o
         -I  I  I I   I    I    I  I
              SEPTEMBER 20

         -  O SEPTEMBER 21
                                   I  I  I  I I  I  I  i  i  I
   0.06 . -
   0.04  -
   0.02
12 123456783 10 11 I  123456789 10
                       NOON

         am.HOUR OF DAY     P'm-      *
                                                       11
                                                        o
                                                        o
                                                        its
                                                        CM

                                                        @l
                                                        m
                                                        E
                                                        X
                                                        ro


                                                        Z
                                                        O
                                                        Z
                                                        o
                                                        o
                                                        LLJ
                                                        Z
                                                        o
                                                        N
                                                        o
            Figure 5-11. Diurnal patterns of 1 -hr ozone

            concentrations, September 20 and 21, 1980,

            Detroit, Ml.


            Source: SAROAD (1985c).
                           5-30

-------
U. \£.
£
g 0.10
Z
o
P 0.08

oc
g 0.06
U
z
S 0.04
Ul
O
N 0,02
O
0
CENTER CITV-CO*MMERCIAL'1 1 1 1 I 1 I

—


—


-
•^
t
i
L
I
i

mmr-L

I ,

"



\ i
\;

0^
. i •

^
i
i
i
r
i
i
i
i
r /
7
/


J
. i

s,
%
^'




"**"**•,






1
	 JULY 8-9"
M —"~"— 1 -month
i —
i
\

1
1







I .




I
<
(
i /
\!
V
L
N —
• 1 i
-»' S






x>
^
r 1 ,

\ y
i
t
i
i
i
*
j
V
, 1 •
/*•"
1
t J
1 /
1
*



m «







k»**\e;
•«.*^





i i I

—
ww

—


—
— ^
^1
^ ^

T
• ' WM
. N-
• I 1 i lilt
24 4 8 12 16 20 24 4 8 12 16 20 24
I* a.m.-
— NOON n.m._ M-!

                      HOUR OF DAY
      Figure 5-12. Diurnal and 1-month composite
      diurnal variations in ozone concentrations,
      Washington, DC, July 1981.
      Source: SAROAD(1985d).
a.
2
O
5
c
z
Ul
U
z
O
U
Ul

1
O
U. !£.

0.10
0.08
0.06

0.04
0.02
0
I I i
1 1 1 1 M i 1 1 1 1 1 1
1 1 1 I 1
RURAL-NEAR URBAN 	 SEPT. 29-30
— — — 1 -month —
— . . —
_ -

—
V
£
j
4
/
I
1
*\ '
7
•
>
| |
/

\
| .
f
S
. i
      24
8  12  16  20  24
8  12  16  20  24
      J-*—a.m.—NOON—p.m.—*•{•*—a.m.—NOON—p.m.—*J
                      HOUR OF DAY
      Figure 5-13. Diurnal and 1 -month composite
      diurnal variations in ozone concentrations, St.
      Louis County, MO, September 1981.
      Source: SAROAD (1985d).
                        5-31

-------


i 0.10
Q.
Z
2 0.08
| 0.06
O
z
0 0.04
UJ
Z
i 0.02
0

0
111 1
1 I i I 1 1 1 1 I I 1 I I I 1 1 i
SUBURBAN-RESIDENTIAL
	 OCT. 11-12 -
	 • 1-month

- -
— _

a
i
— i
f
_ ' *
~ ^?
CS^* A
>£7 ^
\ - f
\

"**^»



, J. , J. , , , •
* ^ -^ ""
T ' % T
* v>^ ~j^.
i ' i y^ "* i-\ •—
V| ^^r^f i*^"* i •**' 'w X ^-
'Jr ^I^L^' ^-^ ^.^
, 1 , 1 , 1 | .. ! , , ! | 1 |s
                24   4   8   12  16   20  24   4  8   12   16  20  24
L»— a.m.— NOON— p.m.
                                          a.m — NOON — p.m — 1~\
                                 HOUR OF DAY
                Figure 5-1 4. Diurnal and  1 -month composite
                diurnal variations in ozone concentrations, Alton,
                IL, October 1981 (fourth  quarter).
                Source: SAROAD (1985d).

an urban area (St. Louis  County, MO), and suburban-residential  but in a crop-
growing area (Alton,  IL).   Several  obvious  points  of interest present themselves
in these graphs:   (1)  at  some sites, at least, peaks can occur at virtually
any hour of  the  day  or night but these peaks may  not show up strongly  in the
longer-terra  average  data;  (2) some sites may be  exposed  to multiple  peaks
during a 24-hour  period;  and (3)  disparities, some of them large, can exist
between peaks (the diurnal  data)  and the 1-month  mean (the composite diurnal
data) of hourly ozone concentrations.  These figures are  given  simply as exam-
ples of the  differences that  can occur between daily and monthly mean concen-
tration patterns.
     The effects  of  averaging are readily apparent when diurnal or short-term
composite diurnal  ozone concentrations are  compared with  longer-term composite
diurnal ozone concentrations.  When  compared with Figures 5-12 through 5-14
(daily values and 1-month averages),  Figure 5-15  (SAROAD, 1985d), based on
3-month averages, demonstrates rather graphically the effects  of  lengthening
the period  of time  over  which Values are  averaged.  This figure shows  a
composite diurnal  pattern  calculated  on the basis  of 3 months.   While seasonal
                                    5-32

-------

•i
1
o
i
fE
z


o
0
LU
2
8




V.IU
0.09
0.08
0.07
0.06

0.05

0.04

0.03


0.02
0.01
0
2

I I 1 I I I I 1 1 I i I i 1 i I 1
- ALTON, IL
—
—
•»

- .*—•. 	 3rd
/ - — N
— ^" x** 2nd Q ^
* ^
"• -^^


[-••-=.•?- 	 < ....-*' 4th Q -.

1 i 1 1 1 i I 1 1 1 1 1 1 I i 1 1
42 4 6 8 10 12 14 16
I*. 	 a.m. NOON
i I i I I 1
. —
—
—
•m*m

Q - -

mm*
\
\
Nk

ss—S

1 I I 1 1 1
18 20 22 2

                                  HOUR OF DAY
                 Figure 5-15. Composite diurnal patterns by
                 quarter of ozone concentrations, Alton, IL, 1981.
                 Source: SAROAD (1985d).
differences are seen, and will be  discussed later, the comparison of 3-month
and 1-month composite diurnal concentrations at the  Alton,  IL, site readily
demonstrates the smoothing out of peak concentrations as  the averaging period
is lengthened.   Thus, a fourth pertinent  point  of interest, related to the
third point cited  above,  emerges  from the data presented above:   that is,
increasing the  averaging time obscures  the magnitude  and time of occurrence of
peak ozone  concentrations.   This  is  an obvious  and familiar result in the
statistical treatment of monitoring data,  but  one  that is highly pertinent to
the protection  of human health and  welfare from the effects of ozone.
     The  significance of  the relationships  of  peak and mean concentrations
depends upon whether the  health  and  welfare effects  of exposure to ozone are
solely  concentration-dependent,  heavily  concentration-dependent,  or  both
concentration-  and time-dependent.   Depending upon whether  acute  or chronic
exposures elicit the health and welfare effects  of concern,  careful  attention
                                    5-33

-------
may have to  be paid to the  relationship of  short-term  (e.g., 1-hour) versus
longer averaging times.
     Quantitative analyses of the relationships among maximum or second-highest
1-hour concentrations  and daylight, diurnal,  monthly,  seasonal,  and yearly
average ozone  concentrations lie  outside  the scope of this  document.   An
example of one study,  however,  serves to describe the kinds of analyses that
may be done  to determine relationships among  various exposure  statistics or
"averaging times."   Lefohn and  Benedict  (1985) have examined  the  relationship
between 7-hour and  12-hour (daylight)  mean ozone concentrations averaged over
a season (7  months).   They also examined the  relationships  between each of
these two exposure  statistics and the 1-hour  values during  the same season
>0.10 ppm.   For their analyses they evaluated data records from all sites in
the SAROAD network  that met  the following criteria:  (1)  existence  of 4 years
of monitoring  data  (1978 through 1981); (2) >4300 1-hour concentrations re-
ported; and  (3) the occurrence of  exposures ^1^.2  ppm-hours, which was an
index of sites exposed repeatedly to ozone concentrations MK10 ppm over the
specified time period.   Sites so selected were then subjected  to additional
criteria that  resulted in the  identification  of monitoring sites located in
areas where agriculturally important cash crops were grown.
     On the  basis  of  calculated integrated exposures  (hours times 1-hour
concentrations >0.10 ppm), the monitoring sites thus identified were subdivided
by Lefohn and  Benedict (1985)  into two categories:   (1) those with 80 to 200
occurrences  of 1-hour values <0.10 ppm  (i.e., 80 to  200 hours at >0.10 ppm
ozone), and  (2) those with 200 to 1612 hours at >0.10 ppm 03.   To describe the
quantitative relationships among the exposure statistics at sites in the two
categories,  a  Pearson linear correlation matrix was  calculated.   For sites
with 200 to  1612  occurrences of 1-hour concentrations >0.10 ppm,  the correla-
tion was 0.88  between the number of occurrences and the 7-hour seasonal mean
and 0.93 between the number  of occurrences and the 12-hour seasonal  mean.  The
correlation between the 7- and 12-hour seasonal means was 0.96.  When correla-
tions were calculated from data obtained at sites with 80 to 200 occurrences
of 1-hour concentrations  >.0.10 ppm, the correlations between occurrences and
the  7-  and  12-hour seasonal means were 0.40  and  0.47, respectively.   The
correlation  for the two seasonal means  at these  sites  was 0.94  (Lefohn and
Benedict, 1985).
                                    5-34

-------
     No attempt  is  made  in this section to document the respective contribu-
tions  of  local  formation of ozone versus  transport  of ozone; however, the
occurrence of multiple peak ozone concentrations within a 24-hour period is
usually construed as  indicating the presence of ozone transported to the site
from elsewhere.  Figure  5-16 illustrates the diurnal variations that can be
seen when transport occurs.    Note the occurrence of dual  peaks on each of 3
successive days  at  this  site,  part of the Sulfate Regional  Experiment (SURE)
network.
     A  familiar  measure  of ozone  air quality is  the number or percentage of
days on which some specified  concentration  is  equalled or exceeded.  This
measure, however, does  not shed light on one of the more important questions
regarding the effects of ozone on both people and plants; that is,  the possible
significance of high concentrations lasting 1 hour or longer and then recurring
on 2 or more successive days.
     In human controlled exposures, attenuation of responses to ozone has been
observed at  about 0.20 to  0.50 ppm in exercising  subjects receiving  repeated,
consecutive-day  exposures  (see Chapter  10).   That attenuation is  lost after
exposures to those  levels cease (see Chapter 10 for the time course of loss of
attenuation).  It becomes  of interest,  therefore,  to examine  how many days  in
a  row  the maximum  1-hour  ozone concentration  reaches  or exceeds .specified
levels  in communities in high-ozone  areas, as well  as  in other parts of the
country.
     The recurrence of high ozone concentrations on consecutive days was exam-
ined in data (SAROAD, 1985b-d) for one  site  in each  of  four cities:  Pasadena
and  Pomona,  CA;  Washington, DC; and Dallas, TX.   The numbers of multiple-day
events were tallied by length  of event (i.e., how many consecutive days) using
data for  the daylight hours (6:00 a.m.  to 8:00 p.m.) in the second and third
quarters of  1979 through 1981.  These sites were chosen because they include
areas  known  to  experience  high ozone concentrations  (California),  and because
they represent  different geographic  regions of the country (west,  southwest,
and  east).   Similar data could be compiled for any city for which sufficient
aerometric data  exist.   The choice of the 14-hour daylight period and of the
second  and third quarters  is consistent with known diurnal  and seasonal patterns
of ozone concentrations and with typical human, crop, and ecosystem exposures.
     In this discussion of  the recurrence of respective specified ozone levels,
a  day  or series  of  days  on which  the daily 1-hour maximum reached  or exceeded
                                    5-35

-------
    0.16
  E 0.14
 2

 tt
 H
 g
 a
0.12

0.10

0.08

0.06

0.04
 Z
 o
 0 0.02
0
tn
i
oo
01
   24      a.m.    NOON    p.m.


           SATURDAY, 27 AUGUST
                    24      a.m.    NOON     p.m.


                             SUNDAY, 28 AUGUST
24      a.m.    NOON    p.m.


      MONDAY, 29 AUGUST 1977
24
                 Figure 5-16. Three-day sequence of hourly ozone concentrations at Montague, MA, SURE
                 station showing locally generated midday peaks and transported late peaks.
                 Source: Martinez and Singh (1979).

-------
the specified  level  is called an "exposure"; the  intervening day or days when
that level was  not reached is called a "respite."  Four ozone concentrations
were chosen:   0.06,  0.12,  0.18,  and 0.24 ppm.   The 1- to 7-day events (i.e.,
exposures or  respites) are  individually tabulated and  events  longer than
7 days are grouped together.
     The summaries  of  "exposure"  and "respite" events  in Tables 5-6  through
5-9 show expected  regional  differences in the successive recurrence of these
respective 1-hour  concentrations.   At the Dallas  station,  for  example,  the
3-year data tally shows 11 exposure events when the daily 1-hour maximum ozone
value equalled or exceeded 0.06 ppm for more than 7 days in a row (Table 5-6).
At first glance, the Pasadena station, with 10 such exposures, appears similar,
but those 10 exposure  events  spanned  443  days;  in Dallas the 11 exposures  in-
volved only 168  days.   At this lowest concentration (> 0.06 ppm),  the Dallas
station recorded more  short-duration  (<  7 days) exposures (45)  involving more
days (159) than the Pasadena station (14 exposures over 45 days), simply because
the daily 1-hour maximum statistic in Pasadena remained above 0.06 ppm for such
protracted periods.  At  concentrations >_ 0.12 ppm  (Table 5-7),  the lengthy
exposures at the Pasadena site resolved into numerous shorter exposures; whereas
in Dallas the  exposures  markedly dwindled in number and  duration.   At concen-
trations >_ 0.18  ppm (Table 5-8), short-term (< 7 days) exposures in Pasadena
were yet more  numerous;  Dallas had only  two 1-day exposures.   At > 0.24 ppm
(Table 5-9), the  incidence rate of exposures at  the Pasadena station finally
decreased.
     Note that  these tables (Tables 5-6 through  5-9) also present the single
and multiple-day periods  when concentrations were lower than  specified levels
("respites");  "event-days"  include all  "exposures" to and "respites"  from
specified concentrations,  such that "total  event-days"  for  a  given  site equal
the total days monitored.
     These tables  give a more extensive  indication of the long-term  severity
of ozone air quality at some of the selected sites  than is given by the use of
a  1-hour exposure  statistic alone.   The  human  responses to  single acute expo-
sures to ozone are well  documented (Chapter 10)5  but  the full  significance of
the attenuation  of human responses to ozone over  the  course of  a multiple-day
exposure, and any possible consequences of the repetition of such multiple-day
exposures within  a smog season or  over  a number of years remain uncertain.
                                    5-37

-------
en
to
Co
                      TABLE 5-6.         OF CONSECUTIVE-DAY  EXPOSURES  OR  RESPITES WHEN THE  DAILY
                           1-hr MAXIMUM OZONE CONCENTRATION  WAS > 0.06 ppm,  IN  FOUR  CITIES
                                     (APRIL THROUGH SEPTEMBER, 1979 THROUGH 1981)
Length of
event,
days
1
2
3
4
5
6
7
Subtotals
Events
(Event-days)
>7
(Event- days)
Total events
(Total event-
days)
Total days
monitored
Source: SAROAD
a
Pasadena
3
5
0
2
2
0
2
14
(45)
10
(443)
24
(488)
532
(1985b-d)
Exposures"
Pomona
6
2
4
3
1
0
1
17
(46)
16
(426)
33
(472) .
542
.
(>0.06 ppm)
Washington
21
10
6
2
2
0
2
43
(91)
5
(55)
48
(146)
442


Dallas
8
10
6
5
8
3
5
45
(159)
11
(168)
56
(327)
451

ii
Pasadena
9
8
2
2
1
0
0
22
(44)
0
(44)
22
(44)
532

Respites
Pomona
11
11
6
2
1
1
0
32
(70)
0
(0)
32
(70)
542

11 (<0.06 ppm)
Washington
17
16
0
3
3
3
1
43
(101)
8
(195)
51
(296)
442


Dallas
25
13
8
4
1
0
1
52
(103)
2
(2.1)
54
(124)
451


-------
CJI
                      TABLE 5-7.   NUMBER OF CONSECUTIVE-DAY EXPOSURES OR RESPITES WHEN THE DAILY
                           1-hr MAXIMUM OZONE CONCENTRATION WAS > 0.12 ppm, IN FOUR CITIES
                                    (APRIL THROUGH SEPTEMBER, 1979         1981)
Length of
event,
days
1
2
3
4
5
6
7
Subtotals
Events
(Event- days)
>7
(Event- days)
Total events
(Total event-
days )
Total days
monitored:
Source: SAROAD
11

Pasadena
11
10
8
4
7
2
_

42
(118)
14
(254)
56

(372)

532
(1985b-d)
Exposures"

Pomona
13
16
8
5
2
3
1

48
(124)
13
(211)
61

(335)

542

(>0.12 ppm)

Washington
3
1
0
0
0
0
0

4
(5)
0
(0)
4

(5)

442



Dallas
20
4
2
0
1
0
0

27
(39)
0
(0)
27

(39)

451

11

Pasadena
21
14
10
1
3
2
3

54
(131)
3
(29)
57

(160)

532

Respites

Pomona
16
18
9
3
6
3
2

57
(153)
5
(54)
62

(207)

542

" (<0.12 ppm)

Washington
0
0
0
0
0
0
0

0
0
7
(437)
7

(437)

442



Dallas
3
2
-
1
2
2
1

11
(40)
18
(372)
29

(412)

451


-------
         TABLE 5-8.   NUMBER OF  CONSECUTIVE-DAY  EXPOSURES OR RESPITES  WHEN THE DAILY
              1-hr MAXIMUM OZONE  CONCENTRATION  WAS  >  0.18 ppm,  IN FOUR CITIES
                       (APRIL THROUGH SEPTEMBER,  1979 THROUGH 1981)
Length of
event,
days
1
2
3
4
5
6
7
Subtotals
Events
(Event-days)
>7
(Event-days)
Total events
(Total event-
days)
Total days
monitored
"Exposures" (>0.18 ppm)
Pasadena
21
9
10
6
3
4
1
54
(139)
7
(90)
61
(229)
532
Pomona
24
12
9
4
4
4
2
59
(149)
2
(20)
61
(169)
542
Washington
0
0
0
0
0
0
0
0
(0)
0
(0)
0
(0)
442
Dallas
2
0
0
0
0
0
0
2
(2)
0
(0)
2
(2)
451
ii
Pasadena
15
8
14
3
7
2
3
52
(153)
11
(150)
63
(303)
532 .
Respites
Pomona
15
11
9
4
5
1
2
47
(125)
16
(248)
63
(373)
542
" (<0.18 ppm)
Washington
0
0
0
0
0
0
0
0
(0)
3
(442)
3
(442)
442
Dallas
0
0
0
0
0
0
0
0
(0)
5
(449)
5
(449)
451
Source:   SAROAD (1985b-d).

-------
                         TABLE 5-9.   NUMBER  OF  CONSECUTIVE-DAY  EXPOSURES  OR  RESPITES WHEN  THE  DAILY
                              1-hr MAXIMUM OZONE  CONCENTRATION  WAS >  0.24 ppm,  IN  FOUR  CITIES
                                       (APRIL THROUGH SEPTEMBER,  1979 THROUGH  1981)
en
                Length of
                  event,
"Exposures" (>0.24 ppm)
"Respites" (<0.24 ppm)
days
1
2
3
4
5
6
7
Subtotals
Events
(Event-days)
>7
(Event-days)
Total events
(Total event-
days)
Total days
monitored
Source: SAROAD
Pasadena
20
13
5
2
2
0
0

42
(79)
1
(9)
43

(88)

532
(1985b-d).
Pomona
21
10
0
2
1
0
0

34
(54)
1
(9)
35

(63)

542

Washington
0
0
0
0
0
0
0

0
(0)
0
(0)
0

(0)

442

Dallas
0
0
0
0
0
0
0

0
(0)
0
(0)
0

(0)

449

Pasadena
7
1
6
4
3
1
0

22
(64)
23
(380)
45

(444)

532

Pomona
8
4
5
2
2
1
1

23
(62)
14
(417)
37

(479)

542

Washington
0
0
0
0
0
0
0

0
(0)
3
(442)
3

(442)

442

Dallas
0
0
0
0
0
0
0

0
(0)
3
(449)
3

(449)

449


-------
5.5.1,2   Seasonal  Variations  In  Ozone  Concentrations.   In  addition  to the
diurnal cycles and between-day variations discussed in the  preceding section,
seasonal variations  in  ozone  concentrations occur (for,the reasons discussed
in Chapter 3  and  Section 5.1) and usually assume characteristic patterns.
     In order to  compile an assessment of  potential ozone  damage to the six
leading commercial crops  in the  United States  (corn,  soybeans,  hay,  wheat,
cotton, and tobacco), Lefohn (1982) surveyed 304 ozone monitoring stations and
identified 24 that (1)  were located  in counties producing significant quanti-
ties of one  or  more  of these six  crops  in 1978;  (2)  reported at  least 50
percent of possible  hourly  data  in 1978; (3) reported  an hourly maximum of at
least 0.1 ppm 0~;  and  (4)  ranked  high  in  cumulative  ozone exposure for the
                <3                      - -• -     .                 ...  •   .
period April to October 1978.   Six of these sites represented counties with
high soybean, wheat,  or hay production.  Quarterly composite diurnal patterns
for six of these  sites with reasonably complete  (>75  percent) 1981 data are
shown  in  Figure 5-17  (SAROAD,  1985d).  The average  levels are  apparently
comparable with the long-term averages at the NAPBN sites previously discussed
(Section 5.4.1).   In addition, the diurnal patterns for these sites clearly
show the  division of the afternoon  ozone  concentrations into two seasonal
patterns, the low  "winter"  levels in the  first and  fourth quarters and the
higher "summer" levels  in the second and third quarters of the year.
     Although averaging causes details to  be obscured,  the average diurnal
patterns  in  Figure 5-17 show that the time of  occurrence  of peaks differs
among sites, and,  to an extent,  between seasons.  The  seasonal differences in
time of day are especially  noticeable for the fourth quarter.  Among the sites
shown in Figure 5-17, ozone concentrations appear to peak at 2:00 to 2:30 p.m.
in Little  Rock in the  higher-concentration second  and third quarters.  At
Bakersfield in the second and third quarters,  there  is  evidence,  even from
these composite data, of two peaks, the first at about 1:00 p.m.  and the second
at 5:00 to 6:00 p.m.   At the Clark County, Ohio, site, the peak concentrations
in the second and  third quarters  center around about 5:00 p.m., but they do not
return to "baseline"  until  after midnight.   The patterns at the Bakersfield and
Clark County sites appear to indicate transport  into  the areas.   It is also
possible that single peaks  that  are shifted to mid- to late afternoon, as at
the Little Rock  site,  are the product  of transport.  Depending upon proximity
to urban  centers  and  wind  speed  and direction,  rural  areas are typically
exposed to their  peak concentrations later than  urban areas,  usually within
daylight hours but not  always.
                                    5-42

-------
Q.
a
<
cc
ui
O
z
o
o
o
o
 a
 a
O
P
K
Z
O
o
1
 a
z
o
H
cc
o
o
IU
1
O
0.10
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01
  0
  I  I I  I I  I I  I I  I I I  M  I I  i M  I
  N. LITTLE ROCK, AR
  (A)
   	 1st Q
   .— — • 2ndQ
   	3r
-------
     Composite diurnal  variations  in ozone concentrations at a rural site in
Argonne, Illinois, over a 7-week period of the third quarter of 1980 are shown
in Figure  5-18 (Kress  and Miller,  1983).   The actual day-to-day variations in
ozone concentration  over  the entire third quarter of  1980  at that site are
shown for  comparison in Figure  5-19  (Kress and Miller, 1983).  As part  of the
National Crop  Loss Assessment  Network, the  site at Argonne monitors ozone
concentrations over  a  7-hour daytime period approximating the period of peak
photosynthesis in crops.   The data  in  Figure 5-18 yield a 24-hour  average
ozone concentration of 0.026 ppm (Kress and Miller, 1983).  The 7-hour average
for the same 7-week period at the Argonne  site is 0.042 ppm (Kress and Miller,
1983).  The  day-to-day variations  in both  the 7-hour and  the 24-hour averages
(Figure 5-19) generally appear to  be greater than the  average difference
within  a day for either the  7-hour  or 24-hour  periods (Figure 5-18).  The
fluctuations in 1-hour values either within a day or from day to day, however,
would be larger than within-day or between-day variations in either the 7-hour
or the  24-hour average.   The 7-hour average will  be higher than  the 24-hour
average because  the former  excludes the  low nighttime concentrations.  As
Figure 5-19 and its data illustrate, the selection of the appropriate averaging
time  is critical  for the accurate description of dose-response relationships
and  for the protection of human,  vegetation, and other receptors from the
effects of ozone.
     It is worth  noting in this context that air pollution exposure statistics
are  of  two basic functional  types:  descriptive and preventive.  Descriptive
exposure statistics  simply define the conditions of concentration and exposure
duration (averaging  time)  under which a specified effect  has  been observed or
detected.   Preventive  exposure statistics are used prescript!*vely  with the
expectation  of keeping a specified  effect from  happening.   To  be effective
bases  for  an air quality  standard,  preventive  exposure statistics  must be
(1)  related to the true time-course  of development of the specified effect(s),
and  (2) based  on observed or predictable  distributions, or both, of exposure
conditions  over  the range  of  concentrations and durations producing  the
effect(s) of concern.
     In Figure 5-20  (A-H), seasonal  variations in ozone concentrations in 1981
are  depicted using  1-month  averages and  the single 1-hour maximum  concen-
tration within the month for eight  sites  across the nation  (SAROAD, 1985d).
The  data from  most of  these  sites exhibit the expected pattern of high ozone
                                    5-44

-------
a
a
DC

Z
111
u
Z
o
u
111
Z
o
N
o
   0.06
0.04
   0.02
      24
                  a.m.
                           NOON
                                           p.m.
                   24
      Figure 5-18. Composite diurnal ozone pattern at a rural
      NCLAN site in Argonne, IL, August 6 through September
      30,1980.
      Source: Kress and Miller (1983).
      0.15
    a
    a
   2  o.io —
   <
   DC
   U

   o
   u
   LU
   Z
   o
   N
   o
   0.05
        JUL
                   AUG
SEP
OCT
                        MONTH OF YEAR
         Figure 5-19. Daily 7-hour and 24-hour average
         ozone concentrations at a rural NCLAN site in
         Argonne, IL, 1980.
         Source: Kress and Miller (1983).
                          5-45

-------
u.oa
0.30
1
u.
2 0.26
g
e 020
ty
U
Z 0.15
O
U
Ul
o a1°
8
0.05
0
0.35
0.30
n.
2 °'25
O
5
5 0.20
us
O
O
UJ
Z 0.10
N
O
0.05
0

_

MM
_
—


_

TUCSON, AZ
(AJ

B MAXIMUM ~
AVERAGE




I
MMVMH

1


11
_

1
••^•K

1
_

1
—

i



»*»***




•i'x
—

.ri

tTO
—

m

—
fff
JFMAMJJASOND

-.
—
—
_

_

—

POMONA. CA
(0 __,







T"
«_



55




1
—




fx
MM*





11






rrr
1
—





WT
P-:
—







S!


—
—






1
-
_

—


Wl-

—


— ~1

_


0.30

0.25
0.20
0.15

0.10
0.05
0
0.35
0.30
0.25
0.20
0.15

0.10

0.05
0
LENNOX, CA
_ (B)

"~
_
—


-n
H
j

"""
~




•J-X
_

9



^




:•:•:•:




I
_



r?9



:JS



1

1 	 1 ~
—
1 W ~
FMAMJJASOND
DENVER, CO
_ (D) _
_
_


-..

"r-i
D
_
_






_


i

—

P
—

^







*!'
••H


((•(•I
—


1
— i
m_

    J  F  M A M  J  J A  S O  N  D

            MONTH OF YEAR
J FMAMJJASOND

        MONTH OF YEAR
Figure 5-20. (A-H). Seasonal variations in ozone concentrations as indicated
by monthly averages and the 1 -hour maximum in each month at selected sites,
1981.
                              5-46

-------
u.oa
0.30
D.
^ 0.25
O
2 0.20
Ul
i 0.15
O
o
Z 0.10
o
s
0.05
0
0.35
0.30
Q.
^ 0.25
g
< 0.20
y
Ul
Z 0.15
O
U
z o.io
1
0.05
0
WASHINGTON, DC
_

(E

)



—

B MAXIMUM —
AVERAGE
_
—
I



n
j


n
F


M
M
MORRIS
—

D

W.
^™
I

i
_
TO

•X**
—
PWafcWi LwJ
AMJJASOND
CO., NJ
_ (6) _

—
—

_

^^










W!








PJJJ3







"*""
m

—
—

__ —
ff™^^

—
ra


M
—


1
^™^^i


W:



I

— __



DO

0.30

0.25
0.20
0.15
0.10
0.05
0
0.35
0.30

0.25
0.20

0.15

0.10

0.05
0

_

•"""
—
_
—
_


_

—
—

_

^^


TAMPA, FL
IF) _

—
_
-

D
rr>

I
-
m

m

IT-
—
1

m
—
xT
""""•
1!
—
M tea
JFMAMJJASOND
DALLAS, TX
(H) _

—
—

— . —_ —


—

—
~
_
I



m



55
—


n



1

1.

^


53
„ 	 „. „„,„
-—
ii

sltei
   JFMAMJJASOND
           MONTH OF YEAR
J FMAMJJ  ASOND
       MONTH OF YEAR
Figure 5-20. (A-H) (continued). Seasonal variations in ozone concentrations as
indicated by monthly averages and the 1 -hour maximum in each month at
selected sites, 1981.
Source: SAROAD (1985d).
                            5-47

-------
levels in  late  spring or in summer and  low levels in the winter.   Data for
Pomona (Figure  5-20C)  and Denver (Figure 5-20D) show  summer  maxima.   Tampa
shows a  late spring maximum but with  concentrations  in the fall (October)
approaching those  of  spring (June) (Figure 5-20F).  Dallas data also tend to
be skewed  toward higher  spring concentrations; but note that November concen-
trations are also relatively high (Figure 5-20H).  Averaging together data for
several  years would give a smoother "characteristic"  pattern but also would
obscure the  fact that local,  and even national, weather in a particular year
plays at least as big a role in the formation of ozone as the regular seasonal
changes in the elevation of the sun and the resulting variations in insolation.
Because of seasonal changes in storm tracks from  year to year, the general
weather conditions  in  a  given year may be more favorable  for the formation  of
ozone and  other oxidants  than  during the  prior or  following  year.   Thus,
short-term concentration  trends  may  not be indicative of real changes in air
quality.
5.5.1.3    Weekday-Weekend Variations in Ozone Concentrations.   Atmospheric
ozone concentrations  represent  the combined effects of emission sources  and
meteorological  conditions.  The  various sections of this  chapter  have been
based on the  assumption  that ozone precursor sources  operate on a generally
steady-state or at least on an average, repeatable  diurnal  cycle.   For the
most partj urban source patterns of oxidant precursors do appear to be reason-
ably  constant;  however,   in most urban areas changes  occur  in  traffic and'
commercial emission  patterns  that are  keyed to a weekday-weekend activity
cycle.   The  effects of  these  changes have been observed in corresponding
changes in ozone concentration patterns.
     Debate  in  the 1960s over the role  of  nitric oxide (NO) in scavenging
ozone led  to  the  examination of whether weekday-weekend differences in ozone
concentrations might occur  in urban areas,  on the assumption that NO emissions
would be lower on weekends  when some NO -emitting sources virtually shut down.
                                       y\
Altshuller (1975) reported  the observation  that no alerts had ever occurred in
the  Los  Angeles Basin on a Sunday.   Likewise, Elkus  and Wilson  (1976),  Levitt
and Chock  (1976), and Graedel et al. (1977) reported the existence of a weekday-
weekend difference in ozone concentrations.  While not reporting lower average
ozone concentrations  on  weekends for the  Los Angeles  Basin, Schuck et  al.
(1966) noted  a  spatial shift of concentrations on summer weekends, away from
the  central  commercial  and urban areas  to  suburban  and coastal  areas.   They
                                    5-48

-------
attributed the spatial shift in concentrations to changes in traffic patterns
resulting from a shift from job-related travel to recreational  travel.
     No newer reports of possible weekend-weekday variations are available in
the literature,  but any exhaustive exposure assessment should take into consid-
eration the possibility that subtle differences  in absolute concentrations or
their spatial and  temporal patterns could  have some effect on total exposure.

5.5.2  Spatial Variations in Ozone Concentrations
     Ozone is commonly thought of as a relatively homogeneous,  regional pollu-
tant, largely because  of  the nature of the sources of its precursors and the
processes that contribute to its formation as a secondary pollutant.   Numerous
factors affect the occurrence  of ozone in ambient air, however, so that dif-
ferences  in the  true exposures  sustained by  respective receptors  (e.g.,  human
populations, crops, natural ecosystems) will vary within regions, airsheds, or
other defined areas.  The  data  presented in  this section  demonstrate a few of
the known spatial variations in ozone concentrations that should be taken into
account when assessing actual or even potential exposures.  The data presented
in this  brief section are  intended  to  be illustrative  rather than  exhaustive,
since numerous meteorological, topographical, and physicochemical factors will
influence the spatial  and temporal  distributions of ozone concentrations at
specific  sites.
5.5.2.1   Urban versus Nonurban  Variations.   Data were  presented in the  1978
criteria  document  demonstrating that peak concentrations of  ozone in rural
areas are generally lower than those  in urban  areas,  but that "dosages or
average  concentrations  in  rural areas are comparable  to  or even  higher than
those  in urban  areas"  (U.S.  Environmental  Protection Agency,  1978).   The
diurnal  concentration  data presented in the  preceding section  indicate that
peak ozone  concentrations  can occur later  in the  day  in  rural  areas than  in
urban, with the  distance downwind from urban  centers generally determining how
much later  the  peaks occur.   The data presented in the preceding section for
Montague, Massachusetts,  in  Figure  5-16 (Martinez and Singh, 1979) exemplify
high late-afternoon secondary  peak concentrations resulting from  transport.
     The  NAPBN and other nonurban data presented in Section 5,4 illustrate the
urban-nonurban  gradient in ozone concentrations that  generally exists, and
support  the statement quoted above  from the  1978 document.  While corroboration
of  that  statement  would require the  calculation of means or ppm-hours, the
                                     5-49

-------
data given  in  Section 5.4.2 tend to support  that  conclusion.   For example,
data presented in that section (Table 5-5) showed that rural sites of the SURE
network were exposed, on the average, to daily maximum 1-hour ozone concentra-
tions in excess  of those to which the  suburban  sites were  exposed (Martinez
and Singh,  1979).   While one rural  site  recorded  21 occurrences of 1-hour
ozone concentrations  >0.12  ppm  during  August through December 1977 (referred
to below as  A-D),  the suburban sites recorded no such occurrences.   All  four
of the  rural  sites in the SURE network recorded two or more occurrences  (and
one recorded 33 such occurrences) of daily maximum 1-hour concentrations  >0.10
ppm; only one  of the five suburban sites recorded the occurrence (n = 10) of
that concentration  over  the  same period (A-D).   All  exposure statistics exam-
ined by Martinez and  Singh  (1979) for the SURE concentration data were higher
at the rural sites, on the average, than at the suburban sites.  The respective
statistics, weighted  by  the number of observations in the data record, were:
average of  24-hour mean  concentrations  (A-D), 0.028  ppm at  rural versus 0.024
ppm at  suburban  sites; mean  of daily maximum  1-hour  (A-D),  0.050 ppm at rural
versus  0.044 ppm  at  suburban  sites; and the 1-hour maximum  for the whole
period  (A-D),  0.121 ppm  at rural sites versus  0.094 ppm at suburban sites.
     In another  example,  data reported  by Lefohn (1984) from sites classified
as rural in the  SAROAD data  bank showed 9 occurrences of daily maximum 1-hour
ozone concentrations  >0.12  ppm  at a rural  Preble  County, Ohio,  site; and 19
such occurrences at a rural site in Camden County, New Jersey.
     In addition to the occurrence of higher mean concentrations and occasion-
ally higher  peak concentrations  in nonurban areas than in urban, it is  well
documented that  ozone persists longer in  nonurban  than in urban  areas  (Coffey
et  al.,  1977; Wolff  et  a!.,  1977;  Isaksen et al.,  1978).   The  absence  of
chemical scavengers appears to be the chief reason.
5.5,2.2 Intracity  Variations.   Despite  relative  intraregional homogeneity,
evidence exists  for intracity variations  in concentrations  that  are  pertinent
to potential  exposures  of human populations  and to  the assessment of actual
exposures sustained in epidemiologic studies.   Two illustrative pieces of data
are presented  in this section,  one  a case  of relative homogeneity in  a city
with a  population  under  500,000 (New Haven,  Connecticut)  and one a case of
relative inhomogeneity of concentrations  in a city of greater than 9 million
population (New York City) (U.S.  Department of Commerce, 1982).
                                    5-50

-------
     New Haven, Connecticut, was the site of an epidemiological study in 1976
by Zagraniski  et  al.   (1979).   Symptoms  recorded in diaries  kept  by  study
participants were correlated with ozone concentrations measured by the chemi-
"luminescence method at a  downtown  New Haven site  characterized  as  Center
City-Residential.   Table 5-10 shows several  percentiles in the distribution of
hourly values  for that site plus two other monitoring stations in the county
that were operating at the time, one  in Derby, Connecticut,  9 miles west  of
New Haven,  and one  in Hamden,  Connecticut, 6 miles north of  New Haven.  The
Derby site also is characterized as Center City-Commercial,  the Hamden site as
Rural-Agricultural.   The general  similarity of  values among  the three sites
appears to  substantiate  the New Haven data used in the epidemiological  study
since there was probably a reasonable temporal  correlation between these close
sites.   Thus,  wherever individuals might have traveled about  the county, they
probably were exposed to similar concentrations  of ambient ozone.   This conclu-
sion is reinforced by the data in Table 5-11, showing the date and time of the
maximum hourly concentrations by quarter at these three sites.  The significant
data are  those for  the second and third  quarters  when the potential  for 0~
formation and for exposure was the greatest.  Differences in peak concentrations
varied from  0.006 ppm  in the fourth  quarter to  0.055 ppm  in the third quarter
among sites.
     The source of much of the ozone found  in the New Haven, Connecticut, area
is the greater New York City area (e.g., Wolff et al., 19/5; Cleveland et al.,
1976a,b)  and an  urban plume transported over the distance from New York City
to New  Haven would  tend to be  relatively  well-mixed  and uniform, such that
intracity variations in New Haven would probably be minimal.
     The highest or second-highest 1-hour maximum ozone concentration reported
from a given station during a given year frequently gives an  indication of the
potential for  repeated human exposure to high ozone levels.  Nevertheless, a
one-to-one correspondence between peak levels and either the  number of days or
the  number  of  hours that a given  level may be  exceeded  does  not  necessarily
exist.  Data obtained in the metropolitan  New  York area in  1980 illustrate
this  latter fact (Smith,  1981); and  illustrate,  as well, that  intracity
gradients can  exist that should be taken  into account  in  exposure  assessment.
The  data  given in Table 5-12 were  obtained at  the monitoring sites shown  in
Figure  5-21.   The second-highest 1-hour ozone readings at the Eisenhower Park
and  Queens  College  stations have values only a few percentage points  apart,
                                    5-51

-------
             TABLE 5-10.  OZONE CONCENTRATIONS AT SITES IN AND AROUND
                           NEW HAVEN, CONNECTICUT, 1976
                 CCHEMILUMINESCENCE METHOD, HOURLY VALUES IN
                                      of values < stated concentration
                       No.
Site (SAROAD No.)  measurements
              50%
           90%
95%
99%    Max concn.
New Haven, CT
(070700123F01)

Derby, CT
(070190123F01)

Hamden, CT
(070400001F01)
4119
5698
3853
0.021    0.035     0,091    0.162      0.274
0.023    0.038     0.071    0.095      0.290
0.030    0.045     0.075    0.098      0.240
Source:  SAROAD (1985a).
            TABLE 5-11.  QUARTERLY MAXIMUM 1-HOUR OZONE VALUES AT SITES
                    IN AND AROUND NEW HAVEN, CONNECTICUT, 1976
                  (CHEMILUMINESCENCE METHOD, HOURLY VALUES IN ppm)

New Haven, CT
No. measurements
Max 1-hr, ppm
Hour of day
Date
Derby^ CT
No. measurements
Max 1-hr, ppm
Hour of day
Date
Hamden, CT
No. measurements
Max 1-hr, ppm
Hour of day
Date

1
10
0.045
11:00 a.m.
3/29
11
0.015
11:00 p.m.
3/31
56
0.050
Noon
3/29
Quarter of
2
1964
0.274
2:00 p.m.
6/24
2140
0.280
2:00 p.m.
6/24
2065
0.240
3:00 p.m.
6/24
Year
3
2079
0.235
2:00 p.m.
8/12
2187
0.290
2:00 p.m.
8/12
1446
0.240
1:00 p.m.
7/20

4
66
0.066
10:00 p.m.
10/3
1360
0.060
7:00 p.m.
12/20
286
0.065
3:00 p.m.
10/7
Source:  SAROAD (1985a).
                                    5-52

-------
                         TABLE 5-12.  PEAK OZONE CONCENTRATIONS AT EIGHT SITES IN NEW YORK CITY
                                            AND ADJACENT NASSAU COUNTY, 1980
en
Site
Susan Wagner H.S.
Mabel Dean H.S.
Woolsey Post Office
Mamaroneck
P.S, 321
Sheepshead Bay H.S.
Queens College
Eisenhower Park
Site
no.

1 •
2
3
4
5
6
7
8
No. 1-hr
averages
>0.12 ppm

20
19
37
0
24
44
51
7
Days
with 1-hr
averages
>0.12 ppm

8
10
6
0
9
12
15
2
Four highest 1-hr daily
values, ppm, and date
1st
0.174
(8/28)
0.155
(7/21)
0.188
(7/20)
0.092
(6/14)
0.148
(7/26)
0.184
(7/31)
0.174
(8/28)
0.175
(8/28)
2nd
0.152
(7/18)
0.154
(7/26)
0.163
(7/21)
0.080
(8/28)
0.146
(8/28)
0.173
(7/18)
0.164
(7/21)
0.158
(7/21)
3rd
0.140
(7/26)
0.144
(7/18)
0.151
(7/22)
0.076
(7/2)
0.145
(7/18)
0.165
(8/7)
0.163
(6/14)
0.119
(7/20)
4th
0.131
(9/1)
0.139
(8/28)
0.148
(8/28)
0.075
(7/26)
0.144
(7/9)
0.164
(7/14)
0.159
(8/24)
0.118
(8/24)
                    Sites monitored during the Northeast Corridor Monitoring Program (NECRMP);  site
                    numbers assigned here are keyed to Figure 5-21.

                   Source:  Smith (1981).

-------
      SITES
1. Susan Wagner High School
2. Mabel Dean High School
3. Woolsey Post Office (Astoria)
4. Mamaroneck
5. Public School 321
6. Sheepshead Bay High School
7. Queens College
8. Eisenhower Park (Nassau Co.)
NEW JERSEY
    Figure 5-21. New York State air monitoring sites for
    Northeast Corridor Monitoring Program (NECRMP).

    Source: Smith (1981).
                          5-54

-------
yet there were 51 hours of ozone concentrations exceeding 0.12 ppm and 15 days
when ozone  levels  exceeded 0.12 for at  least  1 hour at the Queens College
Station; whereas  corresponding values were recorded at  Eisenhower Park for
only 7  hours  during 2 days.   At both  stations, data for about 94  percent of
possible hours were recorded as valid.
     The range  of  first-,  second-, third-, and fourth-highest values,  along
with frequencies  of values >0.12 ppm, establishes  an apparent concentration
gradient in the area from sites 6 and 5 to site 4.  Exposures of human popula-
tions living and working in metropolitan New York City could differ appreciably
if  the  residences  were  located and all activities were centered  in lower.
Brooklyn as opposed to the upper Bronx.  Differences in peak concentrations at
the respective  sites  varied by date  (6/14 to  8/28) and  by  level on the same
day (8/28), when  ozone was 0.080 ppm  at  site 4 and  0.174 at site 7, a differ-
ence of 0.094 ppm (a factor of 2.18).
     Intracity  gradients  in ozone  concentrations  have also been reported by
Kelly et al.  (1986) for a  1981  study in Detroit, MI.   Ozone concentrations
were measured  for about  3  months at 16 sites in the metropolitan Detroit area
and in  nearby Ontario, Canada.   Values at  15 sites  were  correlated with those
at  a  site  adjacent to the  Detroit  Science Center,  about 3  km north  of the
central business  district  'in Detroit.   In general, the  correlation decreased
as  distance from the  Science Center  site increased; and,  in general,  the
actual concentrations  increased with distance  from that  site toward the north-
northeast.  The highest  ozone concentrations  were recorded at sites  about 10
to  70  km  north-northeast of the urban core.  At greater  distances  or  in other
directions, ozone maxima decreased.
5.5.2.3   Indoor-Outdoor Concentration Ratios.   Most  people  in  the   United
States  spend  a large proportion of their time  indoors.   Although  knowledge  of
actual  exposures  of populations to ozone is  essential for optimal  interpreta-
tion  and  use  of the results  of epidemiological studies, essentially  all air
pollution monitoring is done  on outdoor  air.   The modeling  of actual  exposures,
as  opposed to  potential  exposures, therefore  necessitates knowing  general
activity  patterns  and at  least approximate indoor/outdoor  ratios (I/O) of
ozone.
      For  slowly reacting compounds such as carbon monoxide, long-term average
ratios  of  indoor to outdoor  concentrations tend  to be close to unity,  in the
absence of indoor  sources.  'Over  short  time periods, however, the I/O  may be
                                     5-55

-------
significantly different because of non-equilibrium factors (Yocom, 1982).   The
situation for  reactive pollutants such as ozone  is  much more complex, and
reported I/O values  for ozone  are highly variable.   Unfortunately, the number
of experiments and  kinds  of structures examined to date provide only limited
data for  use in  modeling indoor exposures.   Yocom  (1982)  has  presented  a
chronological summary of studies in which either ozone or photochemical oxidant
indoor-outdoor gradients  in  buildings  and  residences were measured.   Studies
summarized by  Yocom have  been conducted over the period  1971  through the
present (one ongoing study) by five research  organizations:  University  of
California,  the  California Institute of Technology,  GEOMET,  Inc.,  Lawrence
Berkeley Laboratory, and  TRC Environmental  Consultants.  Structures examined
have included  hospitals,   schools, office buildings, single-dwelling homes,
"experimental"  dwellings,  apartments, and mobile homes.   Private homes included
those with  and without gas stoves and fireplaces,  and those inhabited  by
smokers versus nonsmokers.   Areas of the country in which the buildings  were
located ranged from  Southern California to Boston,  including such cities  in
between as  Denver,   Chicago, Washington, Baltimore,  Pittsburgh, as well  as
other unspecified locations (Yocom, 1982).   The indoor-outdoor ratios reported
from these  studies   are summarized in Table 5-13 and are discussed later.
     Among newer  reports  of indoor/outdoor gradients in ozone concentrations,
also summarized  in  Table  5-13, are the  studies  of  Stock et  al.  (1983) and
Contant et  al.   (1985), undertaken to provide exposure assessments  for  an
epidemiological study of asthmatics in Houston, Texas,  in 1981 (see Holguin et
al., 1985, in  Chapter  11).   Stock et al. (1983)  found  I/O ratios of nearly
zero to 0.09 in  one  air-conditioned  residence  (maximum  ozone  concentration of
5 ppb) and an I/O ratio of 1.0 in a residence ventilated completely by outdoor
air.  (Indoor  concentrations were monitored via  a sampling manifold  connected
by Teflon lines  to  a chemiluminescence analyzer  in  a mobile  van  parked near
the respective houses.)  Contant et al. (1985), in a continuation and extension
of  the  same study,   used  "personal  monitors"  (i.e.,  portable analyzers) to
measure indoor and  outdoor air in the  immediate  environs  of  participants  in
the asthma study.  Mean, median, and maximal ozone concentrations, respectively,
were 10.8, 5,  and 147 ppb  indoors  and  51.8,  42, and 250 ppb outdoors.  The
respective I/O ratios were 0.21, 0.12, and 0.59.
     Davies et al.  (1984) found I/O ratios of 0.7 ± 0.1 at an art gallery in
rural Norwich, England, where  the outside ozone  concentrations  ranged  from 18
                                    5-56

-------
             TABLE 5-13.  SUMMARY OF REPORTED INDOOR-OUTDOOR OZONE RATIOS
           Structure
   Indoor-outdoor
     ratio (I/O)
      Reference
Residence
  (with evaporative cooler)

Office
  (air-conditioned; 100% outside
   air intake)
  (air-conditioned; 70% outside
   air intake)

Residence

Residence

Residence
  (gas stoves)
  (all electric)

Office

School room
Residence

Residences (1 each)
  (air-conditioned)
  (100% outside air; no
   air conditioning)

Residences (12)
  (ai r-condi tioned)
Art gallery
  (three modes of ventilation
   in each 24-hr period:
   recirculation, mixture of
   recirculated and outside
   air, and 100% outside air)
0.60*



0.80 + 0.10

0.65 + 0.10


0.70

0.50 to 0.70
0.19
0.20

0.29

0.19
(maximum concn.)

0.10 to 0.25
0.00 to 0.09
1.0
0.21 (mean concn.)
0.12 (median concn.)
0.59 (maximum concn.)

0.70 + 0.10
(mean~concn.)
Thompson et al. (1973)
Sabersky et al. (1973)
Sabersky et .al.  (1973)

Moschandreas et al. (1978)

Moschandreas et al. (1981)



Moschandreas et al. (1978)

Berk et al. (1980)


Berk et al. (1981)

Stock et al. (1983)




Contant et al. (1985)



Davies et al. (1984)
 Measured as total oxidants.
                                          5-57

-------
 to 58 ppb in the 3-week study period.   The interior  of the  modernistic  gallery
 had few reactive surfaces (two plate glass walls and two walls and ceiling of
 painted aluminum, with  minimal  other  walls and furnishings).  In  addition,
 ventilation cycled between recirculation air,  a mixture of  outside  and  recireu-
 lation air, and 100 percent outside air over each  24-hour period.
      The results  of all  the  studies  cited above  are  shown  in Table 5-13.
 These results are highly  variable,  to say the least.   The  variability  is  not
 surprising, however,  considering  the  diversity of structures and  locations
 included in the studies.   In  Table 5-13,  the highest  I/O,  1.0,  was reported
 for a residence ventilated  by outside air (Stock  et al., 1983).  The second-
 highest I/O, 0,80,  was determined  by smog-season measurements made  in a multi-
-story, air-conditioned  building  on the  Pasadena  campus of  the  California
 Institute of Technology (Sabersky  et al.,  1973).   Air exchange in this  building
 was at a rate  of  10 changes per hour with 100 percent outside air (i.e.,  no
 recirculation of inside air).   For another Cal  Tech building, in which there
 was a mixture of 70 percent outside air and 30 percent recirculated inside air,
 Sabersky et al.  (1973)  found an  indoor-outdoor ozone ratio  of  0.65.   The
 lowest indoor-outdoor ozone concentration ratios shown in Table 5-13 are those
 reported by Stock et  al.  (1983) for one residence;  the  indoor concentration
 recorded is barely above the limit of  detection of the analyzer.
      A relatively large number  of factors can affect the difference in ozone
 concentrations between  the  inside of  a structure and  the  outside  air.  In
 general, outside air  infiltration or exchange rates, interior air circulation
 rates, and  interior surface composition (e.g., rugs,  draperies,  furniture,
 walls) affect  the  balance between replenishment and decomposition  of  ozone
 within buildings (Thompson  et al.,  1973;  Sabersky et al.,  1973;  Berk et al.,
 1980; Moschandreas  et al.,  1978).   The rate at which  exterior air enters a
 building depends on  local  wind speed  and direction, on  how  well-sealed the
 building is, on how frequently doors and windows are opened,  and  on the operat-
 ing characteristics and cycles of heating/air conditioning/ventilating  systems.
 A significant  factor  that increases infiltration is an increasing temperature
 differential between warm  interior air and cold outside air (Moschandreas et
 al., 1978), although  a  high differential would be unusual  for a photochemical
 smog period.   Moschandreas  et al.  (1978)  reported exterior-interior exchange
'rates ranging  from  ten  changes per hour in an office  building to one change
 every 5 hours  in  a residence.  At the  higher exchange rates, inducted ozone
                                     5-58

-------
remains at a level  indoors that is closer to the outdoor level.   As the exchange
rate decreases, surface  decomposition  processes can result in progressively
lower  equilibrium  ozone  concentrations.  Other factors, such  as  relative
humidity, also affect decomposition.   The half-life of ozone inside residences
has been  estimated at 2 to 6 minutes  (Moschandreas  et al.,  1978;  Mueller
et al., 1973; Sabersky et al., 1973), while its half-life in an office environ-
ment has been  estimated  at 11 minutes (Mueller et al., 1973).  These results
are indicative of  the relatively  rapid  reaction rate  that can be expected  for
ozone in a building or room environment.  The problem of I/O values in buildings
was the  subject of a model  development  program  by  Shair and Heitner  (1974) in
which  they  tried to  account for ventilation and for  losses by reactions and
surface scavenging.  Considering  the research  results  shown in Table 5-13  and
summarized  by  Yocom  (1982), any estimates of  indoor  ozone exposures must  be
considered as having a large degree of probable variability.
     Automobiles and other  vehicles  constitute another indoor environment in
which  people may spend  appreciable amounts of  time.   As with buildings, the
mode of  ventilation  and cooling  helps  determine  the  inside concentrations.
Peterson and  Sabersky (1975)  reported ozone  measurements made  inside and
outside  a  car traveling  a  freeway in southern California.   Concentrations
inside were  higher when  the air-conditioning was operating on "maximum" than
when it  was operating on "normal,"  largely  because at higher air turnover
rates ozone intake exceeded ozone decay.  When windows were opened, an equili-
brium between inside and outside ozone concentrations was established in about
3 minutes,  with inside concentrations  then  remaining  significantly lower than
outside concentrations as the result of the exponential decay of  ozone from
contact with interior surfaces.  Inside ozone concentrations were one-third or
less of  the outside  concentrations (I/O <  0.33),  on  the average, under all
ventilation and cooling conditions.  Ozone concentrations measured outside the
car were  lower than  those measured at  nearby fixed sites, a result attributed
to  scavenging  of  ozone  by  nitric oxide.  In the exposure study of Content et
al. (1985)  in  Houston,  I/O ratios from 49  measurements inside vehicles were  '
0,44 for mean, 0.33  for  median,  and  0.56 for maximum  concentrations  measured.
Driving routes and ventilation and cooling modes were  not described.
     Concentrations  of  ozone inside vehicles  obviously depend  first on the
concentration outside and then on the various effects  of cooling and ventilation
conditions  and  surface  decay.   Thus, the titration of 03  by NO on and near

                                    5-59

-------
roadways is  an  important  factor.  The  titration  reaction  is rapid but depends
on the amount of NO emitted, which in turn depends on the amount of traffic on
a given road or freeway.  Rhodes and Holland (1981) found average ozone levels
on the  nine-lane  freeway between Los  Angeles  and San Diego,  which carries
200,000 cars each day, to be one-ninth to one-sixth the levels found 30 meters
upwind.
     At present there are no long-term monitoring data on indoor air pollutant
concentrations  that  are comparable to the  concentration  data available for
outdoor locations.  Thus, for estimates of the exposure of building or vehicle
occupants to  ozone  and other photochemical oxidants, it is necessary to rely
on extrapolations of very limited I/O data such as those presented above and
in Table 5-13.
5.5.2.4  Altitudinal and Latitudinal Variations.   Concentrations of ozone vary
with  altitude and with  latitude.   These variations occur for  a number of
reasons, including the nature of the interchange mechanisms involved in strato-
spheric- tropospheric exchange, the decay of stratospheric ozone as it traverses
the troposphere, and the known production of ozone in the apparently unpolluted
free  troposphere  at  certain altitude ranges above mean  sea level (MSL).   In
addition, other meteorological,  physical,  and chemical factors contribute to
the concentration gradients found with changes in altitude and latitude at the
surface of the  earth.
     Among specific factors known to contribute to variations in ozone concen-
trations with altitude and  latitude,  and,  to  a much  lesser extent,  longitude,
are the following:

      1.   Incursions of stratospheric ozone;
      2.   Global circulation patterns, along with accompanying altitudinal and
          geographic differences in atmospheric pressure, direction of prevail-
          ing winds, wind speeds, temperature, humidity, and precipitation;
      3.   Intensity and spectral distribution  of sunlight; and
      4.   Meteorological factors, such as mixing heights, frequency and persist-
          ence  of inversions, upslope  flows in mountainous areas, and cloud
          cover and precipitation patterns.

      Most of  these factors were discussed in Chapter 3, as well as in the 1978
criteria  document (U.S. Environmental Protection  Agency,  1978).   A few are
briefly  noted here  to present an  overview  of the macroscale differences in
ozone concentrations that may be expected.  Concentration gradients found with
increases  in altitude at the earth's  surface are presented  in  more  detail
because of their  relevance  for specific  kinds  of  exposure assessments.
                                    5-60

-------
     Latitudinal differences in  ozone  concentrations have been demonstrated.
These are in part attributable to the variations, temporal as well as geogra-
phic, in actinic  irradiance  at the earth's surface.  Both the intensity and
the spectral distribution of sunlight have direct effects on the photochemical
reactions that  initiate and sustain oxidant  formation  (Chapter 3 and U.S.
Environmental  Protection  Agency, 1978).   The  effect of  latitude  on ozone
formation is significant during the winter but not during the summer, since in
summer,  for  the contiguous  United  States, the maximum  light  intensity  is
fairly constant and  the duration of the  solar day varies only slightly with
latitude.  Light  intensity varies somewhat with longitude in the summer, with
the highest intensities occurring in the western states.
     Incursions of stratospheric ozone  also contribute to variations in ozone
concentrations  with  latitude.    These  are expected to be strongest  in  the
mid-latitudes of the northern hemisphere because of the nature of the mechanisms
by which stratospheric-tropospheric exchange occurs.
     The 1978 criteria document presented discussions of the effects of tropo-
pause-folding events  (TF) and  of the seasonal tropopause adjustment  (STA)  and
small-scale  eddy  transport  (SSET)  mechanisms  on stratospheric-trospheric
exchange.  As described previously  in  Chapter 3 and in the 1978 document, TF
events would be expected to produce sporadic increases in ground-level  ozone
concentrations, resulting from strong  incursions  of stratospheric ozone,  in
the  southern and  eastern  United  States  (latitudes of < 37°N and  longitudes of
< 90°)  (U.S. Environmental Protection Agency, 1978).  The STA-produced incur-
sions also  occur  in  these latitudes.   Both STA and  TF produce incursions  in
the winter or early spring.
     Logan et al.  (1981) summarized earlier measurements of ozone in the lower
troposphere.  Hemispheric asymmetry  is  apparent in the  data they presented,
with higher concentrations  occurring in the northern hemisphere than in  the
southern.  Also apparent is a seasonal   increase in lower-tropospheric ozone in
the  summer  at  mid-latitudes  in  the northern  hemisphere.  The  authors noted,
however, that ozone  data for the lower  troposphere are sparse, especially for
the  southern hemisphere.   Compared  with ozone data from two east-coast sites
at  mid-latitudes  in  the  northern  hemisphere, ozone data from one  site  in
Australia showed  (1)  little seasonal variation at <3 km (MSL),  (2)  about 40
percent  lower concentrations at  2  km (MSL) in summer, and (3)  similar concen-
trations in  autumn  and winter.  Altitudinal profiles of ozone concentrations
                                    5-61

-------
for the tropical northern  hemisphere  (Panama, 9°N, and Hawaii, 19.5°N) show a
considerably  lower  gradient with altitude  than  that seen at mid-latitudes
(Oltmans,  1981; cited  in  Logan et a!,,, 1981).   Logan et al.  (1981)  noted the
difficulty, given the sparseness of the data, of separating hemispheric effects
from variations in concentrations attributable to other factors (e.g., oceanic,
coastal, inland meteorology).
     In a  more recent report,  the  same investigators presented other data
showing hemispheric  differences in  ozone concentrations and  the seasonal
distributions of those concentrations (McElroy et al., 1985).   Ozone concentra-
tions in the  lower  troposphere measured at  Cape  Grim,  Tasmania (ca.  41°S),
showed strong similarities, including seasonal  distributions,  to those measured
at three different sites,  at different longitudes, in Canada (latitudes of ca.
53°N to 59°N).  Ozone  concentrations  from sites in West Germany,  Switzerland,
and the United States showed even greater similarity, both in level  and seasonal
distribution.
     Although of interest  and concern when  estimating global ozone budgets,
variations in ozone concentrations with latitude have little practical signifi-
cance for assessing exposure within the contiguous United States.   The effects
on ozone concentrations of latitude, as well as longitude, within the contiguous
states are minor.
     A number of additional reports are available on the increase of tropospheric
ozone concentrations with altitude (e.g., Kroening and Ney, 1962; Galbally and
Roy, 1980).   The data  cited  above showing latitudinal effects also show  alti-
tudinal effects, particularly at mid-latitudes in both the northern and southern
hemispheres  (Oltmans,  1981;  cited in Logan  et al., 1981).  At mid-latitudes,
the  increase  in  background ozone concentrations rises rather sharply (from a
spring-summer mixing ratio of about 40 to  60  ppb to one of about 100 ppb)
between altitudes of  about 8 and 12  km (ca. 5 to 7.4 mi, or ca. 26,400  to
39,600  ft).   Between the  surface and about  2 km  (ca.  1.2  mi, or ca. 6600  ft),
another relatively  sharp  increase in ozone  concentrations is observed for the
mid-latitudes  (from a  spring-summer mixing  ratio of  about  35  ppb to one of
about 50  to  55 ppb)  (Oltmans,  1981;  cited in Logan et al., 1981).  Other data
corroborate these findings.  Seller and Fishman (1981) reported ozone measure-
ments taken on flights in  remote tropospheric air during July and August 1974.
Their data also show that  ozone concentrations increase with increasing altitude
and  in  general  substantiate the accepted view  that  the  lower  atmosphere and
the  surface of the earth act as ozone sinks.
                                     5-62

-------
     Increases in ozone  concentrations  at higher altitudes are of potential
concern relative to  exposures  sustained in-flight by airline passengers and
employees on  high-altitude  flights  (see Chapter 11).   Air-filtration systems
are employed on airplanes, however,  that reduce ozone concentrations.  Increases
in concentrations with altitude at  lower  altitudes (e.g.,  in the zero to 2 km
range) have potentially  greater  importance,  since these gradients may affect
forest ecosystems in mountainous  areas,  (It must be borne in mind,  however,
that atmospheric pressure decreases with altitude and thus, for a given concen-
tration, the mass of ozone per cubic meter of air also decreases with altitude.)
Contributors  to  increased ozone  concentrations  at  the higher altitudes of
relevance for forest ecosystems include stratospheric ozone intrusions and ozone
transported aloft and conserved overnight in nocturnal inversions layers.   The
latter,  in  particular,  may be an important  consideration for the accurate
assessment of the ozone  exposures sustained  by mountain forest ecosystems, as
discussed below (see Chapter 7 also).
     While a  number  of  reports contain data  on  ozone concentrations at high
altitudes  (e.g.,  Coffey  et  a!.,  1977;  Reiter,  1977b; Singh  et  a!., 1977;
Evans, 1985; Lefohn  and Jones, 1986), fewer reports are available that present
data  for  different  elevations at the same locality.   Studies by Berry (1964)
and  Miller  and Ahrens  (1969) are  among  earlier reports  documenting ozone
gradients with  altitude  in forested areas.  Two more recent studies are dis-
cussed  here  because  the  investigators  acquired and  presented  data  on the
diurnal patterns of  ozone at those sites, as well as on differences in ozone
concentrations at the respective  altitudes.  In  addition, the two studies were
conducted in different parts of the country and  presented data that  demonstrate
differences in  the  meteorological influences on the  concentrations  found  in
the  respective studies.
      The  mixed-conifer  forest ecosystem  of  the  San  Bernardino  Mountains in
California  is the most  thoroughly studied  forest ecosystem  in  the United
States  (see Chapter  7).   In conjunction  with vegetation and forest  ecosystem
studies conducted there  in the 1960s and  1970s,  a number of monitoring stations
were  established at  different elevations  on the  south-facing slopes  of the San
Bernardino Mountains.   The mountain range begins about  80 miles  east  of  Los
Angeles and its center is  almost  due east of Pasadena.  The first four monitor-
ing  stations  were established  at Highland, City Creek,  Mud  Flat,   and Rim
Forest, located as shown  in Figure  5-22.
                                    5-63

-------
   2745
                                                                         90
CD
EH
z
2135

1525

 915

 305
   0
                I
            REDLANDS
                 I   I   I
                                      I
                                   CITY CREEK
                                                —N-
                          HIGHLAND
                       I   I   I   I   I   I
        DEEP
        CREEK
           70

           50

           30

           10
           0
                      CD
                0
                          8     12    16   20    24
                                DISTANCE, kilometers
28
32
36
             Figure 5-22. Altitudinal sequence of monitoring sites in the San
             Bernardino Mountains.
             Source: Adapted from Miller et al. (1972).
                                    5-64

-------
     Total oxidant,  temperature,  and  relative  humidity (RH) were measured
continuously during 16 days in July and August at all but the Highland station
(Miller et a!., 1972).  The vapor pressure gradient was calculated from RH and
temperature data.  The  time  of the daily peak oxidant concentration was pro-
gressively later at  stations  of higher elevation, as  shown  in Figure 5-23.
Temperatures and  vapor-pressure gradients were also  progressively  lower at
higher elevations  at the  time of the  oxidant peak.  The average duration of
                                           3
oxidant concentrations  exceeding 200  |jg/m  (0.10 ppm)  was  9, 13,  9, and
8 hr/day  going from  lower  to  higher  stations for June, July, and August  1969.
The longer duration at City Creek (elevation 817 m; 2680 ft) probably coincided
with the zone where the inversion layer most often contacted the mountain slope.
                                                         3
The oxidant concentrations rarely decreased below 98 ug/m  (0.05 ppm) at night
on the slope of the  mountain  crest,  whereas they usually decayed to  near zero
at the basin station  (Highland, elevation 442 m; 1459 ft).
     The vertical and horizontal distributions of oxidant air pollution in the
Los Angeles Basin  have  been described by several investigators and presented
in the 1978 criteria  document (Blumenthal et a!., 1974; Edinger, 1973; Edinger
et al., 1972;  Miller et al., 1972; as cited in U.S. Environmental Protection
Agency, 1978).   In Los Angeles, a marine temperature inversion layer frequently
forms  above the heavily urbanized metropolitan area and often extends inland
as far as 144  km (90 miles),  depending  on season and time of day.   Surface
heating of  air under the  inversion  increases with  distance eastward in the
basin and often disrupts the  inversion by midmorning at its eastern edge.  The
marine temperature inversion layer  encounters  the  mountain slopes, usually
below  1200  m  (4000  ft).   In the morning, the  temperature inversion often
remains intact at  this  juncture, and  air pollutants are confined beneath it.
The heated  mountain  slopes act to vent oxidant air pollutants over the crest
of the mountains and  cause the  injection of pollutants into the stable inversion
layer  horizontally away from the slope.   Oxidant  concentrations  within the
inversion are  not  uniform, but occur  in  multiple  layers and strong vertical
gradients.  In some  cases, the inversion may serve as a reservoir for ozone,
which  may arrive at  downwind locations along the  mountain slopes relatively
undiluted because  of a  lack of vertical  mixing  within the  inversion layer  and
a lack of contact with  ozone-destroying material generated at  the ground.  The
important result of  the trapping of oxidant in these  layers is its prolonged
contact with high terrain  at  night.
                                    5-65

-------
§
Q.

    0.35



    °'3°


    0.25
S   0.20

§
J   0.15

O
»-   0.10
    0.05


       0
    0.35



i   0.30
Q.
Q.

H*   0.25
Z


—,  0.20
    0.15


    0.10


    0.05


       0
                            HIGHLAND
                            (442m; 1459 ft)
   95



   90



   85
              g  80
              Q.

              Ul
              «-  75
                                 6      9     12    15

                                        TIME, hours
                                                           18
                                                    21
   95



   90
".  85
ui
DC


1  8°
ui
a.

S  75
P
                 70



                 65
                                              VPG
                            CITY CREEK
                            (817m; 2696 ft)
                                        I
                                        I
                                                                           24
                                                                           18
                                                                           12
                                 6      9     12     15

                                        TIME, hours
                                                           18
                                                    21
                                                                                E
                                                                               UJ
    111
    CC.

    «>
    co
    Ul
    DC
    Q.

6   .§
    Q.
    <
           Figure 5-23. Relationship between elevation and diurnal patterns of
           total oxidant concentrations, temperature, and vapor pressure at
           four sites (A-D) in the San Bernardino Mountains, CA, July-August
           1969.

           Source: Miller et al. (1972).
                                    5-66

-------
    0.35
 E  0.30
 Q.
 Q.

H*  0.25
g   0.20

O

<   0.15

O

H   0.10



    0.05



       0
LL
O



K
             tt
             Ml
             O.
             2
             UJ
             95




             90




             85




             80




             75




             70




             65
                       MUD FLAT

                       (1110m; 3663 ft)
                                                              24
                                                                          18
                                                              12
                                       _L
                                          I
                                         I
                                       9     12     15


                                         TIME, hours
                                                       18
                                                     21
      en
     X

      E
      E
                                                                   UJ
     UJ
     cc
     3
     (0
     (A
     UJ
     K
     Q.

     K
     O
     o.
a.
Q.
0.35



0.30



0.25
Q   0.20
X
O
-J   0.15
    0.10



    0.05



       0
             Ul
             K
        E
        U4
        Q.

        5
    95




    90




    85



    80




    75




    70




    65
                                        1
                                                 I
                       RIM FOREST

                       (1725m; 5792 ft)
                                         VPG
24
                                                                          18
12
                            69     12     15

                                    TIME, hours
                                                           18     21
      S.
                                                                                UJ
                                                                               CC
                                                                               o
                                                                               01
                                                                               cc
                                                                               3
                                                                               W5
                                                                               W
                                                                               UJ
                                                                               DC
                                                                               Q.

                                                                               CC

                                                                               2
           Figure 5-23 (Cont'd). Relationship between elevation and diurnal
           patterns of total oxidant concentrations, temperature, and vapor
           pressure at four sites (A-D) in the San Bernardino Mountains, CA,

           July-August 1969.


           Source; Miller et al. (1972).
                                   -5-67

-------
     The studies  in  the  Los Angeles  Basin and  the  nearby San Bernardino Moun-
tains  underscore  the importance  of  understanding the interactions between
topography and  local  meteorology for specific areas.  Consideration of those
factors is essential  for accurate exposure assessments at  higher altitudes,
especially in the absence of monitoring sites and adequate aerometric data.
     Wolff et al.  (1986) conducted a study of the effects of altitude on ozone
concentrations  at three  sites located at three  separate elevations  on High
Point Mountain  in northwestern New Jersey.   Ozone concentrations and temperature
were measured at  ground level at the respective sites.  In addition, temperature
and relative humidity were measured above the surface by balloon-borne sensors,
with wind direction and speed data obtained by means of a tracking theodolite.
     Data for several days (July 1975) indicate that in mid-day, when atmos-
pheric mixing was good,  vertical profiles were nearly constant, with concentra-
tions  increasing  only slightly  with  elevation.   Likewise,  the daily ozone
maxima were similar at different elevations.   At night, however, ozone concen-
trations were nearly zero in the valley (the  lowest-elevation  site) and in-
creased with elevation.   Comparison  of the ozone  dosages at the three sites
(number of hours > 0.08 ppm) showed that greater cumulative doses were sustained
at the  higher elevations (Table 5-14).  Comparable  data from  an urban area
(Bayonne, NJ) about  80  km southeast of High  Point Mountain showed that the
cumulative doses  were higher at all  three of  the  mountain  sites than in the
urban area (Table 5-14)  (Wolff et al., 1986).

     TABLE 5-14.  COMPARISON OF OZONE CONCENTRATIONS AT THREE DIFFERENT
                  ELEVATIONS, HIGH POINT MOUNTAIN, NJ, AND
                         AT BAYONNE,  NJ, JULY 1975
Locati on/el evati on
High Point, 500 m
'High Point, 300 m
High Point, 140 m
Bayonne, sea level
Source: Wolff et al.
1-hr
max,
ppb
66
61
59
69
(1986).
24- hr
mean,
ppb
49
38
26
33

1-hr
max,
ppb
130
110
120
119

24- hr
mean,
ppb
81
61
52
48

No. hr
>80 ppb
13
9
9
7

                                    5-68

-------
       The maximum depth of  the  nocturnal  inversion at High Point Mountain was
  about 900 m (ca. 2970 ft).  Extrapolating from the gradient observed from the
  ground sites,  Wolff et al. (1986) estimated that mountain peaks exceeding 900
  m (2970 ft) would sustain exposures to relatively constant diurnal concentra-
  tions that  would approximate the maximum ozone  concentration  found at the
  valley site.  The investigators concluded from their concentration and meteorol-
  ogical data that elevated, mountainous sites in the eastern United States may
A'be expected to be exposed to higher ozone dosages than valley sites throughout
-1 the year.
  5.5.2,5  Vertical Gradients at Ground Level.   Just as macroscale variations in
  ozone concentrations have  been observed  by measuring vertical  and horizontal
  profiles at various altitudes,  so too have microscale vertical variations at
  ground  level  been  observed as a  function of  placement of sampling probes.
  Data drawn  from  a  recent study on rural ozone concentrations  illustrate the
  effects on  concentration  data  of the placement of monitoring  probes.  These
  effects are pertinent  for  vegetation exposures,  in particular, although the
  vertical gradient at ground level has obvious implications for human exposures
  as well.
       Pratt et al. (1983) studied concentrations of ozone and oxides of nitrogen
  in the  upper-midwestern  part of the United States.   Concentration data were
  obtained over 4  years  by means of monitors placed at two different sampling
  heights (ca.  3  and 6,  or  3  and 9 meters) at three  air quality monitoring
  sites:  LaMoure  County,  North  Dakota;  Traverse County, Minnesota; and Wright
  County, Minnesota.  All  stations  were rural sites.  The mean ozone concentra-
  tions did  not differ  greatly among the sites, but in at least some instances
  the mean differences between sampling heights were as large or larger than the
  differences among the  scattered sites.   Table 5-15 presents the  mean ozone
  concentrations measured at two separate sampling heights (Pratt et al., 1983).
  Annual  average concentrations  were  1 to 3 ppb lower at the 3.05-meter height
  than at the 6.10 or 9.14-meter heights, reflecting the depletion of ozone near
  the surface.  As might be expected, the gradient was especially conspicuous at
  night because of continued surface scavenging and a  decrease  in  the  rate  of
  transfer from  layers  aloft.   The concentrations of ozone occurring at these
  sites were  near  background in  all years  measured.  In areas with  higher  ozone
  concentrations,  one  would expect to  see larger absolute gradients between
                                      5-69

-------
                      TABLE 5-15.   MEANS AND STANDARD  ERRORS  OF  OZONE CONCENTRATIONS MEASURED OVER 4 YEARS
                     AT TWO SAMPLING HEIGHTS AT THREE  STATIONS IN  THE RURAL,  UPPER-MIDWESTERN UNITED STATES
                                                           (ppb/v/v)

CJl
1
•-4
o

Site
LaMoure
County, ND
Traverse
County, MN
Wright
County, MN
Sampling
height, m
3.
9.
3.
9.
3.
6.
05
14
05
14
05
10


1977
22.97
(6
24.14
(6
24.44
(9,
26.29
(9,

± 0.20
,413)a
± 0.19
,410)
± 0.19
672)
± 0.19
810)
-

1978
35.30 ± 0.20
(15,218)
35.67 ± 0.19
(15,220)
35.2 ± 0.19
(22,675)
36.77 ± 0.19
(22,624)
34.64 ±0.23
(17,437)
35.61 ±0.23
(17,440)

1979
30.80 ± 0.16
(21,064)
32.25 ± 0.15
(20,470)
30.64 ± 0.17
(19,900)
32.39 ±0.16
(19,289)
28.71 ± 0.18
(17,771)
29.27 ±0.18
(17,775)
Years
1980
34.98 ± 0.25
(17,157)
37.53 ± 0.25
(17,157)
34.66 ± 0.23
(22,629)
37.60 ±0.23
(22,625)
J4.27 ± 0.26
(18,222)
35.16 ±0.27
(18,280)


31.
34.
28.
31.
31.
32.

1981
92 ± 0.26
(6,699)
23 ± 0.26
(6,364)
78 ± 0.23
(7,141)
08 ±0.22
(7,142)
60 ± 0.24
(7,764)
28 ± 0.24
(7,766)

1977-1981
32.26
(66,551)
33.82
65,597
32.09 .
(82,017). .
34.21
(81,550)
32.42
(61,254)
33.21
(61,261)
 The numbers in parentheses refer to the number of hours  of monitoring included in the reported values.   Values are based on
 all valid: data per site.   For each sampling height at each site,  values  for three monitors separated by 76 m are included
 in the calculations.   Monitoring was conducted only during the second half of 1977 and only until  30 June in 1981.

Source:  Pratt et al.  (1983).

-------
monitors at different heights.  In fact, the careful measurement of concentra-
tion gradients over di.stancesrof 1 to 10 meters above a surface is a recognized
method for estimating the scavenging potential of the surface.
     In this  context,  it should be noted that the dry deposition of ozone on
various surfaces  has  been fairly extensively studied.  Summaries on surface
scavenging and other natural removal processes appeared in the previous criteria
document for ozone and other photochemical oxidants (U.S.  Environmental Protec-
tion Agency, 1978) and in the review by the National Research Council (National
Research Council,  1977).   Newer review  articles  on  dry  deposition and  surface
scavenging have  been  published by Wesely (1983) and Gal bally and Roy (1980),
among others.
5.6   CONCENTRATIONS  OF PEROXYACETYL NITRATE  AND  PEROXYPROPIONYL NITRATE IN
     AMBIENT AIR
5,6.1  Introduction                                     '    .
     As noted  in  the introduction to this  chapter (Section 5.1), published
data on the concentrations in ambient air of photochemical oxidants other than
ozone are  not  comprehensive  or abundant.   Much more  is known now, however,
about their  atmospheric  concentrations  than was known when the 1978 criteria
document for ozone and other photochemical  oxidants was published.   Review  of
the data that  follow will  show  that  peroxyacetyl  nitrate (PAN)  and  peroxypro-
pionyl nitrate (PPN)  are the most abundant  of,the  non-ozone oxidants in ambient
air in the United 'States other than the inorganic  nitrogenous oxidants such as
nitrogen dioxide  (NO,,),  and  possibly nitric  acid  (HNO~),  in  some areas.  The
inorganic nitrogenous oxidants found in ambient air are  reviewed  in air quality
criteria documents on the nitrogen oxides and are  not treated in  this document
except for  the review of the role of nitric oxide (NO)  and N02  in atmospheric
photochemistry.
     Available data  show that PAN is toxic to  vegetation  and is potentially
toxic to humans.   No data indicating potentially  adverse  effects of PPN  are
available.    It exists in trace  quantities  and apparently no research has  ever
been  undertaken  to  determine  its potential  toxicity.   At least one study
(Heuss and  Glasson,  1968)  has reported  that a higher  homologue  of the  series,
peroxybenzoyl  nitrate (PBzN),  like  PAN,  is a lachrymator.   No  unambiguous
identification of PBzN in the ambient air  of the  United States  has  been  made,
however.      • '  ,-  .•!        • •      ''.  :  .     ,

                                     5-71

-------
     Given the  information  available  on PAN, the concentrations of PAN that
are of most concern are those to which vegetation could potentially be exposed,
especially during daylight hours in agricultural areas; followed in importance
by those both indoors and outdoors, in both urban and nonurban areas,  to which
human populations could potentially be exposed.   Most of the available data on
concentrations of PAN and PPN in ambient air are from urban areas.   The levels
to be  found  in  nonurban areas will be highly dependent upon the transport of
PAN and  PPN  or  their precursors  from  urban areas, since the concentrations of
the NO   precursors to these  compounds are considerably lower  in nonurban than
       4 ppb around 10:00 a.m. and  one of ~10 ppb between 4:00 and 6:00 p.m.
Seasonal data from Riverside  for June 1966 to June 1967 showed that PAN concen-
trations were highest in September 1966 and in March and June 1967.
     Total oxidants  (by Mast meter) in  these  Los Angeles  sites reached a peak
concentration (the same composite diurnal data as above) of  close to 140 ppb
in September 1965  at about the same hour  of  day as the PAN peak.   In October,
total  oxidants  peaked  at nearly 200  ppb, again coinciding in time with peak
PAN concentrations.   In Riverside, the morning PAN peak preceded the oxidant
peak  (~110  ppb around  noon)  by almost  2 hours but  the  afternoon  PAN peak
trailed  the  afternoon  oxidant peak (~160 ppb around 2:00 to 4:00 p.m.) by
about  2  hours.   The  ozone/PAN ratio  was variable from month  to month but was
generally lower from November through April  than during the  rest of the year,
                                   5-72

-------
i.e., PAN was  generally  a greater percentage of the total oxidant during the
November through April period.
     In the 1978 criteria document for ozone and other photochemical oxidants
(U.S. Environmental  Protection Agency, 1978),  additional  PAN data for  Los
Angeles as well  as  two other cities were presented.  Lonneman et  al.  (1976)
measured PAN and total  oxidants in Los Angeles in 1968.   In 118 samples col-
lected for the period 10:00 a.m. to 4:00 p.m. during the study, the median PAN
concentration was 13  ppb and the average PAN  concentration was 18 ppb.  The
median total  oxidant concentration  (measured  by UBKI) was 97 ppb and the
average was 117 ppb. Thus, the median oxidant/PAN ratio was 7.5 and the average
oxidant/PAN ratio was  6.5;  the median PAN  concentration  was  13.4  percent  of
the  median total oxidant concentration, and the average PAN concentration  was
15.4 percent of the average total oxidant concentration.
     Lonneman et al.  (1976)  conducted  similar  studies in  Hoboken,  New  Jerseyi
in 1970 and in St.  Louis, Missouri, in 1973.  Samples were measured during the
period 10:00 a.m. to  4:00 p.m.  over  the course of the study.   In Hoboken,  PAN
concentrations averaged 3.7 ppb.  Ozone concentrations (measured by chemilumi-
nescence) averaged 90.5 ppb.  In St.  Louis, PAN averaged 6.4 ppb, ozone (meas-
ured by  chemiluminescence)  averaged  50.1 ppb,  and  total  oxidants  (by UBKI)
averaged 74.3 ppb.
     From these  1966  through 1973  urban data,  it is clear that PAN concentr^r
tions in  urban  areas are appreciably  lower than those of ozone, even  in the
winter when PAN is higher relative to ozone than in summer.
     In addition to urban data, the 1978 criteria document (U.S.  Environmental
Protection Agency, 1978) also included PAN  concentration data from one nonurban
agricultural area,  Wilmington,  Ohio  (Lonneman  et al., 1976).   The  maximum  PAN
concentration observed  in 1500 samples taken  during August 1974 was 4.,1 ppb.
The  daily  maximum  PAN concentration rarely  exceeded  3.0  ppb  even  though the
daily maximum ozone concentration frequently exceeded 80 ppb.
     Conclusions reached  in  the 1978 document were  (1) that PAN concentrations
are  much  lower in  the ambient  air of  nonurban than of urban areas; and (2)
that ozone/PAN  or  total  oxidant/PAN ratios  vary with location,  such  ratios
being higher  in nonurban areas than  in urban;  that  is,  PAN concentrations  are
a  lower  percentage  of total oxidant concentrations in  nonurban  areas (U.S.
Environmental  Protection Agency,  1978).   From  data  presented in  the 1970
document,  it may be concluded  (1) that  oxidant/PAN ratios vary  with season;
                                    5-73

-------
(2) that  PAN concentrations  are  lower than ozone concentrations  in  urban
areas; and  (3)  that  PAN  and ozone concentrations exhibit similar but not
identical  diurnal patterns.                                     ,
     The historical data presented above have been given in some .detail because
the information  about  PAN conveyed by those data remains valid. -  Examination
of the more recent data presented in subsequent sections shows that newer data
extend and corroborate, for the most part, the findings of the older literature.

5.6.3    Ambient Air Concentrations of .PAN and Its Homologues in Urban Areas
     Additional   studies on  concentrations of PAN and  its homologues  in both
urban and nonurban areas  are  now available.  Newer data  are presented  in this
section in summary form, where possible; and the individual  studies or examples
presented are merely  a few of many,  chosen to represent current  knowledge
regarding ambient air PAN concentrations and their patterns.
     The existing  literature  on the concentrations in  ambient  air  of the per-
oxyacyl nitrates, PAN and its higher homologues5 has been compiled and examined
in two recent review articles (Temple and Taylor, 1983; Altshuller, 1983).   In
the first, Temple and Taylor reviewed the concentrations of PAN in the ambient
air in Europe, Japan, and North America in the context of the phytotoxicity of
PAN.  Altshuller,  in  the  second, reviewed the  published  concentrations of  PAN
and of PPN  in ambient air, also within and outside the United States.  In
addition, Altshuller  analyzed the relationships  to  ozone  of PAN  and  other
photochemical reaction  products.  The  reader is  referred to these  reviews  for
detailed information and for references therein.                •
     Table  5-16  presents a  summary of PAN  concentrations observed in the
ambient air  of  urban  areas of the United States.  Data  in this table include
the results of studies cited in Section 516.2 above, along with the results of
newer studies.  This table was derived  from the reviews  of Altshuller (1983) and
Temple and  Taylor  (1983), as  well  as  from a few additional sources (Jorgen et
a!.,  1978;  Lewis  et  a!., 1983).   The data are  summarized in  the  table by
region of the United States and by date, with the newer  studies .reported first
for each region.                      ,
     Because of  variations in diurnal patterns  of PAN by location and season,
and because  no  national,  uniform aerometric data base for PAN exists, few of
the data  reported  in Table 5-16 are  really comparable.   Thus, data in this
table lend themselves to general conclusions but  not to  the analysis of trends
                                   5-74

-------
                                                                                                   .3.
                                                                                                               ID
TABLE 1-16.  SUMMARY OF CONCENTRATIONS OF PEROXYACETYL NITRATE IN AMBIENT AIR IN URBAN AREAS OF THE UNITED STATES
PAN concentrations,
Site
West : . • •
Riverside, CA.
• W. Los Angeles, CA
Claremont, CA-
• Claremont, CA
Claremont, CA
en
i
-sj
01 Riverside, CA
Riverside, CA
Riverside, CA
Riverside, CA
West Covina, CA
West Covina, CA
Pasadena, CA
Riverside, CA
Downtown
Los Angeles, CA
Time
of yr

All year
June
Sept. -Oct.
Aug. -Sept.
Oct.
Apr. , May
July, Aug. ,
Oct.
All year
Oct.
Aug. , Sept.
Aug. , Sept.
July
Aug. -Apr.
Sept.-
Nov.
Yr

1980
1980
1980
1979
1978
1977
1977
1975 (May)-
1976 (Oct.)
1976
1977
1973
1973
1967-1968
1968
Hours
sampled

8 a.m. -8 p.m.
6:35 a.m. -1:35 p.m.
. 24 hr/day
Morning to late
evening
Late morning to
late evening
24 hr/day
.Late, morning to
evening
24 hr/day
Late morning to
early evening
23 hr/day
NA
7 a.m. -4:30 p.m.
24 hr/day
10 a.m. -4 p.m. .
No. days
sampled

365
2
.11
8
.5
10
10
533
3
24
NA
3
273
NA
Method9

GC-ECD
LP-FTIR '
• GC-ECD
GC-ECD
LP-FTIR
GC-ECD
LP-FTIR
"GC-ECD -
LP-FTIR
GC-ECD
GC-ECD
LP-FTIR
GC-ECD
GC-ECD
Avg.

NAb
1
'.13
4,
6
1.6
7
3.6
9
9
NA
30
NA
8
Max.

41.6
16
47 :
11
37
5.7
18
32
18
20
46
, 53
58
68
ppb
Monthly
mean

4.9.
NA
NA
NA
NA
NA
NA
NA
NA
NA
8.8
NA
4.6
NA
Original reference

Temple and Taylor (1983)
Hanst.et al. (1982)
Grosjean (1983)
Tuazon et al. (1981a)
Tuazon et al. (1981a;
1981b)
Singh et al. (1979)
Tuazon et al. (1980)
Pitts and Grosjean (1979)
Tuazon et al. ' (1978)
Spicer (1977)
Spicer (1977) .
Hanst et al. (1975)
Taylor (1969)
Lonneman et al. (1976)

-------
              TABLE 5-16 (continued).  SUMMARY OF CONCENTRATIONS OF PEROXYACETYL NITRATE IN AMBIENT AIR IN URBAN AREAS OF THE UNITED STATES
.in
 i
cr>
Site
Salt Lake
City, UT
Los Angeles, CA
Downtown
Los Angeles, CA
Southwest
Houston, TX
Houston, TX
Midwest
Dayton, OH
(Huber Heights, OH)
St. Louis, MO
St. Louis, MO
East
New Brunswick, NJ
New Brunswick, NJe
Hoboken, NJ
Time
of yr
July-
Sept.
Sept.-
Oct.
July-
Oct.
Oct.
July
July,
Aug.
June-
Aug.
Aug.
All year
Sept.-
- Dec.
June, July
Yr
1966
1965
1960
1977
1976
1974
1973
1973
1978 (Sept.)
-1980 (May)
1978
1970
Hours No. days
sampled sampled
7 a.m. -3 p.m.
8 a.m.-l p.m. 35
9 a.m. -Noon 9
8 a.m. -8 p.m.
24 hr/day 22
24 hr/day 20
23 hr/day 26
8 -to 24 hr/day 12
24 hr/day 400
(9600 1-hr values)
8 a.m. -6 p.m.
10 a.m. -4 p.m.
PAN
Method3
GC-ECD
GC-ECD
IR
GC-ECD
. GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
•GC-ECD
concentrations, ppb
Avg. Max.
54
31 214
~20 70
15.6
0.4 11.5
0.7 10
1.8 19
6.3 25
0.5 10.6
10.5
9.9
Monthly
mean Original reference
4.4 Tingey and Hill (1968)
38 (Sept.) Mayrsohn and Brooks
40 (Oct.) (1965)
Renzetti and
Bryan (1961)
Jorgen et al. (1978)
Westberg et al. (1978)
Spicer et al. (1976)
Spicer (1976)
. - Lonneman et al. (1976)
Lewis et al. (1983)
2.7 Brennan (1980)
3.7 Lonneman et al. (1976)
     GC-ECD, gas chromatography with electron-capture detector; LP-FTIR, long-path Fourier-transform infrared spectroscopy; IR, infrared spectroscopy.

     Not available.

    cSubset of data  reported by Lewis et al.  (1983), cited  in table above.


    Source:  Altshuller  (1983); Temple and Taylor  (1983); Lewis et al. (1983); Jorgen et al. (1978).

-------
over the past decade and a half or necessarily to analysis of between-city or
between-region similarities or differences.   For example, concentrations of
PAN in  Los  Angeles  appear  to have been a great deal higher in 1965 (Mayrsohn
and Brooks, 1965) than  in  1980 (Hanst et a!., 1982) for nearly the same time
of day, but the 1965 concentrations were measured in September and October and
the 1980 concentrations were  measured in June.  Other  data  from California
indicate that September and  October are more  likely  to be part of the smog
season  there  than June,  A  comparison by Temple and Taylor  (1983)  of PAN
concentrations in Riverside  in 1980 with those  in  1967 and 1968  indicates
little  difference.  Again, however,  the sampling periods were not identical
relative to averaging time  or time of year.   Tabulated data  from Temple  and
Taylor  (1983) have been plotted in Figure 5-24.
     In addition  to  compiling existing data on  the concentrations of  PAN in
ambient air, Altshuller (1983) also related PAN concentrations to ozone concen-
trations where data for both exist.  It must be borne in mind for this review,
as well, that  sampling  periods (years, time of year, number of measurements)
are not identical and in many cases are  not even similar between studies  (see
Table 5-16).  Neither are the averaging times over which samples were collected
and calculated within respective studies identical  or even necessarily similar.
Nevertheless, the data  of Altshuller  constitute a  comprehensive review and
examination of the  relationships  among respective photochemical oxidants in
urban  areas of the  United States.  Concentrations  of PAN as  a percentage of
ozone concentrations are given in Table 5-17, but Table 5-16 should be consulted
for information on sampling periods and averaging times.
     The existence  of peroxybenzoyl  nitrate (PbzN) in  ambient  air of urban
areas was postulated in the 1978 criteria document for  ozone and other oxidants
(U.S.  Environmental  Protection Agency, 1978), but  PBzN has  not  been  clearly
identified  in ambient air in the United States.  Hanst  et al. (1982) estimated
that as much  as  2 ppb PBzN would  be  clearly  detectable in FTIR  measurements
but  reported  no  clear absorption  band for PBzN in their measurements  during  a
smoggy  period  in Los Angeles.  They  estimated an  upper limit of 1 ppb PBzN
during  their  1980 study (the maximum ozone concentration was 272 ppb and the
maximum PAN concentration was 16 ppb during that period).
     The  only homologue of  PAN that  has been unambiguously identified in
ambient air in  the  United States  is  peroxypropionyl  nitrate (PPN).   In his
                                   5-77

-------
   0.07
   0.06
               J 1967-1968
-   I    I 198°
I 0.05
a

g
fc  0.04
ui
S  0.03
o
o
   0.02
   0.01
         AUG  SEPT  OCT  MOV   DEC   JAN   FEB  MAR  APR


         Figure 5-24. Comparison of monthly daylight average and
         maximum PAN concentrations at Riverside, CA, for 1967-
         1968 and 1980.

         Source: Derived from Temple and Taylor (1983).
                              5-78

-------
          TABLE  5-17.   RELATIONSHIP  OF  OZONE  AND  PEROXYACETYL NITRATE  AT
                   URBAN  AND  SUBURBAN SITES  IN  THE  UNITED  STATES
Site/year
California
Downtown Los Angeles, 1960
Downtown Los Angeles, 1965
Downtown Los Angeles, 1968
West Los Angeles, 1980
Pasadena, 1973
West Covina, 1977
Claremont, 1978
Claremont, 1979
Riverside, 1967-1968
Riverside, 1975-1976
Riverside, 1976
Riverside, 1977
Riverside, 1977
Southwest
Houston, TX, 1976
Midwest
St. Louis, MO, 1973
St. Louis, MO, 1973
Dayton, OH, 1974
(Huber Hts, OH)
East
Hoboken, NJ, 1970
New Brunswick, NJ, 1978-1980

Avg.
8
NA
13
9
10
20
7
4
8
9
5
4
4
3
13
5
2
4
4
PAN/OS, %
At 03 peak
7
7
NAa
6
8
12
6
4
NA
5
4
4
NA
3
NA
5
1
NA
2
Reference
Renzetti and Bryan (1961)
Mayrsohn and Brooks (1965)
Lonneman et al . (1976)
Hanst et al . (1982)
Hanst et al . (1975)
Spicer (1977)
Tuazon et al. (1981a, 1981b)
Tuazon et al . (1981a)
Taylor (1969)
Pitts and Grosjean (1979)
Tuazon et al . (1978)
Tuazon et al. (1980)
Singh et al . (1979)
Westberg et al . (1978)
Lonneman et al . (1976)
Spicer (1977)
Spicer et al . (1976)
Lonneman et al . (1976)
Brennan (1980)
 Not available.

Source:   Adapted from Altshuller (1983).
                                     5-79

-------
review of existing  literature, Altshuller  (1983) compiled data on the concen-
trations of  PPN  in  ambient air in  urban  areas.   In addition, He calculated
ratios  of  the  concentrations of  PPN  and  PAN.,   His data,  expressed  as
percentages [(PPN/PAN) x 100], are presented as Table 3-18 (Altshuller,  1983).
     As Altshuller  has  pointed out, average PPN concentrations are 10 to 30
percent of the average PAN concentrations shown in Table 5-18 except for those
reported for  San  Jose (8 percent) and Oakland (42 percent).   The maximum PPN
concentrations reported are highly variable, however, ranging from 0.13 ppb in
San Jose (August  1978)  to 6.0 ppb  in  Riverside  (month and year as well as
sampling period of day unknown).   Thus, the PPN/PAN ratio at maximum concentra-
tions of PPN is highly variable, as well.  Among more recent data, the maximum
PPN concentration was  5.0 ppb in St.  Louis  in August 1973.   Note,  however,
that the sampling period in St.  Louis was  10:00 a.m. to 3:30 p.m. Depending
upon temperature, concentrations of precursors, and  other factors, a true PPN
maximum may or may not have occurred by 3:30 p.m.
     Table 5-19 presents  PAN and PPN concentrations reported by Singh et al.
(1981) for three  cities,  Los Angeles, Oakland, and  Phoenix; as well  as PPN
concentrations as percentages of PAN concentrations (PPN/PAN x 100).   Comparison
of the  data  from these three cities helps demonstrate the variability of PPN
and PAN concentrations with location.

5.6.4   Ambient Air  Concentrations  of PAN and  Its Homologues  in Nonurban  Areas
     Data on  the  concentrations  of PAN and PPN in  agricultural and other non-
urban areas of the United States are sparse.  They include data from the study
done by  Lonneman  et al.  (1976)  in  Wilmington,  Ohio,  in  August 1974, and  cited
earlier  in Section  5.6.2.  In that  study,  measurements  made by GC-ECD from
10:00 a.m. to 4:00  p.m., local .time,  showed  a maximum  concentration for the
study period  of  4.1 ppb.  The average daily  maximum was 2.0 ppb.  While the
4:00 p.m.  cutoff used by Lonneman et  al. (1976) could possibly have resulted
in missing some peak PAN  concentrations, especially  in transported air masses,
the data are within the range reported by Westberg et al. (1978) and by Spicer
and Sverdrup  (1981) for 24-hour measurements at nonurban sites.   At the Sheldon
Wildlife Reserve, Texas,  Westberg et  al.  (1978)  found  a 24-hour  average PAN
concentration of 0.64 ppb and a maximum  concentration of 2.8 ppb for the study
period  (October).   Spicer and Sverdrup  (1981) found a 24-hour average  PAN
                                   5-80

-------
                               TABLE 5-18.   AMBIENT AIR MEASUREMENTS OF PEROXYPROPIONYL NITRATE (PPN) CONCENTRATIONS
                                     BY ELECTRON CAPTURE GAS CHROMATOGRAPHY AT URBAN SITES IN THE UNITED STATES
CJl
Site
Los Angeles, CA
Riverside, CA
Riverside, CA
Riverside, CA
San Jose, CA
Oakland, CA
Phoenix, AZ
Denver, CO
Houston, TX
St. Louis, MO
St. Louis, MO -,-.
.Chicago, IL
Pittsburgh, PA
Staten Island, NY
Period of
measurement/
no. of days (n)
April 1979 (13)
NAa (1)
April -May
July 1980 (13)
August 1978 (7)
June-July 1979
(13)
April -May 1979
(14)
June 1980 (14)
May 1980 (12)
10 Aug. 1973
(1)
-May- June, 1980
(9)
April -May, 1981
(13)
April 1981 (11)
March- April
1981 (11)
Period
of day
24
NA
24
24
24
24
24
24
24
100
.(LT
24
24
24
24
hr

hr
hr
hr
hr
hr
hr
hr
g-1530
hr
hr
hr
hr
PPJf,
Avg.
0.7
NA
0.3
0.2
0.08
0.15
0.09
0.05
0.11
3.0
0.66
0.05
0.05
0.20
ppb
Max.
2.7
6
1.8
0.9
0.13
0.5
0.33
0.32
0.63
5.0
0.25
0.13
0.07
3.1
PPN/PAN
Avg.
15
NA
21
16
8
28
12
10
14
17
23
12
17
27
,*
Max.
16
12
32
16
10
42
9
3
25
20
28
8
10
80
Reference
Singh
Darley
Singh
Singh
Singh
Singh
Singh
Singh
Singh
et
et
et
et
et
et
et
et
et
Lonneman
Singh
Singh
Singh
Singh
et
et
et
et
al.
al
al.
al.
al.
al.
al.
al.
al.
et
al.
al.
al.
al.
(1981)
. (1963)
(1979)
(1981)
(1979)
(1981)
(1982)
(1982)
(1982)
al. (1976)
(1982)
(1982)
(1982)
(1982)
             Not available.
            DLocal time.
Source:   Altshuller (1983).

-------
en
i
00
ro
                               TABLE 5-19.   CONCENTRATIONS OF PEROKYACETYL AND PEROXYPROPIONYL NITRATES
                                              IN LOS ANGELES, OAKLAND, AND PHOENIX, 1979a

                                                                 (ppb)
Los Angeles
Value
Mean
Std. dev.
Maximum
Minimum
PAN
4.98
4.48
16.82
0.03
PPN
0.72
0.67
2.74
NDb
PPN/PAN, %
14
_
16
-
PAN
0.36
0.42
1.85
0.05
Oakland
PPN PPN/PAN, %
0. 159 42
0.12
0.50 27
NDb

PAN
0.78
0.77
3.72
NDb
Phoenix
PPN PPN/PAN, %
0.09 12
0.08
0.33 9
NDb
 Measurements were made by gas-phase coulometry.

bNot detectable.


Source:  Singh et al. (1981).

-------
concentration of 0.50 ppb and a maximum of 6.5 ppb for the study period  (July
and August)  at Van Hiseville, New Jersey, in the New Jersey pine barrens; and
Spicer et  al.  (1983) reported averages of  0.46  ppb at a  nonurban  site  in
Indiana (Huntington Lake) and 0.74 at a nonurban site in east central Missouri
(42 km west of St.  Louis).
     In contrast with these  values,  24-hour average concentrations of PAN at
one nonurban and four remote sites (see Table 5-20) were reported by Singh et al.
(1979) to range from 0.08 to 0.30 ppb, with maxima at those sites ranging from
0.22 to 0.83 ppb.  The Texas, Ohio,  and New Jersey sites sustained  higher PAN
concentrations than the nonurban sites where Singh et al.  (1979) measured PAN.
The higher  concentrations may reflect the influence at those sites .of nearby
metropolitan areas.
     Concentrations  of  PAN  have recently been reported  by Singh and  Salas
(1983) for  a Pacific marine site, Point Arena, California, at which earlier
measurements were  also made  and reported by Singh  et  al.  (1979) (see Table
5-19).  Data collected  in  August 1982 showed concentrations  of  PAN ranging
from 0.01 to 0.12 ppb during the 5-day study period.   The average concentration
for the period was 0.032 ± 0.024 ppb.  Winds were west-to-northwest 90 percent
of the time  and northerly the rest of the time.   Modeled trajectories confirmed
that air masses  passing  over the site were of a marine origin.  The site is
thought to be free of manmade pollutants..
     Data from two nonurban  sites in Canada are of interest even though they
are outside  the United States.  Cherniak and Corkum (1981; cited in Temple and
Taylor, 1983)  measured PAN at a  nonurban site in Simcoe, Ontario, Canada, for
6 months.   Measurements  made by GC-ECD showed monthly means  of  <2  ppb and  a
maximum concentration during the study of 5.6 ppb.  At a  remote site  in the
Kananaskis  Valley  of Alberta, Canada,  monthly mean concentrations were <1 ppb
for samples  taken at half-hour intervals, 24 hr/day, for 110 days   (Peake et al.,
1983).  The  site,  located at the base of a mountain range and about  50 miles west
of Calgary,  is thought to be free of manmade pollutants, including  transported
pollutants.

5.6.5  Temporal Variations in Ambient Air Concentrationsof Peroxyacetyl Nitrate
5.6.5.1   Diurnal  Patterns.   Concentration data  obtained in the 1960s were
briefly discussed  in Section  5.6.2, where it was noted that the  first  criteria
                                    5-83

-------
                               TABLE 5-20.   CONCENTRATIONS IN AMBIENT AIR OF PEROXYACETYL AND PEROXYPROPIONYL NITRATES AMD OZONE
                                                         AT NONURBAN REMOTE SITES IN THE UNITED STATES
Site
Mill Valley, CA
Point Arena, CA
Badger Pass, CA

Reese River, NV

Jetmore, KA


Wilmington, OH



Van Hiseville, NJ

Reference
Singh et al. (1979)
Singh et al. (1979)
Singh et al. (1979)

Singh et al. (1979)

Singh et al. (1979)

Westberg et al.
(1978)
Lonneman et al.
(1976)


Spicer and Sverdrup
(1981)
Nature of site
Maritime
Clean-Maritime
Remote-high
altitude
Remote-high
altitude
Rural-
continental
Rural -
continental
Rural -
continental


Rural -
continental
Period of
measurement and
no. of days (n)
Jan. 1977 (12)
Aug. -Sept. 1978 (7)
Hay 1977 (10)

May 1977 (7)

June 1978 (7)

October 1978 (9)

August 1974 (9)



July-Aug. 1979 (31)

Period
of day
24 hr
24 hr
24 hr

24 hr

24 hr

24 hr

10:00
a.m.-
4:00
p,n.
24 hr

Average
concentrations
PAN PPN 03
0.30
0.08
0.13

0.11

0.25

0.64

NAb



0.50

0.04
N0a
0,05

0.04

NDa

NDa

NDa



NDa

38
NDa
46

39

31

47

NAb



36

Maximum
concentrations
PAN
0.83
0.28
0.22

0.26

0.52

2.8

4.1



6.5

PPN
0.11
NDa
0.09

0.09

NDa

NO3

NDa



NDa

03
0,55
NDa
54

56

53

148

107



161

Avg. PAN/
Avg. 03, %
0.8
NDa
0.3

0.3

0.8

1.4

NAb


*
1.4

 Not determined.

 Measured, but results not given in the reference.

Source:   Altshuller (1983).

-------
document for photochemical oxidants (U.S. Department of Health, Education, and
Welfare, 1970)  reported  concentrations and patterns  for PAN that remain valid
now.  In that document, the general proximity in time of PAN and oxidant peaks
was shown  in  data from Los Angeles and  Riverside,  California.  Maximum PAN
concentrations, although varying from location to location, generally occurred
in  midday;  i.e.,  late  morning  to mid-afternoon.   Figures 5-25  and  5-26, taken
from the 1970 criteria document (U.S. Department of Health,  Education,  and
Welfare, 1970), graphically present the diurnal patterns of PAN in Los Angeles
in  1965  and in Riverside in 1966.  The  occurrence  of the  second PAN  peak in
Riverside,  which  appears to trail  a  second  total oxidant  peak by  an  hour or
two, was  ascribed to  transport,  as verified by  the occurrence of maximum
oxidant concentrations at three receptor sites east of West Los Angeles (down-
town Los  Angeles, Azusa, and  Riverside),  at times  that corresponded, wind
speed factored  in, with  respective distances from West Los Angeles.
     Examples  drawn  from recent data substantiate  that  the general  diurnal
pattern  (as it appears  in  composite  diurnal data  averaged over a week,  a
month, or  longer) remains the  same as the pattern established  by data obtained
in  the mid-1960s.
     Using  FTIR spectroscopy,  Tuazon and coworkers  (Tuazon et  al.,  1978, 1980,
1981b) measured concentrations of  PAN at Claremont  and Riverside,  California,
over a  5-year period.   Concentrations of PAN  ranged from  about 5 to 40 ppb
over the course  of  the  study.  The diurnal  profiles  for  PAN  and  ozone at
Claremont  are shown  for 2  days of a multiday smog  episode in  October 1978 in
Figure  5-27 (Tuazon  et al., 1981b).  Note the qualitative  relationship of the
two pollutants, with  the peak concentrations  of the two  occurring almost
simultaneously.   The relationship between PAN  and  ozone concentrations and
behavior  in the atmosphere is  neither constant nor  monotonic,  however, as is
borne out  by  the  slight  differences in time  of occurrence  of their peak concen-
trations but  especially  by  the persistence of somewhat elevated PAN concentra-
tions before  return  to "baseline"  levels.  It appears that PAN concentrations,
in  this  instance at  least,  closely parallel  the nitric acid (HNO,) concentra-
tions, persisting after  ozone  concentrations have subsided.  The percentage of
PAN relative  to ozone differed slightly at the peak concentrations of the two
on  the 2 days.  On October 12,  the  peak PAN concentration was close to 6 percent
of  the peak ozone concentration; on October  13, the peak PAN concentration was
nearly 8 percent  of  the  peak ozone concentration.
                                    5-85

-------
£
Q.
Q.
CO*
z
o
I
UJ
o
z
o
o
Q.
Q
Q
X
O

I
0.20


0.18



0.16


0,14


0.12


0.10


0.08


0.06


0.04


0.02


  0
^H MMM ^^ 1
    AVERAGES:
    19 WEEKDAYS,
— I   OCTOBER
 - 116 WEEKDAYS,
      SEPTEMBER
                9     10

               	 a.m—
                       11     12
                           1     2    3

                          —p.m.i.	•	  t»-j
                       HOUR OF DAY, PST

        Figure 5-25. Variation of mean 1 -hour oxidant
        and PAN concentrations, by hour of day, in
        downtown Los Angeles, 1965.

        Source: U.S. Department of Health, Education,
        and Welfare (1970).
                       5-86

-------
Q.
a.
g
S
01
u
I
I
o
X
o
0.18


0.16


0.14


0.12


0.10


0.08


0.06


0.04


0.02
           1IIIT~T
                                 OXIDANT
                      0.010
       Q.
       Q.
                      0.008  2
                      0.006


                      0.004


                      0.002
       O

       5
       ec
       01
       o
       z
       O
       u
       6
         8   10
12
8
             •a.m.-
                              •p.m.
               HOUR OF DAY. PST

    Figure 5-26. Variation of mean 1-hour
    oxidant and PAN concentrations, by hour of
    day. Air Pollution Research Center,
    Riverside, CA, September, 1966.

    Source: U.S. Department of Health,
    Education and Welfare (1970).
                   5-87

-------
en
00
00
                    a.
                    a
cc

z
Ul
u
z
o
u
Ul
z
O
N
O
                       0.50
1000    1400     1800


      OCTOBER 12,1978
2200     0200     0600    1000

         TIME OF DAY, PDT
1400     1800    2200


                  i—J
                                                                                     •OCTOBER 13,1978-
                                                                                                               a
                                                                                                               a
                                                                                                               o
                                                                                                               z
                                                                                                               X
                                      Figure 5-27. Diurnal profiles of ozone and PAN at Claremont, CA,

                                      October 12 and 13,1978, 2 days of a multi-day smog episode.
                                      Source: Tuazon et al. (1981b).

-------
5.6.5.2  Seasonal Patterns.  Seasonal  differences  in PAN concentrations were
alluded to in Section 5.6.3 and mean and maximum PAN concentrations were pre-
sented by month for 2 years (1967-1968 and 1980) for Riverside, California, in
Figure 5-24.   That  seasonal  differences exist in PAN concentrations was also
documented in  the 1970 criteria  document  for photochemical oxidants (U.S.
Department of Health, Education,  and Welfare, 1970).  Total oxidants and PAN
were monitored for 13 months in Riverside,  California.   The oxidant concentra-
tions were obtained by continuous Mast meter measurements, 24 hr/day.   Concen-
trations of PAN  were measured in sequential  samples analyzed  by GC-ECD from
6:00 a.m. to  about  4:00 or 5:00  p.m.   The data are  not  strictly comparable,
since the shorter, daylight averaging time for PAN would be expected to result
in somewhat higher mean concentrations  of PAN than would be obtained across a
24-hour averaging period.   Nevertheless, the patterns given in Figure 5-28 are
of  interest  and demonstrate that peak PAN  concentrations can constitute  a
higher percentage of the  peak ozone concentrations during winter months than
during the rest  of the year.  This observation  is still  valid  (Spicer et al.,
1983) and has  been  attributed by Lewis et  al.  (1983)  and by Holdren et al.
(1984) to greater PAN stability in the winter months because of cooler tempera-
tures (Cox and  Roffey,  1977).   The possibility, however,  that the somewhat
greater NO  emissions of  the winter heating  season  also contribute to this
          JTl
phenomenon should not be overlooked.
     Data from one additional study complement the older data already presented
on diurnal and seasonal  patterns (Section 5.4.2).   Lewis et al. (1983) measured
PAN  and  ozone  concentrations  from September 25, 1978, to May 10, 1980,  in  New
Brunswick, New Jersey.  Average  (10-hr and 24-hr) and maximum concentrations
of both pollutants are given in Table 5-21 by month of the year (Lewis et al.,
1983).  Note  that the highest monthly  mean concentrations,  both the 10-hr  and
24-hr means,  occurred during the smog season (August and September) but that
the  next  highest occurred  in  October  and February, respectively.  Average
diurnal profiles were obtained during this same study and are shown, by month,
in Figure 5-29 (Lewis et al., 1983).

5.6.6  Spatial Variations in Ambient Air Concentrations  of Peroxyacetyl Nitrate
5.6.6.1  Urban-Rural Gradients and Transport  of PAN.  As noted earlier, precur-
sors  to  PAN,  especially N02, are lower in nonurban than in urban areas, such
that  little  local formation is expected in  nonurban  areas.  Available data on
                                   5-89

-------
                            III!)
                             MONTHLY MEANS OF DAILY MAXIMUM
                             1-hour AVERAGE CONCENTRATIONS
                             MONTHLY MEANS OF 1-hour AVERAGE
                             CONCENTRATIONS
                               -O	-O— —
    —.*-• ^   PAN
       I     L
JUN. JUL  AUG. SEP,  OCT.  NOV. DEC. JAN.  FEB. MAR. APR.  MAY JUN.
  I                            MONTH I
  U	1966—	»-U	1967-	
si
  O
  z
  O
  O
   Figure 5-28. Monthly variation of oxidant (Mast meter, continuous
   24-hr) concentrations and PAN (GC-ECD, sequential, 6:00 a.m. to
   4:00-5:00 p.m.) concentrations. Air Pollution Research Center,
   Riverside, CA, June 1966 - June 1967.

   Source: U.S. Department of Health, Education, and Welfare (1970).
                            5-90

-------
           TABLE 5-21.   PAN AND OZONE CONCENTRATIONS IN AMBIENT AIR,
         NEW BRUNSWICK,  N.J.,  FOR SEPTEMBER 25,  1978,  TO MAY 10, 1980
PAN concentration, ppb
Month
January
February
March
April
May -
June
July
August
September
October
November
December
24- hr
average
0.12
0.61
0.36
0.45
0.23
0.09
0.26
1.17
1.04
0.93
0.25
0.57
10-hr
average
0.19
0.69
0.41
0.57
0.28
0.17
' 0.44
1.63
1.41
1.08
0.31
0.62
Hourly
maximum
1.3
4
1.3
2.5
1
0.8
3.5
10.6
7.5
5.8
3.5
2.5
QS concentration, ppb
24- hr
average
11.5
17.2
23
28.5
31.4
NA
37,5
37.4
22.4
15.8
11.6
9.7
10-hr
average
15.5
23.2
29.1
37.3
40.9
NA
57.6
55.9
33.9
22.6
15.8
12.8
Hourly
maximum
34
40
58
80
78
NA
" 130
145
110 .
68.''
40
35
 These results are lower than expected; however, there was no evidence of
 instrument malfunction.                                 ,
Source:   Lewis et al.  (1983).                                          ;

PAN concentrations indicate clearly that they are lower in nonurban areas than
in urban (Section 5.5.3).  It should be noted, however, that few data exist on
concentrations in  agricultural  areas and that  the  possibility that PAN  is
transported  is therefore important in assessing exposures  of  vegetation to
PAN.   Lonneman et  al.  (1976) and Nieboer and  Van Ham  (1976), in studies  cited
in the  1978  criteria  document (U.S. Environmental  Protection  Agency, 1978),
reported the transport of PAN.   The more recent study by Nielsen et al.  (1981)
has confirmed that PAN can be present at relatively  high concentrations in
photochemically polluted  air after long-range transport.  Variations in con-
centrations  of  PAN and  other  oxidants measured  in  Claremont, California
(Grosjean, 1983),  are  consistent with  transport patterns.   The recent work of
Singh and Salas (1983) has shown that significant nighttime PAN concentrations
can occur  aloft,  at least in a relatively clean environment.  It is possible
that the transport of PAN occurs aloft, as with ozone, and that under favorable
conditions PAN can be transported  long distances.
                                   5-91

-------

E
O
Z
O
   0.080
   0.060
   0.040
   0.020
   0.025
   0.015
   0.005
_  OZONE
        _ PAN
              -a.m.-
                         12
                                 -p.m.
                   TIME OF DAY, hour
                                   24
      Figure 5-29. Average daily profile by month
      (July 7 - October 10) for PAN and ozone in
      New Brunswick, NJ, 1979. Numbers refer to
      months of the year.

      Source: Lewis et al. (1983).
                      5-92

-------
5.6.6,2  Intraclty Variations.  Few  data on PAN concentrations at different
sites in the  same  city are available.   One  study  is available for Houston,
Texas (Jorgen et a!.,  1978),  in which PAN was measured on October 26 and 27,
1977, at  three  separate sites, two  in  Houston and one north  of  the city.
Comparison of the levels of ozone and PAN peaks among the three sites on those
2  days  reveals  significant  differences  (Table 5-22).  On  October  26,  the
highest ozone concentrations  (two peaks of 110 and  95 ppb  at ~3:00  p.m. and
~7:00 p.m., respectively) were seen at site 3, in southeast Houston; while the
highest PAN concentration  on  that day was  a  secondary peak  of 15 ppb at ~8:00
p.m. at site  2  (mid-city).  On October  27,  the highest ozone concentration,
180 ppb, occurred  at site I, about  4 miles  north-northeast of Houston, at
~1:00 p.m.   That was accompanied at the same time by the highest PAN concentra-
tion for the day, 16 ppb.
     Examination of  Table  5-22  shows the  lack of  a  consistent, quantitative
relationship between  PAN and ozone as indicated by  intracity differences of
more than  twofold  in ozone concentrations (afternoon peaks,  sites  1 and 2,
October 27) and  differences  of about fivefold in  PAN concentrations between
sites (afternoon peaks,  sites 1 and 2, October 27).
5.6.6.3   Indoor-Outdoor Ratios of PAN Concentrations.   No  recent  studies
appear in  the  literature on indoor  concentrations of PAN or indoor-outdoor
ratios (I/O). In  three school buildings in  southern California, Thompson et
al. (1973) found I/O ratios (expressed here as percentages) of 89, 97, and 148
percent, respectively,  in  the absence of  air  conditioning.  With air condi-
tioning, the  I/O  ratios were 75,  108,  and 117 percent,  respectively.  Total
oxidants were  nearly constant all day,  remaining  about 30 percent  (in air
conditioning) of  the average  outside concentration.   The  higher I/O  gradients
for PAN than for oxidants were attributed  by the authors to the greater break-
down of ozone  ("oxidants") through  its  reaction with surfaces.   The cooler
temperatures indoors  are the  probable cause  of the greater  persistence of PAN
indoors.   Like  ozone,  PAN also decays  indoors, but  over an extended period
(Thompson et al., 1973).
 5.7  CONCENTRATIONS OF OTHER PHOTOCHEMICAL OXIDANTS IN AMBIENT AIR
     Concentrations of nitrogen dioxide (N02) and related nitrogenous oxidants
 are presented  in a recent criteria document on oxides of nitrogen (U.S. Environ-
 mental  Protection  Agency,  1982b) and are  not  given here.   In addition, the
                                   5-93

-------
   TABLE 5-22.   INTRACITY VARIATIONS IN PEAK OZONE AND PAN CONCENTRATIONS IN
                       HOUSTON, OCTOBER 26 AND 27, 1977

1


2
Site
(~4 mi NNE
of Houston)

(mid- city)
Date
10/26

10/27
10/26
o8,
ppb
75

180
50
Time
1:

2:
2:
00

00
00-
of peak
p.

P-
3:
m.

m.
00 p.m.
PAN,
ppb
2

16
2
Time
9:00
p.m.
2:00
2:00-

a
of peak
.m. - 5:00
(plateau)
P
6
.m.
:00 p.m.
(plateau)


3





(SE Houston)



10/26
10/27
10/26
10/26
10/27
10/27
50
70
110
95
90
55
8:
1:
3:
7:
1:
7:
00
P-
00-3:
00
00
00
30
P.
P-
P-
P.
m.
00 p.m.
m.
m.
m.
m.
10
• 3 •
8
12
4
4
7:00
P
12:30-
2:00
5:30
11:30
6:30
P
P

P
.m.
3:00 p.m.
.m.
.m.
a.m.
.m.
Source:  Jorgen et al. (1978).

comprehensive review by Altshuller (1983) also documents available information
on nitrogenous  oxidants  such as nitric acid (HNO-) and peroxynitric acid, as
well as  on  formic acid (HCOOH) and  hydrogen peroxide  (H202).   The  reader  is
referred to  these reviews for information on  these  oxidants.   The two non-
nitrogenous  oxidants,  formic acid  and hydrogen  peroxide,  are appropriate
concerns for this document, however, and the limited information on concentra-
tions of these two pollutants is summarized below.
     Studies on the toxicologic effects  of  formic acid, though limited  in
number,  have shown  only  negligible  effects,  even  in  animals  exposed to  levels
as high  as  20  ppm (aerosol  vapor)  (see Chapter 9).   These  levels are three
orders of magnitude greater than the  concentrations seen  in polluted urban
atmospheres.  For example,  maximum  concentrations of HCOOH observed by  Tuazon
and coworkers,  using FTIR (Tuazon et al., 1978; 1980; 1981b), in Claremont and
Riverside, California, were in the range of 5 to 20 ppb in  a  study covering
5 years.  The ranges of concentrations  of HCOOH measured by  Tuazon et al. were
consistent with those found in  a  long-path  FTIR study  by Hanst et  al.  (1982).
The FTIR method offers a  reliable assessment of the ambient  air concentrations
of HCOOH and reported concentrations are believed to be accurate.
                                   5-94

-------
     Data on HCOOH  concentrations  for 2 days  in  October  1978 are shown in
Figure 5-30 (Tuazon et a!., 1981b).  The diurnal pattern is similar to that of
related oxidants and of some of the other smog products.
     The measurement of  hydrogen  peroxide (H-O^) in ambient  air  is  fraught
with  difficulties  that remain unresolved.  Ozone itself  is  known to be an
interference in virtually  all  of the  past and  current measurement methods for
H?0?  (Chapter 4).  In  his review of non-ozone oxidants and other smog consti-
tuents and photochemical products, Altshuller (1983) examined data obtained in
the South Coast Air Basin of  California  in  the late 1970s and in 1980 for
possible consistency  in the  interference of  ozone  in  H?0p  measurements.
Laboratory experiments  (Heikes et  al. ,  1982)  have  indicated  that 1 ppb H?0?
would be generated  per 100 ppb ozone.  The analysis by Altshuller shows that
this  relationship  does not hold in ambient  air in  the  South  Coast Air Basin
once  H^Op levels  exceed about 5 to 10  ppb;  and Altshuller (1983) concluded
that variations in l-LOj, measurements there remain unexplained.
     Because of measurement  problems, the true levels of hLO^ in ambient air
are unknown, especially  in polluted areas, where multiple interferences may
possibly occur.   Attempts to detect hydrogen peroxide by means of FTIR spectro-
scopy have all  been negative,  even in polluted areas.   The method can measure
HLOp  in  ambient air with specificity at  H^Qp  levels >  40 ppb, which is the
limit of detection  for an  FTIR instrument with a  1-km pathlength  (see Chapter
4>-
     Notwithstanding measurement difficulties,  some  ranges of H?0p concentra-
tions at urban and nonurban sites  have been reported in the literature.   These
are given in  Table 5-23, along with  the  general  type of measurement method
used to obtain the reported concentrations.  It must be kept  in mind, however,
that  the reported  concentrations, though they represent state-of-the-art
measurements, are thought not  to be accurate.
5.8  SUMMARY
     In  the  context of this document, the  concentrations  of  ozone  and  other
photochemical  oxidants  found in ambient  air are  important  for:   (1) assessing
potential exposures of  human and other receptors; (2) determining the range of
ambient  air  concentrations of ozone and  other photochemical oxidants relative
                                   5-95

-------
CJ1
I
ID
CTt
                 E
                 a
                 a
                  •>

                 O


                 I
                 ui
                 U
                 z
                 O
                 O
                 LU
                 Z
                 O
                 N
                 O
     -OCTOBER 12,1978


1000    1400     1800     2200
TIME OF DAY, POT
0200     0600     1000
                                                         	OCTOBER 13,1978

                                                         1400     1800     2200
1000    1400     1800     2200


      -OCTOBER 12,1978	
                               0200     0600     1000

                                TIME OF DAY, PDT
                         1400     1800    2200


                         — OCTOBER 13,1978-
      Figure 5-30. Diurnal profile of HCOOH, along with other oxidants and smog
      constituents, on October 12 and 13,1978, at Claremont, CA.
a
a.

Z
O
Z

-------
            TABLE  5-23.   CONCENTRATIONS11 OF HYDROGEN PEROXIDE IN AMBIENT AIR
                              AT URBAN AND NONURBAN SITES
                                          (ppb)
Location
Hoboken, NJ
(urban)
Riverside, CA
(urban)
Riverside and
Claremont, CA
(urban)
Minneapolis, MN
(urban)
Boulder, CO
(urban)
Boulder, CO
(nonurban,
east of
Boulder)
Tucson, AZ
(nonurban,
54 km SE
of Tucson)
Tucson, AZ
(remote, near
Tucson)
Concns. , and
Date comments
1970 < 40
1970 <180 (during
smog episode
with 650 ppb
oxidants)
July-Aug. 100 max. (ozone
1977 also 100 and
increasing);
10 to 50 on
most days
NAb <6
NA <0.5
Feb. 1978 0.2 to 3
NA <7
NA ~1
Method
Titanium (IV)
sill f ate/8-qui no-
linol
Titanium (IV)
s u 1 fate/8- qui no-
linol
Luminol
chemi 1 umi nescence
Wet chemical
Wet chemical0
Luminol
chemi 1 umi nescence
Luminol
chemi 1 umi nesence
Luminol
chemi 1 umi nescence
Reference
Bufalini et
al. (1972)
Bufalini et
al. (1972)
Kok et al.
(1978)
He ikes et al.
(1982)
Hei kes et al .
(1982)
Kelly et al .
(1979)
Farmer and
Dawson (1982)
Farmer and
Dawson (1982)
aNote:  Since methods used to obtain these data are all subject to positive inter-
 ference by ozone, data presented here are not reliable.  They are included as a
 summary of reported concentrations.

 Not available.
cSee Chapter 4 for method used by Heikes et al. (1982).

Source:  Derived from data in Altshuller (1983).
                                      5-97

-------
to  demonstrated  "effects levels"  (Chapters  6-12);  (3) determining indoor-
outdoor gradients  for  exposure  analyses;  (4) assessing whether the concentra-
tions of  oxidants  other than ozone, singly, collectively, or in combination
with ozone, are cause for concern; and (5) evaluating the adequacy of ozone as
a control surrogate for other photochemical oxidants, if concentrations of the
other oxidants are cause for concern given the effects and the "effects levels"
of those oxidants.

5.8.1  Ozone Concentrations in Urban Areas
     In Table  5-24, 1983 ozone  concentrations  for Standard Metropolitan Stat-
istical Areas  (SMSAs)  having populations >_ 1 million are given by geographic
area, demarcated according to United States Census divisions and regions (U.S.
Department of  Commerce,  1982).  The  second-highest  concentrations among daily
maximum 1-hour values  measured  in 1983 in the 38 SMSAs having populations of
at least 1 million ranged from 0.10 ppm in the Ft.  Lauderdale, Florida; Phila-
delphia,  Pennsylvania;  and  Seattle,  Washington,  areas to 0.37 ppm in the Los
Angeles-Long Beach, California,  area.   The second-highest value among daily
maximum  1-hour ozone  concentrations for 35  of the 38 SMSAs  in  Table 5-24
equaled  or exceeded 0.12 ppm.   The data clearly show, as well,   that the
highest  1-hour ozone  concentrations  in the United  States occurred  in the
northeast  (New England and  Middle Atlantic states),  in  the  Gulf Coast area
(West South  Central states),  and  on  the west coast  (Pacific states).  Second-
highest daily  maximum  1-hour concentrations in 1983 in the SMSAs, within each
of  these  three areas  averaged  0.16,  0.17, and 0.21 ppm, respectively.   It
should be  emphasized  that  these three areas of the United States are subject
to those meteorological and climatological factors that are conducive to local
oxidant formation, or transport, or both.  It should also be emphasized that 9
of  the 16 SMSAs  in the  country with populations :> 2 million are located in
these areas.
     Emissions of manmade  oxidant precursors  are  usually  correlated with
population, especially emissions from area source categories such as transpor-
tation  and residential  heating  (Chapter 3).   Accordingly,  when grouped  by
population,  the  80 largest  SMSAs  had the  following  median  values  for their
collective  second-highest daily maximum 1-hour ozone concentrations in 1983:
populations  >  2  million, 0.17 ppm QS;  populations  of 1 to 2 million,  0.14 ppm
03; and populations of 0.5  to 1 million, 0,13 ppm Og.  As noted above, however,

                                   5-98

-------
   TABLE 5-24.   SECOND-HIGHEST OZONE CONCENTRATIONS AMONG DAILY MAXIMUM 1-hr
  VALUES IN 1983 IN STANDARD METROPOLITAN STATISTICAL AREAS -WITH POPULATIONS
              > 1 MILLION, GIVEN BY CENSUS DIVISIONS AND REGIONS3
Division
and region
SMSA
population,
SMSA millions
Second-highest
1983 03
concn. , ppm
Northeast

  New England
Boston, MA
  Middle Atlantic  Buffalo, NY
                .   Nassau-Suffolk, NY
                   Newark, NJ
                   New York, NY/NJ
                   Philadelphia, PA/NJ
                   Pittsburgh, PA

South
. . >2

1 to <2
  >2
1 to <2
  >2
  >2
  >2
0.18

0.12
0.17
0.25
0.19
0.10
0.14
  South Atlantic
South

  West South
   Central
North Central

  East North
   Central
  West North
   Central
Atlanta, GA                       >2
Baltimore, MD                     >2   :•
Ft. Lauderdale-Hollywood, FL    1 to <2
Miami, FL                       1 to <2
Tampa-St. Petersburg, FL        1 to <2
Washington, DC/MD/VA              >2
Dallas-Ft. Worth, TX              >2
Houston, TX                       >2
New Orleans, LA                 1 to <2
San Antonio, TX        ,    ,     1 to <2
Chicago, IL                       >2
Detroit, MI                       >2
Cleveland, OH                   1 to <2
Cincinnati, OH/KY/IN            1 to <2
Milwaukee, WI                   1 to <2
Indianapolis, IN                1 to <2
Columbus, OH                    1 to <2
St. Louis, MO/IL
Minneapolis-St. Paul
Kansas City, MO/KS
                                         MN/WI
  >2
.  >2
1 to <2
                  0.17
                  0,19
                  0.10
                  0.12
                  0.14
                  0.17
                  0.16
                  0.28
                  0.12
                  0.12
                  0.17
                  0.17
                  0.15
                  0.15
                  0.18
                  0.14
                  0.12
0.18
0.13
0.13
                                   5-99

-------
 TABLE 5-24 (cont'd).  SECOND-HIGHEST OZONE CONCENTRATIONS AMONG DAILY MAXIMUM
        1-hr VALUES IN 1983 IN STANDARD METROPOLITAN STATISTICAL AREAS
           WITH > 1 MILLION, GIVEN BY CENSUS DIVISIONS AND REGIONS9
Division
and region
West
Mountain

Pacific










SMSA

Denver-Boulder, CO
Phoenix, AZ
Los Angeles- Long Beach, CA
San Francisco-Oakland, CA
Anaheim-Santa Ana-
Garden Grove, CA
San Diego, CA
Seattle- Everett, WA
Riverside-San Bernardino-
Ontario, CA
San Jose, CA
Portland, OR/WA
Sacramento, CA
SMSA Second-highest
population, 1983 03
millions concn. , ppm

1 to <2
1 to <2
>2
>2

1 to <2
1 to <2
1 to <2
1 to <2

1 to <2
1 to <2
1 to <2

0.14
0.16
0.37
0.17

0.28
0.20
0.10
0.34

0.16
0.12
0.15
 Standard Metropolitan Statistical Areas and geographic divisions and regions
 as defined by Statistical Abstract of the United States (U.S. Department of
 Commerce, 1982).
Source:  U.S. Environmental Protection Agency (1984a).

coincident meteorology favorable for oxidant formation undoubtedly contributes
to the apparent correlation between population and ozone levels.
     Among all  stations  reporting valid ozone data (:> 75 percent of possible
hourly values per  year)  in 1979, 1980,  and  1981  (collectively,  906  station-
years) in the United  States,  the median  second-highest 1-hour ozone  value was
0.12 ppm, and 5 percent  of the  stations  reported  second-highest 1-hour values
> 0.28 ppm.
     A pattern of concern  in assessing responses to ozone in human populations
and in vegetation  is  the occurrence  of repeated or prolonged  multiday periods
when the  ozone  concentrations in ambient air are in the range of those known
to elicit responses (see Chapters 10 and 12).  In addition, the number of days
of  respite  between such multiple-day periods  of  high ozone is of possible
consequence.  Data show  that repeated, consecutive-day exposures to or respites
                                   5-100

-------
from  specified  concentrations  are location-specific.  At a  site  in Dallas,
Texas, for example,  daily maximum 1-hour concentrations were >^ 0.06 ppm for
2 to  7 days  in  a  row 37 times  in  a 3-year period  (1979 through 1981).  A con-
centration of  >0.18 ppm was recorded  at that site  on only  2  single days,
however,  and no multiple-day recurrences of that concentration or greater were
recorded over the 3-year  period.   At a  site  in  Pasadena,  California, daily
maximum 1-hour  concentrations  >_0.18  ppm recurred on  2 to 7  consecutive days
33 times in that  same 3-year period (1979 through 1981) and occurred, as well,
on 21 separate  days.  These  and other  data  demonstrate the occurrence  in some
urban areas  of  multiple-day  potential  exposures to  relatively high  concentra-
tions of ozone.

5.8.2.  Trends  in Nationwide Ozone Concentrations
     Trends  in  ozone concentrations  nationwide are  important for estimating
potential exposures  in the future of human populations and other receptors, as
well as for examining the effectiveness of abatement programs.   The determina-
tion  of  nationwide  trends requires the  application  of  statistical  tests to
comparable,  representative,  multiyear aerometric  data.   The derivation of
recent trends in  ozone  concentrations and the interpretation of those trends
is  complicated  by two  potentially significant factors  that have affected
aerometric data since 1979:   (1)  the promulgation  by EPA  in 1979  of  a  new
calibration procedure for ozone monitoring (see Chapter 4); and (2) the intro-
duction by EPA  of a quality-assurance program that has  resulted.in improved
data-quality  audits.   The effects of  these factors on ozone concentration
measurements are  superimposed  on  the effects  on concentrations of any changes
in meteorology  or in precursor emission  rates that  may  have  occurred over  the
same time span.
     The nationwide  trends  in  ozone concentrations  for a 9-year period, 1975
through 1983, are shown in Figure 5-31 (U. S. Environmental Protection Agency,
1984a).  The data given are trends as gauged by the composite average of the
second-highest  value among  daily  maximum 1-hour  ozone  concentrations.   Data
from  four  subsets of monitoring  stations,  most of  them urban stations, are
given:   (1)  California stations  only;  (2)  all  stations except  those in
California;  (3)  all  stations  including those  in California; and  (4)  all
National Air Monitoring Stations  (NAMS), which report  data  directly to  EPA.
Only  stations reporting > 75 percent of  possible  hourly values in the  respective
years are represented in  the data.
                                   5-101

-------
    0.18
    0.17
I
Q.
 *
z
o

§
QC


Ul
o
z
o
o
UJ
Z
o
N
O
0.16
0.15
0.14
    0.13
    0.12
                                                       CA (27 Stations) —
                                                                   f
             """"?	-9	
0 NAMS STATIONS (62)

T 95% CONFIDENCE
1 INTERVALS

O ALL STATIONS (176)

J 95% CONFIDENCE

JL INTERVALS
              A CALIFORNIA STATIONS (27)


              V ALL STATIONS EXCEPT

                CALIFORNIA (149)

                I      I       I      I
                                         V
                                      I
              1975  1976   1977  1978   1979  1980  1981   1982  1983



                                        YEAR


         Figure 5-31. National trend in composite average of the second highest value
         among daily maximum 1 -hour ozone concentrations at selected groups of

         sites, 1975 through 1983.


         Source: U.S. Environmental Protection Agency (1984a).
                                  5-102

-------
     For the entire 9-year period, 1975 through 1983, all subsets of monitoring
stations show a  decline  in the composite second-highest daily maximum 1-hour
ozone concentration.  Between  1979,  when the new, more  accurate calibration
procedure was promulgated,  and 1982, a small decline  of 9 to 10 percent in
nationwide ozone concentrations occurred.  From 1982 to 1983, however, concen-
trations  increased  by  about 10 percent  in  California,  by about 14 percent
outside California, and by about 12 percent nationwide (all states).  Recently
published data for  1984  from a somewhat  smaller  number  of sites (163)  (U.S.
Environmental, Protection  Agency,  1986) show a decrease  in nationwide ozone
concentrations from 1983 to 1984, with 1984 levels approximating those recorded
in 1981.  The portion  of the  apparent decline in ozone  nationwide  from  1975
through 1984 that  is attributable to the  calibration change  of  1979 cannot be
determined by simply applying  a correction factor to all data, since not all
monitoring stations began using the UV calibration procedure in the same year.
     Figure 5-32 shows  the nationwide frequency distributions of the first-,
second-,  and  third-highest 1-hour 0,,  concentrations at predominantly urban
stations aggregated for 1979, 1980, and 1981, as well as the highest 1-hour 0.,
concentration at site of the National Air Pollution Background Network (NAPBN)
aggregated for the  same  3 years.  As shown by Figure 5-32, 50 percent of the
second-highest 1-hour  values  from non-NAPBN sites in this 3-year period were
0.12 ppm  or less and about 10  percent were  equal  to  or greater  than 0.20 ppm.
At the  NAPBN sites, the  collective 3-year distribution (1979 through  1981)  is
such that about  6  percent of the  values  are less  than  0.10 ppm  and  fewer than
20 percent are higher than 0,12 ppm.

5.8.3.   Ozone Concentrations in Nonurban Areas
     Few nonurban areas have been  routinely monitored for  ozone concentrations.
Consequently, the aerometric data  base for  nonurban areas  is considerably less
substantial than for  urban  areas.  Data are  available, however,  from  two
special-purpose networks, the  National Air  Pollution Background Network  (NAPBN)
and  the  Sulfate  Regional Experiment network (SURE).   Data on maximum 1-hour
concentrations and  arithmetic  mean 1-hour  concentrations  reveal that maximum
1-hour  concentrations  at  nonurban sites classified  as rural (SURE study,
Martinez  and Singh,  1979; NAPBN  studies, Evans  et  a!., 1983) can  sometimes
exceed  the  concentrations  observed  at sites  classified as  suburban (SURE
study, Martinez and Singh, 1979).  For example, maximum  1-hour ozone concentra-
tions  measured  in  1980 at Kisatchie  National  Forest (NF), Louisiana;  Custer
                                   5-103

-------
in
i
                ^>
                     99.99
                   0.4S
                   0.40
                   0,35
                   0.30
             99.9 99.8
99  38   96   90
80  70 60  60 40 30  20
10
1 0,5  0.2 0,1 0.05  0.01
<   0.25

ui
O
§   0.20
O
ui

O   0.15
O
                   0.10
                   0.05
                        II    I    I     I    I   I  I  I   I   I     II
                      	 HIGHEST
                      	2nd-HIGHEST

                      	•• 3rd-HlGHEST
                      	 HIGHEST, NAPBN SITES
                                                            I   I   I   I   I   I    I
                                                                                   I  I
                      0.01  0.06 0.1 0.2 0.5  1   2    5    10    20  30 40 50 60 70  80    90  95   98  99   99.8 99.9    99.99

                                STATIONS WITH PEAK 1-hour CONCENTRATIONS < SELECTED VALUE, percent

                       Figure 5-32, Distributions of the three highest 1 -hour ozone concentrations at valid sites (906
                       station-years) aggregated for 3 years (1979,1980, and 1981) and the highest ozone
                       concentrations at NAPBN sites aggregated for those years (24 station-years).
                       Source: U.S. Environmental Protection Agency (1980,1981,1982).

-------
NF, Montana; and  Green  Mt.  NF, Vermont, were  0.105,  0.070,  and 0.115 ppm,
respectively.   Arithmetic  mean 1-hour  concentrations  for 1980 were 0.028,
0.037, and 0.032 ppm at the respective sites.   For four nonurban (rural)  sites
in the  SURE study, maximum 1-hour  ozone  concentrations were 0.106, 0.107,
0.117, and  0.153;  and  mean 1-hour concentrations ranged  from 0.021 to 0.035
ppm.   At the five nonurban (suburban) sites of the SURE study,  maximum concen-
trations were 0.077, 0.099,  0.099,  0.080, and 0.118 ppm, respectively.  The
mean 1-hour concentrations at those sites were 0.023,  0.030,  0.025, 0.020,  and
0.025 ppm, respectively.                                                  ;
     Ranges of concentrations and the maximum 1-hour concentrations at some of
the NAPBN and SURE sites show the probable influence of ozone transported from
urban areas.  In one documented case, for example, a 1-hour peak ozone concen-
tration of 0.125 ppm at an NAPBN site in Mark Twain National  Forest, Missouri,
was measured during passage of an air mass whose  trajectory was calculated to
have  included  Detroit,  Cincinnati,  and Louisville in  the  preceding  hours
(Evans et al.,  1983).
     The  second-highest concentration  among  all  the daily  maximum 1-hour
concentrations measured at  the NAPBN sites appear  to  be  about one-half the
corresponding concentrations measured  at  urban sites in  the same  years.' No
trend in concentrations at these NAPBN  sites is discernible in the data record
for 1979 through 1983.
     These data corroborate the conclusion given  in the 1978 criteria document
(U.S.  Environmental  Protection Agency,  1978)  regarding  urban-nonurban  and
urban-suburban gradients;  i.e.,  nonurban  areas may sometimes  sustain  higher
peak ozone concentrations than those found in urban areas.

5.8.4.  Diurnal and Seasonal Patterns in Ozone Concentrations
     Since the photochemical reactions  of precursors that result in ozone for-
mation  are  driven by sunlight, as well  as  by  emissions, the  patterns  of  ozone
occurrence  in  ambient  air depend  on daily and  seasonal  variations  in  sunlight
intensity.  The  typical diurnal pattern of ozone  in ambient  air has a  minimum
ozone level around sunrise (near zero in most urban areas), increasing through
the  morning to  a peak  concentration in  early  afternoon, and  decreasing toward
minimal  levels  again  in the  evening.   The 1978 criteria document  ascribed.th§
daily ozone pattern to  three simultaneous processes:  (1) downward transport of
ozone from  layers  aloft;  (2) destruction of ozone through contact with surfaces
and  through reaction  with nitric oxide (NO) at ground  level; and  (3)  in situ
                                    5-105

-------
photochemical production of ozone (U.S. Environmental Protection Agency, 1978;
Coffey et a!., 1977; Mohnen, 1977;  Reiter, 1977a).   Obviously, meteorology is
a controlling factor;  if strong winds  disperse the precursors or heavy  clouds
intercept the sunlight, high ozone levels will not develop.   Another important
variation on the basic diurnal  pattern appears in some localities as a secondary
peak in  addition  to the early afternoon peak.  This secondary peak may occur
any time from midafternoon to  the middle of  the  night and  is  attributed to
ozone transported  from upwind areas where high  ozone levels have  occurred
earlier  in the day.  Secondary peak concentrations can be higher than concen-
trations resulting from the photochemical reactions of locally emitted precursors
(Martinez and Singh, 1979).   At one nonurban site  in Massachusetts (August
1977), for example, primary peak concentrations of about 0.11, 0.14, and 0.14
occurred at  noon,  from noon to about 4:00 p.m., and  at noon, respectively, on
3 successive days  of high  ozone levels (Martinez  and Singh,  1979).  Secondary
peaks at the  same site for those same 3 days were  about 0.150, 0.157, and
0.130 ppm at about 6:00 p.m., 8:00 p.m., and 8:00 p.m., respectively (Martinez
and Singh, 1979).
     Because weather  patterns,  ambient  temperatures, and the  intensity and
wavelengths of sunlight all play important roles  in  oxidant formation,  strong
seasonal as well as diurnal patterns exist.   The highest ozone levels generally
occur in the spring and summer (second and  third quarters), when  sunlight
reaching the  lower troposphere is  most  intense  and  stagnant meteorological
conditions augment the potential for ozone formation and accumulation.   Average
summer afternoon  levels can be from 150 to 250 percent of the average winter
afternoon levels.   Minor  variations in the smog  season occur with  location,
however.  In  addition, it is  possible  for the  maximum  and second-highest
1-hour ozone concentration to occur outside the two quarters of highest .average
ozone concentrations.   Exceptions to seasonal  patterns are potentially  important
considerations with regard to  the  protection of crops  from ozone damage,
especially since  respective  crops  have different growing seasons in terms of
length,  time of year,  and  areas of the country in which they are grown.
     In  addition  to the seasonal meteorological  conditions  that  obtain  in the
lower troposphere,  stratospheric-tropospheric exchange mechanisms exist that
produce  relatively frequent but  sporadic,  short-term incursions  into the
troposphere  of  stratospheric  ozone  (see Chapter  3).   Such  incursions show a
seasonal pattern,  usually  occurring in  late winter or early  spring.
                                   5-106

-------
     Percentile distributions, by  season  of the year, of concentration data
from all eight NAPBN sites show that the arithmetic mean 1-hour concentration
(averaged over a  minimum  of  3 years of data  at each site,  for 1977 through
1983) was higher in the second quarter of the year (April,  May, June) at seven
of the  eight  stations;  and was only negligibly lower than the third-quarter
value at the  eighth station.  The  maximum 1-hour concentrations at  respective
sites, aggregated  over  3  to  6 years, depending on  the  data record for each
site, ranged  from 0.050  ppm  at  Custer NF, MT  (in  the  fourth quarter) to
0.155 ppm at  Mark Twain NF,  MO (in  the  third quarter).   The second-highest
1-hour concentration among maximum daily  1-hour values  ranged from 0.050 ppm
at Custer NF,  MT  (in  the fourth quarter),  to 0.150 ppm at Mark Twain NF, MO
(in the  third quarter).   The  data also show  that  99 percent of the 1-hour
concentrations measured were  well  below 0.12 ppm,  even  in the second quarter
of the  year,  when incursions  of stratospheric  ozone are expected to affect,
at least to  some  degree, the concentrations  measured  at these stations.
Excursions above  0.12  ppm were recorded in 1979 and 1980 at NAPBN sites; but
none were recorded in 1981 (Evans et a!.,  1983; Lefohn,  1984).
     Because  of the diurnal  patterns of ozone, averaging across longer-term
periods  such  as a month,  a season,  or  longer masks the occurrence  of peak.
concentrations (see, e.g., Lefohn and Benedict, 1985).   This is an obvious and
familiar statistical phenomenon.   It is pointed out, however, because it has
direct  relevance  to the protection of public  health and welfare.   Averaging
times must correspond to, or be related in a consistent manner to, the pattern
of exposure that elicits  untoward responses.                              • . •

5.8.5  Spatial Patterns in Ozone Concentrations                           ,>
     In  addition  to temporal  variations,  both macro- and microscale spatial
variations  in ozone concentrations  occur  that have relevance ranging from
important to  inconsequential  for exposure assessments.   Differences in concen-
trations or  patterns of occurrence,  or  both,  are known to exist, for example,
between  urban and  nonurban areas, between indoor and outdoor air, within large
metropolises,  and between lower and higher elevations.   The more important
variations are summarized below.                                            t
5.8.5.1  Urban-Nonurban Differences in Ozone Concentrations.  Ozone concentra-
tions  differ between  urban  and  rural,  between urban and  remote,  and even
between  rural  and remote  sites,  as discussed in part in  the  preceding  section
                                   5-107

-------
on temporal variations.  The variations with area and type of site are varia-
tions both in the timing and the magnitude of the peak concentrations, and,  in
the case of transported ozone, are related to the temporal variations between
urban and  nonurban  areas  discussed above.  Data from urban,  suburban, rural,
and remote sites (see, e.g., SAROAD,  1985a-f; Martinez and Singh, 1979;  Lefohn,
1984; Evans, 1985;  respectively) corroborate the conclusion drawn in  the 1978
criteria document  (U.S.  Environmental  Protection  Agency,  1978) that ozone
concentrations can  sometimes be  higher in some suburban or even rural areas
downwind  of urban plumes than in the urban areas themselves;  and,  furthermore,
that higher  concentrations can  persist longer in rural and  remote  areas,
largely because of  the  absence of nitric oxide (NO) for chemical scavenging.
     In nonurban  areas  downwind of  urban plumes, peak concentrations  can
occur,  as  the  result  of transport, at  virtually any  hour of the  day or night,
depending  upon many factors,  such as  the strength  of the emission source,
induction time, scavenging, and wind speed (travel  time) and other meteorological
factors.   The dependence of the timing of peak exposures upon these transport-
related factors is  well-known  and  is illustrated here by two studies.  Evans
et al.  (1983)  calculated multiday trajectories  for  air parcels reaching a
nonurban sites in the Mark Twain National Forest, Missouri, during an episode.
Four separate trajectories, all of which passed over the Ohio River Valley and
the  Great  Lakes region, impacted  the  forest site at different  times in a
24-hour period (in  which  the  maximum 1-hour concentration measured was 0.125
ppm).  Subsequently,  regional  cloud  cover and rains produced  shifts  in  air
flow and  also reduced  the potential  for ozone  formation,  alleviating the
impact at  the  site.   Kelly et  al.  (1986)  showed in the  Detroit area that peak
ozone concentrations  occurred  at distances of 10 to 70 km (ca.  6 to 43 mi)
north-northeast of  the  urban  center.  Consequently, it would be possible for
peak ozone concentrations  to  occur  in the late afternoon or early evening in
nonurban areas downwind of Detroit.  Kelly et al.  (1986) also  found that
concentrations diminished again beyond 70 km (ca.  43 mi) downwind of the urban
center.  Thus,  as  illustrated  by these and similar data, beyond the distance
traversed  in the  time required for maximum ozone formation in an urban plume,
ozone concentrations  will  decrease (unless fresh emissions are injected into
the  plume) as  the  rate of ozone formation decreases, the plume disperses,
surface deposition  or other scavenging occurs, and meteorological  conditions
intervene.
                                   5-108

-------
     It is  not  surprising,  therefore,  that in rural  areas  lying beyond the
point of maximum ozone  formation,  for a given set of  conditions, peak concen-
trations are  lower  and  average diurnal profiles are flatter than in urban and
near-urban areas (see, e.g., SAROAD, 1985b-f, for rural and remote sites).   In
remote areas beyond the influence of urban plumes,  average peak concentrations
will be  still  lower and average diurnal  profiles  still  flatter (see e.g.,
Evans, 1985).   Exceptions  to these generalizations occur, of course,  because
of the complex  interactions of topography, meteorology,  and photochemistry.
     Such  temporal  and spatial differences  between ozone concentrations in
urban versus nonurban areas are important considerations for accurately assessing
actual and  potential  exposures for human populations  and for  vegetation in
nonurban areas, especially  since  the aerometric data  for nonurban areas are
far from abundant.
5.8.5.2  Geographic, Vertical, and Altitudinal Variations in Ozone Concentrations.
Although of  interest  and concern when estimating global  ozone  budgets,  demon-
strated variations in ozone concentrations with latitude and the lesser variations
with  longitude  have  little  practical  significance for  assessing exposure
within the  contiguous United States.   The effects on ozone concentrations of
latitude and longitude within the contiguous states are minor,  and are reflected
in the aerometric  data bases.  Of more  importance, ozone concentrations are
known  to  increase with  increasing height above the  surface of the earth.
Conversely,  they may  be viewed as  decreasing with proximity to the surface  of
the earth,  since  the  earth's surface acts  as  a  sink for ozone  (see, e.g.,
Seller and Fishman, 1981;  Galbally and  Roy,  1980;  Oltmans,  1981, cited in
Logan et al.5 1981).  The most pertinent vertical and altitudinal gradients in
ozone  concentrations  are:   (1) increases in  concentration with height  above
the surface  of the ground (regardless of altitude); (2) increases in concentra-
tion with  altitude;  and (3) variations in  concentrations with elevation in
mountainous  areas,  attributable to transport  and  overnight  conservation of
ozone  aloft, nocturnal inversions, trapping  inversions,  upslope flows, and
other, often location-specific interactions between  topography, meteorology,
                                                                  *
and photochemistry.
                                   5-109

-------
     The importance  of  monitoring concentrations at the proper height.above
the surface  of the ground  has  been  known  for a  long time, and EPA guidance on
the placement of  monitoring instruments, (see Chapter 4) is predicated on the
existence of a vertical  gradient as ozone is depleted by reaction with ground-
level emissions of NO or by deposition on reactive surfaces such as vegetation.
Data illustrative of the near-surface gradient were reported by Pratt et al.
(1983), who  measured ozone concentrations at two separate heights (3  and ,9 or
6 and  9  meters)  above the  ground at  three  rural,.vegetated  sites.  Although
the maximum  mean  difference between 3 and 9 meters was  3 ppb, this difference
was similar  to the  mean difference between sites at the same height.  Given
the height of  some  vegetation canopies, especially forests, even such small
differences over a spread of 6 meters should probably be taken into considera-
tion when interpreting reported dose-response functions.
     The gradual  increase  in  ozone  concentrations with altitude  has been
documented by a  number  of workers  (see e.g., Viezee et al., 1979; Seiler and
Fishman, 1981; Oltmans,  1981,  as cited in  Logan et  al.,  1981).  There is a
general  increase  in  concentration with altitude, but as described by Seiler
and Fishman  (1981) and Oltmans (1981; cited in Logan et al., 1981), for example,
two  relatively  pronounced gradients,  exist, one  between the surface  of the
earth  and  2  km  (ca.  1 mi)  and one  even more pronounced between 8 and 12 km
(ca. 5 and 7.5 mi).
     Increases in concentrations  with altitude could potentially be  of some
consequence  for passengers and airline personnel on high-altitude flights in
the  absence  of adequate  ventilation-filtration systems (see  Chapter 11).
Variations with height  above  the surface and with elevation, in mountainous
areas, however, should be taken into  account to ensure the accurate assessment
of exposures and the accurate derivation of dose-response functions,  especially
for forests  and other vegetation.
     Variations in ozone concentrations with elevation, not  always consistent
or predictable, have been reported  by researchers investigating the effects of
ozone  on .the mixed-conifer  forest ecosystem of the San Bernardino Mountains of
         *
California.  Measurements  taken  at  four monitoring stations  at  four different
elevations showed that  peak ozone concentrations occurred progressively later
in the day at progressively higher elevations  (Miller  et al.,  1972). Ozone
concentrations >0.10 ppm  occurred  for average  durations of 9,  13,  9,  and
                                   5-110

-------
 8 hr/day  at  the  four  respective  stations,  going  from  lower  to  higher elevations.
 The  occurrence for 13 hr/day  of concentrations  >0.10, ppm  at  the  station  at
 817  m  (2860  ft)  was probably the result  of contact of that  zone  of the mountain-
 side with the inversion  layer (U.S.  Environmental  Protection Agency,  1978).
 Nighttime concentrations  rarely  decreased  below  0.05  ppm  at the  mountain crest;
 whereas  at the  lowest elevation,  the  basin station at 442 m  (1459 ft), the
 nighttime concentration  usually decayed to near zero.   Trapping  inversions
 were major.contributors to the  elevational  gradients observed i,n  this study,
 which  was conducted in the  1970s.   Oxidant concentrations within the inversion
 were found not to  be  uniform but to occur  in multiple layers and strong vertical
 gradients.   The  important result of the trapping of oxidants in the inversion
 layers was the prolonged contact of high terrain with oxidants at night (U.S.
 Environmental  Protection  Agency, 1978).
      In  a more recent report,  Wolff et al. (1986) described measurements  made
 in July  1975 at  three separate elevations  at High Point Mountain in northeastern
 New  Jersey.   The daily ozone maxima were similar at different elevations.   At
 night, however,  ozone concentrations were  nearly zero in  the valley but increased
 with elevation on  the mountainside.  Greater cumulative doses  .(number  of hours
 at >0.08 ppm) were sustained at the higher elevations, 300 and 500 m,  respec-
 tively (ca.  990 and 1650 ft, respectively).  Wolff et al.  (1986) related this
 phenomenon to the depth of the  nocturnal inversion layer and the contact with
 the  mountainside of ozone conserved aloft at night.            -.••.-
      These concentration gradients with increased elevation are important for
 accurately describing concentrations at which injury or damage to vegetation,
 especially forests, may  occur.   Researchers investigating the  effects  of ozone
 on  forest ecosystems have seldom  measured nighttime ozone concentrations
 because  the  stomates of  most  species  are thought to  be closed at  night, thus
 preventing the  internal  flux  of ozone that produces  injury or  damage (see
 Chapter 6).    If stomates  remain even  partially open  at  night,  however, the
 possible occurrence  of nighttime peak concentrations of ozone, the occurrence
 of multiple peaks in a 24-hour  period, or the persistence  of elevated concen-
 trations that do  not decay to  near zero overnight should  not be overlooked.
 Furthermore, the lack of NO for  nighttime scavenging  in  nonurban areas and the
•persistence  of ozone  overnight at higher elevations will  result in the presence
 of relatively higher  concentrations in such areas at  sunrise when  the  stomates
                                    5-111

-------
open and  photosynthesis begins.   This  possibility requires  that exposure
assessments,  in  the absence of  sufficient  aerometric  data for forests and
other vegetated areas, take such factors into consideration.
5.8.5.3  Other SpatialVariations in Ozone Concentrations.   Other spatial  varia-
tions are  important for exposure assessments for human populations.   Indoor-
outdoor gradients in ozone concentrations are known to occur even in buildings
or vehicles ventilated by fresh air rather than air conditioning (e.g., Sabersky
et al., 1973; Thompson et a!., 1973; Peterson and Sabersky, 1975).  Ozone reacts
with surfaces inside buildings, so that decay may occur fairly rapidly, depending
upon the nature of  interior surfaces and furnishings (e.g., Davies et al., 1984;
Contant et al., 1985).  Ratios of indoor-to-outdoor (I/O) ozone concentrations
are quite variable, however, since cooling and ventilation systems, air infil-
tration or exchange rates, interior air circulation rates,  and the composition
of  interior surfaces all  affect indoor  ozone  concentrations.   Ratios (I/O,
expressed  as  percentages)  in the literature thus vary from 100 percent in a
non-air-conditioned  residence  (Contant  et  al.,  1985); to 80  ±  10 percent
(Sabersky et al., 1973) in an air-conditioned office building (but with 100 per-
cent outside  air  intake);  to 10 to 25 percent in air-conditioned residences
(Berk et  al.,  1981);  and to  as  low as  near  zero  in  air-conditioned residences
(Stock et al., 1983; Contant et al., 1985).
     On a larger scale, within-city variations  in  ozone concentrations can
occur, even  though  ozone  is a "regional" pollutant.  Data show, for example,
relatively homogeneous ozone concentrations in New Haven, Connecticut (SAROAD,
1985a), a moderately large city that is downwind of a reasonably well-mixed
urban  plume (Wolff  et  al.,  1975;  Cleveland et  al.; 1976a,b).   In a large
metropolis,  however,  appreciable gradients  in ozone concentrations can exist
from one  side of the city to  the  other, as demonstrated  for  New York City
(Smith, 1981), and  for Detroit (Kelly et al., 1986).  Such gradients should be
taken into consideration, where possible, in exposure assessments.

5.8.6  Concentrations and Patterns of Other Photochemical Oxidants
5.8.6.1   Concentrations.  No aerometric  data are routinely obtained by Federal,
state, or local  air pollution agencies  for any  photochemical  oxidants other
than nitrogen dioxide and ozone.  The concentrations presented in  this document
for non-ozone oxidants were  all obtained in special field  investigations.   The
limitations  in the  number of locations and  areas of the country represented in
                                   5-112

-------
the Information presented  simply  reflect the relative paucity of data in the
published literature.
     The four  non-ozone photochemical oxidants  for  which at least minimal
concentration data are  available are  formic acid, peroxyacetyl nitrate (PAN),
peroxypropionyl nitrate  (PPN),  and  hydrogen  peroxide (H~0p).  Peroxybenzoyl
nitrate has  not been  clearly identified in ambient air in the United States.
     The highest concentrations of  PAN reported  in the older  literature, 1960
through the present,  were those found in the Los Angeles area:  70 ppb (1960),
214 ppb (1965);  and  68 ppb (1968)  (Renzetti  and Bryan,  1961; Mayrsohn  and
Brooks, 1965; Lonneman et al., 1976; respectively).
     The highest concentrations of PAN measured and reported in urban areas in
the past 5 years were 42 ppb at Riverside, California,  in 1980  (Temple and
Taylor, 1983)  and  47 ppb at  Claremont,  California,  also in 1980  (Grosjean
1981).  These  are  clearly  the maximum concentrations of PAN reported for
California and  for the  entire country in this  period.   Other maximum PAN
concentrations measured  in  the last decade in  the Los Angeles  Basin have been
in the range of 11 to 37 ppb.  Average concentrations of PAN in the Los Angeles
Basin  in the past  5 years  have ranged from 4  to  13 ppb (Tuazon et  a!,, 1981a;
Grosjean, 1983; respectively).  The only published  study  covering urban PAN
concentrations outside  California in  the  past  5  years is  that of Lewis et  al.
(1983) for New Brunswick,  New Jersey, in which the average PAN concentration
was 0.5 ppb and the maximum was 11 ppb during September 1978 through May 1980.
Studies outside California  from the early 1970s through  1978 showed  average
PAN concentrations ranging  from 0.4 ppb in Houston, Texas, in 1976 (Westberg
et al., 1978)  to 6.3  ppb in St. Louis,  Missouri,  in 1973 (Lonneman et al.,
1976).  Maximum  PAN   concentrations outside  California  for the same  period
ranged from 10 ppb in Dayton, Ohio, in 1974 (Spicer et al., 1976) to 25 ppb  in
St. Louis (Lonneman et al., 1976).
     The highest PPN  concentration  reported in  studies  over the period  1963
through the  present was 6  ppb in  Riverside,  California  (Darley et  al., 1963).
The next highest reported  PPN concentration  was  5 ppb at St.  Louis, Missouri,
in 1973 (Lonneman et  al., 1976).  Among more recent data, maximum  PPN concentra-
tions  at respective sites ranged from 0.07 ppb in Pittsburgh, Pennsylvania,  in
1981  (Singh  et al., 1982) to 3.1 ppb  at Staten Island, New York (Singh et al.,
1982).  California concentrations fell within this range.  Average PPN concentra-
tions  at the respective  sites  for the more recent data ranged from 0.05 ppb  at
Denver and  Pittsburgh to 0.7 ppb  at Los  Angeles  in  1979  (Singh et  al., 1981).
                                    5-113

-------
     Altshuller (1983)  has  succinctly summarized the nonurban concentrations
of PAN and PPN by pointing out that they overlap the lower end of the range of
urban concentrations  at sites  outside California,  At remote  locations, PAN
and PPN  concentrations  are  lower than even  the  lowest of the urban concentra-
tions by a factor of 3 to 4.
     The concentrations  of  FLC^ reported in, the  literature  to date  must be
regarded as inaccurate since ozone is now thought to be an interference in all
methods  used  to date  except  FTIR (Chapter 4).   Measurements by FTIR, the most
specific and accurate method now available,  have not demonstrated unambiguously
the presence of HpQ^ in ambient air, even in the high-oxidant atmosphere of the
Los Angeles area.   (The limit of detection for a 1-km-pathlength FTIR system
is around 0.04 ppm.)
     Recent data  indicate the presence  in  urban atmospheres of only trace
amounts  of  formic  acid:  < 15 ppb,  measured by  FTIR  (Tuazon  et al.,  1981b).
Estimates in  the  earlier literature (1950s) of 600 to 700 ppb of formic acid
in smoggy atmospheres were erroneous because of faulty measurement methodology
(Hanst et a!., 1982).
5.8.6.2  Patterns.  The patterns  of formic acid  (HCOOH),  PAN,  PPN,  and H£02
may be  summarized  fairly  succinctly.   Qualitatively,  diurnal patterns are
similar to those of ozone, with peak concentrations of each of these occurring
in close proximity to the time  of the ozone peak.  The correspondence in time
of day  is  not exact,  but is close.   As  demonstrated by the work of Tuazon
et al. (1981b), ozone concentrations return to baseline levels somewhat faster
than the concentrations of PAN, HCOOH, or H202 (PPN was not measured).
     Seasonally, winter  concentrations (third and fourth quarters) of PAN  are
lower than  summer'concentrations (second and third quarters).   The percentage
of PAN  concentrations (PAN/03 x 100)  relative to  ozone, however, is  higher in
winter than in summer.  Data are not readily available on the seasonal  patterns
of the other non-ozone oxidants.
     Indoor-outdoor data  on  PAN are limited to  one  report (Thompson et al.,
1973), which  confirms the pattern to be expected from the known chemistry of
PAN; that  is,  it  persists longer indoors than  ozone.   Data  are lacking on
indoor-outdoor ratios for the other non-ozone oxidants.

5.8.7  Relationship Between Ozone and Other Photochemical Oxidants
     The relationship between  ozone concentrations and the concentrations of
PAN, PPN,  HpOp,  and HCOOH is important  only if these  non-ozone oxidants are
                                   5-114

-------
shown to produce  potentially  adverse health or welfare effects,  singly,  in
combination with each other, or in various combinations with ozone at concentra-
tions correponding to  those found in ambient air.   If only ozone is shown to
produce adverse  health or  welfare  effects in the concentration  ranges  of
concern, then only ozone  must be controlled.  If any  or  all of these other
four oxidants  are shown  to produce potentially adverse  health  or welfare
effects, at or near  levels  found  in  ambient air, then  such oxidants will also
have to be  controlled.   Since ozone and all four of the other oxidants arise
from reactions among the  same organic and  inorganic precursors,  an obvious
question is whether  the  control  of ozone will also result in the  control of
the other four oxidants.
     Controlled-exposure  studies  on these  non-ozone oxidants  have employed
concentrations much higher than those found in ambient air (see Chapters 9 and
10).  Because PAN may have contributed, however,  to the eye irritation symptoms
reported in  earlier  epidemiologies!  studiesj  and because PAN  is  the most
abundant of these non-ozone oxidants, the relationship between ozone and PAN
concentrations in ambient air remains of interest.
     •The patterns of  PAN  and ozone  concentrations are not quantitatively
similar but do show  qualitative  similarities for most  locations at which both
pollutants have been measured in the same study.   That a quantitative, monotonic
relationship between  ozone  and PAN  is  lacking, however, is shown by the range
of PAN-to-ozone ratios, expressed as'percentages, between locations and at the
same location, as reported  in the review of Altshuller (1983);
     Certain other information bears out  the lack of a monotonic relationship
between PAN and  ozone.  Not only are  PAN-ozone  relationships not  consistent
between different urban  areas,  but they  are not  consistent in urban versus
nonurban areas, in summer versus winter,  in indoor versus outdoor environments,
or  even, as  the data  show, in location, timing,  or magnitude of diurnal peak
concentrations within the  same  city.   Data obtained  in  Houston  by Jorgen
et aV.  (1978),  for example, show variations  in peak concentrations of PAN  and
in  relationships  to  ozone concentrations of those peaks among three separate
monitoring sites.  Temple and Taylor (1983) have shown that PAN concentrations
are  a greater percentage  of ozone concentrations in winter than in the.remainder
of  the  year in California.   Lonneman et al.  (1976) demonstrated  that  PAN,
absolutely  and  as a  percentage of  ozone,  is  considerably lower in nonurban
than in urban  areas.  Thompson et al. (1973), in what  is apparently the only
                                   5-115

-------
published report  on  indoor concentrations of PAN,  showed  that  PAN persists
longer than ozone indoors.  (This is to be expected from its enhanced stability
at cooler-than-ambient temperatures such as found in air-conditioned.buildings.)
Tuazon et al.  (1981b)  demonstrated that PAN persists  in  ambient air longer
than ozone,  its  persistence paralleling that of nitric acid, at least in the
locality studied  (Claremont,  CA).   Reactivity  data presented  in the 1978
criteria document for  ozone and other photochemical oxidants indicated  that
all precursors that  give rise to PAN also give rise to ozone.   The data also
showed, however,  that  not all precursors giving rise to ozone also give rise
to PAN, and  that not all  that  give  rise  to both are equally reactive toward
both, with some  precursors preferentially giving rise, on the basis of units
of product per unit  of reactant, to more of one product than the other (U.S.
Environmental Protection Agency, 1978).
     In the review cited earlier, Altshuller (1983) examined the relationships
between ozone  and a  variety  of  other  smog components, including  PAN, PPN,
HpOpj HCOOH, aldehydes,  aerosols,  and nitric acid.  He  concluded that "the
ambient air  measurements  indicate that  ozone  may serve directionally,  but
cannot be  expected to  serve  quantitatively,  as a  surrogate for the other
products" (Altshuller, 1983).   It must be emphasized that the issue Altshuller
examined was  whether ozone could serve as an  abatement  surrogate for all
photochemical  products,  not just the subset of non-ozone oxidants of concern
in this document.  Nevertheless, a review of the data presented indicates that
his conclusion  is applicable  to the  non-ozone oxidants examined in this  docu-
ment.
5.9  REFERENCES
Altshuller, A.  P.  (1975) Evaluation of oxidant  results  at CAMP  sites  in  the
     United States. J. Air Pollut. Control Assoc. 25: 19-24.
Altshuller, A.  P.  (1983) Measurements of the  products of atmospheric  photo-
     chemical  reactions  in  laboratory studies and  in ambient air—relation-
     ships between  ozone and other products.  Atmos.  Environ.  17:  2383-2427.
Berk, J. V.; Young, R.;  Hollowell, C. D.; et al. (1980) The effects of energy-
     efficient  ventilation  rates on indoor air quality in an Ohio elementary
     school.  Berkeley,  CA:  University of  California,  Lawrence Berkeley
     Laboratory;  report  no.  EEB-Vent 80-9.  Available from:  NTIS,  Springfield,
     VA; LBL-10223.
                                   5-116

-------
Berk, J. V,;  Young,  R.  A.; Brown,  S.  R.;  Hollowell, C. D. (1981) Impact of
     energy-conserving retrofits on indoor air quality  in residential housing.
     Presented at: 74th  annual  meeting of the Air Pollution Control Assoc-
     ciation; June;  Philadelphia,  PA.  Pittsburgh, PA:  Air  Pollution Control
     Association; paper no. 81-22.1

Berry, C.  R.  (1964)  Differences in concentrations of surface oxidant between
     valley and  mountaintop  conditions in the southern Appalachians. J. Air.
     Pollut. Control Assoc. 14:238-239.

Blumenthal, D. L,; White, W.  H.; Peace, R. L ; Smith T. B.  (1974) Determination
     of the feasibility of the  long-range transport of  ozone or ozone precursors.
     EPA-450/3-74-061, U.S. Environmental Protection Agency, Research Triangle
     Park, NC.

Brennan, E.  (1980)  PAN  concentrations in ambient air in New Brunswick, NJ.
     Final  report.  Prepared  under  grant  805827 for U.S. Environmental  Protec-
     tion Agency, Research Triangle Park, NC.

Bufalini,  J.  J.;  Gay, B. W., Jr.;  Brubaker,  K.   L. (1972)  Hydrogen  peroxide
     formation  from  formaldehyde  photooxidation  and its presence  in  urban
     atmospheres. Environ. Sci. Techno!. 6: 816-821.

Cherniak,  E. A.; Corkum, R. (1981) Continuous monitoring of PAN in the  polluted
     troposphere. Brock University, St. Catherines, ON, Canada. Cited  in Temple,
     P. J.; Taylor, O.C. (1983).

Cleveland, W. S.; Graedel, T. E.;  Kleiner, B. (1976a) Photochemical  air pollu-
     tion:  transport  from  the  New York City  area into  Connecticut and  Massa-
     chusetts. Science (Washington, DC) 191:  179-181.

Cleveland, W. S.; Guarino, R.;  Kleiner, B.; McRae, J. E,; Warner, J.  L. (1976b)
     The analysis  of the ozone problem  in  the  northeast United States. In:
     Specialty  Conference  on:  Ozone/oxidants—interactions with the  total
     environment.  Pittsburgh,  PA:  Air Pollution  Control  Association; pp.
     109-120.

Coffey,  P.  E.;  Stasiuk,  W. N.  (1975a) Evidence  of atmospheric transport of
     ozone  into  urban areas.   Environ.  Sci. Technol. 9:  59-62.

Coffey, P.  E.; Stasiuk, W. N. (1975b)  Urban ozone:  its  local and extraregional
     components.  Presented at  the 79th annual  national  meeting, American
     Institute  of Chemical  Engineers; March; Houston,  TX. Paper no.  60C.

Coffey,  P.  E.;  Stasiuk,  W. N.;  Mohnen,  V.  (1977) Ozone  in rural  and urban
     areas  of New York State.  In: Dimitriades, B., ed. International  confer-
     ence  on  photochemical oxidant pollution and its control.  Proceedings:  v.
     I.  Research Triangle Park,  NC:   U.S. Environmental  Protection  Agency,
     Office  of  Research  and  Development; pp. 89-96;  EPA report no.  EPA-600/
     3-77-OOla.  Available  from: NTIS,  Springfield, VA;  PB-624232.
                                    5-117

-------
Content, C.  F.;  Gehan,  B. M,;  Stock,  T.  H.;  Holguin, A. H.; Buffler, P. A.
     (1985)  Estimation  of Individual  ozone exposures  using microenvironment
     measurements.  In:  APCA Specialty Conference: Scientific  Basis  for the
     Ozone Standard,  Houston, TX, November 28-December 1,  1984.  Pittsburgh;
     APCA.  In Press.

Cox, R. A.;  Roffey, M. J. (1977)  Thermal decomposition of peroxyacetyl nitrate
     in the  presence of  nitric oxide.  Environ.  Sci.  Techno!.  11:  900-906.

Parley, E. F.;  Kettner, K.  A.; Stephens, E. R.  (1963) Analysis of  peroxyacyl
     nitrates  by gas chromatography with  electron capture detection. Anal.
     Chem.  35: 589-591.

Davies, T. D.;  Ramer, B.; Kaspyzok, G.;  Delany, A.  C.  (1984) Indoor/outdoor
     ozone concentrations at  a  contemporary art  gallery. J. Air Pollut.  Control
     Assoc.  31: 135-137.

Edinger, J,  G. (1973) Vertical  distribution of photochemical  smog in  Los  Angeles
     basin.  Environ.  Sci. Techno!. 7: 247-252.

Edinger, J.  G.; McCutchan, M. H.; Miller, P. R.; Ryan, B. C.; Schroeder,  M.  J.;
     Behar,  J.  V.  (1972) Penetration  and  duration of  oxidant air pollution in  .
     the south  coast  air basin of California. J.  Air Pollut. Control Assoc.
     22: 882-886.

Elkus, B.; Wilson, K. R.  (1976) Photochemical oxidant:  weekday-weekend  differences.
     Draft report.   Cited in U.  S.  Environmental  Protection Agency  (1978).

Evans,  G.  F. (1985)  The National Air Pollution Background Network:   final
     project report.  Research  Triangle Park, NC:  U.  S. Environmental Protection
     Agency,  Office  of  Research  and Development;  EPA  report  no.  EPA-600/4-85-038.
     Available from:  NTIS, Springfield, VA; PB85-212413.

Evans, G.; Finkelstein, P.;  Martin, B.;  Possiel,  N.; Graves, M. (1982)  The
     National  Air  Pollution  Background Network, 1976-1980.  Research  Triangle
     Park, NC:  U.S.  Environmental Protection Agency;  EPA report  no.  EPA-600/
     4-82-058. Available  from:  NTIS, Springfield,  VA;  PB83-100412.

Evans, G.; Finkelstein, P.;  Martin,  B.; Possiel, N.;  Graves, M.  (1983).  Ozone
     measurements  from a network of remote  sites.  J.  Air Pollut. Control
     Assoc.  33:  291-296.

Farmer, J. C.; Dawson, G. A.  (1982)  Condensation sampling of soluble atmos-
     pheric  trace gases.  JGR  J. Geophys.  Res. 87:  8931-8942.

Galbally,  I. E.; Roy, C.  R.  (1980) Destruction of  ozone  at  the  earth's surface.
     Quart.  J.  R. Met.  Soc.  106:  599-620.

Graedel, T.  E.;  Farrow,  L. A.;  Weber, T.  A. (1977) Photochemistry of  the "Sunday
     effect."  Environ.  Sci. Techno!. 11:  690-694.

Grosjean,  D. (1981) Critical evaluation and comparison of measurement methods
     for nitrogenous  compounds in the atmosphere.  Final  report, A 706-05, for
     Coordinating  Research  Council.  Environmental Research  and  Technology.
     Westlake  Village,  CA.

                                   5-118

-------
Grosjean, D.  (1983)  Distribution of atmospheric  nitrogenous  pollutants at a
     Los Angeles  area  smog receptor site.  Environ.  Sci.  Techno!.  17:  13-19.

Hanst, P. L. ; Wilson, W. E.; Patterson, R.  K.; Gay, B. W., Jr.;  Chaney,  L. W.;
     Burton,  C. S.  (1975)  A spectroscopic  study  of California  smog. Research
     Triangle Park,  NC:  U.S.  Environmental Protection Agency;  EPA  report  no.
     EPA-650/4-75-006.   Available  from:  NTIS, Springfield,  VA; PB-241022.

Hanst, P. L.; Wong, N.  W.; Bragin, J. (1982)  A long path  infrared study of Los
     Angeles  smog. Atmos. Environ. 16: 969-981.

Heikes, B.  G.; Lazrus,  A. L.; Kok, G. L.;  Kunen,  S. M.; Gandrud, B. W.;  Gitlin,
     S.  N. ;  Sperry,  P.  D.  (1982)  Evidence  of aqueous phase hydrogen peroxide
     synthesis in the troposphere. JGR J.   Geophys. Res. 87: 3045-3051.

Heuss, J. M.; Glasson,  W. A. (1968) Hydrocarbon reactivity and  eye  irritation.
     Environ. Sci. Techno!. 2: 1109-1116.

Hoggan,  M.   (1986) Comment  on  "Ten-year  ozone trends  in California and Texas."
     J. Air Pollut. Control Assoc. 36: 599-600.

Holdren, M.  W. ; Spicer,  C.  W.;  Hales, J.  M. (1984) Peroxyacetyl nitrate solu-
     bility and decomposition rate in water.  Atmos. Environ.:  in press.

Holguin, A.  H. ;  Buffler, P. A.;  Contant,  C.  F.,  Jr.; Stock,  T.  H.; Kotchmar,
     D.  J.;  Hsi,  B.  P.; Jenkins,  D.  E.;  Gehan,   B. M.; Noel,  L. M. ; Mei,  M.
     (1985) The effects  of ozone  on asthmatics in the Houston area. In:  Proceed-
     ings of  APCA Speciality Conference:   Evaluation  of  the Scientific Basis
     for Ozone/Oxidants Standards.   Pittsburgh,  PA:  Air Pollution Control
     Assoc. Houston, TX, November 28-30,  1984.

Hov,  0.; Hesstvedt, E. ;  Isaksen, I. S.  A. (1978) Long-range  transport of
     tropospheric ozone. Nature  (London)  1:  341-344.

Hunt, W. F.,  Jr.; Curran, T. C.  (1982) National and regional  trends in ambient
     ozone  measurements,  1975-1981.  Research Triangle Park,  NC: U.S.  Environ-
     mental  Protection  Agency,  Office of  Air Quality Planning and  Standards.

Isaksen, I.  S.  A. ; Hov, 0.; Hesstvedt, E.  (1978) Ozone generation  over rural
     areas. Environ. Sci. Technol. 12: 1279-1284.

Jorgen,  R.   T.; Meyer, R. A.; Hughes,  R. A.  (1978) Routine peroxyacetyl  nitrate
     (PAN)  monitoring applied to  the  Houston  Area Oxidant Study.  Presented at:
     71st  annual  meeting  of  the  Air  Pollution   Control   Association;  June;
     Houston, TX. Pittsburgh, PA:  Air Pollution Control Association; paper no.
     78-50.1.

Kelly,  T.  J. ; Stedman,  D.  H.;  Kok, G.  L.  (1979)  Measurements of H202  and HN03
     in  rural air. Geophys. Res.  Lett. 6:  375-378.

Kelly,  N. A.; Ferman, M. A.; Wolff, G. T.  (1986)  The  chemical  and meteorological
     conditions associated with  high and  low ozone  concentrations in south-
     eastern  Michigan and  nearby  areas of Ontario. J. Air Pollut. Control  Assoc.
     36: 150-158.


                                    5-119

-------
Kok, G.  L.;  Darnall,  K.  R.; Winer,  A.  M.;  Pitts, J. N.,  Jr.;  Gay, B. W.;
     (1978) Ambient air  measurements of hydrogen peroxide in the California
     South Coast Air Basin. Environ. Sci. Techno!. 12: 1077-1080.

Kress, L. W.; Miller,  J.  E. (1983) Impact of ozone on soybean yield. J. Environ.
     Qual. 12: 276-281.

Kroening,  J.  L.;  Ney,  E.  P.  (1962) Atmospheric ozone.  J.  Geophys.  Res.
     67:  1867-1875.

Lefohn, A. S.  (1982)  The identification and  description of  ozone monitoring
     sites which  represent high yield  reduction potential for agricultural
     crops. Prepared  for U.S.  Environmental  Protection  Agency,  Corvallis, OR.
     A.S.L. and Assoc. Helena, MT.

Lefohn,  A. S.  (1984)  A  comparison  of  ambient ozone exposures  for selected
     nonurban sites.  Proceedings  of 77th Annual Meeting of the Air  Pollution
     Control   Association.  Vol. 6.  Paper  no.  84-104.1.   Pittsburgh,  PA:  Air
     Pollution Control Assoc.

Lefohn, A. S.; Benedict,1 H.  M.  (1985)  Exposure considerations associated with
     characterizing ozone ambient air quality  monitoring data.  In:  Proceedings
     of the  APCA  Specialty Conference:  Evaluation of the Scientific Basis for
     Ozone/Oxidants Standards,  Houston,  TX,  November 28-30,  1984.  Pittsburgh,
     PA:  Air Pollution Control Assoc.

Lefohn, A. S.; Jones, C. K. (1986)  The characterization  of  03  and S02  air
     quality  data for assessing possible vegetation  effects.  J.  Air Pollut.
     Control  Assoc.  36:  (In Press).  (Presented at annual meeting  of  Air Pollu-
     tion Control Assoc., Minneapolis, MN, July  1986.)

Levitt, S. B.; Chock,  D. P. (1976) Weekday-weekend pollutant  studies on the  Los
     Angeles Basin.  J. Air Pollut. Control Assoc. 26: 1091-1092.

Lewis, T. E.; Brennan, E.; Lonneman, W. A. (1983) PAN concentrations in ambient
     air  in New Jersey.  J.  Air Pollut. Control Assoc. 33:  885-887.

Logan, J. A.; Prather, M. J.; Wofsy, S. C.; McElroy, M.  B. (1981)  Tropospheric
     chemistry: a global  perspective.  JGR  J. Geophys.  Res.  86:  7210-7254.

Lonneman, W.  A.;  Bufalini, J. J.; Seila, R.  L. (1976) PAN  and oxidant  measure-
     ment  in ambient  atmospheres. Environ. Sci.  Techno!. 10:  374-380.

Martinez,  J.  R.;  Singh,  H. B.  (1979)  Survey of the  role  of  NOx  in  nonurban
     ozone formation.  Research Triangle Park,  NC: prepared by SRI  International
     for  U.S. Environmental Protection Agency. EPA report  no.  EPA-450/4-79-035.
     Available from:  NTIS, Springfield, VA;  PB80-122815.

Mayrsohn,  H.;  Brooks, C. (1965) The analysis of PAN by electron capture gas
     chromatography.  Presented at:  Western regional  meeting  of the American
     Chemical  Society;  November;  Los Angeles, CA.  Los Angeles, CA: California
     State Department of Public Health.
                                    5-120

-------
McElroy, M.  B.; Wofsy,  S.  C.;  Logan,  J.  A.;  Prather,  M.  J.  (1985)  Theoretical
     studies of long-range transport of oxidants. Annual technical report to
     Coordinating Research Council. Atlanta, GA: Coordinating Research Council,
     CRC Report No. CAPA-22-83.

Miller, A.; Ahrens, C. P.  (1969) Ozone within and below the West Coast tempera-
     ture  inversion.  San  Jose,  CA;  San Jose State  College.  Report No.  6.

Miller, P.  R. ;  McCutchan,  M. H.; Ryan, B. C. (1972)  Influence of  climate and
     topography on  oxidant air pollution concentrations  that  damage  conifer
     forests in sourthern  California. Mitt.  Forst. Bundesversuchsanst.  Wien
     97: 585-607.

Mohnen, V.  A.  (1977) Review and analysis.  In:  International  Conference on
     Oxidants,  1976 — Analysis of  evidence and viewpoints. Part III.  The
     issue  of  stratospheric  ozone  intrusion.  U.S.  Environmental  Protection
     Agency, Research Triangle Park,  NC; Report no. EPA-600/3-77-115.

Moschandreas,  D.  J. ;  Stark, J.  W. C. ;  McFadden,  J.   E.;  Morse, S.  S. (1978)
     Indoor air pollution  in  the residential environment. Vol. I.  Data collec-
     tion,  analysis,  and  interpretation.  Research Triangle Park,  NC:  U.S.
     Environmental  Protection Agency; EPA report  no.  EPA-600/7-78-229a.  Avail-
     able  from: NTIS, Springfield, VA; PB-290999.

Moschandreas,  D.  J.;  Zabransky,  J.;  Petton,  D.  J.  (1981) Comparison of indoor
     and  outdoor  air quality.  Palo  Alto, CA:  Geomet, Inc. Electric Power
     Research  Institute. Report  no.  EA-1733. Available from: NTIS,  Springfield,
     VA; EPRI-EA-1733.

Mueller,  F.  X.;  Loeb,  L.;  Mapes, W.  H.  (1973)  Decomposition rates  of ozone in
     living areas.  Environ.  Sci. Techno!. 7: 342-346.

National  Research Council,  (1977)  Ozone and other  photochemical  oxidants.
     Washington,  DC; National Academy of Sciences.

Nieboer,  H.;  van Ham, J.  (1976) Perpxyacetyl  nitrate (PAN) in  relation to
     ozone  and some meteorological  parameters  at Delft  in the Netherlands.
     Atmos. Environ. 10: 115-120.

Nielsen, T.; Samuelsson, U.;  Grennfelt,  P.;  Thomsen,  E.  L.  (1981)  Peroxyacetyl
     nitrate  in  long-range  transported  polluted air. Nature  (London)  293:
     553-555.

Oltmans, S. J.  (1981) Surface ozone  measurements  in clean air. J.  Geophys.  Res.
     86: 1174-1180. Cited  in:  Logan  et al.,  1981.

Peaks,  E.; MacLean, M. A.;  Sandhu, H. S.  (1983) Surface  ozone  and  peroxyacetyl
     nitrate (PAN)  observations  at rural  locations in Alberta, Canada.   J. Air
     Pollut. Control Assoc.  33:881-883.

Peterson,  G.  A.;  Sabersky,  R.  H.  (1975) Measurements of pollutants inside  an
     automobile.  J. Air  Pollut.  Control  Assoc.  25: 1028-1032.
                                    5-121

-------
Pitts, J. N.,  Jr.;  Grosjean, D. (1979)  Detailed  characterization  of  gaseous
     and size-resolved participate  pollutants  at  a  south  coast  air basin  smog
     receptor  site:  levels and modes  of formation  of sulfate,  nitrate and
     organic  particulates  and  their  implications  for  control  strategies.
     Sacramento, CA:  California Air Resources Board; report no. ARB-R-5-384-
     79-100. Available from: NTIS,  Springfield, VA; PB-301294.

Pratt, G.  C.; Hendrickson,  R.  C.;  Chevone, B. I.;  Christopherson, D.  A.;
     O'Brien, M. V.;  Krupa, S.  V.  (1983) Ozone and oxides of nitrogen in the
     rural  upper-midwestern U.S.A.  Atmos. Environ. 17: 2013-2023.

Reiter, E.   R.  (1977a) Review and analysis.  In:  International  Conference on
     Oxidants 1976 -- Analysis of evidence and viewpoints. Part III. The  issue
     of stratospheric ozone intrusion.  U.S.  Environmental Protection Agency,
     Research Triangle Park, NC; Report  no. EPA-600/3-77-115.

Reiter, E.   R.  (1977b) The role of stratospheric  import or tropospheric ozone
     concentrations.  In:  International  Conference  on Photochemical Oxidant
     Pollution and  Its  Control. Proceedings. V.  I. Dimitriades, B., ed.  U.S.
     Environmental  Protection Agency,  Research Triangle Park, NC;  Report no.
     EPA-600/3-77-001a.

Renzetti, N.  A.;  Bryan,  R. J.  (1961)  Atmospheric sampling  for  aldehydes  and
     eye irritation  in  Los Angeles  smog  - 1960. J.  Air  Pollut.  Control  Assoc.
     11: 421-424.

Rhodes, C.   E.;  Holland,  D.  M. (1981)  Variations of  NO,  N02,  and 03 concentra-
     tions   downwind of  a  Los Angeles  freeway.  Atmos. Environ.  15: 243-250.

Sabersky, R.  H.; Sinema,  D. A.; Shair,  F.  H.  (1973) Concentrations, decay
     rates,  and removal of  ozone and their relation to  establishing clean
     indoor air. Environ.  Sci. Techno!.  1: 347-353.

SAROAD, Storage  and Retrieval  of Aerometric  Data [data base].  (1985a) Data
     file for  1976.  Research Triangle Park,  NC: U.S.  Environmental Protection
     Agency, Office of Air Quality Planning and Standards. Disc; ASCII.

SAROAD, Storage  and Retrieval  of Aerometric  Data [data base].  (1985b) Data
     file for  1979.  Research Triangle Park,  NC: U.S.  Environmental Protection
     Agency, Office of Air Quality Planning and Standards. Disc; ASCII.

SAROAD, Storage  and Retrieval  of Aerometric  Data [data base].  (1985c) Data
     file for  1980.  Research Triangle Park,  NC: U.S.  Environmental Protection
     Agency, Office of Air Quality Planning and Standards. Disc; ASCII.

SAROAD, Storage  and Retrieval  of Aerometric  Data [data base].  (1985d) Data
     file for  1981.  Research Triangle Park,  NC: U.S.  Environmental Protection
     Agency, Office of Air Quality Planning and Standards. Disc; ASCII.

SAROAD, Storage  and Retrieval  of Aerometric  Data [data base].  (1985e) Data
     file for  1982.  Research Triangle Park,  NC: U.S.  Environmental Protection
     Agency, Office of Air Quality  Planning and Standards. Disc; ASCII.
                                   5-122

-------
SARGAD, Storage and  Retrieval  of Aerometric Data  [data base].  (1985f) Data
     file for  1983.  Research Triangle  Park,  NC:  U.S.  Environmental  Protection
     Agency, Office of Air Quality Planning and Standards. Disc; ASCII.

Schuck, E.  A.;  Pitts,  J.  N., Jr.; Wan, J. K. S. (1966) Relationships between
     certain meteorological  factors  and photochemical  smog.  Int.  J.  Air Water
     Pollut. 10: 689-711.

Seiler, W.;  Fishman,  J.  (19810 The  distribution of carbon monoxide  and ozone
     in the free troposphere. JGR J. Geophys. Res. 86:  7255-7265.

Shair,  F. H.;  Heitner, K.  L.  (1974) Theoretical model for  relating indoor
     pollutant  concentrations  to  those  outside.  Environ.  Set.  Techno!.
     8: 444-451.                           ...           .    <

Singh, H. B.; Hanst, P. W.  (1981) Peroxyacetyl nitrate  (PAN)  in the  unpolluted
     atmosphere:  an  important reservoir  for  nitrogen oxides. Geophys. Res.
     Lett. 8: 941-944.

Singh, H. B.;,Salas, L. J.  (1983) Methodology for the analysis of peroxyacetyl
     nitrate (PAN) in the unpolluted atmosphere. Atmos. Environ. 17: 1507-1516.

Singh,  H.  B.;  Ludwig, F.  L.;  Johnson, W. B.  (1977)  Ozone in clean remote,
     atmospheres.  Concentrations and variabilities. SRI Project No.  5661.  Menlo
     Park, CA:  SRI International.                                 ,,

Singh, H. B.; Salas, L. J.; Shigeishi,  H.; Smith, A, J.;  Scribner, E.;  Gavanagh,
     L.  A.  (1979) Atmospheric distributions,  sources  and sinks of  selected
     halocarbons, hydrocarbons, SF6, and  N20. Research  Triangle Park, NC:  U.S.
     Environmental Protection Agency,  Environmental Sciences  Research  Laboratory;
     EPA  report no.  EPA-600/3-79-107.  Available from:  NTIS,  Springfield,  VA;
     PB80-142706.

Singh, H. B.; Salas, L. J. ; Smith, A.  J..; Shigeishi, H. (1981) Measurements of
     some  potentially hazardous  organic chemicals  in  urban environments.
     Atmos.   Environ. 15: 601-612.

Singh,  H. B. ,  Salas, L.  J.; Stiles, R.;  Shigeishi, H.  (1982) Measurements of
     hazardous  organic  chemicals in the ambient atmosphere.  Research Triangle
     Park,  NC:  U.S.  Environmental Protection Agency,  Environmental  Sciences
     Research  Laboratory.   EPA report  no. EPA-600/3-83-002.  Available  from:
     NTIS,  Springfield, VA; PB83-156935.

Smith, W. J. (1981)  New York State air monitoring data  report for  the  Northeast
     Corridor  Regional  Modeling  Project  (NECRMP). Albany, NY: New York State
     Department of Environment and Conservation.

Spicer,  C.  W.  (1976) The fate of nitrogen oxides  in  the  atmosphere.  Research
     Triangle  Park,  NC:  U.S.  Environmental  Protection  Agency; EPA report  no.
     EPA-600/3-76-030.  Available from:  NTIS,  Springfield,  VA;  PB-248-260.

Spicer, C.  W.  (1977) The  fate, of  nitrogen oxides  in the atmosphere.  In:  Pitts,
     J.  N.,  Jr.;  Metcalf, R.  L.; Lloyd, A.  C.; eds.  Advances in environmental
     science and  technology:  v.  7.  New York,  NY:  John Wiley and  Sons; pp.
     163-261.

                                    5-123

-------
Spieer, C. W.;  Sverdrup, G. W.  (1981).  Trace nitrogen chemistry during the
     Philadelphia oxidant  data enhancement study (1979).  Report on Contract
     no. 68-02-0338. U.S.  Environmental  Protection Agency, Research Triangle
     Park, NC.

Spicer, C. W. ; Gemma, J. L; Joseph, D. W.; Sticksel, P. R. ; Ward, G. F. (1976)
     The  transport  of  oxidant beyond urban  areas.  Columbus, OH:  Battelle
     Columbus Laboratories;  EPA report no.  EPA-600/3-76-018a.  Available from:
     NTIS, Springfield, VA; PB-253736.
Spicer, C. W.;  Holdren,
     acetyl  nitrate  in
     17: 1055-1058.
M. W.; Keigley,
the  continental
G. W. (1983) The
 boundary layer.
ubiquity of peroxy-
 Atmos.  Environ.
Stock, T.  H.;  Holguin, A.  H.;  Selwyn,  B.  J.; Hsi,  B.  P.; Contant, C.  F. ;
     Buffler, P. A.; Kotchmar, D. J.  (1983). Exposure estimates  for the  Houston
     area asthma and runners studies. In: Lee, S. D.; Mustafa, M. G.;  Mehlman,
     M. A.  eds. Adv. Modern Environ.  Toxicol.  Vol.  V.  International Symposium
     on the Biomedical  Effects of Ozone and Related Photochemical  Oxidants.
     Pinehurst, NC;  March 14-18, 1982.  Princeton,  NJ:  Princeton Scientific
     Publishers, Inc.

Taylor, 0.  C.  (1969) Importance of  peroxyacetyl  nitrate (PAN) as a phytotoxic
     air pollutant. J.  Air. Pollut. Control  Assoc.  19:  347-351.
Temple, P. J.; Taylor, 0.
     acetyl nitrate  (PAN)
     17: 1583-1587.
 C. (1983) World-wide ambient measurements  of  peroxy-
  and implications  for plant injury.  Atmos.  Environ.
Thompson, C.  R.;  Hensel,  E. G.; Kats, G.  (1973) Outdoor-indoor  levels  of  six
     air pollutants. J. Air Pollut. Control Assoc.  23:  881-886.

Tingey, D.  T.;  Hill, A. C. (1968) The occurrence  of photochemical  phytotoxi-
     cants  in the Salt  Lake Valley. Utah Acad.  Sci. Proc.  44:  387-395.

Tuazon, E.  C.;  Graham, R.  A.;  Winer, A. M.;  Easton, R.  R.;  Pitts,  J. N.,  Jr.
     (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, E.  C.;  Winer, A.  M.;  Graham,  R.  A.;  Pitts, J. N., Jr. (1980) Atmos-
     pheric  measurements  of  trace  pollutants by  kilometer-pathlength  FT-IR
     spectroscopy.  In:  Pitts,  J. N.; Metcalf,  R. L., eds.  Advances  in environ-
     mental sciences and technology: v. 10. New York,  NY:  John Wiley and Sons;
     pp. 254-300.

Tuazon, E.  C.;  Winer, A.  M.;  Graham, R. A.;  Pitts, J.  N.,  Jr. (1981a)  Atmos-
     pheric  measurements  of  trace  pollutants: long path Fourier transform
     infrared  spectroscopy.  Research Triangle Park,  NC:  U.S. Environmental
     Protection  Agency,  Environmental  Sciences Research Laboratory. Available
     from:  NTIS, Springfield,  VA; PB81-179848.
                                    5-124

-------
Tuazon, E. C.; Winer, A. M.; Pitts, J. N., Jr. (1981b) Trace pollutant concen-
     trations in a  multiday smog episode  in  the  California South Coast Air
     Basin by long path length Fourier transform infrared spectrometry. Environ.
     Sci.  Techno!.  15: 1232-1237.

U.S. Department of Commerce, Bureau of the Census. (1982) Statistical Abstract
     of the United States, 1982-1983. 103rd Ed. Washington, DC.

U.S. Department of Health, Education, and Welfare. (1970) Air quality criteria
     for photochemical oxidants. Washington, DC: National Air Pollution Control
     Administration; publication no. AP-63. Available from: NTIS, Springfield,
     VA; PB82-242231.

U.S. Environmental  Protection  Agency.  (1978) Air  quality  criteria  for ozone
     and other photochemical oxidants. Research Triangle Park, NC: U.S. Environ-
     mental  Protection  Agency,  Environmental  Criteria and Assessment Office;
     EPA  report  no.  EPA-600/8-78-004.  Available from: NTIS, Springfield, VA;
     PB80-124 753.

U.  S.  Environmental  Protection Agency.  (1980) Air quality data—1979 annual
     statistics. Research  Triangle Park, NC:  U.  S.  Environmental Protection
     Agency,  Office of Air Quality Planning  and Standards; EPA  report  no.
     EPA-450/4-80-014.

U.  S.  Environmental  Protection Agency.  (1981) Air quality data—1980 annual
     statistics. Research  Triangle Park, NC:  U.  S.  Environmental Protection
     Agency,  Office of Air Quality Planning  and Standards; EPA  report  no.
     EPA-45Q/4-81-027.

U.  S.  Environmental  Protection Agency.  (1982a) Air quality data—1981 annual
     statistics.  Research Triangle Park,  NC: U.  S.  Environmental Protection
     Agency,  Office of Air Quality Planning  and Standards; EPA  report  no.
     EPA-450/4-82-007.

U.S. Environmental  Protection  Agency.  (1982b) Air quality criteria for oxides
     of nitrogen.  Research Triangle Park, NC:  U.S.  Environmental Protection
     Agency,  Environmental Criteria  and Assessment Office; publication no.
     EPA-600/8-82-026.  Available from:  NTIS,  Springfield, VA; PB83-131011.

U.S. Environmental  Protection Agency.  (1984a) National air quality and emission
     trends  report, 1983.  Research Triangle Park,  NC:  U.S.  Environmental
     Protection Agency.  Office  of  Air Quality Planning and Standards.  EPA
     report  no. EPA-450/4-84-029.

U.S. Environmental  Protection  Agency.  (1984b) National  air pollutant emission
     estimates,  1940-1983. Research  Triangle Park,  NC:  U.S.  Environmental
     Protection  Agency, Office  of Air Quality Planning and  Standards.  EPA
     report  no. EPA-450/4-84-028.

U.S. Environmental  Protection Agency.  (1986)  National air  quality and  emissions
     trends  report,  1984.  Research Triangle  Park,  NC:  Office of Air Quality
     Planning and Standards. Report no.  EPA-450/4-86-001,
                                    5-125

-------
Viezee, W.;  Johnson, W.  B.;  Singh, H.  B.  (1979) Airborne measurements of
     stratospheric  ozone intrusions into  the troposphere  over the United
     States.  Final  report.  SRI  Proj.  6690, CRC-APRAC Proj.  no.  CAPA-15-76
     (1-77), Coordinating Research Council, Atlanta, GA.

Walker, H.  M.  (1985) Ten-year ozone trends in California and Texas. J, Air
     Pollut. Control Assoc. 35: 903-912.

Walker, H.  M.  (1986) Author's reply [to Comment  on  "Ten-year ozone trends  in
     California and Texas"]. J. Air Pollut. Control  Assoc, 36: 600-602.

Wendland, W. M.; Bark, L. D.; Clark, D.  R.; Curry, R. B.; Enz, J. W.; Hubbard,
     K. G.; Jones, V.; Kuehnast, E. L.;  Lytle, W,; Newman, J.; Nurnberger,  F. V.;
     Waite, P.  (1984)  A climatic review of summer 1983  in the upper Midwest.
     Bull. Amer. Meteor. Soc. 65(10).

Wesely, M. (1983) Turbulent transport  of ozone to surfaces common in the eastern
     half of the United  States. In: Trace  atmospheric constituents: Properties,
     transformations, and fates. Schwartz,  S.  E., ed. New York, NY: John Wiley
     & Sons, Inc.

Westberg, H.;  Allwine,  K.;  Robinson,  E.  (1978) Measurement of light hydrocar-
     bons and  oxidant  transport:  Houston  area  1976.  Research Triangle Park,
     NC: U.S.  Environmental Protection Agency, Office of Research and Develop-
     ment; EPA report no. EPA-600/3-78-662. Available from: NTIS, Springfield,
     VA; PB-285891.

Wolff, G. T.;  Lioy, P. J. (1980) Development of an ozone river associated with
     synoptic  scale episodes  in  the  eastern  United States.  Environ.  Sci.
     Techno!.  14: 1257-1260.

Wolff, G.  T.;  Stasiuk,  W. N.;  Coffey, P.  E.; Pasceri,  R.  E.  (1975) Aerial
     ozone  measurements  over  New Jersey, New York,  and Connecticut.  Presented
     at  the 68th Annual  Meeting  of the Air Pollution  Control  Association,
     Boston,  MA:  June.   Pittsburgh,  PA:  Air Pollution  Control  Association.
     Paper no. 75-58.6.

Wolff,  G.  T.; Lioy, P.  T.; Wight,  G.  D.;  Meyers,  R.  F.;  Cederwall, R. T.
     (1977a)  An investigation  of  long-range transport  of ozone  across the
     midwestern  and eastern United States.  In:  Dimitriades,  B., ed.  Inter-
     national  conference on photochemical   oxidant pollution  and  its control.
     Proceedings:  volume I.  Research  Triangle Park, NC: U.S.  Environmental
     Protection  Agency,  Office of Research and  Development;  pp.  307-317; EPA
     report  no.  EPA-600/3-77-001a.  Available  from:  NTIS,  Springfield, VA;
     PB-264232.

Wolff,  G.  T.;  Lioy,  P.  J.;  Wight, G.  D.; Pasceri, R.  E. (1977b) Aerial
     investigation  of photochemical oxidants over the  northeast.  In: Bufalini,
     J. J.;  Lonneman,  W. A.,  eds.  Proceedings of symposium on  1975  northeast
     oxidant transport  study,  Research Triangle  Park,  NC,  January 20-21, 1976.
     Research  Triangle  Park,  NC:  U.S.  Environmental  Protection Agency.  Report
     no.  EPA-600/3-77-017.
                                    5-126

-------
Wolff, G.  T.;  Kelly,  N.  A.;  Ferman, M. A. (1981) On the sources of  summertime
     haze  in the eastern United  States.  Science (Washington, DC) 211: 703-705.

Wolff, G.  T.; Lioy, P. J.; Taylor,  R.  S.  (1986) The diurnal variations of ozone
     at  different  altitudes  on a rural mountain in the eastern United  States.
     J. Air Pollut. Control  Assoc.  36:  (In Press)

Yocom, J.  E.   (1982)  Indoor-outdoor  air quality  relationships:  a critical
     review. J. Air Pollut.  Control  Assoc.  32:  500-520.

Zagraniski, R. T.; Leaderer,  B.  P.;  Stolwijk,  J.  A. J.  (1979) Ambient sulfates,
     photochemical oxidants,  and acute adverse health effects: an epidemiolog-
     ical  study. Environ.  Res. 19:  306-320.
                                    5-127


                                         U.S. GOVEMMENT PRIWOTG OFFICE! 1986- 6 % 6- 1 1 6/ i»0661

-------

-------

-------
Environmental Protection
Agency
Information
Cincinnati OH 45268
Official Business
Penalty for Private Use, $300
                                                                  Please make all necessary changes on tha abovt label,
                                                                  detach or copy, and return to the address in the upper
                                                                  left-hand corner.

                                                                  if you do not wish lo receive these reports CHECK HERE o;
                                                                  detach, or copy this cover, and return to the address in the
                                                                  upper left-hand corner.
                                                                EPA/600/8-84/020bF

-------