^&nn
               "•';^f
U. S. ENVIRONMENTAL PROTECTION AGENCY

-------
                                    PROCEEDINGS
                                                  OF
                                      SYMPOSIUM
                                                 ON
                              MULTIPLE-SOURCE
                              URBAN DIFFUSION
                                           MODELS
                                    Editor: Arthur C. Stern
Sponsors:  National Air Pollution Control Administration
and North Carolina Consortium on Air Pollution
U.S. ENVIRONMENTAL PROTECTION AGENCY
Air Pollution Control Office
Research Triangle Park, North Carolina, 1970
      For sale by the Superintendent of Documents, U.S. Government Printing Office
                 Washington, D.C., 20402 - Price $1.75

-------
The AP series of reports  is issued by  the  Air  Pollution  Control  Office to
report the results of scientific and engineering  studies, and information of
general  interest in the field of  air pollution. Information reported in this
series includes coverage of  APCO intramural activities and of  cooperative
studies  conducted in conjunction with state and local  agencies, research insti-
tutes, and  industrial organizations. Copies of AP reports are available free of
charge to  APCO staff members, current contractors  and grantees,  and non-
profit organizations —  as  supplies permit — from the Office of Technical
Information and  Publications,  Air Pollution  Control  Office,  Environmental
Protection  Agency, Research  Triangle Park, North Carolina  27709
Air Pollution Control Office Publication  No. AP-86

-------
                                                            PREFACE
This symposium was conducted under the terms of a contract between the
Meteorology  Division  of the National  Air Pollution Control Administration
(NAPCA)  and the Department of Environmental Sciences and Engineering,
School of Public Health, University of North Carolina at Chapel  Hill  (UNC).
The contract officer for the sponsor was Mr.  Lawrence  E.  Niemeyer, As-
sistant Director, Division of  Meteorology,  NAPCA. The responsible  officer
for the  University was  Arthur C. Stern, Professor of Air Hygiene. The sym-
posium  was held from  October 27 to 30, 1969, at the  Carolina Inn  on the
University campus. On  December  4,  1970, the functions  of  NAPCA  were
transferred to the Air Pollution Control Office (APCO) of the Environmental
Protection Agency. All references to the former, therefore, now  refer to the
latter.
Although  UNC  was the contractor, it was agreed that the  symposium would
be  sponsored by the  North Carolina Consortium  on  Air Pollution,  com-
prising:  Duke University; North Carolina State University; the Office of Man-
power Development,  NAPCA; Research Triangle Institute; and the University
of North  Carolina  at Chapel  Hill.  The detailed  planning for the symposium
was  done by a  steering committee  representing the members  of the  Con-
sortium  and the contract officer.
All the  papers,  both  invited and volunteered, that were presented during the
symposium, are  included in  this volume. In  addition  to the invited  partici-
pants, some  attendees were  selected from persons who  responded  to the
public announcement  of the  symposium. Almost all papers were preprinted
and  distributed to participants in advance of the meeting. Although  there
was  open  discussion  after  every presentation except the Keynote and Ban-
quet speeches, the only discussions incorporated in the Proceedings are those
subsequently submitted in writing.
Questions  from the floor and authors' responses do not appear unless those
same questions and  answers were also included  in the written discussion.
Every author questioned was given an opportunity to submit a rebuttal. The
nature of  such  an arrangement made it necessary, for the sake of coherence,
to incorporate all discussion in a separate chapter, divided  into two sections:
speaker-directed discussions and discussions submitted by the participants.

-------
Particular thanks are given  to  the graduate students in Air Pollution in  my
department  at  UNC, who acted as floor  monitors during  the  symposium —
particularly  to  Harvey Jeffries  and  Douglas McKay, who managed the audio-
visual arrangements  throughout the symposium. The registration of partici-
pants and the  preparation of the  symposium program and information kits
were  ably handled  by the Continuing Education Department of the School
of Public  Health, UNC. I  am especially appreciative of the excellent services
of my secretary, Martha Davis, for  her help  in the preparation and conduct
of the symposium and  in the coordination of these proceedings.

                                                      Arthur C. Stern
                                                      Chapel Hill, N. C.
                                                      November   1970.

-------
                      STEERING COMMITTEE
Dr. Jay Apple, Director, Institute of Biological Sciences, North Carolina State
    University.

Dr. Robert Barnes, School of Forestry, Duke University.

Mr. James L. Dicke, Meteorologist, Office of Manpower Development, Air Pol-
    lution Control Office.

Dr. James S. Ferrell,  Department of Chemical Engineering, North Carolina
    State University.

Dr. Kenneth Knoerr, School of Forestry, Duke  University.

Dr. Harry Kramer, Director, Office of Manpower Development, Air Pollution
    Control  Office.

Mr. Lawrence Niemeyer, Assistant Director, Meteorology Division, Air Pollu-
    tion  Control Office.

Professor Arthur C. Stern,  Department of Environmental  Sciences and Engi-
    neering, School of Public Health, University of North Carolina at Chapel
    Hill.

Mr. James J. B.  Worth, Associate Director of Engineering, Research Triangle
    Institute.
                       SESSION CHAIRMEN

Mr. James J.  B.  Worth,  Associate Director of Engineering, Research Triangle
    Institute.

Dr. Lester Machta, Director, Air Resources Laboratories, National Oceanic and
    Atmospheric Administration.

Mr. Robert A. McCormick,  Director, Meteorology Program, Air Pollution Con-
    trol Office.

Dr. Kenneth Knoerr, School of Forestry, Duke University.

Dr. Walter J. Saucier, Department of Geosciences, North Carolina State Univer-
    sity.

-------
                                                        CONTENTS
 7.  John T. Middleton
    Keynote Speech: Diffusion Modeling for Urban Air Pollution
    Abatement and Control
 2. Heinz H. Lettau
   Physical  and Meteorological  Basis for Mathematical Models
   of Urban Diffusion Processes                                       2-1
 3. Frank Pasquill
   Prediction of Diffusion over an Urban Area—Current Practice
   and Future Prospects                                             3-1
4. Kenneth L. Calder
   Some  Miscellaneous Aspects of  Current  Urban  Pollution
   Models                                                         4-1
5. Warren B. Johnson, Francis L. Ludwig, and Albert E. Moon
   Development of a Practical, Multi-purpose Urban  Diffusion
   Model for Carbon Monoxide                                       5-1


6. John J.  Roberts, Edward J. Croke, and Allen S. Kennedy
   An Urban Atmospheric Dispersion'Model                            6-1
7. Glenn R. Hilst
   Sensitivities of  Air Quality Prediction to Input Errors and
   Uncertainties                                                   7-1
8. Shin'ichi Sakuraba in collaboration with M. Mariguchi and I.
   Yamazi
   Elevation of Tracer Cloud over an Urban Area
                                                                  8-1
9. Heinz G. Fortak
   Numerical Simulation of the Temporal and  Spatial  Distri-
   butions of Urban Air Pollution Concentration                       g ^
                                  VI

-------
10. Liau J. Shieh, Ben Davidson, andJ. P. Friend
    A Model of Diffusion in Urban Atmospheres: 862 in Greater
    New York                                                      10-1
11. James  Ft. Mahoney,  William  O. Maddaus,  and  John C.
    Goodrich
    Analysis of Multiple-Station Urban Air Sampling Data               11-1
12. Morris Neiburger
    Banquet Speech: Progress + Prof its + Population = Pollution         12-1

13. Arthur C. Stern
    Symposium Summary: Utilization of Air Pollution Models            13-1

14. Discussions                                                     14-1

15. Appendix: Attendance List                                       15-1
                                  VII

-------
      PROCEEDINGS
               OF
       SYMPOSIUM
               ON
  MULTIPLE-SOURCE
   URBAN DIFFUSION
           MODELS
IX

-------
AUTHOR
      JOHN T.  MIDDLETON, Acting Commissioner of the Air Pollution Control Office,
      was Professor and Director of the University of California State-wide Air Pollu-
      tion Research Center where his pioneer studies of the environment as a regulating
      factor in  disease development in agricultural crops enabled him to first recognize
      photochemical air pollution as an  adverse economic  factor to California agricul-
      ture in the mid- 1940's.
      Dr.  Middleton has  served  in  the  past as:  Chairman, California  Motor  Vehicle
      Pollution  Control Board; member. Governor's Interagency Committee on Air Pollu-
      tion,  and Executive Task Force  on Waste Management;  consultant, U. S. Public
      Health Service, Office of Science and Technology, and the World Health Organi-
      zation; and as a Director for the Air Pollution Control Association.
      Dr. Middleton has - B.S. from the University of California and a Ph.D. from the
      University of Missouri.

-------
Keynote Speech
           1.   DIFFUSION  MODELING FOR  AIR POLLUTION

                                   ABATEMENT  AND CONTROL


                      JOHN T. MIDDLETON
        Acting  Commissioner, Air Pollution  Control Office
                 Environmental Protection Agency

I am very pleased to be with you this morning, and to have the chance to talk
to  you about the urgent need for developing systematic solutions to one of the
most ironic incongruities we face in this country  today, the  problem of air
pollution.
Some months  ago,  we received from the National Aeronautics  and  Space
Administration photographs of portions of the  United  States taken  from
Apollo 8 as it streaked toward the  moon.  NASA sent us these particular
photographs because one of the more interesting phenomena they featured was
the pall of  air  pollution hanging over a number of large urban areas. Here, it
seems to me, is the irony for our time, a spaceship, carrying men to the moon
and back, capturing with its cameras our tragic failure to prevent the contami-
nation of the earth's most precious resource, the air that all of us breathe.
Of all  the threats to our environment, air pollution is potentially the most
dangerous; certainly it is the most difficult to solve. We cannot replenish the air
as  we replenish soil  and trees; and we cannot purify the air, as we do water,
before  we use it. We must breathe the air as it comes to us—and each day it
comes to us more heavily laden with the by-product wastes of our industri-
alized society.
Today, this Nation  is paying  billions of dollars each year for the dubious
privilege of living with—rather than controlling—air pollution. Air pollution
soils, corrodes, and  damages a wide variety of structures in  our cities and
towns—from the despoiling of our  homes  to the corrosion  of  suspension
bridges that span our rivers.  It threatens forests  and farmlands  as well.  It
constitutes  a national health  hazard of  greater significance than  many  may
appreciate. It contributes to the rising incidence of such important respiratory
diseases as lung cancer, asthma, bronchitis, and emphysema. Emphysema is the
fastest growing cause of death in America today.
                                 1-1

-------
The ways in which we seemingly may be changing our environment through air
pollution are  endless; and they are  international, as well as national, in their
dimensions. An  acid  tongue  of combustion products from the industrialized
area  of Western Europe is transported over the southern part of Scandinavia;
and,  particularly in the winter months when combustion activities are high, the
acidity of rainfall in  the area increases substantially.  The ability of the soil to
accommodate these  international acid rains is certainly limited,  and  the con-
sequent effects on vegetation and the ecosystem can only be guessed at.
Man  introduces 10,000,000  tons of nitrogen into the  atmosphere annually.
This  is comparable to the input of the natural  nitrogen cycle. The main source
of nitrogen contamination is  inorganic agricultural fertilizers. The nitrates in
the rivers and drinking water of some  Illinois towns  are now approaching the
upper  safe limit  of  10 parts per million  set  by the Environmental  Control
Administration.  As agriculture expands  to keep the pace with world food
needs, the problem of nitrates in water supplies will become  more serious.
Although we are not sure of the health effects of the current levels of lead in
our environment, we do know that a person's body burden of lead corresponds
roughly to the levels  in his environment  and that those levels are going up as
the use of  the principal source of lead, the automobile, continues to increase.  I
have yet to see any  data that show  any  useful purpose served by lead in the
body.
The  toxicity  of  mercury  compounds has also  been well  established. The
inhalation of only small  amounts of mercury or its derivatives can result in
exaggerated emotional response, muscular tremors, and gingivitis.  This poison-
ing has occurred in most of the major industries that use mercury. Vegetation
has been damaged when kept in  a greenhouse painted with a mercury fungi-
cide, and fish  in some Swedish lakes have absorbed and concentrated methyl
mercury  from pesticides  to such a high degree that fishing in those lakes has
been prohibited.
Estimates differ about the potential effects on world temperature and climate
due to increased atmospheric  carbon dioxide  and particulate concentrations.
Estimates also differ  on how  much air pollution  can affect precipitation and
condensation.  Questions  are  also being  raised about the  ultimate effect  of
man's conversion of  energy to heat. The opinions differ about the details of
processes involving temperature trends, climate, melting of  polar ice caps, sea
level, photosynthesis, and the distribution of fish, to name a  few.
There are, of course, many other examples. The point is, that when man alters
the balance of Nature, it  is like tossing  a  pebble  into a pond:  the resulting
ripples spread  out  concentrically from  the entry point until they  touch  every
point on the shore. Continued small alterations of our environment may have
drastic effects later, effects we  cannot foresee now.
Today  we  stand at a point in our scientific progress where our capacity  to
 1-2

-------
enhance—or to degrade—the quality  of our environment is almost  literally
beyond reckoning. From the standpoint of our sheer survival as a nation and a
people, it is essential that the talent and  imagination of all of us—and especially
of the scientific community—be engaged in enhancing it.
If it is obvious that one way to halt the contamination of the air we breathe is
to prohibit automobiles,  stop  the generation  of  electricity, and  shut  down
industry—it is just as obvious that this way is impossible. Air pollution is woven
throughout the fabric of our daily lives. It is a by-product of the way we build
our cities. It is a waste, left over from the ways we transport ourselves and our
goods,  and  from the ways we generate the energy to heat and light the places
where we live  and work. It is part and parcel of the ways in which we produce
and package a multitude of manufactured goods, and, among other things, the
ways in which  we use those goods and dispose of the remains.
The  problem of air  pollution, then, affects  not  one segment of society, but  all
segments.  To  deal  adequately with  such  a  problem  has implications for
thousands  of  daily decisions that are  made  in every  activity of our  busy
modern life.
Air pollution is a direct threat to human health and welfare that  no responsible
person can put aside; but it is also a social, political, and economic problem.
We will  not truly come to grips with air pollution unless we derive our remedies
from a careful  consideration of what is truly in the public interest.
Here, then, is  perhaps the most challenging frontier of our space-age society.
Air pollution,  like the other problems  of our environment, is  no longer the
private preserve of the professional engineer, either in government or out. It is
rather a problem that requires attention from all facets of our society if we and
our institutions are to survive its solution. Governments at all levels must rise
above their jealous  disputes as to what  jurisdictional rights are  being violated
by whom,  and work together side by side as regional  partners in a public
enterprise that no single government possibly can carry off alone. The manu-
facturer will have to look beyond his assembly  line and take responsibility for
the performance of his  product  in the hands of  the sometimes  neglectful
consumer.  The industrialist is going to  have to take responsibility  for his use
and misuse of  those resources which sustain not only the industrial revolution
in this country,  but also  our very lives. The scientist, hardly last and hardly
least, must provide the technical solutions to those by-product  problems that
have unfortunately accompanied the modern miracles his science has  created.
In the  Clean Air Act, as amended, the Congress has attempted to give NAPCA*
the legislative  apparatus that we need to remove the malignancy of air pollu-
tion. This Act requires that we carefully consider all the ramifications of every
step  we take-and every step we do not take-in the control process. The Act
charges us to protect the public in those places where air pollution has reached
acute proportions and to prevent the  problem from occurring in those places
*Now APCO.

                                                                     1-3

-------
 where fresh  air  is still enjoyed. On the  other  hand, the Act reflects  an
 awareness that air pollution is the result of many of the activities that sustain
 our  modern  way of life.  It  says that we must  optimize use of currently
 available technology to control and abate air pollution without disrupting these
 activities.
 The  Federal government has the responsibility for setting the machinery of the
 Clean Air Act in motion. Under the Act, the Department of  Health, Education,
 and Welfare is required  to draw the boundaries around each urban area of the
 country where we believe air pollution is a problem. So far we have drawn the
 boundaries around 20 major urban areas, and by the end of next summer we
 expect that we will have drawn the boundaries for a  total of 57 air  quality
 control  regions,  involving all  the 50 states, the District of Columbia,  Puerto
 Rico, and the Virgin Islands.
 Last  February we issued to the states criteria and control technique reports on
two  of  the major families of  air pollutants, the sulfur oxides and gross
particulate  matter. The states are now on notice  to use this  information in
developing  air quality  standards  for  the  regions we have  designated  and  to
design plans for implementing the standards.
 It is at this later stage of the  Clean Air Act procedure,  the regional implemen-
tation plan, that the technique of diffusion modeling plays such an important
role.  I would  like in the time remaining to me this morning to discuss that role
in more detail.
 In  1958,  the year of  the First  National  Air Pollution  Conference, there
appeared to be some element of doubt at  least  in the minds of some people as
to  whether there was  indeed  an air pollution  problem. . .  At that time,
 NAPCA's meteorologists were already  engaged  in developing  an empirical
diffusion equation. Weather Bureau wind summaries and an IBM 650 computer
were used to  furnish  mean  monthly sulfur dioxide concentration distributions
 for an entire urban area. The data from  Nashville, Tennessee was used  to test
 the model. Although the absolute magnitude of the concentrations could not
 be verified, patterns  that were predicted and observed were generally  similar
 for the 5-month test period.  This was the  first step  by  NAPCA scientists
 indeed the  first by any group following Frenkiel's pioneering effort in  1955, to
 delineate mathematically transport and dispersion processes on an urban scale.
 We have come a long way since  1958, and  NAPCA's Division of Meteorology
 has not been  alone in  its modeling efforts. In recent years its  work has been
 complemented by modeling work conducted by meteorologists at New York
 University, the  University  of California at  Los  Angeles,  the  University of
 Florida, and  ESSA's Air Resources Laboratory. Studies have  also been con-
 ducted  by  Travelers Research  Center and the Stanford Research Institute
 Work also goes on abroad.  There has been diffusion  modeling by scientists in
Germany, Hungary, Japan, England, and Canada. Some of their development
                                                                       :s
 1-4

-------
will, I understand, be presented  at this symposium, and that is good. The need
for abatement of air pollution is a strong international tie.
As we proceed with the implementation  of the Clean  Air Act,  the meteo-
rologist will be called upon more often to provide a mathematical description
of the transport of air pollution  in the research, survey, and operational phases
of control activities.  In essence,  the meteorologist is charged, because of time
limitations and paucity of air quality measurements, with the chore of pre-
senting convincing evidence  of the  magnitude and  frequency of pollution
transport and dilution. The diffusion models we have been discussing are the
yardsticks for furnishing quantitative determinations of ambient  air  concen-
trations of pollutants. These determinations must be credible enough to satisfy
the members of other scientific and technical disciplines, as well as represen-
tatives  of legal, economic, and political interests. These models should also be
flexible enough to assimilate feedback from source  emissions for obtaining
solutions to both chronic and  acute pollution problems.
Future diffusion  models that incorporate air pollution climatology are ex-
pected to  be used  more  frequently  in long-range  air-resource-management
programs for  the  evaluation  of  future air quality and source locations with
regard  to  projected emissions rates and population distributions.  In order to
interpret  the relationships between  proposed  or  existing  source  emission
standards and ambient air quality standards, we need a quantitative account of
the meteorological processes that cause pollutants to be diluted, chemically
changed,  or  dispersed between  source and  receptor. Our hope is that  by
applying  diffusion models to  calculate the  integrated  effects  of  multiple
sources of  the same  pollutant, we will be able  to interpret the relationships
between  proposed or existing  source  emission standards and  ambient  air
quality standards.
The advantages of this computer age technology have already been recognized
by  responsible segments of commerce and  industry,  and more recently  by
community planning agencies. In the words  of  Jimmy  Durante,  "everybody
wants to get into the  act."
Another application of diffusion models will  come with the evaluation of the
effectiveness of individual control or abatement  procedures for a  given source
or complex of sources. By use of appropriately  programmed models,  such an
assessment may be made without the need for an intricate and dense network
of air quality measurements  taken at the right  place in  the right time at the
right season. Using diffusion models, the meteorologist can greatly reduce the
time and  cost of such evaluations by answering the questions of what actually
constitutes "the right place, length of time and season." Moreover,  under
operating situations one may expect that the short-term-period diffusion model
will be employed  to determine  the amount  of  emission decrease required to
meet air  quality standards. Such models will also be helpful in the strategic
                                                                      1-5

-------
deployment of mobile air quality and meteorological monitoring units during
air pollution alert situations.
So far as a warning scheme is concerned, the heart of any effective air pollution
control  strategy will be meteorological forecasts that can provide the necessary
lead times and  guidance required  to deploy and  implement abatement and
control  devices.  The sine  qua non of  such  forecasts  is the capability  of
meteorological  techniques  to forecast  meteorological  conditions within  the
urban boundary layer on urban scales appropriate to the air pollution problem.
There are some meteorologists, and  I must agree with them, who  suggest that
the utility of present diffusion models  is limited  by the inability to make
accurate forecasts of wind  and temperature profiles.  I think it is abundantly
clear that the accurate description within the planetary boundary layer of the
profiles of wind, temperature, and  moisture and their variability  in time and
space as a  function  of  surface conditions and of synoptic-scale features  is a
fundamental requirement for improved  meteorological services not only to air
pollution problems, but also to many other interests. We have an urgent need,
it seems to me, for greatly  improved mesoscale forecasting from the national
weather services.
I do not ask for a doctoral  dissertation on a weather forecast; all  I ask is that
we  be supplied with the facts that  we need to do our job, the facts that you
gentlemen  need to make your diffusion models more efficient. I think we agree
that the quality of your results depends upon the information you start with.
To  alleviate to some degree the inadequacies  in mesoscale  observations and
forecasts in the United States, we  are  supporting the Environmental  Science
Services Administration  program of  low-level soundings in five urban areas. The
program provides  for a full-time  meteorologist to concentrate on mesome-
teorological-scale forecasts in these five metropolitan areas, particularly as they
relate to air pollution.  Hopefully,  ESSA will be able to assume responsibility
for  such services in  these five cities and to extend them to an  additional 15
cities  in the next fiscal year.
The  programs that we  are  supporting in the National  Air  Pollution Control
Administration  to improve  our knowledge in  the  atmospheric sciences area
may seem to some of you at first glance to be a far cry from the usual business
of a Federal regulatory  agency. A closer look, I think, will reveal  (1) that these
and other research programs in NAPCA  are a vital link in the control  process
and fully congruous  with the fundamental thesis of the Clean Air Act  and (2)
that our efforts  to  control air pollution should  be derived from  a  careful
evaluation  of the latest scientific evidence of the effects of air pollution on
health and welfare. This systematic approach to control, this linking  of the
control  effort  to scientific  evidence of the need for control, seems to me to
lend a vitality to the research  area that it might never otherwise acquire. I am
hopeful that the research programs we have found it necessary  to support in
1-6

-------
NAPCA, in  order  to give us that information so vital to the planning of our
control  program directions, will  stimulate other agencies and institutions to
begin to explore this  as yet largely untouched area of the meteorological
sciences.
In this regard, diffusion models such as have been developed up till now,  and
the generations of  models yet to come as we refine, expand, and develop our
techniques,  can  play  a vital role in our attempt to systematically apply air
pollution control theory on a  regional basis. For this reason, I wish all of you
Godspeed in your deliberations at this symposium.
                                                                     1-7

-------
ABSTRACT
      Diffusion processes in the volume of air above a geographical area can
      be described in spatial detail with two- and three-dimensional vector
      models drawn  from boundary  layer  and fluid dynamics concepts. A
      simpler scalar model,  in particular, can directly account for temporal
      trends  in  average pollutant  levels due to simultaneous diffusion of
      mass, energy, and momentum from multiple sources.
      By an application of climatonomy, the qualitative box model of scalar
      diffusion is  expanded quantitatively with fluid dynamic equations.  The
      theoretical model has been  applied  to prediction  of both long-and
      short-range,  thermal and paniculate pollutant trends in large and small
      cities.  The use of the box model  to  evaluate aerodynamic momentum
      drain over the built-up area of a city is also  described.  City types are
      specified in terms  of their  aerodynamic  roughness parameter  with a
      formula based on results of micrometeorological experiments. Knowing
      only the  horizontal  pressure gradient in  the  lower  troposphere the
      model makes  it possible to estimate the characteristic wind speed for a
      specific city type.
      In comparison with Davenport's power law,  a Universal Wind Spiral
      Theory  described here has  the  advantages  of  yielding quantitative
      estimation  of directional  shear and  surface  roughness  effect on the
      wind profile.  Examples of scalar  and  vector application of the latter
      law to thermal pollution are discussed.
AUTHOR
     HEINZ H. LETTAU is Professor of Meteorology and Civil  Engineering at the
     University of Wisconsin, where his current research  is directed toward mode/ing
     earth-air interaction processes. His model of the hydro/ogical cycle for restoring
     the level  in water sheds  was one of  the earliest attempts at environmental
     mode/ing.  He has been a research scientist at the Geophysics Research Director-
     ate U.S. Air Force, Cambridge, Massachusetts; Fellow of the American Meteoro-
     logical Society; corresponding member  of the  Bavarian  Academy of Science-
     member of the NAPCA Research  Grants Advisory Committee, and is main editor
     of the two-volume Exploring the Atmosphere's First  Mile, 1957.

     Dr. Lettau received his Ph.D. in geophysics from the University of Leipzig in  1931.

-------
              2.   PHYSICAL AND  METEOROLOGICAL  BASIS
                                FOR MATHEMATICAL  MODELS
                          OF  URBAN DIFFUSION  PROCESSES


                        HEINZ H,  LETTAU
                University of Wisconsin at Madison
INTRODUCTION
I  shall discuss some  general  aspects of processes taking place  in the atmo-
spheric boundary layer when wind passes over urban terrain. Emphasis will
be on  simultaneous diffusion of  mass, energy, and momentum with special
consideration of aerodynamic resistance due to  roughness of urban surfaces
and resulting feedback on the diffusion potential of the urban atmosphere.

I  shall start with an illustration  familiar to you.  Figure 2-1 shows what I
would  like to call the "vector model" of diffusion from multiple sources.
This illustration comes  from a report by Moses1  demonstrating the conven-
tional approach that  assigns one grid to the sources and another grid to the
receptors, with a composite of "plumes" in  between. The merits of this basic
model  for the study  of the details of complex urban diffusion  processes are
obvious, and I  trust  that its implications and desirable  refinements will be
competently discussed by other participants at this Symposium.

Let me contrast  the vector model with a  "scalar model" (Figure  2-2) in
which  the volume  of "city-air" is  somewhat  crudely  defined.  Pollutant
emission is attributed to a quasi-uniform area source at the lower boundary.
The major flushing agency  is horizontal air motion, additionally supported
by vertical eddy flux  at the level h* that tops the volume. If no other city is
immediately upstream, the wind enters the volume relatively clean but leaves
it  loaded  with  emission products.  The  level  h*  may coincide  with  an
inversion  of temperature, in which case vertical exchange through this level
will be of  minor importance. Such scalar or "box-models" of urban diffusion
have been mentioned occasionally in the literature. They cannot be used for
the study  of spatial  detail,  but produce interesting, specific results,  if em-
phasis  is placed on time-series (either short- or long-range trends). Moreover,
the box-models permit discussion of similarities between  air diffusion and
other  natural processes such as the loading of a  watershed by precipitation,
discharge into  river channels, and  losses by evapotranspiration.  It is not my
                                2-1

-------
       SCHEMATIC
      SOURCE GRID
     POLLUTANT
     ISOPLETHS
                                 POLLUTANT ISOPLETHS

                                      JI.6 ppm
                                                      SCHEMATIC
                                                    RECEPTOR GRID
Sources are located at A, B, C and D of source inventory grid.  A section of the
monitoring grid is given by squares 15, 16 and 17.  The concentration enclosed
by the 0.6 ppm isopleth of source B and 0.3 ppm isopleth of source D contribute
a concentration exceeding 0.9 ppm for square 15.  See hatched area.  Diagram as
published by Moses.'
Figure 2-1.
diffusion.
              Illustration of conventional or vector model of urban
2-2

-------
                                                     FLUSHING BY VERTICAL EDDY FLUXES
                                                                                             FLUSHING
                                                                                               BY
                                                                                              HORI-
                                                                                             ZONTAL
                                                                                              WINDS
NJ
00
                                                     -CITY DIAMETER, D-
                          Figure 2-2.  Schematic illustration of box or scalar model of urban diffusion.

-------
intention,  however, to discuss the descriptive concept of an "air shed.    I
prefer instead  to consider a basic similarity between air pollution dispersal
and the  hydrologic  cycle  that permits me to utilize a theoretical solution of
the water balance equation recently described by  Lettau.

Parameter Considerations
Before moving to that  point, let me insert a general statement: The entity of
transport and  flux  processes must be taken  into account in fluid dynamics
problems. Thus, in atmospheric boundary layer studies we should not restrict
our consideration to transport  of mass  (either intentionally or inadvertently
loaded on the moving fluid) but should investigate concurrent generation and
dissipation  of  both momentum and energy. The scalar  model  (Figure 2-2)
has the advantage over the more specialized vector  model of Figure 2-1, that
it  can,  without alteration, express  a  city  area  as  a  heat source (essentially
due  to  the combustion of  fuels), as  well  as a  momentum sink or  negative
momentum source (due to the efficiency of buildings as aerodynamic rough-
ness elements).  A coherent  consideration of  the  transport  and  flux  of  mass,
momentum and  energy in any  atmospheric boundary  layer is  also a logical
starting point for the discussion of mechanical and thermodynamic feedback
since the horizontal dimensions  involved in  urban  boundary layer structure
are comparable to  meso-meteorological  and  synoptic-scale  atmospheric pro-
cesses.

Aerosol Effects on Sunlight Absorption—Example
Let  me  begin  discussion  of  the combined  aspects of long-term trends and
feedback on meteorological phenomenon, by  considering the attenuation of
sunlight by pollutants in the volume of city air.

McCormick and  Ludwig4  have  recently  discussed  climatic  modification by
atmospheric aerosol. Let me illustrate  my approach with Figures 2-3 through
2-5,* taken from a recent  publication  by  K. and H. Lettau.5  A common
*On some graphs the word "climatonomy" will appear. This word is also found in the
 title of the references by Lettau3 and K. and H. Lettau.5 It is a new term that we use
 at the University of Wisconsin to  indicate a systems approach to problems of climate.
 System in  engineering  problems is conventionally used to encompass the following
 components:  (1) a  process, (2) an input to the process, or a  forcing function, (3) an
 output from the process, or a response function, (4)  feedback,  and  (5) control of
 inputs, or management  of the system.  Typical  of  climatonomy  is the use of a mathe-
 matical-physical model  of a specific atmospheric  process, such as the response of a
 climatic element to a defined forcing  function, which may be an insolation cycle or
 another defined input  like pollutant  emission. The climatonomic approach  considers
 physically significant parameterization  of the problem so that, even though control or
 management  may not be possible, one could  reasonably and  realistically  answer such
 questions as what response can be predicted if, for reasons beyond control, the inten-
 sity of the forcing function, and/or  the value of  a response-controlling parameter,
 would be halfed, doubled, or increased by any specified multiplication factor. A char-
 acteristic result of a climatonomic  case study is documented on Figures 2-4 and 2-5.
 2-4

-------
scheme appears on all three  graphs. It  is intended  to  bring out the contrast
in attenuation  of insolation  by  typical aerosol  between a big city [repre-
sented by  Kew (London),  England] and that  of  an extremely  dry desert
[represented  by the Pampa de la Joya, Peru]. We selected these data  from
the published literature  because  the measurements include not  only direct
sunlight or "beam  radiation" but  also scattered or diffuse light. Note that
external conditions (sun's elevation angle, optical  air  mass, etc.)  happen to
be such that  the  attenuation of beam radiation was nearly the same at both
stations.

In Figures 2-3  to  2-5 the upper  horizontal line  represents the atmosphere-
space  border, across which  solar  rays enter the  atmosphere and  reflected
light  ("earth-shine")  returns  to space.  The lower  horizontal  line represents
the earth—air boundary, at which the attenuated solar beam  and  the diffuse
light scattered down  from  the sky arrive. The dashed arrows show  the sun-
light that  is reflected  by the ground and attenuated in  the air  before entering
space;  part is scattered back down and part is absorbed. Numbers along, and
at the  tips of the  arrows show fractions of scattering and absorption as
decimal fractions of the unit  incident-extra-atmospheric irradiation. These are
always separated  into downward  and upward components.  Numbers above
and below the  braces indicate combined radiation intensities as decimal frac-
tions of the unit solar radiation budget.

Figure 2-3 shows  that beam  attenuation was nearly  identical in  both  cases
with reductions from  100% to 61.8% in Kew, and  to 61.9% in La Joya. The
city air absorbs 22.5 percent and scatters 15.7  percent  (5.2% outwards and
10.5% downwards),  or about 2/3  of  the absorption.  The desert air on the
other hand absorbs 14.3 percent and scatters 7.9% +  15.9% = 23.8%, about
5/3 of the absorption.  Figures 2-4 and 2-5 suggest that the assumed aerosol
increase, first by  an  arbitrary  factor  of two  and then  of five,  leads to
extinction  of direct  sunlight in  both  locations. This means that  the sun,
although visible in the sky, will  not throw a shadow.  Obviously, the "gloom
at noon"  under the opaque smog of the city where only 26.9 percent of the
incoming  energy reaches the ground will be more frightening  than under the
strongly-scattering, mineral-particle-containing desert dust where 55.6 percent
of the energy reaches the ground. If an aerosol increase by a factor of five is
possible, what  can  be done to control the increase once the factors are
known? Let  us discuss it first with the  aid of the box  (scalar) model  of com-
munity air pollution.
                                                                      2-5

-------
1. KEW, ENGLAND, MAY 1948, AFTER ROBINSON (1963)        2. LA JOYA, PERU, JULY 1964, AFTER STEARNS (1966)

                    0.129                     1.000                0.175
1.000
             0.052
                    0.225
           0.105
                  0.005
                    l
                            0.077
                                                        0.079
                            0.027
                                                                0.143
                                   0.252
                                                       0.159
                                                                ,	,
                                                              0.030
                                                                       0.096
                                                                            • 0.162
                                                                        0.019
      0.618
                0.110     -0-109
                                                 0.619
                                                            0.189
                                                                     -0.145
           0.728
                                                       0.808
                0.619
                                                           0.663
 Figure 2-3.  Shortwave radiation budget —air over city versus air
 over desert.
1. KEW; AEROSOL ABSORPT. TO SCATT. = 1.66

1.000                 0.122
2. LA JOYA; AEROSOL ABSORPT. TO SCATT. =0.20
 1.000                 0.194
              0.073
                 J   0.328
            0.146
                   0.013
                    I
                   JL
                            0.049
                                                         0.124
                                  0.358
                                                                 0.170
                            0.030
                                                       0.249
                    /
                   0.042
                                                                       0.070
                                                                              0.193
                                                                        0.023
       0.453
                 0.159      -0-092
                                                 0.457
                                                             .291
                                                                    -0-135
            0.612
                                                       0.748
                0.520
                                                            0.613
 Figure 2-4.  Shortwave radiation  budgets for city versus desert
 after aerosol increase by factor of 2.  Other conditions same as in
 Figure 2-3.  (Note:  no change  in  ratio of scattering to absorption
 efficiency)
 2-6

-------
1. KEW; SMOG DENSE ENOUGH TO OBSCURE SUN

1.000               0.147
                      2. LA JOYA; DUST DENSE ENOUGH TO OBSCURE SUN

                       1.000               °-269
            0.133
                   0.600
           0.267
   OPAQUE
    SMOG
0.002
 i
 l
 I
Y
                         0.014
       "0.024
                                0.624
                                                    0.250
                                                          0.250
                                                  0.500
OPAQUE
 DUST
                                                          l	^
                                                         0.056
                                                     i  i
                                                                 0.019
                                                                 ->• { 0.275
                                                                ' 0.025
      0.000
               0.269
                      -0.040
                                            0.000
          0.269
                                                  0.556
               0.229
                                                      0.456
Figure 2-5.  Shortwave radiation  budget- city versus desert-after
aerosol increase by factor of 5.
BOX MODEL OF URBAN AIR DIFFUSIONS

Qualitative answers  can be  readily provided  by a schematic  "budget box"
such as  the  one  illustrated  in  Figure 2-6.  It is assumed  that  the source
strength  of pollutant release per unit area of the lower boundary  is the same
for two  cities of different size;  and two different flushing rates  are applied
to both  cities.  These rates could correspond to two  weather situations,  one
with relatively  strong  versus  one with relatively weak winds. The particle
loading  of a comparable unit volume of air  is illustrated in Figure 2-6 by the
number  and  arrangements  of  "dots" carried by  the outgoing air  current,
which in  each  case  equals the number of "dots"  released by the  area source
into the volume of city air. Obviously, a small city with a weak flushing  rate
would  be no  worse off than a  big  city  with a good flushing rate.  For
example,  an  innocent activity  like  the burning  of  leaves could  be  well
tolerated  in a small city, but could generate intolerable smog conditions once
the city became large.
Quantitative answers require consideration of the basic fluid dynamics equa-
tions. Let

     s   = specific admixture to air, (g meter"3)
     V=iu+jv  +  kw  (meter sec"1)
           i u
          ->•

     S  =  strength of internal source, or .sink if negative,
           of considered admixture  (sec"1 meter"3)
     v  =  molecular diffusion coefficient (meter2 sec"1)
                                                                       2-7

-------

           SMALL
                                            LARGE
                            CITY SIZE
 Figure 2-6.  Schematic  illustration  of  effects of city  diameter
 and atmospheric flushing rate on pollutant concentration in box
 model of  urban diffusion;   assumes uniform area source of pol-
 lutant  and  removal  from   city volume by  horizontal air motion
 only.   Dots  (.) indicate output of pollutant sources to city air
 volume.   Arrows  indicate  ingoing   (shaded)  and  outgoing
 (white) volumes of air per unit time.  Forcing function, Q.
                               S  = S
                     (1)
The principle of conservation yields

         —  +  V-V
          5t       H>S
Eddy fluctuations are removed from the instantaneous values by time averaging
and by introducing  s — s'  = s,  S -  S'  =  S,  and  V -  V1 =  V,  whereupon
the primitive Equation (1) transforms into
          6t
             4  v. V s  +  V .V's1   -
=  S
                                                                  (2)
It appears legitimate to  neglect  molecular diffusion, the eddy fluxes in the
horizontal direction, and the transport by mean vertical motion, in compari-
son with vertical eddy flux  and  the transport divergence by mean horizontal
wind components. This amounts  to saying that
[(w s)2  +  (ITs7),,
                                                        +(wTsMz]
                                                                  (3)
where subscripts denote partial derivatives with respect to the three indepen-
dent spatial variables (x,y,z).
2-8

-------
From Equations  (2) and  (3), the simplified version of the primitive equation
becomes

         •||+(us)x  +  (vs)y  =  S-  (w^z                      (4)

It  is  characteristic  for the  box  model  of  urban  diffusion  that  we  are
primarily interested  in representative  area averages over city  diameter  D  (m)
and  thickness  h* (m)   of the  boundary  layer  (Figure  2-2).  Let  such an
average  of any function, 0  or representative function value, 
-------
Horizontal  advection  will significantly control the flushing frequency, espe-
cially when the level  h* coincides  with  a  "lid" such as an inversion  layer.
The  transformation of Equation (4) with the aid  of  Equations (5)  through
(6d), yields the equation governing the box  model of urban diffusision,
Which  is  the equivalent of the  governing  equation of  evapotranspiration
climatonomy, see Lettau.3  For further transformation let us introduce the
reduced source  strength or quasi-equilibrium value of pollutant concentra-
tion, q* (nrf3):

          *  _   Q                                                   (8a)
        q   -  h*f*

and a dimensionless time (f)  which serves as the new independent variable,

        t'   =  /  f*  dt ; or,                                         (8b)
        dt1  =  f* dt

In  its final  form  the  characteristic  equation of  the box  model,  or the
governing  equation  of air  pollution  climatonomy,  may  be  written  as  an
ordinary differential equation,
 which is solved by
                                  fl
                      [q0  +  /  er  q*  dt1]                         (9b)
                              0
where q0 is the initial value of concentration at  time i =  0.  Representative
values of a parameter having the physical units of seconds per meter were
estimated for a  variety of  cities  by Holzworth,7 and  I  suggest that for
practical applications, Holzworth's  theoretical  Native pollutant concentra-
tion can be identified with 1/h*f* in Equation (8a).


Prediction of Pollutant Trends by Box Model
It is evident that  Equation  (9a) and its solution, Equation (9b), correspond
to an  initial-value problem only if q* is a quasi-constant. In reality, even for
time-independent  source  strength, Q,  the value  q*  will vary considerably
because, according to  Equation  (8a), it  is  controlled by the  atmospheric
flushing frequency, which changes with the weather. An example of practical
numerical application of the theory  is illustrated in Figure 2-7.
2-10

-------
                      FORCING FUNCTION • POLLUTANT RELEASE
                                                     12
                                                    HOURS
                                                    CONTROL FUNCTION-HORIZONTAL ADVECTION

                                                       AIR FLUSHING RATE t*~U*D, hourl
                      RESPONSE FUNCTION - POLLUTANT DENSITY
                                                    DAYS
Figure 2-7.  Calculated trends of pollutant concentration in response to  indicated forcing function for
day-to-day variation of atmospheric  flushing parameters.  Calculation by  box-model for three different
city diameters to demonstrate adverse  effects of urban sprawl.

-------
Over a period of 5 days, the area source is assumed to have the same daily
average rate on which is superimposed a semi-diurnal variation with peaks at
the hours of 8 am and 6 pm. While the first and fifth day have normal wind
speed conditions, the second day is  stormy and the third and fourth days,
relatively calm. The bulk value  U* according  to  Equation (6c) is depicted in
the middle part of Figure 2-7, showing slight diurnal variation with one peak
per day induced by the  normal insolation cycle. To bring out the effect of
f*-dependency on  the city diameter-see  Equation (6d)-three D-values were
employed,  10 mi for a moderately large  city, 20 mi for  one of  intermediate
size, and 40 mi for a metropolis.
For  practical  numerical   evaluation  of the model, a fixed time-interval At
equal to 2 hr, was  employed for the calculation shown  on Figure 2-7.  For
any period beginning at  time tj and ending  at time  tj + A t, representative
values were assumed for source strength, Qj, (the forcing function)  and for
atmospheric  parameters.  Representative  values  qj*  and   fj* were thus  ob-
tained  with the aid  of  Equations (6c) and 6d). Knowing qj at time tj, the
concentration, A t later  is q, +  1 and  was calculated  numerically using  the
following version of Equation (9b), with f,* At = At|:

   q.+|  =  e-At'i  [qt  +  q*(eAt!,-D  ]  = qt  +   (qs - q?)  e'^i     (10)

The numerical calculation may begin  at time  fy  with  an arbitrary  initial
concentration value  q0.  The  only prerequisite is that the chosen q0 has the
same order of magnitude as  a typical q* value for the city under considera-
tion.  For example,  all three q-series  shown in the  lower  part of Figure 2-7
begin with q0  = 1  (in  relative  units) and are continued  for a total of  120
hours, which means 60  steps in At. We see that within a few  time steps the
three q-functions align themselves to the individual forcing functions quali-
fied by the respective flushing frequencies. A characteristic tendency toward
lagging and  smoothing-out details of  the  forcing function is evident as a city
gets larger. The three curves on the  lower part of  Figure 2-7 are self-explan-
atory. The important feature is that, other factors being the same, the larger
city must suffer considerably more  than the smaller city during periods of
nearly stagnant  air.  This follows  as  the  direct  result of  reduction  of  the
flushing rate in  proportion to city diameter.  The buildup of pollutant levels
during a calm weather spell  in a large city is more severe, and  last for longer
periods than in  a smaller city. This quantitative result of numerical calcula-
tion supports the qualitative statements made previously  in connection with
the discussion of Figure  2-6.

While the box model excludes the possibility of investigating spatial detail of
pollutant distribution, the emphasis on temporal variations is applicable to a
number of  important problems. Before we begin to talk about  long-range
trends,  let  us  discuss another  application of  the box  model  to relatively
 2-12

-------
                                             FORCING FUNCTION • POLLUTANT RELEASE
                   0     12     0      12

                  Ml PARAMETER - AIR FLUSHING RATE
                                                    DAYS

                                          RESPONSE FUNCTION - POLLUTANT DENSITY
       14  20  2  8  14 20  2   8   14  20  2   8  14  20
                  10
                    2   8  14  20   2
                                                     10
                   THU
FRI
SAT
SUN
MON
                                                   DAYS
Figure 2-8.  Trends of pollutant concentration in response to week-end  lull of forcing function, calcu-
lated for constant diurnal cycle of atmospheric flushing parameter.  Response function illustrated for
city diameters  of  10 miles  and 30 miles, but normalized  (to equal daily mean value) to  bring  out
smoothing effect of larger city.

-------
short-range fluctuations.  Specifically, we refer  to  a time-variation on  the
forcing  function  caused  by  the  weekday versus weekend  cycle. While on
Figure 2-7  the forcing function remained the same  from day-to-day, Figure
2-8 schematically  depicts a  weekend  decrease  of  emission. Two cities of
different D-values were assumed, and  the response functions normalized to
bring  out the damping and lagging predicted for the larger metropolitan area.
Direct measurements documenting  diurnal  peaks  and  weekend lulls  that
compare with the theoretical results shown on Figures 2-7 and 2-8 have been
reported for a  number of cities.
Examples of observed weekly cycles in a  variety  of pollutants  including
carbon monoxide,  sulfur  dioxide,  smoke,  and  oxidants  are  discussed by
McCormick  and Xintaras;6  W. Johnson;7 Schuck,   Pitts, and Wan;8  Harris,
Huffman, and Weiland;9 and  others.

Obviously,  any  application  of  the results  of  air  pollution  climatonomy,
represented  by  Equations  (9b)  or  (10),  requires  information  not only on
the input normally  available from  source inventory of urban  emitters,  but
also  on  the atmospheric parameters defined in Equations (6c) and (6d).
Applicability will  be limited if, for instance, the  parameter h*, the  mixing
depth, is unsatisfactorily  known. In principle h*  for downwind change in
surface  roughness and energy supply could be determined from  the dynamic
and thermal aspects of boundary layer  development given the state of the
lower atmosphere at the upwind edges. At the present time, however, it
appears  more   practical  to   rely  on  statistical-climatological  evaluation of
meteorological measurements in the urban area.  Reference is again made to
Holzworth.10

Application of Box Mode! to Thermal  Pollution
Equations (7)  and  (9a) permit an  evidently steady  state if q = q.* This is
used to predict the order of magnitude of the effect of thermal pollution in
the air of a metropolis. New York  City  is a self-explanatory example of an
urban complex for which heat input  and pertinent atmospheric parameter
values have been reported  (Table 2-1).

It follows that the reported  heating rate of 39 calories per second per square
meter, or about 0.2 langley  per minute, will raise the average temperature of
the volume  of  city  air  approximately 2° C above its surroundings. Such  a
temperature-volume average  suggests that the excess may reach 4° to 5°  C at
the center of the city and roof-top or stack levels. According to Bornstein,1'
this value is not uncommonly encountered in temperate zone cities such as
New York.

It is evident that fuel combustion represents only  one among several causes
of the urban heat island. Atmospheric  feedback, from city structures and air
2-14

-------
      Table 2-1   EXAMPLE OF AIR POLLUTION CLIMATONOMY -
         LEVEL OF THERMAL POLLUTION IN NEW YORK CITY,
        CALCULATED FROM INPUT AND PARAMETER VALUES
Element
Area source — fuel
combustion in winter
Parameter value
Predicted excess of
air temperature
(quasi-equilibrium)

Symbols
Q,
Q/CPP
1/h*f*
Q/cpp h*f*



Numerical values
39 cal m~2 sec"1
0.1 3° Cm sec"1
1 6 sec m"1
2.1° C



References
Bornstein11
Holzworth10
Holzworth10
Climatonomy
Equation (7)
for 9p_ o
9t
pollution, to dynamics of the lower atmosphere also involves both insolation
and terrestial or long-wave radiation, and the soil moisture budget. Petersen12
gives more detailed documentation.

To the best of my knowledge there have been no previous attempts to  use
the box  model  of urban diffusion for the calculation of thermal pollution
levels. The application documented in Table 2-1  is, however, a logical step
and it  conforms with the philosophy outlined in  my introductory  remarks
regarding the desirability of a simultaneous treatment  of transfer of masses,
momentum, and energy in boundary layer problems.
AERODYNAMIC ROUGHNESS AND
MOMENTUM DRAIN IN  URBAN AREAS

Having shown examples of the application of the box-model to diffusion of
mass and  thermal energy, we can now apply  it to the drain of momentum
encountered by the wind  in its passage over a city.

First, it  will  be necessary  to determine  the representative  aerodynamic
roughness  length, ZQ  (cm), of the urban area. I propose to use  a formula
that has been  derived from  a series of out-of-doors micrometeorological
experiments done over a period of years at the University of Wisconsin.
In late winter, hundreds of commercial bushel  baskets were distributed over
a section of level, almost undisturbed  ice on Lake Mendota. Changes in the
vertical wind profile as a function of the roughness modification upwind of
an  anemometer  mast  were  analysed. The z0  value of the undisturbed ice
(about 0.01  cm) was increased systematically and in a controlled fashfetn to
nearly 10  cm. It is established13  that the z0  value can be predicted  ks\ng
similarity concepts  if we  know the effective height H (cm), of the  rough
elements; their silhouette area seen by the wind, a (m2) and their lot are.
                                                                 2-15

-------
A   (m2).  Lot area refers to the total area of the test field divided  by the
number of elements.
        zn = H  a / 2A
                                                                     (11)
The  factor  1/2  approximates the average drag coefficient of the roughness
elements.  A set of  urban background data  for  use in  Equation (11) is
summarized  in  Table  2-2.  The  values  are  arbitrary  within an  order  of
magnitude. More specific information will be required for actual case studies
of specified  urban  boundary layers.  The  same applies to  the data in Table
2-3,  where the estimated z0 values for three building types were considered
in a  pilot study intended to show how the  characteristic wind speed  U can
be determined for a specific city type, knowing only the overall atmospheric
dynamics represented by the intensity of  the  horizontal pressure gradient in
the lower troposphere.  U can be  predicted and, in turn, used to obtain the
velocity, U*, in  Equation (6c).
    Table 2-2  AERODYNAMIC ROUGHNESS LENGTH CALCULATED
               FROM EQUATION (11) FOR ESTIMATES OF
          BUILDING CHARACTERISTICS IN AN URBAN AREA
Building type (schematic)
Building height (approximate)
Assumed lot area
Assumed silhouette area
Calculated ZQ,
Low
4 m
2000 m2
50m2
5 cm
Medium
20m
8000 m2
560m2
70cm
High-rise
100m
20000 m2
4000 m2
1000cm
           Table 2-3.   APPLICATION OF LOGARITHMIC LAW
                    TO WINDSPEED CALCULATION3
zo
5
70
1000
5
70
1000
G (m/sec)
5
5
5
20
20
20
log10(G/z0f)
6.00
4.85
3.70
6.60
5.46
4.30
C
0.0378
0.0475
0.0590
0.0342
0.0420
0.0538
U (m/sec)
3.3
2.9
1.9
12.8
10.5
6.8
 aWindspeed U calculated from Equation (12)  at level of 150 meters for given  surface
  roughness (z0), given geostrophic speed (G) and subsequent geostrophic drag coefficient
  (C); derived from universal relationship between C  and scaling factor for atmospheric
  boundary layers,  Ro0 = G/z0f
    ROQ = surface Rossby number
      f   Coriolis parameter  =  1.0 x 1CT4  sec"1 for these calculations assuming 43°
         geographic latitude.
 2-16

-------
To  follow the procedure, take the calculated or estimated geostrophic speed
G (cm  sec"1) as  the ambient  flow.  Geostrophic  speed  is independent  of
roughness.  Then,  the  surface  Rossby  number,  Ro0  = G/z0f,  where f  =
Coriolis parameter  (sec"1)  can be  calculated for a given z0. Lettau14 was
the first to demonstrate that this Ro0  serves as a  scaling factor for atmos-
pheric boundary layer  structure and that  it plays a role equivalent to the
role of the Reynolds number that is used in engineering flow  problems to
determine the friction  factor. The  factor  C,  named the  "geostrophic drag
coefficient,"14' ls  was shown to be a  unique  valued function of Ro0 for
thermally  neutral  states of the lower troposphere.  If height z(cm or m)  is
restricted  to the lowest part of the boundary  layer,  a first approximation
value for U(z) in the model is provided by the logarithmic law, which can  be
written as

       ^U(z) = 5.5 C G Iog10 (1 + z/z0)                               (12)

Application of Equation (12) is determined by the Karman constant and the
natural-to-common logarithm conversion factor.  Note that z0-effects appear
both  directly, under the logarithm, and indirectly, through  its Ro0  depen-
dency in the coefficient, C.
Tables 2-4 and  2-5 are self explanatory. Although the calculated values of
the parameter  1/h*f* are comparable with actual conditions,10  it must be
remembered  that  h* in Table 2-4 was arbitrarily kept constant at 300 m.
Nevertheless  contribution of  the three essential factors   (1)  lowering of
overall wind  speed,  (2) increase of city diameter, and  (3) change from low-
to high-rise buildings, can equivalently increase the net pollutant concentra-
tion. The last two effects favoring the increase of relative pollutant concen-
tration (or the value of 1/h*f*) could  be of interest to  the city planner.
Table 2-5 presents a result of  air pollution climatonomy, comparable with
the example shown  in Table 2-1.  The box model of urban diffusion applied
to momentum  predicts that for the conditions of a big metropolis such as
New York City the  average wind  speed  could  be 2 m  sec"1 lower than in a
comparable  situation over  open  country.  The  "negative  area  source,"  or
momentum  drain  due to the increase of  surface roughness, is estimated  to
equal a ground drag difference of 1.5 dynes cm"2, based on geostrophic drag
coefficients and the  relationship defining surface stress r0 as pC2 G2, where
p indicates air density. Obviously, if the number of high-rise  buildings in a
central city is  increased and  the  lot area  is not enlarged, the resulting  z0
increase will favor stagnant  air and the likelihood of pollution episodes. That
effect is additive to the city-diameter effect. (Figure 2-7).
                                                                     2-17

-------
      Table 2-4.  ESTIMATE OF BOX-MODEL PARAMETER 1/h*f*
  FOR TWO DISTINCT CLASSES OF AMBIENT (GEOSTROPHIC) SPEED
         AND THREE CITY SIZES WITH TWO BUILDING TYPES3
Geostrophic speed,
m/sec
20
20
5
5
20
20
5
5
20
20
5
5
City diameter,
km
4
4
4
4
20
20
20
20
80
80
80
80
Building type^
Low
High-rise
Low
High-rise
Low
High-rise
Low
High-rise
Low
High-rise
Low
High-rise
1/h*f*,c
sec/trf1
1
2
4
7
5
10
20
35
21
39
80
140
   aCompare Tables 2-2 and 2-3.
   bSee Table 2-2.
   cFor h* = 300 meters.
      Table 2-5.  EXAMPLE OF AIR POLLUTION CLIMATONOMY:
       LOSS OF AIRFLOW MOMENTUM OVER AN URBAN AREA3
Element
Momentum area sink
(by aerodynamic roughness)
Parameter value
Predicted speed deficit
Steady-state
Symbols
AT0
A T0p
1/h*f*
AT0/ph*f*
Numerical value
1.5 dynes cm"2
0.13m2 sec"2
16 sec rrf1
2.0 m sec"1
References
Geostrophic
drag coefficient
Holzworth10
Climatonomy
Equation (7)
aAssuming an ambient (geostrophic) speed  of  10  m/sec,  an  Aerodynamic Surface
 Roughness z0 of 100 cm (hence, C = 0.050) for the City, and 5 cm (hence, C = 0.035) for
 the Upwind Region, and a Parameter Value of 1/h*f*   16 sec/m  (Momentum  Drain
 Conditions in New York City*
 2-18

-------
 WIND PROFILE  OVER CITIES
 Boundary  layer  concepts  lend themselves readily  to application in urban
 diffusion  problems.  To emphasize this, let me use the results of  one of the
 bushel basket experiments  performed on the ice of Lake Mendota (Figures 2-9
 and 2-10). The experiment was designed to  study both  control of surface
 roughness (Rossby number) and effect of thermal  stratification (Richardson
 number).  A  total of  420  bushel baskets were used, half of them  painted
 white, the other  half, black. A lot area of  about 2 m2 was selected  in  a
 layout of two adjacent fields, 21 m cross wind, and 20  m downwind. The
 albedo  values  measured were 0.15 for  the  black  and 0.65  for the white
 obstacles. The albedo of the  aged snow on the ice averaged  about 0.52. In
 the March sunshine, the effective heating of the field with the black obsta-
 cles was  estimated  at 0.07  langley min"1, and  that of the field with  the
 white  baskets at 0.03 langley min"1, giving relatively differential heating rate
 of  0.04 langley min"1.
 Figure 2-9 illustrates the analysis of wind profile observations at a variety of
 x-positions for the anemometer  mast, in the  form of velocity defects (AV)
 and vertical  motion computed for two-dimensional mean  flow with  the aid
 of  the continuity equation.  The  velocity defeats and  updrafts  are  more
 pronounced  for  the  black than  for the  white field.  Temperature measure-
 ments suggested  that  the  difference  in  Richardson numbers between black
 and white fields varied from about -0.005 to +0.005. For cases of relatively
 strong changes  in vertical  shear, however,  it  is questionable whether  the
 Richardson number  retains its significance as a scaling parameter.

 Figure 2-10  shows the momentum  balance for adjacent "budget boxes" in
 the x-z plane between the surface and 1.6  m. All  values are  in dynes cm"2,
 double arrows refer to transport by mean  (organized) velocity components
 (u  and w), thin arrows show the eddy flux, p(u' w'),  in the vertical direction
 at  z=0 and  z=1.6 m. The black field drains 2.9, and  the white field  2.3
 dynes  cm"2,  and  the  loss  due to transport to higher  layers amounts to  3.0
 and 1.8  dynes cm"2,  respectively.  These values  indicate  an  increase  in  the
 mixing depth  due to  Richardson  number effects, similar to  the dynamic
 roughness effect  in boundary  layer development over cities in Equation (11).
 Further experiments of this type are desirable.
Although  the horizontal scale of our out-of-doors or free  air experiments is
significantly larger than the scale of wind tunnel studies, the similarity to the
prototype of  boundary layer  structure over urban  areas  is still incomplete
because on the larger, real  scale the effect of Coriolis forces appears, causing
directional  shear  of horizontal  air motion. The two dimensionality of actual
wind profiles  over cities is not accounted  for by the power-law  V 'v z a,
frequently suggested for transport of  material in the lower troposphere in
urban pollution studies; reference is made to  Davenport17 and Turner.18 In
                                                                    2-19

-------
NJ
O
                1

                0L
                -20
                         5m/sec
         W         aV
        5 cm/sec     2 m/sec
-10
                        0           10
                                   x, meters
BUSHEL BASKET EXPERIMENT ON MENDOTA ICE
                                                                  20           30

                                                                     MARCH 23,1963
40           50

11:17-14:45 CST
                V: Undisturbed velocity               W: Completed vertical motion               aV: Velocity Defect
                     Q WHITE BASKETS (11:17-12:45)                    M BLACK BASKETS (13:32 -14:45)
        Figure 2-9.  Analyzed  results  of micrometeorological experiment by Stearns and Lettau on March 23,
        1963,  using fields of bushel baskets on  ice of Lake Mendota.  Mimicks conditions found  in boundary
        layer development over city.  Undisturbed wind profile, extreme left. Vertical profiles of velocity de-
        fects and  calculated  vertical  motion plotted for different mast positions at indicated downwind dis-
        tances.  Note stronger momentum drain and deeper mixing layer over black obstacles illustrating pos-
        sible Richardson  number effect.

-------
           IM

1T 1 Till 11
•1.4 0.5
=*1.7
•0.8
1 n
•1.8 0.8
=>3-3
•2.3
n nln r
1.1 1.6
•1.7=3
•1.0
\ 1
0.8 0.0
0.0=3
•0.8
X

NJ
ro
[ f
•2.0

0 -10
i ir i
0.6
•1.5
lH

•3.0 0.8
•2.9
• •!• •
0 10
x, meters __
XX X
1.0 0.9
•1.0 =*
•0.9 *
i X
1.1
•0.4
X

20 30
»
1
1.2
•1.9=3
i . i
40 50
          BUSHEL BASKET EXPERIMENT ON MENDCTA ICE    MARCH 23,1963     11:17 • 14:45 CST
Figure 2-10.  Budget boxes showing horizontal momentum balance for experiment illustrated
in Figure 2-9.  Double arrows represent budget contributions due to  mean components u and
w; single arrows, contributions  due to eddy  flux in vertical  direction, including  surface
stress.

-------
  addition to the disregard for directional shear, the  power law has two other
  disadvantages:  (1) at low elevation the logarithmic  law  (which alone expres-
  ses the surface roughness effect rationally) is not approached as an asymp-
  totic case, and (2)  at high  elevation, there is no asymptotic approach to the
  gradient or high geostrophic  wind.  Clearly, these  two shortcomings mean
  that the  power  law is merely an  interpolation formula, with difficulties
  expressing upper and lower boundary  conditions. More realistic models of
  wind profile structure  in atmospheric  boundary layers  are available in the
  literature.  Let  me restrict discussion to one of my earlier contributions to
  this problem.  It  concerns a universal wind spiral theory  of  the barotropic
  (thermally neutral)  lower atmosphere.13 The  tabulated  functions for  the
  wind and stress components that this model generates are universally appli-
  cable to a wide  range  of external  conditions. In order  to verify a  wind or
  stress spiral in  the  atmospheric boundary  layer  for a special  case,  the only
  scaling factor  needed  is the surface  Rossby number. Figure 2-11 illustrates
   1,000
                                      APPLICATION OF UNIVERSAL WIND SPIRAL THEORY
                                                 LETTAU (1962)
                                               VECTOR WIND RATIOS V/V,
                                              Zfl= l.OOfl cm
                                                 5 cm     Roo = 2.0 x l.O4
Figure 2-11.  Universal Wind Spiral Theory verified for three values
of surface roughness (Rossby numbers, Ro0) characteristic of urban
areas.  Calculated for thermally neutral atmosphere.  Normalized to
geostrophic speed (Vg). Symbols: ZQ,  aerodynamic roughness length
                "tor (1-°x 1o'4meters sec'1  f°r 43°  -£
2-22

-------
         1,000
          800
          600
          400
   I    I    I    I   I
  UNIVERSAL WIND SPIRAL
  LETTAU (1962)
 -Vg/f=107cm
                                   1    I    T
         1,000
          800
  H   I    I    I   I    I
_ POWER LAW V/Vg = (z/H)a
  DAVENPORT (1963)
  (1) a =0.40  H =560 meters
_(2) a =0.28  H= 430 meters
  (3) a = 0.16  H = 300 meters
Figure 2-12.  Comparison between Universal Wind Spiral Theory
and  Power Law treatment of vertical profiles of normalized hori-
zontal wind speed  (V/Vg).  Symbols: Vg  -geostrophic speed;  f  =
Coriolis factor of 1.0 x 10~4meters sec ~1;  t - surface stress:/? -
air density;  C - geostrophic drag coefficient;  C H r height coef-
ficient;  H -effective height  of roughness  elements  (centimeters);
z - geometric  height; a   silhouette area seen  by the wind (meters2).
                                                              2-23

-------
verification  for  three  different Rossby numbers  that  correspond  to the
urban z0-values  shown in Tables 2-2, 2-3, and 2-4, a geostrophic speed of 10
m sec"1, (denoted by the symbol Vg in Figures 2-11 and 2-12, and a Conolis
parameter f = 1.0 x  1CT4 sec"1  (for 43° geographic latitude).

The  two  parts   of  Figure 2-11  illustrate  scalar and  vector wind ratios,
normalized  relative  to  geostrophic speed.  For  scalar wind speed, plotted
linearly against a geometric scale  in height,  straight sections at small height,
indicate that the logarithmic law is an asymptotic case. The scalar and vector
graphs also  document an asymptotic approach to  the gradient wind at high
elevations. There are reports in the literature of  wind direction changes of
the order of  10 degrees,  at  heights of about  60 m, between  upwind  and
downwind towers. For reference see Graham's1 9 analysis of data from Fort
Wayne, Indiana.

Figure 2-12 compares the power law17 profile  with the scalar wind profile
derived from the Universal Wind Spiral  solution.15  The latter provides all the
necessary  numerical  scale  factors,  including  the  geostrophic drag coefficient
C, and the  height coefficient CH, as unique-valued functions of the surface
Rossby number.  The  power  law  is deficient in  this respect and its applica-
tion,  with  the  assumptions  listed on the lower part  of  Figure 2-12, is
arbitrary or empirical.
SUMMARY
I  hope to have  convinced  some of you that  fluid-dynamical boundary-layer
concepts are promising physical  and  meteorological bases for  mathematical
models of urban diffusion  processes; and that mass, momentum, and energy
transfers and transports can be effectively handled by a  systems approach
with: defined forcing functions, parameterization of response functions, and
feedback and control functions. Much  detailed  research still  has to be done,
especially with  regard  to  the combined  feedback processes of turbulence
generation by surface roughness and  heating over urban areas.
 2-24

-------
REFERENCES
 1. Moses, H.  Mathematical Urban Air  Pollution Models. National Center for Air
    Pollution Control, Chicago  Dept. of  Air Pollution  Control,  Argonne  National
    Laboratory. Argonne, III. ANL/ES-RPY-001. April 1969. 69 p.
 2. Lieber, H. Controlling Metropolitan Pollution  Through Regional Airsheds: Admini-
    strative  Requirements  and  Political  Problems. J.  Air  Pollution Control Assoc.
    18:86-93, February 1968.
 3. Lettau, H. Evapotranspiration Climatonomy. 1. A New Approach to Numerical Pre-
    diction of  Monthly Evapotranspiration  Runoff and  Soil  Moisture Storage.  Mon.
    Weather Rev. 97:691-699, October 1969.
 4. McCormick, R. A. and J. H.  Ludwig. Climate  Modification  by Atmospheric  Aero-
    sols. Science. 755(37801:1358-1359, June 9, 1967.
 5. Lettau, H. and  K. Lettau. Shortwave Radiation Climatonomy. Tellus. 27:208-222,
    February 1969.
 6. McCormick, R. A. and C. Xintaras. Variation of Carbon Monoxide Concentration as
    Related  to  Sampling Internal, Traffic and Meteorological Factors. J. Appl.  Mete-
    orol. 7(21:237-243, June 1962.
 7. Johnson, W. This  Symposium, Chapter 5.
 8. Schuck, E. A., J. N.  Pitts, Jr., and  J. K. S. Wan. Relationships Between Certain
    Meteorological   Factors  and   Photochemical  Smog.  Int.  J.  Air  Water Pollution.
    70:689-711, October 1966.
 9. Harris, D. N., J. R. Huffman,  and J. H. Weiland. Another Look at New York City's
    Air Pollution Problem. J. Air Pollution Control  Assoc. 78:406-410, June  1968.
10. Holzworth, G.  C.  Mixing Depths, Wind Speeds and Air Pollution Potential for Se-
    lected Locations in the  United States. J. Appl. Meteorol. 5(61:1039-1044, Decem-
    ber 1967.
11. Bornstein, R. D. Observations of the  Urban Heat Island  Effect in New  York City.
    J. Appl. Meteorol. 7(4):575-582, August 1968.
12. Petersen, J. T. The Climate of Cities  (A Survey of Recent Literature). National Air
    Pollution Control  Administration.  Raleigh,  N. C.  Publication Number AP-59.  1969.
    48 p.
13. Lettau, H. Note on Aerodynamic  Roughness-Parameter Estimation on the Basis of
    Roughness Element Description. J. Appl. Meteorol. S:828-832, November 1969.
14. Lettau, H. H. Wind Profile, Surface Stress and Geostrophic Drag Coefficients in the
    Atmospheric Surface Layer.  In: Advances  in  Geophysics, Landsberg,  H. E. and J.
    Van Mieghem (eds.) Vol. 6 New York, Academic Press, 1959. p. 241-257.
15. Lettau, H. H. Theoretical Wind Spirals in the Boundary  Layer of a Barotropic At-
    mosphere. Beitr. Phys. Atmos. 35(3/41:195-212, 1962.
16. Lettau, H. H. Problems of Micrometeorological Measurements (On Degree of Con-
    trol in Out-of-Doors Experiments). In: The Collection and Processing of  Field  Data;
    A CSIRO Symposium,  Bradley, E. F. and  O.  T. Denmead (eds.). New York,  Inter-
    science Publishers, 1967. p. 3-40.
17. Davenport,  A.  G. The  Relationship  of Wind  Structure to Wind  Loading. In: Pro-
    ceedings of International Conference  on Wind  Effects on Buildings and  Structures.
    National Physical  Laboratory, Teddington,  England, June 26-28, 1963. London, H.
    M. Stationery Office, 1965.
18. Turner, D.  B. Workbook of Atmospheric Dispersion Estimates. National Air  Pollu-
    tion Control Administration.  Cincinnati, Ohio. PHS Publication Number  999-AP-26.
    Revised  1969. 84  p.
19. Graham,  I. R.  An  Analysis  of Turbulence  Statistics at Fort Wayne,  Indiana. J.
    Appl.  Meteorol. 7(1):90-93, February  1968.
                                                                            2-25

-------
             APPENDIX - GLOSSARY OF SYMBOLS

a       silhouette area of roughness element
A      lot area of roughness element
C      geostrophic drag coefficient
CH     height coefficient
D      city diameter
f       Coriolis parameter
f*      flushing frequency of volume of city air
G      ambient air flow, gradient wind, or  geostrophic wind
h*      upper boundary, or thickness, of volume of city air
H      effective height of roughness elements
q       average pollutant concentration for  box model
q*      quasi-equilibrium value of average pollutant concentration for box
        model
q0      initial value of q at t = 0
q(t)     response function, of air pollution climatonomy
Q      area source of pollutant emission
Q(t)     forcing function of air pollution climatonomy
s       specific admixture to air
S       strength of internal source (sink, if negative) of considered admixture
t       time
t'       dimensionless time
At      time interval
U*     bulk value of windspeed
V      velocity vector =  i u + i v + k w
->•                      ->->->
Vg      geostrophic windspeed
(w1 s1 )   vertical eddy flux of admixture
x       downwind coordinate
y       crosswind coordinate
z       height
z0      aerodynamic roughness length of urban area
v       molecular diffusion coefficient
p       air density
       function
ij5~      average of function
2-26

-------
ABSTRACT
      777e method currently in practical use in the United Kingdom Meteoro-
      logical   Office  for  predicting  levels  of  air  pollution  from  source
      strengths  and  meteorological data  is  reviewed briefly.  A  study of
      6-hour average 5O2  concentrations  in  the town of Reading, recently
      published by the British Petroleum  Research Centre, shows that, while
      the overall average may be predicted reasonably well by the foregoing
      method,  the  individual predicted  values have  large  deviations from
      those observed.

      Some aspects of the theoretical and empirical basis for the estimation
      of  the  diffusive  spread of  material  are considered—particularly  the
      applicability  of the  gradient-transfer approach  and the indications of
      Lagrangian similarity arguments regarding the effects of surface rough-
      ness and thermal stratification.

      The vertical spread for neutral flow over a  roughness typical of urban
      areas predicted on  Lagrangian similarity is found to be compatible with
      that recently estimated  from tracer experiments in St.  Louis, Missouri,
      for evening, when  neutral thermal stratification was indicated by tem-
      perature observations in the  city. It has  to be admitted, however,  that
      the actual flow conditions were  most unlikely to have  been in equilib-
      rium with the urban terrain over  the whole depth of spread of the
      tracer.  Furthermore, there is some  indication  of substantially greater
      spread  during  the  day even  when similar (neutral) stratification  pre-
      vailed in the city; this feature remains  to be explained.

      Elementary aspects of the influence of  the  urban heat island  in  aug-
      menting vertical  spread are considered. Examination of the St Louis
      data reveals some evidence to support  the expected augmentation, but
      there is no obvious support for the notion that vertical spread into an
      initially stable atmosphere is controlled solely by the urban heat input
      as in Summer's model.'

      Finally, brief consideration  is given to the prospects  of and require-
      ments for improvement in  the systems for predicting air pollution.
AUTHOR
      Frank Pasquill took an Honours degree in physics at the University of Durham in
      1935 and was awarded the D.Sc. in  1950. He joined the United Kingdom Me-
      teorological Office in  1937 and since 1954 has  held a special appointment for
      individual research.  He is also  head of the Meteorological  Office's  recently
      formed  Boundary Layer Research  Branch. Throughout his career he has been
      concerned w,th aspects of diffusion and turbulent transfer in the atmosphere. He
      is author of the book Atmospheric Diffusion.

-------
                                3.   PREDICTION  OF DIFFUSION
            OVER  AN  URBAN  AREA - CURRENT  PRACTICE
                                      AND  FUTURE  PROSPECTS

                        FRANK PASQUILL
             Meteorological Office, Bracknell, England
INTRODUCTION
There are three basic  components  in  the development  of  any system of
estimating the concentration of air pollution solely from a knowledge of the
sources  and  of the meteorological conditions. The first  is a theory within
which the relations between the spread of the material and the  properties of
the airflow are at  least implicit, if not entirely explicit. Secondly, for opera-
tional use, this theory  must then  be expressed in a suitable  form  (so-called
diffusion formula  or dispersion formula) giving the concentration of pollu-
tant in terms of the  source-strength and meteorological parameters. Finally,
there is the requirement for observations of the  levels  of air pollution pro-
duced by  known sources, against  which  the theoretical  system can be tested
and from  which generalizations and  predictions may then  be made. It hardly
needs to be  said that the effective union of these components has proved to
be a slow and complicated  process,  in which considerable  care is still needed
to distinguish the essential  features and to avoid unrewarding expenditure of
research effort.

The  first  purpose  of the present  paper is to note the  stage  reached in the
United Kingdom in the provision  and testing of  prediction systems in  prac-
tice. Although  it is evident that for urban situations  there are  complexities
that  seem intractable to precise theoretical analysis, it is nevertheless impor-
tant  to  extract as much guidance  as possible from  basic ideas about the
diffusive  action of idealized airflow. Only by such a  process can we derive
the greatest  benefit from such practical  experience as is now available, and
hence make  the most  sensible  appraisal of the further work required. Ac-
cordingly, the second purpose of the paper  is to recall  the basic concepts
and to note some  ideas and results  that seem likely to be particularly rele-
vant  to future development. From this point, attention  is  turned specifically
to the factors determining the average rate  of vertical spread in urban airflow
and then,  briefly, to  the prospects of achieving significant improvements in
the ability to predict the distribution of air pollution from specified sources.
                                 3-1

-------
CURRENT PROGRAM IN UNITED KINGDOM
In the Meteorological Office of the United Kingdom, the practical estimation
of pollutant concentrations  from  specified sources currently follows lines
that  were formulated  some 10 years ago, largely under the  stimulus of the
accident  at  the Windscale Works of the Atomic  Energy Authority. Although
this system  does  not contain  any specific application to urban conditions, it
is nevertheless relevant to our present considerations in  two respects.  It has
been  widely used  as  a  basis  for examination  and  comparison of  many
measurements, including some of air pollution in urban conditions. Also, it is
now  in the process of a revision that has a bearing on the extension to  urban
conditions.

The details of the current system are set out and discussed  in two articles2'3
and  here it  is proposed  to  do little  more than  emphasize a  few crucial
features:
1. The original  aim was to provide a practical system with  which,  as far as
   possible,  crosswind and vertical spread  of  material  could  be estimated
   from data on  the corresponding fluctuations of the wind.
2. For the case  of short-range  vertical  spread from a ground level source,
   and,  in general, when data on wind fluctuation  were not  available,  broad
   estimates of  spread were provided for categories of atmospheric  stability
   specified  in terms of surface wind speed and state of sky.
3. The broad estimates for short-range spread, < 1 kilometer  (km)-referto
   level, unobstructed  terrain with rather small aerodynamic roughness, z0
   about 3  centimeters  (cm).  (Standard notation  is  followed—see end  of
   paper for list.)
4. The  broad  estimates for  longer range, partially based on  experimental
   data, automatically include some  influence of  the minor topographical
   and urban features present in the area of Central Southern England.
5. No means of quantitatively allowing for either major topographical fea-
   tures or urban influences  were  provided, though some  qualitative appre-
   ciation  of the effect of topography was briefly attempted in the  latter of
   the two  publications.
6. As far  as the magnitude  of spread was concerned, no  distinction was
   drawn between surface and elevated sources.

In operational use,  the system has been  limited  and inadequate in a number
of respects; so far,  no attempts have  been made to allow for these  inade-
quacies beyond  the expression of rather broad qualifications and  specula-
tions.

Experimental and observational  work on the distribution of  air pollutants  is
carried out  in the United Kingdom by  the Warren Spring  Laboratories, the
Central   Electricity  Research  Laboratories, and  the British  Petroleum  Re-
search Centre. It  appears, however, that the only study  published so far on
3-2

-------
the use of diffusion  formulas in prescribing urban air pollution is the one
carried  out by the last-named  organization in the town of Reading. That
study included extensive measurements of the sulphur dioxide concentration
distribution and detailed estimation of the source inventory over a 15-month
period during 1964 and 1965. Aspects of the study have been discussed  in
papers by  Marsh and  Foster4;  Marsh,  Bishop, and Foster5;  and  Marsh and
Withers.6   The  last  of these  publications compares  the observations with
estimates derived from  the source inventory and meteorological data, using a
number of methods,  including  those used  in the Meteorological  Office and
referred to above.

The  pollution data consisted of sulphur dioxide concentrations averaged over
6-hour sampling periods,  for 40 sites in and  around  the town. Continuous
records were available for  wind  direction, wind  speed, air  temperature,
humidity, and  the intensities of the vertical and  lateral components of tur-
bulence. The wind  speed  and turbulence  measurements were made at  a
height of 14 meters  (m) in a  relatively open  space,  100 m or more from
buildings, about 10m high,  and about 700 m from the taller  buildings in the
center of  the  town.  For the source  distribution, three categories of fuel-
burning  installations were considered—domestic heating, industrial  and com-
mercial  installations  for space-heating,  and   processing plants.  Domestic
(space-heating) sources  were divided into 251  housing  areas with a diameter
of roughly 400 m, and the  strength  of source in each area was based on
official estimates of the average fuel consumption in private houses. Installa-
tions  burning more  than   10  tons  of  fuel  per  year  were  considered
individually.

For  evaluation  of the diffusion formulas, Marsh  and Withers6 defined  67
area  sources that were treated as  individual  sources  if their  centers were
farther than  1000 m from  the sampling site, and 7 major sources, that were
treated as individual sources regardless of distance. If the center of  an area
was  less than 1000 m from the sampling site, the individual  housing areas or
installations  were  also  treated as individual sources.  In  some  of the calcu-
lations, the  housing areas and  source areas were treated as  equivalent line
sources.  For  the domestic sources, a standard  height of  15  m was adopted,
but  for  individual sources  the  effective  heights were estimated  from the
chimney height and  the plume rise calculated from the CONCAWE empirical
formula.7

Marsh and Withers7 give an  extensive  series of comparisons of observed and
calculated  concentrations from which  we  shall reproduce only a selection
that  most  directly reflects  on the present point of interest. The calculated
concentrations were obtained in three ways:
   1.  Using the  familiar Gaussian  form  for  the distribution from  a point
       source with the standard deviations {of particle displacement) cry and
       a, obtained from wind  fluctuation data.
                                                                      3-3

-------
   2.  As in (1.)  but with CTy and  az  obtained from the broad estimates of
       spread.
   3.  From a  regression of the observed values against a parameter invol-
       ving  the  air temperature  (as representing the variation  of effective
       overall source strength) and the reciprocal of the wind speed.
The  selection of  comparisons, in  which the  observed  concentrations have
been corrected for 'background' concentration using measurements made up-
wind of the town, is given in Table  3.1
In Marsh  and Withers' analysis, attention is focused primarily on the root-
mean-square  (r.m.s.) differences between the observed and calculated 6-hour
average concentrations. In Table 3-1, these r.m.s. differences are always large
and are of similar magnitude to the corresponding overall mean of the obser-
ved concentrations.  Furthermore, the r.m.s. differences expressed  relative to
the observed mean are lowest for the  regression method and highest for the
broad  estimates of  spread.  On this basis. Marsh  and Withers conclude that
diffusion  formulas do not  provide a satisfactory basis  for calculating  the
pollution  in  a  town  and  that  a  regression  equation  directly  representing
observations  is to  be preferred.
        Table 3-1. COMPARISON OF OBSERVED AND CALCULATED SO2
      CALCULATIONS AT SIX SELECTED SITES IN READING, ENGLAND6






All periods3




119 selected
periods^



calculation
a. Wind fluctuation
b. Broad estimates
of spread
(stability categories)
c. Regression on tem-
perature and wind
a. Wind fluctuation
b. Regression on tem-
perature and wind
Mean of 6-hour
concentration
M9 nrf 3
Obs. C
68


68

68
66
66
Calc.C
38


88

-
53
_


r.m.s.difference
between calculated
and observed C
84


132

68
63
57



r.m.s.
obs. C
1.2


1.94

1.0
0.95
0.86
3  Methods a. and b. follow Pasquill2
b  Steady wind direction and speed  (6 a.m. to 12 noon and 12 noon to 6 p.m. only)
   for which most confident estimates of emission were made.
3-4

-------
It  is not  entirely  surprising, however,  to  find that the 'regression  method',
which  uses  the  actual observations,  provides the  least  scatter between  ob-
served  and calculated values. On the other hand, it is encouraging, that the
'wind  fluctuation'  method  provides independent  estimates with  individual
departures that are  not very  much  larger than those of  the  regression
method.  Also,  the fact that the  least agreement in the  individual  6-hour
values  is provided  by the broad estimates of spread is  not surprising, since
the latter  refer to aerodynamically smoother  terrain  and, as will be seen
later,  the rougher  nature of the terrain is reflected in the wind fluctuation
measurements.  Finally, Marsh and Withers are inclined to attach little signifi-
cance to the overall mean concentration.  In view  of the very wide range of
the component values,  it is particularly  interesting that the observed  and
calculated mean values are as close  as  they are, and that their agreement  is
even closer when  occasions are  selected for greater reliability in  wind  and
emission estimates. Before discounting  the agreement on  the score  of  the
discrepancies in individual  values,  it would seem desirable to examine in
more detail the significance of,  and possible reason for, such discrepancies.
(This point will  be  referred to again in the final  discussion.)

BASIC CONCEPTS
IN THEORETICAL ESTIMATION OF DIFFUSION
The theoretical treatment of  diffusion in a turbulent  fluid  has  developed
along three  main lines.
1.  The gradient-transfer relation,  in which the  eddy  transfer of material
    across a plane  is represented as a  product  of the  gradient of material
    (normal to the  plane) and an eddy diffusivity, K.
2.  Statistical  theory,  involving statistical  descriptions and  laws for  the
    velocities and trajectories of typical  particles  in the fluid.
3.  Dimensional analysis and similarity theory.
Discussion will  be confined to  some  specific   features that  seem to have
particular relevance to future development of methods for estimating  urban
air pollution.

Applicability of Various Approaches
The fundamental validity of the gradient-transfer  approach has always been
regarded as a matter of debate, and it is important to make clear the circum-
stances in which it is most acceptable.  Any qualitative basis for the method,
as is involved in the classical 'mixing length' theory or in direct analogy with
kenetic theory  of  molecular motion, inevitably implies the notion that the
physical  scale and effective range  of  action of the turbulent elements be
small compared  with  the domain over which the diffusible material is spread.
Obviously, this  condition may be  expected to  be  approached as the spread
of the material  increases with time or distance of travel from the source.
                                                                      3-5

-------
Consistent with this physical idea  is the formal result from statistical theory
that,  in the  limit of long  time, T, the  spread  of  particles in homogeneous
turbulence is proportional to TV2, as in the Fickian (constant  K) solutions of
diffusion.
Recognition  of the  existence of the foregoing parabolic law  is confused  by
two  practical  features  of atmospheric  flow.  The  first  is the width of the
spectrum  of  eddy  motion,  which  is  such  that  diffusive  spread  over  di-
mensions  larger than all effective eddies may  not be  achieved  in  practice.
The  second  is the  lack  of  homogeneity in the atmosphere regarding such
properties as mean wind speed and direction, and the intensity and scale of
turbulence. No demonstration has  yet appeared  for the existence of the
parabolic   relation  in  horizontal  spread.  There  is,  however,  evidence  in
Hogstrom's8   experimental   study  using an   elevated  source,  of  a  close
approach  to  this relation for vertical spread in stable conditions. This seems
entirely consistent with the idea that, in such  conditions, the  vertical scale of
the eddies will be significantly limited.  It  is not unreasonable to expect this
limitation  in vertical scale  to be effective irrespective of stability, provided
the cloud  or  plume of material is not  clear of the ground (i.e.,  as  from  an
elevated  source).  In other  words, vertical diffusion  from a source that  is
effectively at ground level  may reasonably  be expected to be in accordance
with the gradient-transfer theory,  without necessarily leading to the parabolic
law,  since the  effective  K must then be expected to be a function of height.
On the other hand, it  is  evident that the  early stages  of spread in the
horizontal, or  in  the vertical, when  the source is  elevated, cannot generally
be attributed to a constant diffusivity  or to  a function only of position in
the flow.  In  these  cases, the statistical theory appears to  be  a preferable
alternative.

The  Lagrangian similarity theory  was originally put forward as  a means of
predicting  the variation  of a  continuous  point-source  concentration  with
distance  downwind. Since, crosswind diffusion is implicity  involved, there is
some controversy regarding the fundamental  validity of the approach, but
this  qualification  does  not apply  to  vertical  spread. In any  case, there is a
basic  restriction of the height range over which the horizontal shearing  stress
is effectively constant.

General Explicit Form for K

Well-known forms for K  have been specified for two limiting  conditions, i.e.,
for the constant-stress layer, when, for vertical  momentum transfer

       K   =  ku*x/0  (f)                                             (1)

           =  ku*z, at small  (^                                         /2\
 3-6

-------
and for particle  spread in the limit of long time in homogeneous turbulence

         K  =  <& tL                                                 (3)

Invoking dimensionally  acceptable and empirically determined relations be-
tween the variables crw, u*, tL, tE, Xm, and z, Equations (2) and (3) may be
reduced to a common form

         K  = Tjy  aw *m                                            (4)

where Xm is the equivalent wavelength for peak nS(n) in the spectrum of the
vertical component.
Invoking a further relation between  CTW, Xm, and  e  (see Pasquill9), this may
be further transformed to
                           4/3
                         m
Note that equations (4) and  (5) allow K to vary with height  in an arbitrary
way, according to the variation with height of the properties aw, e, and Xm,
and prescribe the magnitude  of  K in  terms of  any two of these three quan-
tities. The expressions for  K  are suggested as plausible universal forms, to be
used irrespective of height and stability  in the  atmosphere, in future studies
of the vertical spread  at medium- and  long-range from a surface source.
Estimation  of  the quantities  Xm, aw,  and  e requires  special  turbulence
measurements, either  for direct use or for the specification of climatological
values  for  subsequent  use.  Recent  studies  such  as those of  Busch and
Panofsky10,  Pasquill9,  and  Readings  and Rayment11 are relevant in  the
'climatological'  respect. The  first  of  these provides information on aw and
Xm in  the lowest hundred meters or  so of the  atmosphere, and  the second
and third are first  steps towards a more comprehensive  specification of aw
and e over the first kilometer.

Effect of Aerodynamic  Roughness and Thermal Stratification on
Vertical Spread

Examination of  the effect of aerodynamic roughness and  thermal stratifica-
tion  on vertical  spread can  be  considered in terms of the gradient-transfer
method,  in so far as  the effective K is appropriately specified. The forms of
K  suggested  above  do, in  fact,  reflect  the  roughness  and stratification
directly,  in  the  case  of  Equations (1) and (2), and indirectly, through CTW,
Xm, and  e, in Equations (3) and (4). Such forms may, therefore, be useful in
making  theoretical  estimates,  subject  to the  basic acceptability  of  the
method.  For those circumstances  in  which the gradient- transfer  method is
                                                                     3-7

-------
inapplicable,  and in which the statistical theory is more appropriate (notably
the calculation  of  crosswind spread or of vertical  spread at relatively short
range from an elevated  source),  the effects of roughness and  stratification are
reflected  in  the magnitudes of  aw and tL. Even for  the case of vertical
spread from a  ground level source at relatively  short  range,  however, the
gradient-transfer method  is complicated  by  the  logarithmic  form of wind
profile applicable in neutral conditions; the Lagrangian similarity theory now
seems to  provide an  approach  that is  neater analytically and more satisfying
physically.
For  neutral  flow,  the  Lagrangian similarity theory  provides the  following
relation between the  mean height reached  by successively released particles
(Z)  and  the mean  horizontal  distance that they  have travelled in a given
time, (X).

        Y  = W    [ln 1^   -1   +  ^  d-lnc)]                 (6)

It has been  argued that the dimensionless constant  b  is identical  with von
Karman's constant,  k,  and  there  is some observational evidence for  this.12
On the other  hand,  the constant c has  not yet been well-determined. This
constant  is, in  fact, the ratio of the mean horizontal velocity of the particles
to the  mean fluid  velocity  at  the height Z, and in the earlier forms of the
treatment was  assumed to be  unity. Since  it seems likely to  be between 1/2
and  1   [a  recent theoretical  estimate  by Chatwin13  gives 0.56],  the re-
maining uncertainty in its  precise value  is not particularly important.  Given
the constants b  and c, and  a specification of the characteristic aerodynamic
roughness through  z0,  Equation  (6) can provide  the magnitude  of vertical
spread as  a function of distance.

In the  testing  and practical use  of Equation (6),  there are two other com-
plications not   yet mentioned, which,  fortunately, seem unlikely  to  be of
major  importance in terms  of  the expected accuracies:  1.) Most  published
results on vertical spread  have  been expressed in  the form  of  r.m.s.  spread,
0Z or of  'height of cloud', conventionally defined as the height at which the
concentration  or dosage is one-tenth  of the gound-level value.  Conversion
from Z to one  or the other of  these more useful forms requires the shape of
the vertical  distribution to be specified. In practice it  is customary to  assume
the  simple  Gaussian  form.  While  there  is no  obvious reason to prefer any
other form  in   the case of crosswind spread from  a point source  or vertical
spread from  an elevated  source, there  is  some doubt about its suitability in
the  case  of  vertical   spread  from  a   ground-level  source.  On  theoretical
grounds, the exponent  s  in  the exponential form,  exp(- bzs), seems likely to
be somewhat less than the Gaussian value of 2.0;  and experimentally deter-
mined values lie in  the  range 1.1 to 1.5.1 4- l 5

Measurements  of vertical  spread  of air and  the associated concentration of

3-8

-------
material from a continuous source refer to particles that have all travelled to
a specified position at a distance x downwind  from the source. On the other
hand, in  the foregoing theoretical  result, the vertical spread  is related to the
mean of all distances travelled by all particles in a particular but  otherwise
unspecified  time.  In practice,  the assumption  is made, so far without any
formal or observational demonstration, that the two specifications of vertical
spread are identical when x = X.

Finally, it should  be  noted that although Equation  (6)  has  a fairly con-
vincing theoretical  background, it has so far  been tested  in a limited  way,
only for diffusion  over surfaces with very small z0  (1 cm to 3 cm).

Turning  to  the effect  of  thermal  stratification,  it may  be noted that an
extension of  the  Lagrangian  similarity theory to stratified conditions has
been  developed and,  to some extent, tested.16 Here  again, there are con-
troversial  features,  with specific respect to  the vertical spread.12  Neverthe-
less, the  notion that the effect of stability on properties of vertical spread is
somehow characterised  by  the Monin-Obukhov length-scale has  considerable
appeal, and could  be used  as the stability parameter in  correlating measure-
ments of  vertical spread.  The  writer has  recently  been reviewing  this
possibility  in  relation to the prairie grass measurements at  O'Neill,  Nebraska,
in  1956,  also with the  objective of assigning values of  L  to the  qualitative
stability  categories referred to earlier in this paper. A  reasonably  close cor-
relation between az and Panofsky's  L'(="Kj^"L)  has already been demon-
strated12  (Figure  3-1),  but L' (as distinct from L) has  the  complication of
being a function  of height even  in  the constant-stress  layer. A completely
satisfactory assessment  of the  values of L, to be associated  with  the prairie
grass data, has not yet been achieved. At present, the most reasonable assess-
ment is  based on assuming Ri =-£- as a working approximation in unstable
conditions.  This suggests that the minimum  value of  |L|  in the  particular
prairie grass data summarized  by  the writer2 was between 3 and 4, and that
the values of  |l_|  to be associated with stability categories A,  B, and C  are
roughly  1,  3, and  10.

ESTIMATE OF  EFFECT  OF URBAN  ROUGHNESS
Although the concept that the enhanced  roughness in  an  urban area would
cause an  increase in  vertical spread,  as compared with open and relatively
smooth country, seems to have been accepted, no  theoretical estimate of the
magnitude of the effect  has yet been offered.  For relatively short distance of
travel, we  can easily make such  an estimate from the Lagrangian similarity
treatment for neutral conditions.  In Equation  (6),  b =  k =  0.4; if c is 0.6, it
then remains to specify the roughness length, z0, for an urban area.

Davenport1 7 quotes a value of 3 m for z0,  derived from wind profile obser-
vations  by  Jensen, in  the center  of Copenhagen,  and  by  Shiotani and
                                                                      3-9

-------
l,000r
      500
      100
       50
        10
                                                         T
                                                     (Z=2230)
                                                        A
                         i   i  i  i i i  i
         0.1                        1                         10
                          DISTANCE DOWNWIND, km

            — —  FROM LAGRANGIAN SIMILARITY THEORY,
                              z0 = 3m (upper curve), 0.03m (lower curve) C = 0.6
            	  EARLIER SEMI-EMPIRICAL CURVE FOR OPEN COUNTRY2
 Points from St. Louis dispersion study experiments No. 16 (0), 24 (•), 35 (A),
 37 Q,  42 O< 43 (A).  Open points for daytime, solid for evening experiments.

  Figure 3-1.  Vertical spread Z as a function of distance from a
  source near ground level.
3-10

-------
Yamamoto,  in  the suburbs of Tokyo.  In a later paper on  the wind profile
and  spectrum  of wind  speed  in  London, Ontario, Davenport18  gives  a
further  estimate of 2.3 m. An independent rough indication is  provided by
Marsh's 19  measurements  of turbulence in the rather open site near the
centre  of  Reading. These give ffw/tr =  0.17 at a height of  14 m in neutral
conditions. Adopting the value 1.25, which now seems to be well established
for the  ratio aw/u», and substituting in the usual form kTT/u* = In(z/z0) for
the wind profile, we obtain z0 = 0.7m. This may be an overestimate since no
allowance  has been made for  a zero-plane displacement  in the wind profile.
There is,  however, probably another underestimation arising  from the fact
that aw was measured over an area more open, and  hence, smoother than
the town  in general. On the whole, it would appear  that values in the range
1 to 3  m may be taken as typical for modern urban complexes.

Equation  (6) has accordingly been evaluated for distances in the  range 0.2 to
5  km,  using z0  = 3  m and, for comparison with  relatively smooth open
country, = 0.03  m. The results  are shown graphically  in Figure 3-1. Note
that the difference  of 100-fold  in roughness length brings an  increase in
vertical  spread  that ranges  from 4-fold  at the shortest distances to just over
2-fold  at  5  km. The growth  of vertical, spread  with distance  may  be
approximated by a simple power  lawZ  <* Xp ; the curve can be  fitted  at the
greater distances  with  p equal to 0.85  for z0  = 0.03 m  and 0.72 for z0 = 3
m. For comparison, the corresponding section of the semi-empirical curve for
vertical  spread  in  neutral conditions in open country2  has  been  reproduced
in  Figure 3-1, after converting from  a 'height of cloud' to  Z, assuming a
vertical  distribution of the form  exp  (-bz~1-S). As already pointed out, this
empirical  curve corresponds to a z0 of about  0.03 m,  and  is  remarkably
consistent with the Lagrangian similarity prediction. This latter agreement  is
not  entirely  new, having  already  been demonstrated  for a distance of 100
m,1 2 but  here  is additionally  verified over the considerable range of distance
up to  5 km.  In  view  of  the magnitude of  vertical  spread at the extreme
distance,  however, implying diffusion  over  some  hundred of  meters,  i.e.,
outside  the  likely depth of the constant-stress region, the  extension of the
range of agreement may be rather fortuitous.

No useful direct measurements are yet  available for  vertical spread  over  an
urban area, but indirect estimates have recently been  published  by McElroy
and  Pooler20 on  the basis  of a tracer  study in St. Louis, Mo. The St. Louis
'dispersion study' was  carried  out from  1963  to  1965, using a point source
of fluorescent  particles (zinc  cadmium  sulphide) near ground level.  In each
of about 40 experiments,  the source was maintained for about  1 hour, and
measurements of total dosage at the  surface were obtained on three nearly
circular arcs at distances  between about 1  and 10 km.  (A  few  vertical
distribution measurements were made from a tethered balloon but, evidently,
these did not provide any definitive information.)
                                                                    3-11

-------
By  an  application of mass conservation, McElroy and Pooler derive values of
equivalent  az  from  the observed  surface  distribution  of dosage  and  the
amount of tracer emitted.  In  certain stages of their calculation,  a  Gaussian
distribution  in  the  vertical  is  assumed  and,  in  adjusting  the  values  for
comparison with the Z of Figure 3-1, a correction has been included on the
assumption that  the vertical  distribution  was  actually  of the Jorm exp
(-bz1'5).  Actually,  the effects of  transforming  from  az  to  Z  and  of
correcting  for the  shape of  the  vertical  distribution  are  both  small and
largely  compensatory,  and the  net adjustment is less than  10 percent;  so
that,  for  practical purposes, the magnitudes adopted  for Z are essentially
those  given  for  az  in  Table  2  of  McElroy  and Pooler. The  data  for
effectively neutral conditions, in which  the bulk Richardson Number, RiB,
was within the range ±0.01,  i.e.,  those of experiments 16, 24, 35, 37,  42,
and 43 are plotted on Figure 3.1. Note that the data from the four 'evening'
experiments  are all  above the  z0 = 0.03 m curve, and their average trend is
obviously more in accordance  with, though slightly lower than,  the curve for
z0  = 3 m. Closer examination of the variation with z0  suggests that the best
fit  to  the  evening data would  occur with  a z0 of about 1 m. In considering
this result, it must be  remembered that  the theoretical  treatment  assumes
horizontal uniformity and that the flow over the city  is in equilibrium with
the surface characteristics, at least over the height range encompassed by the
vertical spread of the tracer. The values  of Z/x range from  about 0.06 at 1
km to about  0.04 at 5 km from the source. The corresponding ratio of Z to
the upwind  extent  of  flow over urban terrain  is presumably smaller  (since
the source was not  located at  the  upwind  edge of the  city)  but it  seems
unlikely that  it could be less than say, 0.03, at the  longer sampling distance.
Since  the  currently  suggested  ratio  of the height  of  the new equilibrium
layer to the  fetch over  a new terrain is  <0.01, the vertical distributions of
the tracer cannot  be assumed to have been entirely within the equilibrium
layer.

Also,  most  of  the  estimates  of  vertical spread  from  the two 'daytime'
experiments  are  substantially  larger than  those from  'evening'  experiments.
In considering this difference, it  is to be noted that McElroy and Pooler used
rather  different procedures  in deriving their  az.  They  assumed  a  Gaussian
distribution from the start,  with a  wind constant with  height  for 'daytime1
experiments.  For  the evening experiments, the wind was assumed to have a
simple power law variation with height and an equivalent 'height' of cloud,
h, was evaluated on the assumption that  concentration  was constant up to h
and then  fell to  zero.  The  height,  h,  was converted  to  an  'equivalent
Gaussian'  r/z  by writing
       h x0
3-12

-------
and  assuming that the  effective  mean wind  speeds, ue  and ud, were  the
same. Since  ud is strictly  a mean of the wind distribution over the  real
height of the cloud weighted according to the concentration, it follows that,
if  the  cloud distribution did extend  above h  (as it would  in a  Gaussian
distribution), ud > ue, and the az derived from h would be an overestimate,
suggesting that the discrepancy between the 'evening' and 'daytime' results
could be a consequence of the difference in analytical procedure.

The  tentative conclusion is that, in evening conditions with neutral thermal
stratification  within the city, the  enhancement of the average vertical spread
by urban roughness  is consistent with similarity theory. In practice,  this
vertical  spread  is about two  to four times that normally ascribed to neutral
conditions  over level, unobstructed, open country.  In  daytime conditions,
however, which are apparently  identical  as  regards thermal stratification in
the city, the vertical spread implied by surface concentration downwind of a
source appears to  be substantially greater  than can be  accounted for by
roughness. The reason for this discrepancy  is not yet evident.

EFFECT OF URBAN HEATING
There  is abundant evidence  that  the temperature regimen within an urban
area  is measurably different  from  that  in  open country.  Mid-urban air
temperatures are  known  to  reach values  higher  than those  in  surrounding
rural districts,  by  amounts  ranging from  a few  tenths  to  several degrees
Centigrade,  the maximum effect  (in  UK  cities) being found on calm clear
nights  in  summer. There are probably two important contributions in  the
production of  this heat island. The first  is concerned with the  natural  heat
transfers;  the  balancing of  the  net supply  or loss by radiation  must be
provided predominantly by ground conduction and turbulent transfer in the
air;  whereas in  vegetated  areas,  the loss  of  heat  by  evaporation  is  an
important and, at times, controlling contribution. As a consequence, surface
temperatures may reach appreciably higher values in towns and cities than in
country districts.  Cities also  have a second  'man-made'  supply of heat that
will  contribute  to enhance urban temperatures.

It  is  to be expected that the  urban heat island  will affect both the horizontal
airflow  pattern and  the  vertical  mixing  over a city,  especially  when  the
airflow  is  naturally  weak.  This,  in  turn, may  affect  the  distribution  of
pollution. Summers1  has  proposed a model  of urban air pollution  based on
the  concept  that, with  a stable  incident airstream, the effective depth of
mixing over a city is essentially determined  by the heat released in the city
from domestic  heating and industrial operations.

In the first instance,  it is useful to note the  magnitudes  of heat transfer that
are likely to  be involved. Summers estimates that, over  the densely built-up
area  of  Montreal,  the heat output from space-heating and industrial activities
                                                                     3-13

-------
is, on  the average, about 4 x  1CT3  cal  cm'2  sec"1 and  points out that in
winter this amounts to  an appreciable  fraction (about  one-fourth) of  the
heat received from the sun. In open country, even with  strong insolation,  the
rate of transfer of heat from the surface by turbulent and convective transfer
is  probably  not  much  larger  than  the  urban  heating  rate quoted  by
Summers-for example, measurements of turbulent  heat flux by Swinbank21
range up to  about 8 x  1CT3 cal cm"2  sec"1.  On  the  other  hand, on clear
nights  the turbulent  transfer of heat to the surface is probably only about
10~3 cal  cm"2 sec"1
It  is probably a reasonable assumption that, in the  absence of unnatural heat
supplies, the turbulent transfer of heat over a city  would be larger than over
open country by day, and of similar or smaller magnitude by night. The  net
effect of also including a fraction (say half) of the unnatural output of heat
might, therefore, range from a  modest  increase in the  (upward)  turbulent
heat flux, H, during  the  day and a marked decrease, or even  reversal, of  the
(downward)  heat flux at  night. A first-order estimate may then  be attempted
by regarding the effect on the  rate of vertical  spread as characterised by  the
length-scale,  L, which is proportional  to u»3/H.   From the  trend of  the
prairie grass  data on  short-range vertical  spread (see Figure 1 of Pasquill12),
it  seems that a  substantial  reduction or  reversal of the downward H (hence
of 1/L)  at night is likely to be  more effective in proportionately  increasing
vertical   spread than  a  corresponding increase  of the upward  H  during  the
day.

The  St.  Louis  data on vertical spread have been classified by  McElroy and
Pooler20  in  terms  of qualitative  stability  categories equivalent  to those
defined  by the writer,2 as  well as in terms of Richardson number and wind
gustiness. The former classification  provides  an opportunity to examine  the
effect of urban heating in the following way.

Figure 3-2 (reproduced from McElroy and Pooler's20  Figure 10) shows  the
mean curves of az against distance, for the  various stability  classes. Here,
classification D, for example, is  not representative of effectively neutral con-
ditions;  rather  it represents the net effect of an incident neutral atmosphere
plus any  effect  of urban  heating. Plotted on the same diagram are two mean
curves for RiB = ±0.01, one reproduced from McElroy and Pooler's  Figure 2,
the other based on  evening data  only.  If  the latter curves  are  taken  to
represent neutral conditions in the urban airflow, then  their departures from
the McElroy and  Pooler  curves  represent the  net  effect of  natural atmos-
pheric thermal  stratification and urban heating. For comparison, we may  use
the 'open-country'  curves of the writer's2 Figure 2, here  referring the curves
to the D(1)  curve as one truly representative of naturally neutral conditions.

Table  3-2 has  been constructed from  Figure 3-2 and  the corresponding
 3-14

-------
      104
      103
  fcS4
      102
          — FIGURE 10 OF MCELROY AND
               POOLER20
          "• RJB = ±0.01 FIGURE 2 OF
               MCELROY AND POOLERZO
          - -RiB -±0.01 EVENING EXPERI
               MENTS ONLY
                                               E-F
102
103
                                           104
105
                               DISTANCE, m
      Figure 3-2.  Vertical  spread from ground-level source in St.
      Louis, Missouri.
Table 3-2.  RATIO OF Oz TO ITS VALUE IN EFFECTIVELY NEUTRAL CONDITIONS,

          FOR OPEN COUNTRY STABILITY CATEGORIES B TO F
Vertical
distance,
km
1

10

Location
St. Louis3
St. Louis5
Open country
St. Louis3
St. Louis
Open country
Ratio of az to value in effectively neutral conditions
for stability categories,
B
4.5
4.0
3.2
9
11
6
C
2.7
2.4
1.9
3.4
4.1
2.4
D
1.7
1.5
1.0
1.0
1.2
1.0
E-F
0.7
0.6
0.5
0.3
0.4
0.3
 a Using McElroy and Pooler's curve for Rig ± 0.01 in their Figure 2.

 b Using data for Rig ± 0.01 in evening conditions only.
                                                                3-15

-------
curves for  open country  to display the apparent effect of natural stability in
the two  cases at distances of  1  and 10 km. The figures in the table are the
ratios of az to  its value  in effectively neutral  conditions. A modification of
the effect  of natural stability  by urban heating  should  result in larger ratios
over a city than over open  country. For the  shorter distance, this effect is
evident consistently at all stabilities,  but, for the longer distance,  it is more
obvious  in the B  and C  than  in the  D and  E-F categories. Thus, the specu-
lation  above that the effect of urban heating  might be expected to be more
obvious  in naturally stable than in unstable conditions does not seem to be
supported  by  the figures,  which,  in fact,  imply a  difference at the larger
distance  and no obvious  difference at the shorter distance. Related  to this, is
the fact  that in Figure 3-2 the slopes of the E-F and D  curves are apparently
quite different from the slopes of  the  curves for  effectively neutral  con-
ditions within  the city. A possible  implication is that the artificial sources of
heat act in two ways. Those producing heat  at low level may lead to en-
hanced mixing  at  low level  (hence,  at short  distances  from the  source  of
material),  whereas those effective  in heating  relatively high layers  (elevated
sources  and rising  plumes)  may  reinforce  the stability that is ultimately
restrictive  as  regards vertical  spread to  greater heights (i.e.,  at longer  dis-
tances from the source of material). Such  a restriction  would not  apply to
material  released with the heat; however  this type of release presumably was
not the case in the tracer experiments under consideration.

As regards categories  E-F and D, the temperature gradient data for St. Louis
(obtained from 40- and  150-m levels on  a  'downtown' television tower and
represented in  RiB) provide additional  evidence. Of  the five E-F  cases, three
have  specified  values of  RiB, which  are +  0.14, — 0.03, and  +  0.12. From
these there is  no strong indication  of a general  elimination  of the stable
classification,  consistent  with  the  indication in Table 3-2 that at the longer
distance  the influence of urban  heating had not brought the vertical spread
in E-F conditions near the magnitude expected  in effectively  neutral condi-
tions.  On  the  other  hand,  for those  occasions when  a D classification is
indicated  consistently  at all  three  of the  meteorological sites  employed
(McElroy and  Pooler's Table 3,20 the values for RiB are in the range -0.05
to  0,  i.e.,  no positive values, suggesting conversion to somewhat unstable
thermal  stratification  in the urban airflow, which is generally consistent  with
Table 3-2.

It would be unwise to attach too much significance to these results alone,
bearing in  mind  the considerable scatter in the data that lead  to the curves
on Figure  3-2,  and the  qualitative nature  of  the stability classification. In
particular,  there  is  some  doubt  about  the  precise comparability of  the
present urban and open country  curves  when so classified.

At this point,  some comment  on Summers' model1  for vertical spread in the
3-16

-------
flow over an  urban area of an  initially stable  atmosphere seems desirable.
The  essential  principle of  this  model  is that  any heat  released  into the
atmosphere  from domestic  and industrial processes is instantaneously mixed
in the vertical, generating a growing layer with  a dry adiabatic temperature
lapse  rate (or constant potential temperature) extending  from the ground.
This layer is identified with the  mixing layer  for pollutants, and the corres-
ponding assumption of uniformity of concentration with height provides an
easily deducible relation between rate of emission of pollutant, wind speed,
and distance alongwind.

Initially Summers adopts linear  profiles of  temperature and  wind, but for
much of the ultimate  analysis  he eliminates the wind variation. The analysis
can easily be  generalized, using power law forms for the height-dependence
of temperature  and wind  and also for the  variation  of  heat  input  with
downwind  distance, x, from the upwind edge  of  the  city. Taking for the
potential temperature distribution
      0(z)  =  azp at  x  < 0
           =  ahp at  x  > 0, 0  < z  < h                            (8)
           =  azp at  x  > 0,      z  > h

for the wind distribution

      u =  bza   =  -^  za                                           (9)

uniform  in  the  horizontal  both  upwind of  and directly on the urban  area,
and for the  surface heat input  from urban processes

      q(x) = cxT  per unit area and unit time                          (10)

conservation of heat requires
                  1            1 + 7
      h oc  U  1  + a + p    x  1 + a + P
Simplification to the case of p = 1,  a = y =  0 gives

      h  oc  Ul-»«   X1/2                                              (12)

as obtained by Summers, while more realistic values, say 0.3 for a and p,
give a result that is only slightly different

      i.e.  h oc Ul-o.6    x0-6                                        (13)

As  a  matter of practical interest, three examples of stable temperature and
wind  profiles observed  near midnight  at  Cardington,  England,  have  been
                                                                     3-17

-------
analysed,  by numerical integration, to get the heat required to convert to
uniform potential  temperature over various depths. The results are displayed
in Figure  3-3. Of  particular interest is the fact that over a considerable range
there is, on  the average, a fairly close fit to a  relation
                UX     "I 1/2
                q(x)dx                                              "4)


which for uniform heat  input  implies h  <*  x'/2,  as predicted by the simple
model.
The heat  input required to produce a value of h similar to the vertical spread
of  tracer  material observed  in  St.  Louis has been considered nominally for
these cases  where conditions of stability prevail  (E-F). If we take the total
vertical spread of the tracer cloud to be, say 2az, this gives, accord ing to
Figure  10  of  McElroy  and  Pooler20, a  value of about 350 m  at  10 km
downwind  from the  source. Assuming uniform  heat  input.  Figure  3-3  in-
dicates that such  a value for h at 10 km would  require 7.4 x 10"3 cal cm~2
sec"1, not  unlike, though  somewhat  larger  than, the figure of 4  x 10~3
estimated by Summers1  for Montreal. Moreover, the  temperature data  al-
ready discussed for  St.  Louis  imply that mixing to  the stage of  uniform
potential  temperature did not extend over the full depth of the tracer-cloud.
In  other words, the extent of vertical mixing is not determined solely by the
urban heat input as  implied  in  Summers'  model but,  as implied  by the
analysis of  urban  roughness, is  probably  controlled to a large extent by the
surface roughness  of the area.

CONCLUSIONS AND  FUTURE  PROSPECTS
In  order  to predict levels of air pollution in any condition  of terrain, one
must distinguish  between two main classes of  meteorological conditions-
those in which  the airflow is relatively well  defined,  with geostrophic wind
speeds, say  > 5 m sec"1  and those in which the wind is calm or very light
and the general pattern of flow is not under any large-scale control. For the
second  class, the foregoing treatments are almost certainly inappropriate and
a special  approach recognizing the local and  individual character of the flow
pattern would appear necessary.  For the  first class, however, the preceding
discussion  suggests that  the 'idealized'  (open-country) treatment  may  be
extended  to urban situations to  give  at  least a  crude specification of the
effect of  increased  roughness  and, to some extent,  the  effect  of  urban
heating on  the general rate of vertical diffusion and the consequent average
level  of  concentration  from an  array  of sources. There are,  of course, still
many deficiencies in the system and it would be unrealistic to expect further
improvements to appear rapidly.
Current  and  foreseeable investigations  of  the  turbulent  properties  and
 3-18

-------
               103
            12
               102
                                                                                             8TH NOV. 1968
                                                                                             26TH DEC. 1968
                                                                                             16TH JAN. 1969
                                                                                             hccql/2
                                                                                       I	I
                  102
                  103
105
                                                   HEAT INPUT, cal  curl sec-l
CO
CD
Figure 3-3.  Heat input required to adjust 2300 temperature profile
(Cardington-England) to uniform potential temperature e(h) through
layer 0-h.

-------
diffusive action  in  the  upper parts of the planetary boundary  layer seem
likely  to  be quite  relevant to the matter  of  vertical  diffusion  at  medium
range over an  urban area. With  increasing  height in the  boundary  layer, it
seems clear that the character of the turbulent motion will have progressively
greater  dependence on  buoyancy effects and  progressively less dependence
on  the  aerodynamic character of the surface. The latter feature  is,  to some
extent, reflected  in the  theoretical  results discussed  earlier,  in which  the
proportionate effect of  a  change of roughness on Z decreases  with  distance
of travel. When vertical spread is large, whether as a  result of instability or
of long-distance travel, observations of turbulence at the  greater  heights  will
be particularly meaningful.

The continuous  evolution of  the boundary layer in  response to changing
surface roughness and surface heating is a  matter attracting increasing atten-
tion in meteorological  studies. Such studies should bring further insight both
into the process  of modification  of  an  airstream as it  leaves rural  sur-
roundings and  flows over an urban  area and  into the analogy of the modifi-
cation of an  initially stable airflow by an urban heat input with the temporal
modification of the nocturnal boundary layer by  the sun's heat  early in the
morning. Except for complications that may arise from the particular varia-
tions of the  heat input in time or space, and from the effects  of wind shear
in the advective  case, we may take  -—- as  equivalent to u -j^. Reference to
existing information on the diurnal  development  of the vertical temperature
profile  may well be rewarding, but it is also evident that special observations
on  these evolutionary  aspects  are desirable.  Examples  that spring  to mind
immediately  are  the post-dawn evolution of  the  turbulence profile and the
formation of an evening 'heat plume' over a large  urban complex.

It may appear that much could  also be done directly to  improve our know-
ledge of the turbulent  and diffusive structure over an  urban area. Amid the
complexities that are inevitably  present, the problem  is  to  isolate  the con-
trolling features  and subject them to theoretical  analysis  and  critical obser-
vation.  For certain  purposes, tracer experiments have an immediate appeal in
that they  provide  a direct answer to the ultimate questions, but they may
involve expensive repetitions unless  supported and generalized  by a rational
theoretical approach. Thus  it would seem  no more than common  sense to
ensure, by examinations of the kind described  previously,  that the maximum
information  is extracted  from a  preliminary experiment,  such as  that con-
ducted  in  St.  Louis, before  planning and  embarking on  repetitions or ex-
tensions.  One  of the outstanding prerequisites for the  theoretical  specifica-
tion of the rate  of vertical  mixing is the provision of  characteristic  profiles
of the  intensity and scale of turbulence. For the effective  vertical diffusivity,
we recall that such a  specification  is provided by data on any  two of the
quantities aw, e, and Xm,  and  it would clearly be  a step of major importance
to extend  to the higher parts of the boundary layer the generalizations  that
3-20

-------
are now available for these same quantities in  the lower layers. The experi-
mental program now evolving at the Cardington station of the Meteorological
Office is expected to provide such data, ultimately. It will be  characteristic
of open country with  variable  but  relatively minor roughness  from both
natural  and urban features. It would, of course,  be valuable if  at least some
of the simpler observations were  to be carried out over rougher  terrain and
specifically  over a  major urban  complex.  Measurements of  the relatively
high-frequency contribution  to the intensity  of  turbulence,  from which  to
derive estimates of e, seem particularly worthy of consideration.

In the  meantime, the attention  of the Meteorological Office is focused on
the exercise of making the most realistic estimates of the K profile, commen-
surate with  the  limited theoretical  understanding and  observational data
currently available on the effects of natural thermal stratification  and rough-
ness.  The intention  is then to examine the consistency of these profiles with
the independent  data available on the vertical spread  of windborne material
over  medium  (10  km)  and  long distances (100  km). This exercise  is an
essential part  of  the  previously  mentioned revision  and extension of the
practical system for estimating pollutant concentration, and will be reported
in detail elsewhere.
Most of this discussion  has been  concerned with the  general  rate of vertical
mixing  in  the atmosphere and  as such  is meaningful  only in  relation  to
spatial  integrals  of  pollutant  concentration.  As  regards the  detail of the
variation of concentration in space  and time,  it is  obvious,  even from  a
consideration of relatively ideal  conditions of diffusion, that  there are  ad-
ditional  limitations  to  be faced.  From general  estimates  such  as  those
provided by the  writer,2 it is easily demonstrated that the estimated concen-
tration  arising  from  a point source, at a given  position relative to the source,
may change by several-fold as a  consequence  of  a 5-  to 10-degree change in
the assumed wind direction,  especially in  neutral and stable conditions. It
would obviously be unrealistic to expect any  better  precision  in specifying
the  effective wind  direction from the  observations  available  in an  urban
experiment such as  that reported by  Marsh and  Withers.6 From  this uncer-
tainty alone, it should not be surprising that the  individual 6-hour dosages in
that experiment  deviated  widely from  the  corresponding calculated values.
Discrepancies introduced in  this way and from other uncertainties, notably
those in the magnitude of the spread and in the strengths of the sources, can
be reasonably  expected  to contain  a  large  random element. It  would  be  of
interest to  examine whether  there is, nevertheless, some measure of  agree-
ment in the statistical  distributions of pollutant concentration as measured
and  as  calculated  (i.e.,  the  frequency of  occurrence of various  levels  of
concentration in  space or time).  It may be significant  in this connection that
in Marsh and  Withers'  analysis6   (Table  3-1) the r.m.s.  differences between
calculated  and  observed  values  are  not  greatly  in  excess  of  the  r.m.s.
                                                                     3-21

-------
variations of  the observations themselves (as represented  in the simple regres-
sion  analysis).  A  more direct examination  of  the compatibility  of the
distributions  would  be desirable,  however, and  if successful, would be an
encouraging  pointer to the prospects of predicting  more than  the  average
level of pollutant concentration.

ADDENDUM
Since this  paper  was completed,  Mr. Marsh has followed up the suggestion
made in the last paragraph  and  has derived frequency distributions of observed
and  calculated  6-hour concentrations for  three  of the  sites in the Reading
experiment. I am  grateful  to Mr.  Marsh for his permission to reproduce the
combined data here, in the form of a plot on a log-normal scales (Figure 3-4).
6-hr CONCENTRATION, jug/m3
i—
i— • r*o c_n G
-- r-o en o o <=> c
Z5 C3 O o CZ5 O C




	




	


OBSE
(CORREC
BACK
	 .__


IRVED
;TEDF
3ROUN
V
N
j
CALCULATED
(BROAD ESTIMATES

OR
°) .
4
A
-*
OF

//
1
i
SPRE/i
\
k
/

ID) >
'* ^>.
X \
CAL(
^

v
:ULAT
— —
**»>
ED
(WIND
FLUCTUATIONS)
	
	

             0.1     1       10    30   50  70    90      99
                       FREQUENCY DISTRIBUTION, percent
99.9 99.99
   Figure 3-4.  Marsh's  reanalysis  of  Marsh and Wither's data.6
   Reading sites 1, 5, and 11; all  periods-
The observed  and calculated  distributions  are  roughly in  agreement and
conform fairly well to a  log-normal shape over a considerable range. The fact
that, over  most of the  range, the broad estimates of spread provide better
agreement  must be fortuitous.  Furthermore - and probably most  important
- it appears that the calculated distribution provided a statistical estimate  of
the incidence of the highest concentrations (say  the  concentration exceeded
on  a  few percent of occasions), which  would  be satisfactory  for  many
3-22

-------
practical purposes. Obviously, a great deal of further work and discussion are
required on  this  point, and  the matter is, of course, taken  further in  Dr.
Fortak's paper. It seems to me, however,  that there is already some reason
to anticipate that a  useful  statistical  prediction  of the probability distri-
bution  of pollutant concentration in an urban area  is realizable in  practice.
                                                                      3-23

-------
REFERENCES

 1  Summers  P  W. An  Urban Heat Island  Model; Its Role in Air Pollution Problems
    with Applications to  Montreal  (abstract only). In: Proceedings 1st Canadian Con-
    ference Micrometeorology. 1967. 435 p.
 2. Pasquill,  F. The  Estimation  of the Dispersion of  Windborne  Material.  Meteorol.
    Mag. 30(10631:33-49, February  1961.
 3. Pasquill,  F. Meteorological Aspects of the Spread of Windborne Contaminants. In:
    Seminar  on the Protection  of the Public  in the  Event  of Radiation Accidents,
    Geneva,  November   18-22,  1963.  Geneva,  World  Health  Organization, 1965. p.
    43-53.
 4. Marsh, K.  J. and M. D. Foster. An Experimental Study  of the Dispersion  of the
    Emissions from Chimneys. In Reading - I   The Study of  Long-Term Average Con-
    centrations of Sulphur Dioxide. Atmos.  Environ. 7(5) :527-550, September 1967.
 5. Marsh, K.  J.,  K. A. Bishop and  M.  D. Foster. An  Experimental  Study  of the
    Dispersion  of  Emissions from Chimneys. In Reading — II. A Description  of the
    Instruments Used. Atmos. Environ.  7(5) :551-559, September 1967.
 6. Marsh, K.  J. and V.  R. Withers. An Experimental Study  of the Dispersion  of the
    Emissions from Chimneys. In  Reading  — III. The  Investigation of Dispersion  Cal-
    culations. Atmos. Environ.  3(3) :281-302, May 1969.
 7. The  Calculation of  Atmospheric Dispersion from  a Stack. Stichting  CONCAWE.
    The Hague, Netherlads. August  1966. 57 p.
 8. Hogstrom,  U.  An   Experimental  Study on  Atmospheric Diffusion.  Tellus.  16
    (2).205-251, May 1.964.
 9. Pasquill,  F. The Vertical Component of Turbulence at Heights up to  1200 Metres.
    Atmos. Environ.  7(4) :441-450, July 1967.
10. Busch, N.  E. and  H. A. Panofsky. Recent Spectra of Atmospheric  Turbulence.
    Quart. J. Roy. Meteorol. Soc. 94(4001:132-148, April 1968.
11. Readings, C.  J. and D.  R. Rayment. The High-Frequency Fluctuation  of the Wind
    in the First Kilometre of the Atmosphere (in press). Radio Science. 1969.
12. Pasquill,  F  Lagrangian Similarity and Vertical Diffusion from a Source at Ground
    Level. Quart J.  Roy Meteorol. Soc.  52(3921:185-195, April  1966.
13. Chatwin, P  C.  Dispersion of a  Puff of  Passive Contaminant in the Constant Stress
    Region. Quart. J. Roy. Meteorol. 34(4011:350-360, July 1968.
14. Pasquill,  F. Atmospheric Diffusion.  London, D. Van Nostrand  Co., Ltd., 1962. 297
    p.
15. Elliott, W.  P  The Vertical Diffusion of  Gas from a  Continuous Source. Int. J. Air
    Water Pollution. 4(1/2) :33-46, June 1961.
16. Gifford,  F. A.  Diffusion  in the  Diabetic  Surface  Layer.  J. Geophys.  Research.
    67(8):3207-3212, July 1962.
17. Davenport, A. G. The Spectrum of  Horizontal Gustiness Near the Ground  in High
    Winds. Quart. J. R. Meteor. Soc. 87: 194. 1961.
18. Davenport, A.  G. Instrumentation  and  Measurements of Wind Speed Spectra  in  a
    City.  Proceedings of  the First Canadian  Conference on  Micro-meteorology. Part III.
    Department of Transport. Toronto, Canada. 1967. 361 p.
19. Marsh, K.  J. Measurements  of  Air  Turbulence in Reading  and Their Relation to
    Turner's Stability Categories  (in  press). 1969.
20. McElroy, J. L.  and  F Pooler, Jr. St. Louis Dispersion Study, Volume  II - Analysis.
    National  Air  Pollution Control  Administration. Arlington,  Va. Publication Number
    AP-53. December 1968.  51 p.
21. Swinbank,  W.  C. The  Exponential Wind Profile  Quart  J  Roy  Meteorol  Soc
    90(3841:119-135, April 1964.
 3-24

-------
ACKNOWLEDGMENT

This  paper is published with the permission of the Director-General of the
United Kingdom Meteorological Office.
                                                             3-25

-------
             APPENDIX - GLOSSARY OF SYMBOLS


b, c    dimensionless constants in the  Lagrangian similarity treatment.

k      von Karman's constant (0.4).

H      vertical eddy flux of heat.

K      eddy  diffusivity;  subscripts  M  and H  refer to momentum and heat,
              respectively.
                                     pCp T  uj
L      Monin-Obukhov length scale, —   ^ u

q      rate of heat input from urban  processes.

Ri     Richardson  number; subscript  B denotes 'bulk' form  using tempera-
              ture difference over specified height interval.

S(n)   spectral density of wind component variance at frequency n.

tL,tE  Lagrangian and Eulerian integral time scales.

T      time of travel from source; absolute temperature.

IT      mean wind speed  in x (alongwind) direction.

u*     friction velocity.

X     mean distance of  travel of particles in  a given time.

Z      mean displacement of particles  in vertical (z) direction.

z0     roughness parameter in wind profile.

e      rate of dissipation of turbulent  kinetic  energy per unit  mass.

Xm     equivalent wavelength ( = U7nm) at which nS(n) is maximum.

p      air density.

a      standard  deviation;  subscript w refers to vertical component of wind;
              y and z, to particle spread in respective directions.


-------
ABSTRACT
      The  basic  structure  of many  current  urban  pollution  models is
      examined  from  the  point  of  view of underlying assumptions and
      physical basis. Initial attempts to extend steady-state models to vari-
      able  conditions and long-term  predictions are indicated,  with a brief
      discussion  of stochastic  simulation of  the concentration  patterns and
      average seasonal and annual  distributions. In view of the complexity of
      most computer-oriented models, a sensitivity analysis is recommended
      to identify the input parameters that most critically affect the concen-
      tration predictions.
AUTHOR
      KENNETH L. CALDER obtained honors degrees in physics and mathematics at
      the  University of London before joining the British Meteorological Office at its
      Porton Research Establishment in  1936.  He left Proton in 1949 to accept an
      appointment  with  the U. S.  Army Chemical Corps as  Chief of the Meteorology
      Research  Division  at Fort Detrick. In  1968  Mr.  Ca/der transferred  to  the Air
      Resources Laboratory of NOAA where  he currently serves as Chief Scientist to
      the  Division  of Meteorology of the  National Air  Pollution  Control Adminis-
      tration.

-------
      4.   SOME MISCELLANEOUS  ASPECTS  OF CURRENT
                                   URBAN  POLLUTION MODELS


                      KENNETH  L. CALDER
               Air Pollution Control Office, NOAA*
INTRODUCTION
The  last  decade has  witnessed an unprecedented interest  and  concern  with
the development of mathematical-physical models for predicting air pollution
concentrations in an  urban environment. It is not surprising that, in  most of
these efforts, we see  the  underlying influence of the pioneering thoughts of
Francois  Frenkiel on  the  subject, and  it is fitting that he  can  be with us at
this  symposium.  The  ever  increasing  current  interest in urban pollution
modeling derives, of  course, from  its  obvious  potential value in nearly  all
practical  problems involving quantitative consideration  of air quality relative
to the sources of pollution. These include air pollution abatement strategies,
the forecasting of the  occurrence of undesirable levels of urban pollution,
and  problems  of urban planning for the future. Each one  of these topics
constitutes  a  large well-recognized area of  study  that is being  vigorously
investigated by many groups,  although these  matters lie outside the scope of
the present symposium.

As  a newcomer  to  the  field of  urban pollution modeling  who  was at-
tempting, with no little  trepidation,  to select an appropriate topic  for  a
paper to be  presented before experts,  I was  primarily influenced  by two
considerations.  First,  was the knowledge that the fundamental physical and
meteorological aspects of  the  subject would certainly be covered at the  sym-
posium in an authoritative fashion  by  Drs. Lettau  and Pasquill. Second, was
the fact  that  several  quite  comprehensive general  papers and  bibliographies
on urban diffusion modeling have recently been written. Here it will suffice
to mention the  article by Wanta in Stern's book1 and two comprehensive
review articles prepared this year by Turner2 and Moses.3  With these recent
articles, still  another review seemed out of place  for this meeting. Under
these circumstances,  it seemed more worthwhile to mention  miscellaneous
questions that come  to mind in  reading accounts  of some of  the currently
'National Oceanic and Atmospheric Administration, U. S. Department of Commerce.
                                4-1

-------
available models-and, more particularly,  those developed  in the  past few
years by members of a  U. S. air pollution meteorological team  under the
direction of Mr. R. A. McCormick.  The somewhat fragmentary observations
contained in  this paper are made at a time when a rapidly  expanding effort
in modeling  gives optimism  for  expecting a second  generation of  improved
models  to  become available in the  near future. While it is  appreciated that
the present paper  raises  a  lot more questions than it answers,  it is certainly
not  done with the feeling that the  problem is  so hopelessly complicated  as
to defy useful analysis.  Instead,  it is believed that an explicit recognition  of
some of the  difficulties  will lead to more mature appreciation of the limita-
tions of modeling  techniques among  those who may be wanting to apply
them. This recognition of  the difficulties  is also a prerequisite to the devel-
opment of improved  urban pollution models.

A MULTIPLE-SOURCE URBAN POLLUTION MODEL
In the  following discussion, we  shall be  concerned  almost exclusively with
modeling the effects  of  a chemically stable gaseous pollutant that is released
from a  stationary  source  in a city. Sulphur dioxide that  results  from the
combustion of sulphur-containing fossil fuels  in domestic,  commercial, and
industrial  operations  is normally regarded as the classical example. We shall
also  restrict  attention to physically based  models of the pollution distribu-
tion. These  are explicitly  constrained by the  law  of the conservation  of
matter,  in  contrast to empirical  models that are based on  statistical correla-
tions established  between observed concentrations of pollutant  and  simul-
taneously  observed  meteorological parameters, and that  do  not  involve
considerations of material balance.

Steady-State Equation

The  starting  point  of the majority of urban pollution models is the postula-
tion  of  a quasi-steady state. Thus, in  spite of the  obvious long-period varia-
bility of diffusion  and  meteorological conditions  over  an  urban  area, it is
assumed that this variability can  be  treated as though it  resulted from a
sequence of  different steady-state situations. The  sequence interval may  be
some relatively short-term  period—perhaps  only of the order of  an hour.
During  each  "steady-state period," it  is further assumed that  a unique hori-
zontal mean  wind  direction can  be defined  for the entire urban  area, and
that  if  this is used to define the x-axis of a rectangular coordinate system,
with y-axis also horizontal, and  z-axis vertical,  then  the  mean concentration
at a receptor location at  ground level (x, y, 0) resulting  from an elevated
continuous point source  (0, O, z0), release of pollutant, of constant strength
Q, per unit time has the  function form Q R (x,  y; z0)

For  a "given" meteorological situation, R  denotes a fixed, i.e., nonstochastic,
4-2

-------
function, independent of the location of the origin of the coordinate system,
i.e., is the same irrespective of the horizontal location of the source. Further-
more, R does not depend on the horizontal mean  wind direction that defines
the orientation  of the coordinate  system relative to the urban  area. The
assumption  that  the form of R  is independent  of the source location,  is
tantamount to the more formal statement that the concentration distribution
function is invariant under arbitrary  horizontal  translations of the source-
receptor pair.  This might  conceivably  be a reasonable approximation over a
very extensive  area  of  uniform, level, open  countryside. Considering, how-
ever, the edge effects that inevitably are associated with the relatively small,
finite size  of  all cities, and the rather obvious horizontally inhomogeneous
nature of the buildings and their distribution, the  above assumption can only
be  a  very  crude  approximation  for  an urban diffusion  environment. The
same  comment would seem to apply with equal force to the assumption that
the form of R is unaffected by the direction of the airflow over the city.

The specific functional form of R(x, y; z0) that  is used in  the so-called
Gaussian plume dispersion  models is, for x > o,
                                     a(x,S)   +  a(x,S)
        R(x,y;z0)   =  - - - - -         (1)
                             TT  U  av (x,S) az(x,S)

        with R = 0 for x < 0, i.e., no upwind travel of material,
            U = a representative, constant, and spatially uniform, wind
                speed;
            S = atmospheric  stability category first introduced by Pasquill.4
                Here S =  1  may denote  extremely  unstablity and  S = 6,
                extreme stablity, and
cry(x,S),az(x,S) = horizontal and  vertical  diffusion functions— the horizontal
                and vertical  standard deviations  of the Gaussian concentra-
                tion distribution at distance x from the source. These dif-
                fusion functions are  normally assumed to be parameterized
                by the stability  category  and are assumed  independent of
                release height.5

In addition, in some models, a  crude attempt may be made to allow  for the
known fact that diffusion in the vertical direction is frequently limited, such as
when the pollutant  is trapped between the ground surface and a stable layer
aloft. When this occurs, a uniform vertical distribution of concentration would
be expected throughout the effective mixing layer for sufficiently large values
of x. One simple method of approximately  allowing for this was suggested by
Pasquill.6  If the mixing layer is of height  L, then  Equation (1) is used for
downwind  distances, x, such that  az(x,S) < VzL. When crz(x,S)  =  L, the
                                                                      4-3

-------
resulting concentration becomes approximately uniform with height and of a
magnitude appropriate to az = L. At greater distances than this, the ground-
level concentration is  no longer reduced by vertical diffusion and  the  equa-
tion can  continue  to  be used with CTZ  = L = constant and with the appro-
priate value of Oy (x, S).
In the original diffusion  prediction scheme based on the atmospheric stability
categories suggested by Pasquill,4'6 the diffusion functions ay{x,S) and az(x,S)
were  tentatively established  from  limited  experimental data  on  diffusion
over open level  country. There was thus no  allowance  for the possible dis-
turbing  effects of buildings and topography. With some minor modifications,
Turner5  has utilized these functions in an urban diffusion model and has
indicated both a working scheme for objective determination of the stability
category and  a  crude  qualitative method of adjusting the values to allow for
urban conditions. More  recently, the results of ad hoc experimental diffusion
studies  made  with tracer material in metropolitan St. Louis, Missouri, have
been reported by  McElroy and Pooler,7 and McElroy.8 These studies have
shown  that, for  low-level  point  sources, the crosswind diffusion in  cities,
although  initially greater than in open  country, soon converges to the  latter
value. Significantly greater  vertical diffusion occurs in the city  than in open
country;  and  it  is  most pronounced  under  stable meteorological conditions.
As is to be  expected,  however, many quantitative  aspects of the  urban
point-source diffusion problem remain to be  resolved, although  it  is encour-
aging that  these recent urban experiments  have generally  confirmed the
appropriateness of at least some aspects of the simple Gaussian model.

Area Source Application
With all the provisions mentioned in the preceding discussion, it is now only
a small  step to a formal  mathematical formulation of a multiple-source urban
pollution model. For any  "steady-state period," choose  the  x-axis of the
rectangular  coordinate system  along  the horizontal mean wind  direction for
the urban area. Furthermore, let the urban pollution emission be regarded  as
a  steady, horizontally distributed, area source such that Q(x0, y0, z0) dx0
dy0 is the total  amount of pollutant emitted  per unit time at an "effective"
height,  z0 from  a horizontal element of area  dx0 dy0 surrounding the point
(x0, Yo>- 't is implied here that the function Q(x0, y0, z0) can be specified
on the   basis  of appropriate emission  inventory techniques for the entire
urban area,  and will have the value  of zero outside this area. Then, assuming
that at  any ground-level receptor location (x, y), the total pollutant concen-
tration  is a  simple additive function  of  the concentration contributions from
all the individual elements of area, it is evident that
                  /OO    -1X3

                        I    a(Xo,y0.Zo>R(x-x0,y-yo;z0)dx0dy0      (2)
                  •oo  J -oo
4-4

-------
The  area source concept has been justified as an appropriate  idealization in
urban  diffusion  modeling  when  there are too  many sources  (e.g., from
domestic heating units) to be considered individually. It has sometimes been
used, however,  (e.g., Turner5) even  when larger commercial  and industrial
sources are included in the emission  inventory. For the primary  purpose of
reducing the computational effort,  a single effective height of  emission is
sometimes  selected,  also.  More  realistically,  large pollution  emitters,  like
power plants with tall stacks, are treated as individual  elevated point sources,
and  the effective height is calculated from the actual height plus the plume
rise  associated  with  upward momentum of discharge  and  thermal buoyancy
effects. Consideration of plume rise is a somewhat controversial topic and far
from straightforward, although it becomes an  important element in realistic
pollution models. A comprehensive and, hopefully, definitive, critical  review
of the subject  has  recently been prepared by Briggs.9  [We  note that  the
point  source situation  may be regarded as covered by Equation  (2)  by  the
formal  device  of  the  Dirac delta  function. Thus, for a continuous point
source  at  height  z0,  of strength, q,  per unit time, and  with horizontal
coordinates (a,b), we have  Q (x0, yo, z0) = q 5(x0 — a) 5 (y0 — b) so that
Equation (2) gives x = q R(x—a, y-b; z0).
The  integral of  Equation (2) has to  be evaluated by  numerical methods, of
course, once the source function Q (x0, yo, ZQ) ar|d  the  diffusion function
R (x  — x0, y —  yo; ZQ) have  been specified. The calculations that  are
involved are normally  so extensive  that they necessitate use of a high-speed
computer.  Various and  somewhat  arbitrary devices  have  been used  in  the
past for the approximate numerical  integration, with no proper study of  the
errors incurred. For example, Pooler,10 in developing  a multiple-source S02-
prediction  model,  used a source inventory for Nashville, Tennessee, based on
1-mile-square area sources;  then  replaced  each square (with  a single  excep-
tion) with  a  point source at its center. Again, Turner5 assumed that  1-mile-
square  area  sources  could  be  approximated by  appropriate  Gaussian-
distributed, line sources oriented crosswind and passing through the center of
each  area.  These approximations appear to be very crude and consideration
of more sophisticated  integration techniques  should  be worthwhile  in  the
future.

It is, perhaps,  obvious that an  urban  pollution model of the type  we  are
considering, which  requires as a quantitative  input, a source-strength func-
tion, Q (x0, y0, z0), cannot yield concentration predictions  that are more
accurate than the  pollutant source  inventory itself. Unfortunately, the latter
may  be quite  crude  and  is  frequently  based on indirect  reasoning. The
accuracy may also  be expected  to  diminish rapidly as the time period over
which the  emissions are averaged  becomes shorter.  Sulphur dioxide, pro-
duced by the burning of coal and oil  for space heating, for example, requires
indirect estimation  (see, e.g., Turner11)  of diurnal and day-to-day variations
                                                                      4-5

-------
of fuel consumption.  The  estimations are made  by using such empirical
concepts as "heating-degree-days" (the  depression of a suitably defined daily
temperature below some  limit like 65° F). While an  annual emission inven-
tory derived from the total  annual  consumption  where  fuel  has a known
sulphur content may be quite accurate, large errors can be expected in daily
and hourly  inventories. These are probably unavoidable due primarily to the
fact that,  in  general,  no adequate records of fuel consumption with fine
enough resolution  in  time and  space  are  available. This fact, itself,  could
impose an intrinsic limitation on the accuracy  of urban pollution models as
applied to short-period  predictions.

EXTENSION OF MODEL TO VARIABLE CONDITIONS
AND  LONG TIME PERIODS
The basic model  of  the  preceding discussion  relates to  meteorological and
emission  conditions  that can be  regarded as steady-state and  that  have
prevailed  for a sufficiently long period  of time to permit full development of
the concentration fields associated  with the individual sources. In considering
extension of  the model  to  variable conditions, we will assume  that the
steady-state assumption is normally  an adequate working approximation for
time periods of the order of an hour, and  that the general nonsteady-state or
variable situation  can  be  resolved  into a  succession of steady-states of this
duration.  In general, both the pollutant source distribution and the meteoro-
logical conditions must be regarded as varying  between consecutive periods.

Quasi-Steady-State  Applications

Since  the x-axis for Equation (2) is determined by the mean urban wind
direction, while the  source distribution is  specified in  a coordinate system
that is fixed relative to the  city, and is also desirable for the concentration
distribution, it will be  necessary  in utilizing Equation (2) to use the standard
formulas  for  rotation  of coordinate axes.  In  addition  to  the mean  urban
wind  direction, the change  in meteorological  conditions occurring between
consecutive 1-hour periods should reflect changes of the mean wind speed,
U,  the stability category, S, and the height,  L, of  the mixing  layer, because
these all will affect the concentration function  R(x, y; z0). It  is then evident
that by repeated applications of Equation  (2), each with an appropriate set
of  hourly parameter values,  and an emission distribution estimate; we can
simulate the pollutant  concentration changes occurring  at any  receptor loca-
tion over  any  specified  period of time.

Precisely  one  such urban  pollution model  was developed  by Turner5 for the
sulphur dioxide pollution in Nashville, Tennesee. This analysis was based on
an  emission inventory  prepared  for  mile-square areas forming a rectangular
array  17  by 16 miles.  Source strengths and meteorological  parameters were
then estimated by 2-hour periods (instead of the 1-hour  period above). The
4-6

-------
mean wind direction, to 16 compass points, and wind  speed, to 0.1 meters
per second  (m sec"1), were estimated from instrumental  records at several
wind stations in the area. The corresponding 2-hour stability category, S, was
obtained by using an objective technique proposed by Turner (in this model,
no allowance was made  for a finite mixing height, i.e.,  L  = °°). In  terms of
these inputs to  the  model, 2-hour  SO2  concentration  values  were  first
estimated at a number  of  receptor locations at 1-mile intervals. From the
2-hour  values,  average  24-hour concentrations were  then obtained. Com-
parison  of the calculated 24-hour concentration values with those obtained
by  actual  chemical  sampling during a  number of  test  periods occurring
throughout  a  1-year period  showed  quite  reasonable  agreement.  In  fact.
Turner showed that even for 2-hour periods the calculated values were in fair
agreement with corresponding observed values.

It is evident  that, in principle,  the above approach could be used to provide
elaborate information on the characteristics of the concentration distribution
in terms of forecast or observed meteorological and emission  variations.  If
frequency statistics are available for the relevant meteorological parameters,
then a  probabilistic model  is possible and the  actual frequency distribution
of the pollution  concentration at a receptor can be estimated.  (It should be
noted  that  this  is  strictly true  only  if  temporal variations of the source
strength function,  i.e.,  Q  (x0,  y<>,  z0), of Equation  (2), are disregarded.
Otherwise, it would be necessary to know the joint frequency function for Q
and the meteorological parameters.)

Thus, after calculating the  concentration values for all possible combinations
of the  three  meteorological  parameters: wind direction, wind speed,  and
stability category,  and  for simplicity  assuming L = °°,  i.e., diffusion is un-
limited  in the vertical direction, we can  get  the frequency of occurrence  of a
particular concentration  value from the sum of the joint  frequencies of all
the combinations of meteorological parameters that give rise to this concen-
tration value. If hourly observations of wind direction are rounded off to the
nearest  10 degrees,  and  there are, say, ten classifications of wind speed, then
with only  five possible  stability categories, we  obtain 36  x 10x5= 1800
possible  combinations;  hence, 1800 meteorological situations for which the
integral of Equation (2) has to be evaluated numerically  for each  receptor
location  of interest.
Since each of  the  meteorological  "parameters"  in reality has a continuous
distribution, this use of only a finite number of combinations will, of course,
mean that the real ensemble of  diffusion  situations is only being  approxi-
mated   in  the  mathematical simulation. Such extensive  computation  can
quickly  involve  prohibitive time,  even  with present high-speed computers.
Nevertheless, extensive  calculations  designed  to characterize  the concen-
tration  frequency distribution relating specifically  to  ground level  S02 for
the  urban  area  of Bremen,  Germany,  have  already been  reported  by
                                                                       4-7

-------
Professor  Heinz  Fortak.12  We shall probably hear the  latest  news of  this
pioneering effort  later at this symposium. It  might be noted that  the same
approach  could  be used to  provide probabilistic  information  on  the  two-
dimentional concentration  distribution, where the joint  frequency distribu-
tion for concentrations at  two  or more  receptor  locations could  be deter-
mined by the same enumerative procedure.
On the basis of extensive air  pollutant concentration data for  several pollu-
tants and  several  U. S. cities,  Larsen13  has  been able to show that the
empirical  frequency  distribution of concentrations is  closely approximated
by the logarithmic-normal  distribution. With  the possibility discussed above
for stochastic simulation, it would be  of considerable interest to see whether
the  log-normal  distribution can  be generated by  appropriate  mathematical
simulation.

A less  ambitious objective than the determination  of a concentration  fre-
quency distribution  is that of determining just the long-term  average value
over some specified period—usually a  month, season, or  year. Although  this
can obviously be obtained  in  calculations of  the type just considered, much
simplification is  possible when  interest  is restricted to  average values.  This
simplification is achieved by a change in the order of integration and summa-
tion.  In the previous consideration of an aggregate of 1800 possible meteoro-
logical  situations corresponding to wind directions 0, (i = 1,2,3,. . .36), wind
speeds, Uj (j  = 1,2,3,. . .10), and  stability categories Sk(k = 1,2,3.. .5), let
/(i, j, k) denote the corresponding joint frequency function. Also let C(i, j, k)
now  denote the  concentration  for a  particular meteorological situation as
given by the integral  of Equation (2) for  a given receptor location.  Then the
long-term average concentration,  say C, is evidently  given by the triple sum

        C =  2  2 2  C(i,j,k)/(i,j,k)                                 (3)
               i   j   k
and would involve 1800 numerical evaluations of the integral. (Providing the
source-strength function and  the  meteorological  conditions are independent,
the  average  concentration  is calculated  from  an average value  of  source
strength.) Consider now an elementary source of unit area at (x0, y0), that
for a particular  meteorological situation   (i,  j, k)  produces a concentration
G(x0, y0; i, j, k) at some given  receptor  location. Then, for this source, the
average concentration, say G (x0, y0),  at the  receptor will be given by

             "G(x0,y0>  =  222 G(x0,y0;i,j,k)f (i,j,k)
                           i   j    k                                    (4)
Evidently
                   ° J      J     G
-------
1800 terms, considerable simplification arises from the physically obvious fact
that, for  a given source-receptor  pair, it  is only a  limited range of  wind
directions that really contributes to the concentration at the  receptor. For
example,  Meade  and Pasquill,14  in their  paper analyzing  the average dis-
tribution  of  SO2 pollution  around Staythorpe  power  station in  England,
have shown  that, with  the assumption of a  Gaussian type concentration
distribution given by Equation (1), the average concentration at a  receptor,
produced  by  a  fixed-point  source and corresponding  to all possible  wind
directions, can be approximated  by a simple analytic formula. This actually
involves only  the wind direction frequency for  the  single  wind sector (10
degrees in our example) that contains the  source-receptor direction.  In this
case, the  sum in  Equation (4)  would reduce to a double sum extending over
all  combinations of  wind speed and  stability category and would  involve
only 10 x 5 = 50 terms.

Multiple-source urban pollution models that essentially utilize this simplifica-
tion and  provide estimates  of long-term average concentrations have  been
proposed  by Pooler10 and, more recently,  Martin and Tikvart.1 s The Pooler
model  mentioned previously for  the average monthly SO2 concentrations at
Nashville, Tennessee, is an example for which sampling data from an exten-
sive air   pollution   study  were  available.  Pooler  disregarded  atmospheric
stability  changes and, in  effect, considered only a single stability category.
The meteorological frequency  function only involved consideration of  wind
direction  and wind  speed  and  was obtained  from monthly summaries of
hourly observations  at the airport.  In addition, the Pooler model assumed  no
limitation of  vertical diffusion,  i.e., the mixing  height, L = °°. The source
inventory was based on  1-mile-square areas; then, for each receptor location
in the numerical  integration  over the entire urban area, each square  (with a
single  exceptional case)  was replaced  by  a point  source  at  its center. A
reasonably close  agreement  was  obtained  between  calculated  and  observed
patterns of average monthly SO2  concentrations.

The model proposed by Martin and Tikvart15  differs only  in its details and
actually incorporates the simple  Meade-Pasquill averaging formula. The mete-
orological  frequency function  involved the  additional variable  of stability
category,  S,  and was thus established  in terms of an  array of 480 possible
meteorological combinations from  the  long-term  records of  hourly meteoro-
logical observations  that  are  normally available from  Weather  Bureau sta-
tions. A  difficulty arising here is that  the  long-term records refer to  airport
locations  and not to the  nearby urban area that is of primary interest. In
addition,  a finite mixing  height,  L, was  assumed,  and a  crude procedure
proposed  in an attempt to reflect  the  functions  major  diurnal  changes. The
calculations for  both the  above models become so extensive  that they  re-
quire the use of a  high-speed computer. This is so  in spite of the drastic
mathematical  approximations that  have  been  adopted  in order to simplify
numerical evaluation of the surface integrals involved.

                                                                      4-9

-------
Instantaneous Gaussian Puffs Approach
The  preceding models  are all based  on the postulated equivalence of the
general  long-term  variable situation to  a  sequence of different quasi-steady
states. This assumption allows model  development in terms of the relatively
simple steady-state concentration formula of  Equation  (1). To  more accu-
rately  portray the general  nonsteady-state situation, it will be necessary to
regard the pollution emission from any point source as resulting  in a series
of "instantaneous  puffs" that may  follow different trajectories  and  have
different  rates of diffusion. Such  a basis for  a more general  type of urban
diffusion  model  in cases where  meteorological  conditions  could  not be re-
garded  as constant and  uniform  for the whole  urban area  was  explicitly
indicated  by  Frenkiel16  in his now classic paper. Such a modeling procedure
implies the ability to specify separately the trajectory in space of  every puff
center and the puff  concentration distribution at any time  relative to the
puff center. A sophisticated  urban pollution model of this character, based
on  the assumption of  a  three-dimensional  Gaussian  distribution of concen-
tration  in the individual  instantaneous puffs,  was recently  proposed by the
late  Professor Ben Davidson17 and his associates  at New  York University.
The  Gaussian puff approach has also  been applied by Start and Markee18 to
the calculation of average  concentrations resulting from  a  prolonged point-
source release of pollution  in a valley where the winds underwent a marked
diurnal  cycling. Start and Markee also indicated the  potential application of
their  model  to urban  air pollution problems.  A further detailed application
of the  Gaussian  puff  concept to  urban pollution  modeling has been made
recently at the Argonne  National Laboratory  under a contract supported by
the National Air Pollution Control Administration, and a  paper on  this work
is to be presented at this  symposium.  Models that utilize the Gaussian instan-
taneous puff  rather than  the steady-state  Gaussian plume distribution of the
continuous point  source  obviously possess greater flexibility and  may well
mark the  advent  of the  second generation of  urban air pollution models.
Increased   flexibility is, of course, only  bought  at  the price of the very
sophisticated  knowledge required to predict, for an urban environment, both
the trajectories in space  and  the diffusive spreading of the  individual puffs.
In principle,  the  important  meteorological situation corresponding to the
"calm" wind  condition, which approximates the situation  frequently encoun-
tered in episodes of high urban air pollution,  could be discussed in terms of
the Gaussian  puff-type  model. This is not possible in terms of the Gaussian
plume of Equation (1), for which  wind speed U = 0 represents a degenerate
case.

SOME MISCELLANEOUS QUESTIONS
Prediction Limitations
The  preceding discussion may have pointed to some of the difficulties of
4-10

-------
urban pollution  modeling. In the past, most theoretical  attempts to model
the diffusive conditions in an urban environment have been based simply on
a judicious adjustment—guided perhaps by a few ad hoc tracer experiments—
of our knowledge  of what happens in the much simpler situation over open,
level country. Nevertheless, a  reasonable agreement  has  been obtained be-
tween the  predicted and  observed concentration values in a number of dif-
ferent comparisons. In view of the somewhat tenuous nature of the physical
basis of  these models, however, it  should be emphasized that they need to
be  applied with proper  caution  and, where  possible, only  by professional
meteorologists fully  conversant with  the practical  aspects of  urban atmos-
pheric  diffusion.  In  a slightly different connection, this point has  been
emphasized by Scorer19,  who says: "There are NO universal  formulae which
can be used to describe what will happen to pollution emitted into the air,
and this  needs to  be said because formulae have been so misused in  the past
by  people who have been secretly frightened of realities and have dispensed
according to the prescription of the most  fashionable  wizard. Responsibilities
cannot be so easily shed. Pollution problems are certainly local, almost per-
sonal." It seems, however, to  be a point worth reemphasizing, particularly in
view of  the current urgent demands for  urban  models to provide a quanti-
tative basis for a  variety  of critical  management decisions relating  to urban
air  quality. It seems likely that any major improvement beyond the present
somewhat empirical  approach to urban  diffusion modeling will require a
much improved  understanding  of the dynamics  and thermal  structure of the
continuously  developing urban boundary layer. Detailed observational studies
of the structure of the urban boundary  layer of the  type described  recently
by  Clarke and McElroy20  may provide the basis for  such understanding. In
any case, it should be evident that present  models are based on some  ideal
assumptions and that considerable departures from the model predictions can
be  anticipated in the presence of any special local phenomena  such as lake
breezes and mountain-valley effects.

Parametric Sensitivity  Analysis
In spite of the conceptual simplicity of the urban pollution models that have
been discussed, the superposition of the effects produced by a large number
of pollutant souces and their  evaluation for  complex  sequences of meteoro-
logical conditions  involves  such  large  computational effort  that,  in  most
cases, it  can  only  be handled with a high-speed computer. Under these cir-
cumstances, the more general features of the  models may be obscured by the
massive details of  particular applications.  At the present  time, there appears
to be a real need  to study the sensitivity  of the concentration predictions to
variations in  inputs to the models. In some cases, large  variations  in  input
may have little  effect on output. With others,  the reverse may be  true. It,
therefore, appears  desirable to conduct a sensitivity analysis  for some of the
                                                                    4-11

-------
current models  in  order  to  identify the sensitive parameters and the condi-
tions under which they are sensitive.
Two typical questions that come to  mind, on  which  a sensitivity  analysis
would  throw  some light, relate to the degree  of  accuracy and detail that is
necessary for an  urban  pollutant emission  inventory, and  the accuracy re-
quired to  specify the  concentration field  corresponding to a single source,
Ad hoc  tracer  diffusion experiments in  cities,  of the type described by
McElroy  and Pooler,7  are  difficult and  expensive to   perform. Since they
must,  to  some extent, reflect localized conditions, they may not add signifi-
cantly to  the predictive accuracy  of the model in general urban  situations
after a certain stage of specification  is reached.

The sensitivity analysis is also valuable from a rather different point of view,
Because  it  is not  very difficult  to  identify a large number of the  possible
sources of error  and unrealistic simplifications in existing models, in  a quali-
tative  way, there is a temptation to make rather arbitrary and simple adjust-
ment  for  these  in order to improve the  agreement between computed and
observed  concentrations. This can be a dangerous procedure because, until a
good understanding of the  response of the  model  to its many inputs exists,
there  can be no  guarantee that the adjustments necessarily  reflect a real
expression of improved physical understanding. Thus the apparently improved
agreement could  be fortuitous. In  the above  sense, some recently proposed
models  may already  be  over-complicated.  We  mention  here the  recent
notable revelation  by Dr. Frank  Gifford  of NOAA Atmospheric Turbulence
and Diffusion Laboratory at Oak Ridge,  that, at least for the calculation of
isopleths  of annual-average air quality  in terms of a  typical urban source
inventory, the calculations can be organized and simplified to such  an extent
that they can be carried out quickly and easily by hand.

Of those  models for  which  accounts have  appeared in the open  literature,
the only other one specifically developed  in a simple enough form  to permit
calculation from  a source inventory without the aid of  an electric computer,
is  that of  Clarke.21 At a time when computer-based models can  be expected
to  proliferate rapidly.  Dr. Gifford's recent success can serve as a reminder
that simple models should  not be  overlooked since they  might still be of
great value.
4-12

-------
REFERENCES
 1. Stern, A. C. (ed.), Air Pollution. Vol. 1, 2d ed. New York, Academic Press, 1968.
    694 p.
 2. Turner,  D.  B. Workbook of  Atmospheric Dispersion Estimates—Past,  Present, and
    Future.  Presented at  meeting of Mid-Atlantic State  Section of the Air Pollution
    Control  Association. Philadelphia, APCA 9946. June 1969.  17 p.
 3. Moses, H. Mathematical Urban Air Pollution Models.  National Center for Air Pollu-
    tion  Control, Chicago  Dept. of  Air  Pollution,  Argonne  National  Laboratory.
    Argonne, III. ANL/ES-RPY-001. April 1969. 69 p.
 4. Pasquill, F. The Estimation  of the Dispersion of Windborne Material Meteorol. Mag.
    90( 10631:33-49 February 1961.
 5. Turner,  D.  B. A Diffusion Model for an  Urban Area. J. Appl. Meteorol. 3(1):83-91,
    February 1964.
 6. Pasquill,  F. Atmospheric Diffusion.  London, D.  Van  Nostrand  Co.  Ltd.,  1962.
    297  p.
 7. McElroy, J. L. and F.  Pooler,  Jr. St. Louis Dispersion  Study, Volume II — Analysis.
    National Air Pollution Control Administration. Arlington,  Va. Publication Number
    AP-53. December  1968. 51  p.
 8. McElroy, J. L. A  Comparative Study of Urban Rural  Dispersion. J. Appl. Meteorol.
    5(11:19-31, February 1969.
 9. Briggs, G. A. Plume Rise: A Critical Survey. Air Resources Atmospheric Turbulence
    and  Diffusion Laboratory,  Oak  Ridge,  Tenn. U. S.  Atomic  Energy Commission,
    Division  of Technical Information.  USAEC  Critical Review Series. Publication
    Number TID-25075. 1969. 81  p.
 10. Pooler, F.,  Jr. A Prediction Model  of Mean Urban  Pollution for Use with Standard
    Wind  Rose. Int. J. Air Water Pollution. 4(3/41:199-211, September 1961.
 11. Turner,  D. B. The Diurnal and Day-to-Day  Variations of Fuel Usage  for  Space
    Heating in St. Louis, Missouri. Atmos. Environ. 2:339-351,  July 1968.
 12. Fortak,  H.  G. Rechnerische Ermittlung der SC>2  — Grundbelastlung aus Emissions-
    daten — Anwendung  auf die Verhaltnisse des Stadtgebietes von Bremen. Institute
    for Theoretical Meteorology, The Free University of Berlin. 1966.
 13. Larsen,  R.  I. A New Mathematical  Model of Air  Pollutant Concentration  Averaging
    Time and Frequency. J. Air Pollution Control  Assoc. 79:24-30, January 1969.
 14. Meade,  P. J. and  F.  Pasquill. Study of Average Distribution of Pollution Around
    Staythorpe. Int. J. Air Pollution.  7(1-2) :60-70, October 1958.
 15. Martin,  D.  O. and  J. A. Tikvart.  A  General  Atmospheric Diffusion Model for
    Estimating the Effects of One or More Sources.  Presented at 61st Annual Meeting
    of the Air Pollution Control Association. APCA-68-148. St. Paul. June 1968. 15 p.
 16. Frenkiel, F. N. Atmospheric Pollution and Zoning  in  an Urban Area. Sci. Monthly.
    S2.-194-203, April 1956.
 17. Davidson, B. A Summary of the New York Urban Air Pollution Dynamics Research
    Program. J. Air Pollution Control Assoc. 77:154-158, March 1967.
 18. Start, G. E. and E. H. Markee, Jr. Relative Dose Factors  from Long-Period Point
    Source Emissions  of  Atmospheric  Pollutants. In: Proceedings of the  USAEC Me-
    teorological  Information  Meeting,  Chalk  River  Nuclear  Laboratories, September
    11-14,  1967, Mawson, C. A.  (ed.). Atomic Energy of Canada, Ltd. Chalk  River,
    Ontario. Report Number AECL-2787. 1967. p. 59-76.
 19. Scorer, R. S. Air Pollution. New York, Pergamon  Press, Inc. 1968. 151  p.
 20. Clarke, J. F. and  J.  L.  McElroy.  Experimental Studies  of  the  Nocturnal  Urban
    Boundary  Layer.  In:  Proceedings  of WMO Symposium  on Urban Climates and
    Building Climatology, Brussels, Belgium, October  15-25, 1968.
 21. Clarke, J. F. A Simple Diffusion Model for Calculating Point Concentrations from
    Multiple Sources. J. Air Pollution Control Assoc.  74:347-352, September 1964.
                                                                            4-13

-------
5. A PRACTICAL MODEL  FOR  CARBON MONOXIDE
 ABSTRACT
      This paper reports our progress in the development of essentially two
      different types of diffusion models for  carbon monoxide (CO): (1) a
      "synoptic" model,  which calculates hour-to-hour concentrations, for
      vertification  studies  and possible operational use, and (2) a  "climato-
      logical" model, which calculates  arithmetic mean concentrations and
      frequencies of extreme concentrations, for planning use.  The design of
      these two models, as well as the  results of limited evaluation trials of
      an early version of the synoptic model are described.
      The synoptic model design  is based upon a modification of Clark's1
      receptor-oriented  model,  and emphasizes practicality, generality and
      completeness. The present version of the  model includes provision for
      handling extra-urban as well as intra-urban  sources. In addition,  a
      "street submodel" is being developed. Based largely upon experimental
      work, it converts the intra-urban background to street-level concen-
      trations.  When incorporated,  this  should substantially improve  the
      agreement between calculated and  observed  values at the Continuous
      Air Monitoring Projects (CAMP) Stations.
      The climatological model is similar in basic form to the synoptic model
      but uses the joint frequency function for all combinations of specified
      classes of the input parameters,  derived from  a large body of hourly
      meteorological data.  The arithmetic mean concentration  thus obtained
      is  then used in statistical relationships, derived by Larsen2  from carbon
      monoxide (CO)  concentration data,  to  calculate  the magnitude  of
      high-percentile and maximum concentrations  for any averaging time.
      Alternately,  such  calculations  can  be performed  directly using  the
      predicted concentration joint frequency function.
      Implementation of the model required  the development of an objective
      technique for converting  traffic data to  a time- and space-dependent
      CO emission inventory, as  well  as methods for  estimation of the
      meteorological variables appropriate to the urban area from  routinely
      available (airport) weather data.  The accuracy of these methods effec-
      tively establishes an upper limit to the model prediction accuracy.
      Sample  computed area-wide concentration patterns and time-varying
      point concentrations for St.  Louis and Washington, D.C., are presented.
      As expected,  the preliminary calculated concentrations  are, for the
      most part, lower  than those observed at the CAMP Stations,  reflecting
      the need  for incorporation  of the  street submodel into the program.
      The need for better treatment of light-wind conditions is also indicated
      by these evaluation trials; possible improvements are discussed.

-------
AUTHORS
     WARREN B. JOHNSON, JR., Manager of Environmental Meteorology Program, at
     SRI received a 6.5. degree in electrical engineering from Texas A&M in 1957 and
     M.S. and Ph.D.  degrees in meteorology from the University of Wisconsin in 1962
     and 1965.  At the University of Wisconsin, under  Professor  H.  H. Lettau and
     subsequently at ESSA's Atmospheric Turbulence and Diffusion Laboratory in Oak
     Ridge, Tennessee, Dr. Johnson  studied atmospheric  boundary  layer problems. At
     SRI, he has been developing lidar (laser radar) techniques for studying the behavior
     of pollutants in the atmosphere.

      FRANCIS  L. LUDWIG, Meteorologist,  Environmental  Research  Department, at
      SRI received B.A. and M.A. degrees  in meteorology  from UCLA in 1957 and
      1958, respectively. While at UCLA he participated  in studies  of Los Angeles air
      pollution  problems.  Since  coming to SRI  in  1959, Mr.  Ludwig's interest has
      centered on atmospheric aerosols and urban meteorology.

      ALBERT E.  MOON, Senior Operations Analyst, Management Systems Division,
      at SRI received B.S. and M.S.  degrees from the University of Texas in 1951 and
      1952. He worked on the design of electrical and electronic control systems.  He
      received  a  degree  of Master of Business Administration from the Harvard Grad-
      uate School of Business Administration, in 1964 shortly before joining Stanford
      Research Institute. Since that  time,  he has evaluated public  and private invest-
      ments, including benefit-cost  of mass  transit systems  in Los Angeles and San
      Mateo County,  California.

-------
                      5.   DEVELOPMENT OF A PRACTICAL,

                MULTIPURPOSE  URBAN DIFFUSION  MODEL

                                      FOR  CARBON  MONOXIDE



WARREN B. JOHNSON, F.  L. LUDWIG, AND ALBERT E. MOON

              STANFORD RESEARCH INSTITUTE


INTRODUCTION

Scope and Objectives
There is a need  for an objective, practical method of  simulating the  current
and future impact of motor vehicles  upon air pollution in urban communi-
ties. Such a technique could be used  to assess the characteristics of  current
concentration  patterns of primary vehicular  pollutants in cities, as well as to
predict the effects of exhaust  emission  controls and expressway routings
upon  future concentration distributions.  We  are currently engaged   in the
development of a simulation model of this nature, and  this paper reviews our
progress midway through  the first year's effort. In  this task,  we have been
aided  considerably by previous urban diffusion modeling studies, which have
been  aptly reviewed and  summarized  by Wanta,3  Stern,4   Moses,5  and
Neiburger.6  An  investigation of  carbon monoxide (CO) diffusion in Washing-
ton, D.C., by Ott et  a/.7 could be considered a prototype for our  study.

For simplicity, this initial  modeling effort is restricted to CO, since  (1) the
gas  is relatively  inert in the  atmosphere with no known significant  natural
sources or sinks  in urban areas/ (2)  motor vehicles  are known  to  furnish
large percentages of  the CO in urban air (discussed  in a subsequent  section
of this paper), and  (3) CO is believed to be an important pollutant in terms
of health effects. Later, it is planned  to extend the  model developed  for CO
to phutoreactive, automobile-generated pollutants.
*Research sponsored by the Coordinating Research Council and the National Air Pollution
 Control Administration (Division of Meteorology).
tEvidence  is mounting, however, that such sources and sinks may be found in the bio-
 sphere (Swinnerton,8 Went,9 and Robinson and Bobbins.10).
                                5-1

-------
Our  goal  is to develop a versatile  model that can predict both urban back-
ground  and street-level CO concentrations at any point within a city. Cal-
culations of the following types are required:
1.  Hourly concentrations as a function of  time, for  verification, use, and
   possible operational  applications (with the "synoptic" model).
2.  Long-term  average concentrations and statistical  predictions of high (and
   maximum)  concentrations for  planning  purposes  (with the  "climato
   logical" model).
We  are  aiming to  complete a  practical model that  is convenient to use,
economical to run, generally applicable to  most cities,  and,  of  course, ac-
curate.  To  achieve  the latter, we are attempting to  develop capabilities for
assessing  concentration contributions  resulting  from  diffusion  on various
scales, including the treatment of:
1.   Extra-urban transport and diffusion, mainly from upwind cities.
2.   General intra-urban diffusion from arterial and feeder streets.
3.   Local  (microscale)  diffusion within streets.

Fundamental Concepts and Assumptions
Our basic approach  has been  to develop individual components of the model,
and  then to arrange these components into  submodels, which can be appro-
priately  combined  to suit particular needs. This procedure affords maximum
flexibility.
Other key elements in our rationale have been to
1.   Bring the  best features of previous efforts together into a  single model.
2.   Maintain  simplicity until additional complexity  is  shown  to  be clearly
     required to achieve greater accuracy, and is justified by the accuracy of
     model inputs.
3.   Strike an  appropriate balance between the complexity of the model and
     the computing costs (program  running time).

Input data requirements for the model include  two major types, traffic and
meteorological. For the synoptic model, it  is necessary to assume that diur-
nally adjusted, average  daily traffic data (which are all that are available on a
network basis) are  compatible  with  hour-to-hour (synoptic) meteorological
data.

EMISSION INVENTORY DESIGN

Introduction
Estimates of vehicular emission of  carbon monoxide range from 98 percent of
the total for the Washington, D.C.,  area' ' and 91 percent in Los Angeles,'2 to
75 percent in  the San  Francisco  Bay  region.13  (The  latter estimate is  for
automobiles  only.)  Hence,  the  principal   effort thus far  (discussed  in
5-2

-------
detail  below)  has  been  to  model the  mobile  sources.  Significant  fixed
sources, such as airports, will later be included in the emission inventory.
The  inventory  of motor-vehicle CO emissions  is represented as a primary
network  of the major  arterial streets and  freeway links,  plus a secondary
background of emission from vehicles on the many  local streets that are not
included  in the primary network.

Two kinds of data  are  used  for the emissions inventory, (1)  historical data,
to re-create  past  conditions for  comparison with recorded  concentration
measurements for  model  verification; and  (2)  forecast  data, to  estimate
future  concentrations.

Calculation of Primary Traffic Network Emissions  Using
Historical Data
Figure  5-1  illustrates the  primary network used in the  Washington,  D.C.,
area.*  The amount of travel  on the major network is in excess of  90 percent
of the total travel  in  the  metropolitan  area. Emission  from each  straight
segment,  or link,  of  the  network  is determined  from the emission per
vehicle-mile and the number of vehicle-miles traveled on the link.

CO emission, e (grams per vehicle-mile), is determined from the equation

     e=  c/S*3                                                         (1)

where  S  is the average speed  over the link, in miles per hour (mi hf1), and c
and 0  are constants. For vehicles in use before exhaust control systems were
used, c = 103  and |3 = 0.8, as determined  by Rose et a/.14 from observations
on vehicles in several locations. For conditions where a majority  of vehicles
have exhaust-control devices, alternative values will be used for  the coeffi-
cients.

To define the network  and  compute the  emissions, the following input data
are needed:
     • Network description
     • Node coordinates
     • Link distances
     • Link volumes
     • Link speeds

    Historical  data for  the  network  description,  the node coordinates, and
the link distances are obtained from  highway and street maps. The network
is  described by  assigning numbers  to the  intersections, or nodes, of the
network  and by identifying the pairs of nodes that are connected by  links.
The  coordinates of the nodes establish the locations of the traffic  links  in
"Computer-generated displays of this nature are very useful for examining the accuracy of
 the coding of the traffic links.
                                                                      5-3

-------
the network. Link distances are measured along the links, rather than as the
shortest distance  between the  nodes, although  the  assignment of the emis-
sions assumes that the link is a  straight line between the nodes.
                                                             TA-7874-18S
  Figure 5-1.  Computer-generated  display of traffic links in 24-by
  24 mile central portion of primary network for Washington,  D.  C.
 Historical  link-volume data  are  obtained  from  traffic  departments in  the
 cities, towns, and counties of the region being studied. These agencies sample
 traffic  volumes by means  of  portable  counting units  and a  few fixed
 installations. The  frequency of  their observations  may vary from almost
 continuously to as infrequently as once in two or three years. Because traffic
 varies  according to seasonal, weekly,  and daily  cycles, an  observation of
 volume for one day must be adjusted  for the weekly and seasonal fluctua-
 tions. The  resultant corrected  value  is recorded as the  average daily traffic
 (ADT) for that  location.
 5-4

-------
For use  in the synoptic model, the emissions must be expressed in hourly
averages, so the average daily traffic must be allocated among the hours of
the day. Allocation can be accomplished  using data on the hourly distribu-
tion  of trips, which  are  compiled by the traffic study agencies of many
areas. A typical diurnal  pattern of trips  is shown  in Figure 5-2.  There  are,
however, limitations to the accuracy of  the assumption that the traffic on
each  link is  distributed  like  the  hourly trip  pattern.  We  know that trips
           00        04        08       12        16
                               HOURS OF DAY, LSI
20       24

TA-7874-15R
  Figure 5-2.  Hourly distribution  of  trips  in Washington, D. C.16

taken  during  the morning  and  evening  peak hours are longer than those
taken during other times of the day, and  that the pattern of use of different
kinds of  streets  in different parts of the city is different, as shown by the
comparison  of  hourly volume on a freeway and a downtown arterial  in the
St. Louis area, shown  in Figure 5-3. Accordingly, we are planning to use all
data available to develop different hourly volume patterns for  each facility
type.

Associated with the computation of  hourly traffic volume from  average daily
traffic is  a feature  of the model that  accounts for the fact that emissions at
                                                                     5-5

-------
                                  RADIAL EXPRESSWAY
                          	CIRCUMFERENTIAL
                                   ARTERIAL
                          08        12        16
                             HOUR OF DAY, LSI
20       24
  TA-7874-19S
Figure 5-3.  Hourly distribution of traffic for two facility types
in St. Louis.50

different times from different source  locations affect the concentration at a
given  time at a given receptor point.  For  example,  with a 2-m sec"1 wind,
the average concentration at a receptor point from, say, 0800 to 0900 Local
Standard Time (LST) will result from  emissions generated in the hour from
0700  to 0800  at a  location  7.2 km  upwind, and from 0600 to 0700  at a
location  14.4 km upwind, and  so  on. A computational  loop in the model
uses wind  speed to  determine  the displacement in time needed  for  each
emission computation.

Link speeds for historical data are determined by using averages for  peak and
off-peak travel hours on the various kinds of route facilities in the area.  Data
for the average speeds are obtained from the traffic survey.  Speeds used  in
the Washington,  D.C. area were based  upon the study by W.  Smith and
Assoc.1 6 shown  in Table 5-1.

Accuracy of  the historical  data is limited by the frequency with which it is
recorded and  by the accuracy of the factors used  to  correct  it to average
daily  traffic.  However,  since locations  with  heaviest  traffic are  generally
5-6

-------
Table 5-1. AVERAGE SPEEDS FOR WASHINGTON, D.C., HIGHWAY FACILITIES
                               (mi hr
                                    -l,
Facility type
Suburban freeway
Downtown freeway
Suburban arterial
Downtown arterial
Local or feeder street
Peak traffic hours
(0700 to 0900, 1600 to 1800)
42
33
30
24
10
Off-peak traffic hours
(0000 to 0700, 0900 to 1 500,
1800 to 0000)
48
39
36
30
12
monitored most often, greatest confidence can be placed in data with highest
volumes. In  addition, very few data are available for Saturdays and Sundays,
since traffic planners are concerned with  recording peak  demands, and peak
demands on  streets  usually occur  on weekdays. This  lack of data impedes
efforts to run weekly cycles, however, because peak traffic usually occurs on
weekdays, peak pollution  concentrations from  motor-vehicle emissions also
usually occur on weekdays.

Calculation of Secondary Traffic Emissions Using  Historical  Data

The  number of vehicle-miles  traveled  on  streets not represented  by the
primary  network is computed from an estimate of the total  vehicle-miles
traveled  in the area and the total vehicle-miles on the links of the primary
network.  The  local street  mileage  is distributed  over the study  area  by
estimating the  relative density of  local streets as opposed  to parks,  open
spaces, and  primary streets, for each 4-square-mile area in a 2-mile by 2-mile
grid covering the area. The average emission  from the local street travel in
each square  is then assumed to emanate uniformly from  that square. Al-
though  the  emission per  mile  is  high  because of low  speeds, the overall
contribution  is small because of the small  fraction of the total area travel
that is carried out on these local streets.

Calculation of Future Emissions Using Forecast  Traffic Data

Most urban  areas in the United States have completed or are conducting an
area-wide transportation study to determine traffic demands and transporta-
tion facility  needs  for  a  specific future. Such  studies are required for par-
ticipation in  Federal highway  programs.  Two  important outgrowths of  these
studies are:  (1) a design for a future traffic network for the area  and  (2) a
forecast  of the traffic volumes on  the links of  the network. The procedure
for conducting these investigations  has been  highly developed  and  partially
                                                                      5-7

-------
standardized. We are designing the emission inventory components of the
model so  that the magnetic tape  form of network description, link volume,
and  link speed  data, of widely used traffic-planning computer programs can
serve as input for the diffusion model.  In most cases,  the only manual step
required will be  the measurement and  coding  of node coordinates for the
network.
Accuracy  of the  forecast data is difficult to establish. One study that sought
to appraise  the accuracy  of earlier forecasts was unsuccessful in finding a
case where the planning recommendations had  been followed closely enough
to  make  the actual situation equivalent to the  conditions  that  were fore-
cast.16  Checks made to establish the adequacy  of  calibration of the fore-
casting  models are  usually  considered successful if the forecast for present
conditions gives  link volumes in  wide corridors within 10 to  20  percent of
counted volumes. Variations between  the volumes on individual streets with-
in the  corridor  may be  much wider since the  models do not  distinguish
between parallel routes  as  readily  as do drivers.
 IIMTRA-URBAN  TRANSPORT AND DIFFUSION  SUBMODEL

 Spatial  Partitioning of Emissions

 Our diffusion model is a modified form of the receptor-oriented model de-
 veloped  by Clarke.1 As  shown in Figure 5-4, we use logarithmic spacing of
 the area segment boundaries to permit more precise location assignment for
          16km
                                                   2  1
                                                           RECEPTOR
                                                             POINT
                                          1,000m
                   EXPANDED VIEW OF
                  ANNULAR SEGMENTS
                     WITHIN 1km OF
                       RECEPTOR
                                                 500
250
   125
      RECEPTOR
        POINT
                                                           TB-7874-ls
      Figure 5-4.  Diffusion model area-source configuration.
5-8

-------
i emissions near the receptor and to tend to equalize the contributions from
I each segment to  the concentration at the receptor point. The closest segment
1 extends from the receptor to 125 meters, roughly comparable to the size of
; a city  block, while the farthest segment extends to  32 kilometers, approxi-
 mately the diameter of a large city.

 The 22.5° sector width  is convenient and fits reasonably well the angular
 plume  widths (between  ±2a points), predicted by  Gifford's17 model,  for
 slightly unstable  conditions. The sector width is expanded to 45° within  the
 closest  1 kilometer to allow for the  large initial  lateral dispersion observed
 during the St. Louis tracer studies.18'19

 Emissions from the line sources (links) in the traffic  network are assigned to
 one of the nine area-segments as shown by  Figure  5-5.  To simplify these
 computations, the  segment  boundaries  are taken as straight-line  fits to  the
 curved  boundaries  illustrated in  Figure 5-4. The computations transform  the
 rectangular  node coordinates into  polar coordinates originating at the recep-
 tor  points and then determine whether all or  part  of a  link  falls  within a
 given  segment.  The  link  is  assumed to  be  a  straight line connecting  the
 nodes.  Where part  of  the link is within a segment, an equivalent fraction of
 the  link emission is included in  the computation. The  total emission within
     CHECK FOR LINK FALLING WITHIN
   DIFFUSION SECTOR; DETERMINE AREA
    SEGMENTS CONTAINING END POINTS
     LOCATE INTERSECTIONS OF LINK
       WITH BOUNDARIES DIVIDING
            AREA SEGMENTS
     LOCATE INTERSECTIONS OF LINK
       WITH SIDE BOUNDARIES OF
           AREA SEGMENTS
LOCATES
   A's
                                    LOCATES
                                      B's
     COMPUTE PROPORTIONATE LINK
        LENGTH  FALLING WITHIN
         EACH AREA SEGMENT
     COMPUTE AND ACCUMULATE CO
   CONCENTRATION CONTRIBUTION FROM
      LINK TO EACH AREA SEGMENT
        REPEAT FOR OTHER LINKS
                                                                   1000m
                                                                 TA-7874-4I
    Figure 5-5.  Schematic of traffic  link  assignment subroutine.
                                                                       5-9

-------
each  segment is assumed  to be uniformly  distributed  and consists of the
total  emission  from  all the  links and  parts  of links  that  lie within that
segment, plus the secondary traffic emission in  the segment.
As  previously  discussed  in  this  paper, the  secondary traffic emission is
described in  a  2-mile-square grid system. The values at the  centers of the
squares  represent the  average  emissions  over the squares.  The  secondary
emission assigned  to each segment in the diffusion model is computed by
locating  the  center  of  the area segment, then interpolating among the four
adjacent secondary  emission grid centers. The value thus determined  at the
center of the segment is assumed to prevail over the entire segment.

The contributions of each area segment to the concentration at the receptor
point are computed separately and then totaled. The calculation of pollutant
concentrations  from  each  of these individual  segments is based on a com-
bination of two simple mathematical models, the Gaussian and the "box." A
similar approach has been employed by Miller  and Holzworth.20

Basic Gaussian Model
Gifford17  has developed a generalized diffusion model to describe  the plume
from  a continuous ground-level  point source,  assuming  perfect  reflection at
the ground.  In  its  simplest  form, for the ground-level concentration, the
equation is

        C=F^TU    exp   (-y'/2o>)                           (2)

where
      C =  ground-level  concentration  (g m~3)
      Q =  source strength (g sec"1)
      u  =  average wind speed (m sec"1)
      y  =  lateral  (crosswind) distance from the plume axis (m)
CTy,  az,  =  lateral  and vertical standard deviations of plume concentration
           (m);  these parameters are functions of source-receptor distance
           and atmospheric stability.

This equation takes  the following basic form for ground-level concentrations
emitted from a ground-level line source:
                /9\1/2
      c  =  QL   if)     
-------
McElroy.19  At  present,  we  are  using  Gifford's* values,  but the model  is
flexible and can  use any  set of  functions  that  can be reasonably approxi-
mated, over the  intervals  between  segment boundaries, with expressions of
the form:
               b,j
     az =  BJJ r                                                         (4)

Where  r is  the travel  distance, the subscript  i  refers to annular segments
upwind of the receptor point (see  Figure 5-4), and j refers to the stability
classes.  The  az  functions are, thus  approximated  by  a series  of  power
functions, each  appropriate  to  some upwind  distance interval  and stability
class. This allows us to  select different representations for different distances,
if  we choose. Short ranges, in particular, may require special  treatment to
account  for the initial  rapid mechanical mixing found  in  urban areas.18'19
We are testing two ways  of  handling this  initial dispersion: (1) adding 25
meters to az over all distances, as suggested by Pooler18 and  by McElroy;19
and  (2) using  a  value of az throughout the closest segment corresponding to
Gifford's values at a distance of 125 meters.

Substituting Equation (4)  into the line-source equation and integrating with
respect to r from r = r, to r = ri+1  gives the contribution for stability class j
from the /th segment area  source QA:
where r-t  and  ri + 1   are, respectively, the  distance  to  the downwind  and
upwind boundaries of the /th segment. In this expression QA has the units (g
rrf2  sec"1). This  basic  Gaussian model  applies  when  there  is no effective
limitation to vertical  mixing or when the cloud has not spread  sufficiently to
be affected by  such a limitation.

Transition from Gaussian to Box  Model
When the  layer into which the pollutants  are being dispersed is  restricted,
they will tend  to  become  uniformly mixed  at a  sufficient distance from the
source.  Under  these  conditions, the box model  is  used. The concentration
arising from the uniform area source in the /th annular segment is
where h  is the depth of the layer into which the pollutants are mixed.
•Gifford's work is a modification of that by F. Pasquill (unpublished) and Meade22 The
 Gifford-Pasquill  curves are based upon data from  open country, and probably under-
 estimate the diffusion in  urban areas. We are examining possible improvements on this
 point.
                                                                      5-11

-------
We have chosen  to change from the Gaussian model to the box model at
that point  where the two  equations  (in their respective line source formula-
tions)  give  equal  concentration  values. The distance to that point,  rT, is
given by:
                                1/bij
              rT =   (O.8h/a,j)                                        (7)

The part of  the  annular  segment  inside rT  is treated according to  the
Gaussian model, and the part that is outside, according to the box model.
For this application rT is substituted for ri+1  in  the Gaussian equation  and
for T| in the box equation.

Determination of Model Inputs
The model  requires  meteorological  inputs  of  mixing depth,  stability,  and
wind  speed  and direction. At present, the model  uses winds observed  at the
nearest airport. Central  urban daytime wind speeds above the buildings are
known  to be  on  the order of 30 to 40 percent less than airport speeds,  and
a  correction for  this effect is planned, but  not  yet incorporated into  the
computations.
The case of  calm winds  cannot  be handled by the model  without some
adjustment. At present, the model assumes that  a  wind that is reported as
calm  has  a speed of 2.0 m  sec"1  and that its direction is the same as the
most recently measured  direction. Other  approaches to  this problem  are
possible, but they require  special models to be developed for the calm-wind
case, as will be discussed later.
The stability and  the mixing  depth are  not directly measured quantities  and
require  estimation methods. We are using Turner's23 criteria for determining
stability from routinely available data. This procedure estimates the stability
class on the basis  of cloud  cover, ceiling height, wind speed, and solar eleva-
tion.  The stability class is  used  to select the Gifford-Pasquill az  function.
Following Turner, we  are  using five stability classes  ranging from  "very
unstable" (Class 1)  to  "slightly  stable" (Class  5),  eliminating the "stable"
and "very stable"  classes. These  conditions are known to occur very seldom
over cities because of the  heat-island effect. Mixing depth over the  urban
area is determined for nighttime cases from Summers'24 simple model of the
urban  heat  island  and  from an  empirical  relationship involving heat  island
intensity,  city population,  and  low-level  rural (airport) lapse rate, as  dis-
cussed in detail in  Appendix A.
Afternoon  mixing depth is determined from  the height of intersection of the
morning sounding and  the dry adiabat passing through  the maximum tem-
perature for the day. Other daytime values are interpolated on the basis of
the observed hourly temperatures (see Appendix). The model currently uses
a uniform mixing depth throughout the area.
5-12

-------
 Model Limitations and Prospective Refinements
 The simple model that has been outlined here has a number of limitations.
 Refinements are  possible, but one must first determine whether such refine-
 ments are warranted  in  view of the  limitations  of the available meteoro-
 logical  and  traffic  data  in  terms of accuracy and  representativeness. In
 addition,  the increased  prediction accuracy  must  be  worth  any  additional
 computing costs. We are currently considering several potential model refine-
 ments in this context.

 Suitability of Input Parameters
 The model currently takes the airport wind speed to apply over the entire
 urban area. We plan to adjust this value to give a  more realistic wind speed
 over the city. Graham25  found an urban/rural average wind speed ratio for
 Ft. Wayne of 0.67 at 15 meters height and 0.64 at 60 meters. In Nashville,
 Schnelle et a/.26  found the roof-level  wind speed, Ur, to be related to the
 airport wind speed, Ua, as

     Ur= 0,33  Ua+ 1.0 msec"1

 where the overbars refer to daily averages. Graham's observations also show
 the wind direction over the city to  be backed  by 10° to 15° from that at
 the  outskirts, which could  also  warrant a correction to the airport data.
 These results are mostly for daytime  conditions. At  night  the wind speed
 over the city may be higher than that at the airport.

 Incorporation of nonuniform meteorological  parameters  over the city would
 be more  complex. Spatial variations in  nocturnal  mixing depth  have  been
 observed  by  Clarke27  in Cincinnati,  and Bornstein28  in  New York,  and
 modeled in limited fashion by Summers24 and Leahey.29 The desirability of
 inclusion  of such  effects is  under  study. f Efforts  are  also  being made to
 improve the form of the traffic data, to  include (1) diurnal  volume patterns
 for each road type, and (2) weekend data.

 Nonsteady Conditions
 The steady-state nature of the model is a more  important  limitation.  The
 reasonably uniform spatial distribution  of  the traffic network, however, helps
 to offset  this  situation.  Wind  direction changes from  hour  to hour are
 included simply  by  switching  to a new diffusion sector  in the model, which
 serves reasonably well for strong  winds on the  order of 30 km hr"1, or  8 m
 s"1 (the diffusion sector extends to  32 km), A special feature of the model
 accounts for the finite.travel time of emissions under light  wind  conditions
 (see page  5-2),  but no provision is made for directional changes in trajec-
tories, as observed by Angell ef a/,30 and Hass  eta/,31 from tetrgon flights.
We are considering  a modification that would calculate  the contribution of
                                                                    5-13

-------
each previous hour's emission  to  the  concentration  at  the receptor for a
given hour,  and  would accommodate wind directional changes by relocating
the apparent receptor  point and diffusion sector with respect to the traffic
network.

The Stagnation  Case
The most important pollution situation, the stagnation with  very light winds,
is virtually  impossible  to treat with our present model. The nocturnal urban
circulation  for  this case has  been reasonably well documented32'33  and
takes the form  of a convective cell driven  by  the urban  heat island. One
approach that we are considering  for night  hours involves treatment of the
circulation  as that  of  a height-limited, ground-based "thermal"  or vortex
ring, symmetrically  oriented over  the  urban  area. The wind field would be
taken as convergent toward the center of the city. Allowing  for recirculation,
a  time-dependent  concentration distribution  would  be calculated  by means
of  a cylindrical version  of the box model (which  might be called a "pan"
model).  A critical parameter would be the entrainment rate for clean air at
the top  and circumference of  the  "pan," which might be estimated on the
basis  of convection theory.34' 3S  A  more  realistic treatment would allow
superposition  of  this circulation  upon a weak gradient flow, giving the
buoyant heat plume observed by Clarke.27
 DIFFUSION SUBMODELS FOR OTHER SCALES

 Microscale (Street)  Diffusion
 Because of the finite  spacing of sources within a city, area source simulation
 such  as is  used  in the  intra-urban  model  is best applied at scales above a
 certain lower limit. This minimum spatial scale can  be considered  to be on
 the order of a city block, hence the choice of 125 m as the finest resolution
 in the intra-urban model. For shorter source-receptor distances, an alternative
 technique is needed.
 Additional  complications  arise because,  contrary  to the usual  situation in
 nonurban diffusion studies, the scale of the largest urban roughness elements
 (buildings,  etc.)  is very large  compared  to the local scales of emission and
 reception.  This means that the aerodynamic  effects  of  structures become
 important.
 Models that do  not include the effects of microscale diffusion will normally
 undercalculate  concentrations  in comparison  with  those  measured  at CAMP
 Stations, which are usually located near streets.  For example, the model used
 by Ott et a/.7  gave average concentrations that amounted to 36 percent of
 the CAMP average.
5-14

-------
A  "street submodel" that can be  used to  calculate curbside CO concentra-
tions  is currently under development. The effort  thus far has been based
largely upon the  extensive measurements  of Georgii36  in Frankfurt/Main,
Germany, although  the work of  McCormick  and  Xintaras,37 Schnelle et
a/.,26  and Rouse38 have also been considered.

Georgii's experiment involved extensive measurements of CO concentrations
and wind speeds at different levels above three different streets in built-up
areas, along with  occasional traffic counts.  One major finding was that the
CO concentrations on the leeward sides of buildings were considerably higher
than those on the windward sides, implying a helical cross-street circulation
component  in the opposite direction  from  the roof-level wind. In  addition,
the averaged data  showed  that  (1)  the vertical  concentration profiles  on
either side of the street  assume an  exponential  form, (2)  the mode of  air
circulation above the  street  apparently changes  when  the roof top wind
speed exceeds about 2  m sec"1, and (3) the concentrations are  exponentially
related  to traffic density.  Examination of the measurements reported  by
Schnelle et  a/.26  also  indicates general agreement with  (1) and (2) above;
their data are insufficient for verifying (3).
Georgii's  observations  show that the roof-level  concentrations differ only
slightly  from  windward to leeward  sides of the  street-side buildings. It is
reasonable to  assume that the urban background  concentrations  calculated
by the intra-urban model are approximately equivalent to  roof-level concen-
trations. The  effect  of the street  is  to  increase  incoming roof-level con-
centrations by  a factor that depends  upon  the receptor  location relative  to
the wind  direction,  the height above the street, the wind  speed at average
roof level, and the traffic density.

Georgii's averaged data for a  traffic density of 1400 vehicles per  hour (the
mean traffic density  during  his experiment) are well  approximated  by the
empirical relationships

      Q
      — = exp [(0.55)   (1    z/zr)]
      Cr
      CL
      	 =  exp [(0.87 +  0.098 ur)  (1 -  z/zr)],                    (8)
      Cr

where Cw and CL are concentrations on the windward and leeward sides of
the street-side buildings,

      Cr =  rooftop concentration (g nrf3)
      ur =  rooftop wind speed (m  sec"1)
      zr =  average rooftop  height (m), and
      z  =  receptor  height (m).
                                                                    5-15

-------
The  windward concentration  ratio (Cw/Cr) turns out to be  approximately
independent  of  wind speed. This  is  apparently because the  wind effect at
this  more  exposed  location is  mostly absorbed  by changes  in Cr On the
leeward  side, the concentration ratio increases with  increasing wind speed,
reflecting the relatively greater  effect of sheltering at  street  level as the  wind
increases.
Although changes in traffic density, thus emission rate, in the adjacent street
are  not reflected directly  in  these  equations, some justification for  their
application exists if it is assumed that  the  local traffic at a given time is
always proportional to that in  the city as a  whole. For this case, the traffic
changes will be reflected in the rooftop concentration, Cr,  calculated by the
intra-urban model, and hence, in Cw  and CL.

An  additional  problem  in applying  the above  empirical  results  is that of
defining leeward and windward cases on the basis of receptor  location, street
orientation,  and wind  direction. The problem is  compounded further when
the  receptor  is  located at the corner of a street intersection, as is frequently
the  case with  CAMP  stations.  In addition, in the vicinity of many  inter-
sections are  buildings  with widely varying heights, as well as open spaces
such as parking  lots.  These complications  make it difficult to apply the
empirical  relationships  on a   general basis.  Additional  work is necessary
before treatment of microscale  effects is  incorporated into the model.

Macroscale  (Extra-urban) Diffusion
In addition  to  considering sources in the  immediate urban area,  the model
also takes into  account  the background  pollution that is  transported  from
one  urban area  to another. We have  accepted, a priori, that the treatment of
extra-urban transport can and  should be considerably more gross than the
treatment  of nearby sources.

The first problem to be faced in the development of an extra-urban diffusion
model is that there is no convenient way of knowing the upwind trajectory
of the air  arriving at a city. We can either determine the trajectory from past
and  present meteorological data or  we  can  assume  that  the air has  come
from somewhere  within  a  very large upwind  segment. The latter approach
seems more practical,  particularly if  we ever  hope to apply  the model to
climatological data. Even the  synoptic  problem  of trajectory determination
by objective means  seems too complicated to be warranted.

We  apply  the box  model to a quadrant centered on the  upwind direction
and  extending  from 32  to 1000 km upwind  of the receptor. The source
strength is assumed to  be constant in time and space throughout  the sector,
and   is  estimated  from  yearly fuel  consumption39  for  those  states and
Canadian  provinces whose  centers fall within  the sector. To determine the
total emission rate within the area, we use the total amount of motor vehicle
 5-16

-------
fuel consumed and  the  conversion factor:  1 gallon of fuel yields 1.32 x 103
grams of carbon  monoxide. If the  yearly (3.15 x 107  sec)  consumption  of
fuel within the quadrant extending to  1000 kilometers is given  by F, then
the average CO emission rate Q in the area  (7.86 x 1011  m2) is given by:

     Q =  5.32 x 10"17 F  (g m"2 sec"1}.                              (9)

The model input includes a table  of  16 values  of  F,  one for each  of the
wind directions used. This table is  changed for each city. According  to the
box model, the concentration C from some upwind segment  is
     C =
                                                                    (10)
               uh
where  r2 and  rx are the distances to the outer and inner segment boundaries
(in this case 106  and 32 x  103  m), u is the wind speed, and h is the depth
of  the layer  through  which  the  material  is mixed.  Substituting Q from
Equation (9)  and  the  values  of t\ and  r2  gives the following equation for
concentration  from extra-urban sources, Ce:

     ce =  5J5  X  1Cr'lp                                          (11)
                uh

The afternoon mixing depth, calculated in the intra-urban diffusion model, is
used in Equation  (11)  because the pollutants are likely to have traveled for
long enough periods that they will have  had time to be mixed through  the
depth of the afternoon layer.  We are aware that the local mixing depth will
not necessarily be appropriate to  the region upwind of the city, but more
detailed treatment  of  this  problem  does  not  seem to be warranted.  To
approximate the  normal vertical wind  shear  in  the lower  atmosphere, we
have taken  the average transport  wind  velocity through the depth  of  the
mixing layer to be 1.5 times the maximum airport winds for the day.
In addition to the material generated within the segment extending to 1000
km, the  general   "worldwide"  background  is  included.  Robinson  and
Robbins10  have  estimated  this to  be  about 0.2  part per million  (ppm)
(about 2.4  x  10~4  g nrf3)  at  sea level in the mid-latitudes of the northern
hemisphere.  This  value is added to  that calculated  from Equation (11) to
give the total  extra-urban concentration, which  generally amounts to a few
tenths  ppm. This  extra-urban  contribution is, in turn,  added to the concen-
trations from  urban  sources  calculated for each  point in the  city.  The
extra-urban contribution is calculated  once for each 24-hour period, and that
value is applied from one midnight to the next.
                                                                    5-17

-------
SYNTHESIS OF BASIC MODEL TYPES
The  previous sections have described the  individual  portions or components
of the  diffusion model.  Here we discuss  the  organization  of these com-
ponents into models capable of furnishing the desired outputs.

Synoptic  Model
The  synoptic model  is designed to take the data for some  hour, calculate
and  display  point concentrations or concentration isopleths within the city
for that hour, and then proceed to the next hour. This is repeated until the
time  period of  interest  has  been covered  on an  hour-by-hour  basis. The
temporal changes  of  concentration for a  selected location can  be compared
with the measured values during the same period.
The  organization of the model is shown in Figure 5-6. In this  flow chart it is
assumed that the traffic data for the city  are already stored  and  that the
inputs consist of the meteorological parameters and those parameters neces-
sary for the application  of the  stability  and mixing depth subroutines (see
Appendix  A). The  calculations  start with  determination of the afternoon
(maximum)  and  night  (minimum)  mixing depths from the  appropriate
sounding. Then  the meteorological parameters for the  starting hour are read
and  the stability class determined. Mixing depths are calculated by different
methods according to  whether  it is  day or night. Concentrations are  cal-
culated and then the  next hour's data are read.
The  diurnal  variation  of mixing depth estimated by  the  model is schemat-
ically  illustrated in Figure 5-7. A constant,  the night (minimum) mixing
depth,  is assigned  to  the hours from  midnight until the first  hour  after
sunrise. This value is based on the morning temperature sounding.  The day
values are  interpolated between  the  night (minimum)  and afternoon (maxi-
mum) values on  the basis of surface temperature (see Appendix  A). The
mixing  depth from sunset to midnight is interpolated  by time between the
sunset and the  new night (minimum) value based on the next morning's
temperature  sounding. This process makes the transition to the night mixing
depth somewhat smoother.
The  meteorological  data used  with  this model  consist  of  hourly surface
observations and 1200  Greenwich Mean Time (GMT)  (early morning) upper-
air soundings, both on punched cards in edited form.
The  traffic data are stored  in  the computer in the form described in an
earlier section. Each traffic link  is identified  by the location of its endpoints.
Data  regarding length, daily traffic volume,  and type (arterial, freeway, etc.)
are associated with each link. This information, when combined with hour of
the day, receptor  location, wind direction, etc., is converted by the emission
model to appropriate source strengths.
Sample results of calculations with the synoptic model are presented later  in
this paper.
5-18

-------
                                        READ POPULATION.  LATITUDE,  INITIAL TEMPERATURE
                                     SOUNDING, DATE. MAXIMUM AND  MINIMUM TEMPERATURES,
                                             AND  LAST OBSERVED WIND DIRECTION
                                      CALCULATE INITIAL VALUES OF AFTERNOON AND NIGHT
                                                MIXING DEPTHS {SEE APPENDIX!
                                       READ HOUR OF DAY, CEILING HEIGHT. CLOUD COVER,
                                           TEMPERATURE, WIND SPEED. AND DIRECTION
                                         IF WIND IS CALM ASSIGN SPEED VALUE 1.0 m
                                              USE THE LAST OBSERVED DIRECTION
                                                  DETERMINE STABILITY CLASS
IS
IT DAY?
                                                    HAS SUNSET OCCURRED WITHIN THE PAST HOUR'
                          DETERMINE MIXING DEPTH
                          FROM AFTERNOON VALUE
                          BY TEMPERATURE INTER-
                          POLATION  ISEE APPENDIXI
                                                             DETERMINE MIXING DEPTH
                                                             FROM AFTERNOON VALUE
                                                             BY TEMPERATURE INTER-
                                                             POLATION ISEE APPENDIX)
                                                                            READ  NEXT MORNING'S
                                                                           SOUNDING AND MAXIMUM
                                                                          AND MINIMUM TEMPERATURE1
                                        DETERMINE MIXING  DEPTH
                                         BY INTERPOLATION (ON
                                         BASIS OF TIME) BETWEEN
                                        NIGHT VALUE AND VALUE
                                         AT FIRST HOUR AFTER
                                                SUNSET
                                                    USE
                                                   NIGHT
                                                  MIXING
                                                   DEPTH
 CALCULATE AFTERNOON
AND NIGHT MIXING DEPTHS
     (SEE APPENDIX!
MORE DATA>
CALCULATE CONCENTRATION
 FOR ONE  OR  MORE POINTS
  AND DISPLAY RESULTS
     Figure  5-6.   Flow chart  of  synoptic model  calculations.
                                                                                           5-19

-------
                              12    14    IE
                              LOCAL TIME, hour
Figure 5-7.  Modeled diurnal variation of mixing depth (schematic).
Climatological  Model
Our computational  work, thus  far,  has  been  limited  to calculations of
hour-to-hour concentrations with  the synoptic model, based on hourly mete-
orological  observations and average daily  traffic data  adjusted  for diurnal
conditions.  We also are developing  a capability to calculate average concen-
trations, as well  as high-percentile and maximum  concentrations over various
time periods.  That capability requires what we have termed a climatological
model.  In this  section  we discuss  how we  plan  to  proceed in this task,
drawing upon recent work by Larsen,2 Larsen and Burke,40 and Martin and
Tikvart.41  A similar approach has been used by Frankel.42

Basically,  a two-step procedure is involved: (1) an arithmetic mean concen-
tration  (C),  for  a  1-hour averaging  time  is  calculated from  the diffusion
model, and (2) either the statistical  relationships developed by Larsen40 are
applied  to  this  value to  yield estimates  of  high-percentile and maximum
concentrations, or  else the predicted concentration-joint-frequency function
is used for direct estimate.

 Calculation of Arithmetic Mean Concentrations

Computation  of  C  at a  point  requires a  different  treatment  of the input
data,  as  suggested  by  Martin and  Tikvart.41 This  computation requires a
large body of hourly meteorological  data,  preferably for a period of at least
5-20

-------
\ year. The discrete joint probability  density  function  f(xi, x2,. ..xk)  is
tabulated*  for all  combinations  of  classes of hourly values of the model
input parameters KI  to xk. By definition,

      S S  ... E     f(xlf x2,  ...xk)  = 1                              (12)
      rii n2    nk

where the n's denote the number of class intervals for each parameter.

The judgment as  to  the  number of  classes  to  be established  for  each
parameter must be based on a compromise between expected accuracy and
required  computing  time. The  latter can become critical when the diffusion
model, G(XJ, x2,  ... xk), is used  to compute  C at several hundred receptor
points, where the calculations at each point take the form

      C =  S S ...S  f(xlf  ...xk)   C(xi,  x2,  ...xk)                  (13)
           r\i n2   nk
Here the concentration computed for each combination of parameter classes
is weighted  by the  probability of occurrence of  that combination, and all
such  fractions are  summed to   give the  average (arithmetic  mean).  The
quantity (N)  of  individual concentration  computations  for  each point  is
determined by the size of the k-dimensional joint-density-function array:

      N =  nj x n2 x ...x  nk

Obviously the number of  parameters,  as  well  as the number  of classes,
affects the computing cost.

For our  study, the traffic input  is currently identified uniquely by the time
of  day,  since  available detailed   data are  only sufficient to establish the
diurnal cycle. As indicated by  Figures 5-2  and 5-3, the source emissions can
be  adequately  represented by  five time  classes, two  during  peak and  three
during off-peak traffic periods: 0000 to 0600, 0700 to 0900,  1000 to  1500,
1600  to  1800, and  1900 to 2300 LST. Wind  speed  is divided into  6
categories, wind direction into  16, and stability into  5. At times it may be
feasible to reduce the number of classes for wind direction to 8.

The remaining parameter, mixing depth h, is the  most difficult to obtain,
since its determination requires a morning  temperature sounding  (see Appen-
dix  A). Frankel42  handled this problem by adjusting Holzworth's15> 43'44
climatological mixing-depth values on the  basis of the stability class,  effec-
tively  eliminating  mixing  depth as a separate  parameter in the joint proba-
bility  function. This  seems to   be  a logical,  if  rather  inexact approach.
Deletion of the requirement for  temperature  soundings has a considerable
computational  advantage.  To  simplify  his  model. Pooler45  also  used the
principle of minimizing the number of variables, by eliminating stability and
Special  computations and tabulations  of this nature  are available from the National
 Climatic Center, Asheville, N. C.
                                                                     5-21

-------
taking the vertical  diffusion parameter to be a function only of wind speed.
We are currently  searching for  a  relationship between  mixing depth and
other  meteorological  parameters.  Since obtaining this relationship on the
basis  of physical  reasoning alone  does not  appear promising, we plan to
analyze 4 months  (1  month per  season) of surface and radiosonde data for
both  Washington, D.C., and St. Louis. The mixing-depth values computed by
the method developed for the synoptic model (see Appendix A) will then be
stratified  by time  classes and  correlated with the other meteorological para-
meters, particularly  stability.  If  the  correlations  are  significant,  the  result
should  be a  general  empirical expression  for  mixing depth  for  each time
class. If not, we will resort to the best alternative, adjustment of climatologi-
cal values.

If the planned number of classes and parameters are feasible, we will end up
with the following  computation:

            5      5     6   16
      C=  2    S     2    E     fdj.S,, Uk,0m)  Cdi.S,. Uk,0m). (14)
            i=l   j=l    k=l   m=l

This amounts to (5 x  5 x 6 x  16) = 2400 concentration calculations for each
point. Since  the diffusion model  orientation is  constant, for each  point, the
traffic links need only be scanned once, saving computer time.

Calculation of High-Percentile and Maximum Concentrations
The arithmetic mean  concentrations  (C)  for  a 1-hour averaging  time, ob-
tained by  the  method just described, can be used to  calculate estimates of
high-percentile and maximum concentrations by  means of Larsen's2  statis-
tical  model.*  Larsen  analyzed   3  years  of  concentration  data  for  seven
pollutants from CAMP stations in six  cities, and found that
1. Concentrations  are  approximately  log-normally distributed  for all  pollu-
   tants, in all  cities, and for all averaging times.
2. The median  concentration  (50  percentile) is proportional  to a power of
   the averaging time.

These conclusions were derived mainly from the concentration data from the
important region above the median  concentration.

Larsen  discussed certain significant  applications  of these  general findings.
Since  arithmetic mean  concentrations, ~C,  for  all averaging times  are  equal,
and  since the Concentration distribution is log-normal, the geometric mean
concentration Cg, for any averaging time t can be calculated from:
	 (Cg)t = C/(Sg), 0-5 ln(Sg)t.                               (15)
*Such calculations could also be performed directly using the predicted concentration
 joint frequency function, if the  latter were based upon a sufficiently large body of
 Meteorological data.
5-22

-------
C is available from the diffusion model calculations, and (Sg)t, the standard
geometric deviation  for a given averaging time, may be  obtained for CAMP
station  measurements of carbon monoxide from the tabulations by Larsen
and Burke.40*
The concentration (Cp)t associated with any percentile  p, of the frequency
distribution, for any averaging time t, may  be calculated from

      (Cp)t  =(Cg)t  (Sg)tq                                            (16)

where the number of standard deviations, q, is found  from  tables of the
standardized cumulative normal distribution, where:

      p/100  =   =ti     /*q  e~*2/2dx                            (17)
The maximum  concentration during a 1-year period for an averaging time to
t hours is designated by (Cmax)t; it is given by
                    	          o 01
      (CmaX)t   =   (Cg)l.hr  (Sg)?Ihr  tm                              (18)

where values of m as a function of Sg are those tabulated by Larsen.2

Using  these relationships  and the arithmetic mean concentrations furnished
by  the diffusion model, we can calculate,  for any averaging  time, (1) the
concentration expected for  any percentile and  (2) the maximum concentra-
tion expected once a year.  It is also straightforward to generalize (2)  above
to  the calculation  of maximum concentrations  expected  once during  any
given period.

PRELIMINARY RESULTS TO DATE
Status
At  the present time the basic  designs for both the synoptic model  and the
climatological model are fairly complete. The synoptic model,  less the street
and extra-urban  submodels, is programmed and  running.  Traffic data for
Washington, D.C., and St.  Louis have been  coded, and  data for three other
cities are  in the process of  reduction. We are currently  evaluating the model
output in terms  of its sensitivity to parametric changes and its response to
design refinements. Sample results of some of these trial computations with the
synoptic model are presented here, to furnish an idea of its eventual capabili-
ties. The  other models and submodels  and necessary  input data are in the
preparation stages.
*The values of (Sglj.f,,. for  CO measured at the CAMP Stations  in eight  cities range
 from  1.44 to 1.95, with a mean of 1.63. Application of the model to non-CAMP cities
 will require  an  estimate of Sg or its calculation  using  CO data. The probable factors
 having the greatest effect upon Sg are (1) the receptor siting, and (2) variability of the
 meteorological parameters. Consideration of  these factors could allow estimation of Sg
 for any citv. ')ased upon those calculated for the  CAMP cities.
                                                                     5-23

-------
Analysis and Display Techniques
To  facilitate  model development, we  are taking advantage of modern com-
puter  techniques,  particularly  objective  contour  analysis  and  graphical
display. In  this work we are  using special subroutines called CONTOOR and
GRAPH  4* with the CDC 6400 computer and the  CDC 280 cathode-ray
tube  (CRT) peripheral display system. The CONTOOR subroutine furnishes
objective isolines describing the  CO concentration distribution over an urban
area  on  the basis  of  a grid of point values calculated according  to the
diffusion  model. The  GRAPH  4/280 combination  permits such  contour
analyses  (or the results of any other calculations — see Figure 5-1, page 5-4) to
be  displayed on the CRT and photographed. This facility significantly expe-
dites the trial and evaluation procedure.
A  sample  product of  these techniques is given  in Figure 5-8.  Here the
                                                         I     I
    12

    10

     8

     6

     4

     2

     0

    -2

    -4

*   -6
x>
5   -8

   -10

   •12
     -12   -10   -8   -6   -4   -20    2    4    6    8   10   12
                  DISTANCE EAST OF CAMP STATION, miles     TD ,„, Qc
                                                           TD-/0/4-3S
 Figure 5-8.  Calculated  intra-urban concentration distribution
 of carbon monoxide (ppm)  over Washington, D.  C. - computer
 generated  contour analysis and display.
 to
 Q.
 O
 u.
 O
 cc.
 O
 O
            O.,
                 J	I..
_L
    0700-0800 LSI
    WIND270°/4ms-l
    MIXING DEPTH 200 m
    NEUTRAL STABILITY
J	I	I     I-	L
 "These subroutines were developed by S. Briggs and B. Sifford of Stanford Research
  Institute, respectively.
 5-24

-------
GRAPH 4/280 display of the CONTOOR objective analysis is superimposed
upon the primary traffic network for Washington, D.C. Good correspondence
between the CO concentrations and road density is evident. The analysis was
based upon  concentrations over a 25 x 25  section grid  of 625  points,  as
calculated by an early version of the diffusion model from simulated mete-
orological data.  The objective  contouring compares  well with  a  manual
analysis of the identical  data, presented in Figure 5-9.  Differences between
the two patterns were traced to plotting and  drawing errors by the analyst.
The  smoother hand  analysis  may  look better, but the objective analysis is
substantially faster, cheaper,*  and more accurate.
O
u.
O

o*
o
o
 10

  8

  6

  4

-2

  0

 -2

 •4

 •6

 •8

 -10

 -12
                                III!
           _L
_L
J	L
                                               0700-0800 LSI
                                               WIND 27074 m s-1
                                               MIXING DEPTH 200 nr
                                               NEUTRAL STABILITY
J	L
       •12  -10-8-6-4-2    0    2    4    68    10   12
                 DISTANCE EAST OF CAMP STATION, miles
                                                           TB-7874-8S

       Figure 5-9.  Same as  Figure 5-8, but hand contoured.
'Computer costs for the diffusion calculations in Figure 10 were about $20. The objective
 was an additional $5.
                                                                    5-25

-------
Model Sensitivity and Evaluation Trials
Figure 5-10 illustrates the type of tests  of  model  sensitivity to parameter
changes that  are  in  progress.  (This  and  all subsequent  graphs were also
produced  by the GRAPH 4/280 system.)  Tests  of this nature are helpful in
the  refinement  of the  model design.  The  figure  shows that  the  highest
concentrations at the Washington CAMP station  should occur for  west winds,
all other parameters being equal.
 o     _
 •a:
 cc.
 LU
 CJ
 Z
 o
 o
 o
 CJ
50
                        100      150      200       250
                          WIND DIRECTION, degrees North
300     350

      TA-7874-6
   Figure 5-10.  Effect of dilution factor and wind direction upon
   computed  concentrations at Washington, D. C.  CAMP  station.

 Figures 5-11  and 5-12  show calculated  CO  concentration  patterns for St.
 Louis  during  the morning  and  afternoon  of 16 October 1964, using  real
 meteorological data. Calculations over a 24- x 24-mile grid  centered on the
 CAMP  station are shown.  At the  CAMP  station,  observed concentrations
 were 19 and  5  ppm for 0700 to 0800  LST and  1500 to 1600  LST. These
 are substantially  larger than the values calculated at  the  center of the grids,
 (approximately 6 and 3 ppm for morning and afternoon). This is as  it  should
 be,  since these calculations are for the intra-urban  component (roof-level
 concentrations) only.

 The typical drop in concentrations  from early  morning to mid-afternoon  is
 apparent. It is also of interest to  note the reasonably high concentrations  in
 5-26

-------
                                       0600-0700 CENTRAL
                                        DAYLIGHT TIME
                                       15 OCTOBER 1964
                                       WIND 280V2 m sec-1
                                       MIXING DEPTH 122 m
                                       STABLE
               •8-6-4-20246
                 DISTANCE EAST OF CAMP STATION, miles
12
                                                             TA-7874-25S
Figure 5-11.  Calculated  intra-urban CO concentration distribution
(ppm) for St.  Louis area during early morning.

the morning near the junctions of freeways on the western side of St. Louis.

The same data as  Figure 5-12 were used earlier in  an attempt to furnish an
idea of the capability of the basic  intra-urban model to compute detailed
concentration  patterns  over  small areas. Those results are given in Figure
5-13, where the horizontal scale has been decreased by a factor of ten, and
the box size is now only 2.4  x 2.4 miles, still centered on the CAMP station.
The concentration pattern  shows a large spatial variability that  is related to
the distribution of downtown streets represented by the underlay.

The results of  trials using the synoptic model to calculate hourly concentra-
tions continuously  for  a week are presented in Figures 5-14 and 5-15. To
provide a  true test of the  model's  capabilities,  weeks  with  a  variety of
                                                                   5-27

-------
                                                1500-1600 CENTRAL
                                                  DAYLIGHT TIME   -\
                                                15 OCTOBER 1964
                                                VKIND 310V1.5 meters
                                                 1.5msec-1
                                                 MIXING DEPTH 1670m
                                                 UNSTABLE
    '--12  -10   -8    -6    -4    -2     0    2    4     6     8   10  12
                     DISTANCE EAST OF CAMP STATION, miles
                                                             TA-7874-26S

      Figure 5-12.  Same as Figure 5-11, but for late afternoon.
observed concentrations were selected. To account for a suspected  instru-
mental  zero-calibration  problem, the observations, indicated  by the dashed
curves in the figures,  have all been  adjusted  downward by  1  ppm. In an
attempt  to  more accurately  treat the effects of nearby sources, the  closest
diffusion area segment was divided into two sections  for these calculations,
One segment extends from the  receptor to 62.5 m in range, and the other
extends from 62.5 out to 125 m.

The CO observations during  the January  period  (Figure 5-14) were anom-
alously  high, and the model  generally gives underestimates  for this week. A
variety  of weather conditions prevailed; cold fronts passed  St. Louis on the
evenings of  17,  19, and  23 January.  The poorest performance of the model
occurs on 19, 21, and 22 January, when the wind was generally southerly.
5-28

-------
      •1.2
-1.0  -0.8  -0.6  -0.4   -0.2   0    0.2   0.4   0.6  0.8
         DISTANCE EAST OF CAMP STATION, miles
1.0   1.2
                                                              TA-7874-24S
    Figure 5-13.  Same as Figure 5-12,  but scale reduced by a
    factor of ten to give detailed CO concentration pattern  in
    central portion of St. Louis business district.

This suggests that the high observed concentrations are caused  by the nearby
sources on the adjacent streets, since the St. Louis CAMP  station is located
on the northwest corner of an intersection.
Weather conditions during  the October period  were not so variable, and
more  typical CO  concentrations were  observed.  In contrast to  the  January
period, the predicted concentrations are mostly too high. The poorest agree-
ment  occurs when the reported airport winds are very light or calm. For
these  cases, the  model  uses a wind  speed of  2 m sec"1  and the wind
direction  last observed. It is probable that  much of  the prediction error is
due to this uncertainty  in  wind  direction, part of  the general  difficulty
encountered in  applying  airport observations to  urban areas. Better results
                                                                    5-29

-------
might be  achieved for very  light-wind conditions if a symmetrically conver-
gent urban circulation, driven  by the heat island effect, were included in the
model.

No agreement was expected between the calculated and observed weekend
values, since no  weekend  traffic distribution  has  yet been  incorporated into
the model.  It is interesting, however, to note  the  low concentrations ob-
served on Sundays, and the late peak in the observed concentration at about
1100  on  Saturday,  23 January (Figure  5-14),  probably  due to shopping
traffic.
  25
  20
o.
z~
o
yio
O
o
o
o
          1700  08 17
      LSI LST
i  i  i
08 17
I!   I
08 17
08 17
08 17
 I   I
                                     I
17
                              OBSERVED  _
                              CALCULATED

                              WEEKEND—I
           20      40     60
      —Won -4—Tues—4—Wed-
       18 Jan    19 Jan    20 Jan
        80     100     120     140     IGOhours
       -Thurs—I-  Fri—4-— Sat  -I-  Sun —J
        21 Jan    22 Jan    23 Jan    24 Jan
         1965                      TA-7874-34S
Figure 5-14.  Hourly predicted and observed CO  concentrations
at St.  Louis CAMP station for period 18-24 January 1965.
5-30

-------
  25
  15
  10
   0
      08001700
      LSI LSI
08 17
I l  I
08 17
08 17
l  I  ' I
 08 17
  i   I
08 17
^i  IT
 08 17
                                                        OBSERVED  -
                                                        CALCULATED
     0       20     40      60
     k- Won -4— Tues 4— Wed
       19 Oct    20 Oct    21 Oct
         80     100
         Thurs—|— Fri -
         22 Oct    23 Oct
         1964
              120     140
              4- Sat-4
                  24 Oct
                    160 hours
                 I— Sun -4
                   25 Oct

                  TA-7874-23S
 Figure 5-15.  Same as Figure 5-14,  but for period 19-25 October
 1964.

SUMMARY AND CONCLUSIONS
The  model design described here is  believed to be a realistic and practical
approach toward developing capabilities for accurate and economical calcula-
tion  of: (1) hour-to-hour  CO  concentrations,  using synoptic weather data,
and (2) average, high-percentile, and maximum  concentrations, using climato-
logical weather  data and forecast traffic  data. The simplified  model design
makes full use of several  previous modeling efforts, with the incorporation of
necessary  and  appropriate modifications.  Unique  features of this model
include (1) an objective treatment of the  traffic data to give a  time-  and
space-dependent emission  inventory,  (2) a  technique for estimating  hourly
values of mixing depth,  and (3) a method (under development) for calculat-
ing  street-level concentrations.  So  far  we  have not used  adjustable "cali-
bration" constants in the model.

Final evaluation of the model must be deferred until all the sub-models,  and
any necessary refinements  are incorporated. To date the preliminary evalua-
tion  trials  have  been  instructive in  showing model  performance. Two
                                                                   5-31

-------
expected results are already apparent: (1) agreement with  observed (CAMP|
values  depends critically  upon  proper treatment of microscale effects, and
(2)  the steady-state model  in  its present form  gives poor results for light
wind conditions.

Since the model is designed to be general, only routinely available input data
are used. This means reliance  upon airport  weather data and rather gross
traffic  data, placing an  upper limit  on  the accuracy  obtainable with the
model. The substantial  effort expended in the development of techniques, to
estimate  the required model input variables,  using such  largely inappropriate
input data, underscores the need for  additional, routine, pollution-oriented
meteorological and  traffic measurements within cities.
ACKNOWLEDGMENT
This research  was jointly  sponsored by  the  Coordinating  Research Council
under  Contract  No. CAPA 3-68  (1-68)  and  by the  National Air Pollution
Control Administration under Contract  No.  CPA  22-69-64. We are grateful
for the assistance of the following individuals at Stanford Research Institute:
Mrs. Shirley  Reid,  who secured and supervised the reduction of the  traffic
data;  Robert Mancuso and Hisao Shigeishi, who programmed the model and
carried out the computer trials; and Elmer Robinson, who aided in planning
the model development.

We also  thank the  members of the CRC-APRAC Urban  Diffusion Project
Group for  their guidance and suggestions, and for supplying useful technical
information.
5-32

-------
 APPENDIX
DETERMINATION OF MIXING DEPTHS
 :OR USE WITH SYNOPTIC MODEL

                           F.  L.  Ludwig
 Initial attempts to estimate mixing depths for the synoptic model from surface
 observations only, were unsuccessful. The method finally chosen for estimating
'daytime mixing depths follows the commonly accepted approach20  and  re-
 quires knowledge of the early morning temperature sounding. The maximum
'afternoon mixing depth (hmax) is calculated as the height at which the dry,
'adiabatic lapse rate from the maximum surface temperature (Tmax) intersects
 the morning sounding. The calculations of mixing depth are made in pressure
 coordinates, and then converted to heights with the hydrostatic equation and
 the gas law.
 For nighttime hours, a minimum mixing depth (hn ) is estimated as follows. For
> daylight hours other than the time of maximum temperature, the mixing depth
 hd is obtained by linear interpolation between hmax and hn, according to the
 temperatureT:
                                hmax-hn    +   hn
              \ ' max ~' min

 The estimation of nighttime mixing depth over the city is  based on Sum-
 mers'24 model of the urban  heat island and Ludwig's48'49 empirical relation-
 ship between  rural lapse  rates  and the intensity of urban heat islands. If we
 assume T to be a linear function of the log of pressure (very nearly the same as
 Summers' assumption of  T as  a linear function of height), then the tempera-
 ture, Th , at  some height, h, outside the city is given by:

                     dT                     dT       / ph \
        Th  = T°  +  IT-  Alnp  ~  TO  +  P 7T~  '"         I     (A'2)
                    dlnp                    dp       \  Po /


 where T0 and p0 are the  rural  surface temperature and pressure, respectively,
 Ph is the pressure at h, and p  is the average of p0 and ph.
                                                                 5-33

-------
If air moving  into a city  is heated from  below to the extent that complete
mixing takes place  through a depth h, then the temperature at  height h will
remain Th and the lapse rate below h will become adiabatic. The dry, adiabatic
lapse rate 7d (in pressure coordinates) is given by
               TR             T               T
        7d  =  	   =  0.287  -    =   0.287  —
               pcp             p               P
                                                           (A-3]
where  R and cp are the gas constant and specific heat at constant pressure for
dry air, T  and p are absolute temperature and pressure, and T and p" are the
respective average values within the layer from the surface to  height, h. Forthe
urban situation, Equation (A-2) becomes
        Th =  T,,
                      In
_Ph
Po
                                                           (A-4]
where Tu  is the surface air temperature in the center of the urban area.

Subtracting Equation (A-4) from Equation  (A-2) gives:
Tu  -
                    = p
                           In
      dT
      dp
(A-51
Substituting Equation (A-3) into Equation (A-5) gives:

                            Pn
            -  T0  =
                In
   dT            _\
 p  	  -  0.287 T
(A-6)
Ludwig48 has shown that (Tu -  T0) can be approximated within about 2° C
by the following equation:
        Tu -  T0  =  p ' 4   0.0633  -  0.298
                                       dT
                                       dp
                                 (A-7)
where  P  is the P9pulation of the urban area and dT/dp is expressed in "Cper
millibar.  The equation is based on 85 sets of data from 18 different cities.

Substituting Equation (A-7)  into Equation (A-6) gives:
             Pn
             Po
                  _  pl/4
                                         dT
                       0.0633 - 0.298   —
                        	     dP
                        -dT           _
                        P-	0.287  T
                          dp
                                 (A-8)
5-34

-------
The value of dT/dp is determined  on the basis of the lowest portion of the
morning sounding between the surface and the first reported level above the
surface; JD and T are the averages  of the pressures and temperatures  at these
points.
To convert Inl —   to mixing depth, we can use the thickness equation.
             Vv
If temperature is assumed equal to T throughout the layer then:
                                                                  (A-9,
and Equation (A-8) becomes
        hn =
29.3 T  P1/4 (0.0633 -  0.298 —
     	dp
      p 4^~  ~  0.287 T'
        dp
 In applying the preceding equations to the calculation of mixing depth for the
 diffusion model, we use Equation (A-1) for the first hour after sunset. Between
 this hour and midnight, the mixing depth is interpolated by time between the
 values given by Equations A-1 and A-10. The latter is the nighttime minimum
 mixing depth, hn,  which is assumed  to apply for all other night  hours. The
 values of the  parameters in Equation (A-10) are determined from the sounding
 for the next morning.

 The  assumption of a constant  urban mixing  depth throughout the  early
 morning hours is based on observations presented by Ludwig and  Kealoha49
 for Dallas, and Ft.  Worth, Texas. These observations show that the urban heat
 island develops rather quickly in the evening. If  the urban heat island and the
 urban mixing  layer are  related,  as suggested by Summers' model, then the
 mixing depth  should reach  its  nighttime  value by  the middle of the  night
 similar to the heat-island  intensity.

 To take care  of  those instances  when the  calculated  mixing depth  is
 unreasonably  high or low, we assigned limiting values of 4000 m maximum or
 50 m minimum.
                                                                   5-35

-------
REFERENCES
 1. Clarke, J. F. A  Simple Diffusion Model for Calculating Point Concentrations from
    Multiple Sources. J. Air Pollution Control Assoc.  74:347-352, September 1964.
 2. Larsen, R. I. A  New Mathematical Model  of  Air Pollutant Concentration, Averag-
    ing Time and Frequency. J. Air Pollution Control Assoc. ?:24-30, January 1969.
 3. Wanta, R.  C. Meteorology and Air Pollution. In: Air Pollution, Stern, A. C. (ed.);,
    Vol. 1, 2d  ed. New York, Academic Press,  1968.  p. 187-226.
 4. Stern, A.  C.  The  Systems  Approach  to  Air Pollution  Control.  North Carolina
    University,  Dept. of  Environ. Science  and  Engineering. Chapel  Hill.  Publication
    Number 216. 1969. 51 p.
 5. Moses, H.  Mathematical Urban Air Pollution  Models. National Center for Air Pollu-
    tion Control, Chicago  Dept. of  Air Pollution Control Argonne National Labora-
    tory. Argonne, III. ANL/ES-RPY-001. April 1969. 69 p.
 6. Neiburger, M.  Diffusion Models  of Urban Air Pollution.  In:  Proceedings of WMO
    Symposium  on  Urban Climates  and  Building  Climatology,  Brussels,  Belgium,
    October 15-25,  1968. to  be published  by World  Meteorological  Organization.
    Geneva, Switzerland. In press. 17 p.
 7. Ott, W., J. F Clarke, and G. Ozolins,  Calculating  Future Carbon Monoxide Emis-
    sions and Concentrations from Urban Traffic Data. National Center for Air Pollu-
    tion Control. Cincinnati, Ohio. PHS Publication Number 999-AP-41. June 1967. 40
    P.
 8. Swinnerton,  J.  W.,  V. J.  Linnenbom,  and C. H. Cheek.  Distribution  of Methane
    and Carbon Monoxide Between the Atmosphere and Natural Waters. Environ. Sci.
    Technol. 3:836-838, September 1969.
 9. Went,  F.  W.  On  the Nature of Aitken Condensation  Nuclei.  Tellus. 78(2-3):
    549-556, 1966.
10.  Robinson,  E.  and  R. C.  Robbins. Sources, Abundance and Fate  of Gaseous
    Atmospheric  Pollutants.  Stanford  Research  Institute.  Menlo Park, Calif.  Final
     Report,  SRI Project PR-6755. 1968. 123 p.
11. Washington,  D.  C., Cincinnati, Ohio., Metropolitan Area  Air Pollution Abatement
    Activity. National Center for Air  Pollution Control. 1967. 46 p.
12. Chass,  R.  L. et  a/.  Total  Air Pollution Emissions in Los  Angeles County. J. Air
    Pollution Control Assoc. 70(51:351-366, October 1960.
13. Report  for Consultation  on the  San   Francisco Bay  Area  Air  Quality  Control
    Region.  National Air Pollution Control  Administration.  1968.
14.  Rose,  A. H.  et al.  Comparison of Auto Exhaust  Emissions from Two Major Cities.
    J. Air Pollution  Cc-ntrol Assoc. 75:362-366, August 1965.
15.  Holzworth, G. C. Estimates of Mean  Maximum  Mixing Depths in the Contiguous
     United States. Mon. Weather Rev. 32:235-242, May 1964.
16. Smith, Wilbur  and  Associates.  Transportation  Survey—National Capital  Region.
    Prepared for National Capital Planning Commission.  New Haven, Conn. 1958.  125
    P.
17. Gifford, F. A.,  Jr.  Use  of  Routine  Meteorological  Observations  for  Estimating
    Atmospheric Dispersion. Nucl. Safety. 2:47-51, June 1961.
18. Pooler, F., Jr.  A Tracer Study of  Dispersion Over a City. J. Air Pollution Control
    Assoc.  76:677-681, December 1966.
19. McElroy, E J.  L. A  Comparative  Study  of  Urban and  Rural Dispersion.  J.  Appl.
    Meteorol. S(1):19-31, February 1969.
20.  Miller,  M.  E. and  G. C.  Holzworth. An Atmospheric Diffusion Model for Metro-
    politan Areas. J. Air Pollution Control Assoc. 77:46-50, January 1967.
21. Smith, M.  E.  and I.  A. Singer. An Improved  Method of Estimating  Concentrations
    and Related Phenomena  from a  Point  Source Emission. J. Appl  Meteorol  5(5):
    631-639, October 1966.
 5-36

-------
22.  Meade, P. J.  Meteorological  Aspects of the Peaceful Uses of Atomic  Energy. Part
    1. Meteorological Aspects of the Safety and Location of Reactor Plants. World
    Meteorological Organization.  Geneva, Switzerland.  Publication  Number 97 TP41.
    Technical Note Number 33. 1960. 33 p.
23. Turner, D. B. A  Discussion Model  for  an  Urban Area. J.  Appl. Meteorol. 3(1):
    83-91, February 1964.
24. Summers,  P.  W. The Seasonal, Weekly, and Daily Cycles of Atmospheric Smoke
    Content in Central Montreal. J. Air Pollution Control Assoc.  75:432-438, August
    1966.
25. Graham, I.  R. An Analysis of Turbulence Statistics at Fort Wayne,  Indiana, J.
    Appl. Meteorol. 7(1):90-93, February 1968.
26. Schnelle,  K.  B.,  F. G.  Ziegler,  and P. A.  Krenkel.  A Study of the Vertical
    Distribution of Carbon Monoxide and Temperature Above an  Urban Intersection
    (APCA Paper No. 69-152). 1969. 65 p.
27. Clarke, J.  F. Nocturnal  Urban  Boundary  Layer Over  Cincinnati,  Ohio.  Mon.
    Weather Rev. 37(81:582-589, August, 1969.
28. Bornstein, R. D. Observations of the Urban Heat  Island Effect in New York City.
    J. Appl. Meteorol. 7(4):575-582, August  1968.
29. Leahey, D. M. A  Model for Predicting the Variation of Pollution Within the Urban
    Heat Island (APCA Paper No. 69-107). 1969. 45 p.
30. Angell, J.  K. ef al.   Tetroon  Trajectories  in  an Urban  Atmosphere. J.  Appl.
    Meteorol. 5(51:565-572, October 1966.
31. Hass, W. A. ef al. Analysis of Low-level, Constant  Volume (Tetroon)  Flights Over
    New York City. Quart.  J. Roy. Meteorol. Soc. 33(3981:483-493, October 1967.
32. Pooler, F.,  Jr. Airflow  Over  a  City   in Terrain  of Moderate  Relief. J.  Appl.
    Meteorol. 2(4):446-456, August 1963.
33. Hilst, G. R. and N. E. Bowne. A Study of the Diffusion of Aerosols Released from
    Aerial  Line Sources Upwind of an Urban Complex. Travelers Research Center, Inc.
    Hartford, Conn.  Final  Report. Vol.  I. Contract DA 42-007-AMC-37(R).  1966. 241
    p.
34. Turner, J. S. The  Motion  of Buoyant Elements  in Turbulent Surroundings. J. Fluid
    Mech.  70:1-16, May 1963.
35. Lilly, D. K. Models of Cloud-topped  Mixed Layers Under a Strong Inversion.  Quart
    J. Roy. Meteorol. Soc.  94(401 ):292-309, July 1968.
36. Georgii, H. W., E. Busch, and E. Weber. Investigation of the Temporal  and Spatial
    Distribution of the Emission Concentration of  Carbon Monoxide in  Frankfurt/
    Main.  Frankfurt/Main  University,  Institute  for  Meteorology  and  Geophysics.
    Report Number 11; Translation Number  0477, NAPCA.  1967. 66  p.
37. McCormick, R. A. and C. Xintaras.  Variation of Carbon Monoxide Concentrations
    as  Related to Sampling  Interval, Traffic, and Meteorological Factors. J.  Appl.
    Meteorol.  7(2):237-243, June 1962.
38. Rouse, H.  Air Tunnel Studies of Diffusion in Urban Areas.  Meteorol. Monogr.
    7(4):39-41, November 1951.
39. Highway Statistics for 1966. Bureau  of Public Roads. Washington, D. C.    1968..
    186 p.
40. Larsen, R.  I. and  H. W. Burke. Ambient  Carbon Monoxide Exposures. Presented at
    62d Annual   Meeting  of the Air  Pollution  Control Association.  New  York.
    (NAPCA) June 22-26, 1969. 38 p.
41. Martin, D. 0. and  J.  A. Tikvart.  A General Atmospheric Diffusion  Model  for
    Estimating the Effects  on Air Quality of One or More  Sources. Presented at 61st
    Annual Meeting  of  the Air  Pollution Control  Association,  for NAPCA, St. Paul.
    June 1968. 18 p.
42. Frankel,  M.  L. Regional  Air Pollution  Analysis—Phase I. TRW Systems Group.
    Redondo Beach, Calif. Status  Report. Sales Number 11 130 000, 1965. 467 p.
                                                                           5-37

-------
43.  Holzworth, G.  C.  Mixing Depths,  Wind Speeds and Air Pollution Potential for
     Selected  Locations  in the  United States.  J.  Appl.  Meteorol.  6(6): 1039-1044,
     December 1967.
44.  Holzworth, G.  C.  Large-scale  Weather  Influence on Community Air  Pollution
     Potential in  the United States.  J. Air Pollution Control Assoc. 79:248-254. April
     1969.
45.  Huber,  M. J.,  H.  B.  Boutwell, and  D. K.  Witheford.  Comparative Analysis of
     Traffic  Assignment Techniques with  Actual  Highway  Use. Highway  Research
     Board. Washington,  D. C. National Cooperative Highway  Research Program Report
     58.  1968. 85  p.
46.  Miller, M. E. Forecasting Afternoon  Mixing  Depths  and  Transport Wind Speeds.
     Mon. Weather Rev. 95:35-44, January 1967.
47.  Ludwig, F  L. Urban Air Temperatures and Their Relation to  Extra-urban Meteoro-
     logical Measurements. Presented at Meeting of the American Society of Heating,
     Refrigerating,  and  Ventilating  Engineers for  Stanford  Research  Institute.  San
     Francisco. January 1970.
48.  Ludwig,  F  L. Urban  Temperature Fields. In: Proceedings of  WMO Symposium on
     Urban  Climates and Building Climatology, Brussels, Belgium, October 15-25, 1968.
     To  be published  by World  Meteorological  Organization, Geneva, Switzerland. In
     press. 23 p.
49.  Ludwig,  F.   L.  and  J.  H.  S.  Kealoha. Urban Climatological Studies.  Stanford
     Research  Institute.  Menlo   Park,  Calif.  Final  Report.  Contract OCD-DAHC-
     20-67-C-0136. (Work Unit 1235A). March 1968.  191  p.
50.  Gilman  and Co. St. Louis Metropolitan  Area Traffic Study. Prepared for City and
     County of St. Louis. New York. 1959. 154 p.
5-38

-------
     AN URBAN  DISPERSION MODEL
ABSTRACT
     A  computerized,  multiple-source,  atmospheric  dispersion model de-
     signed for operational use in air resource management has been formu-
     lated and  programmed for the IBM 360-75  system.  This "integrated
     puff" model provides a more realistic physical simulation of  the pro-
     cesses of smoke plume dispersion than has  hitherto been employed,
     since it provides  for simulation of near-zero wind  speed conditions,
     models three-dimensional wind  vector variation and atmospheric diffu-
     sion, including upwind-downwind diffusion, and permits variation of sta-
     bility and mixing layer depth with time.
     This paper describes the development and preliminary validation testing
     of the model with  data  from a  3-year, computerized inventory of
     sulfur dioxide air quality data recorded by  the receptors of  the Chi-
     cago, Illinois,  telemetered air monitoring system.
     A  detailed, 2-year inventory  of  Chicago coal and  oil  burning SO2
     emission sources was acquired, data storage formats were designed, and
     computer algorithms were developed  to  generate  hourly average esti-
     mates of emissions from  major utility, industrial, residential, commer-
     cial,  and  institutional sources. Small emitters  were aggregated into
     square-mile area sources.
     The model consists of a series of algorithms assembled around a kernel
     that  represents the  transport and  diffusion  of pollutant species  from
     point and area sources according  to  a  Gaussian  distribution  in three
     dimensions. This kernel,  which represents a three-dimensional puff of
     smoke, is  integrated according to  a  time-series of piecewise constant
     wind vectors  and piecewise constant  atmospheric stability parameters
     to simulate  the  transient  behavior   of  a continuous smoke  plume.
     Smoke plume rise and downwash phenomena are simulated.
     The model is incorporated in a master air pollution  data management
     system, which is  employed  to store, retrieve, process,  analyze, and
     display emission, meteorology, and air quality data. This system is used
     to study  micrometeorological  dispersion, to  form and input  data ar-
     rays, and to validate its own theoretical predictions against recorded air
     quality data.
     At present, the model is  in the early  stages of validation and has been
     tested against  1 month of hourly-average S02 data  from five Chicago
     air quality monitoring stations. A statistical  sample  of approximately
     2300 data points  is presently available. The ratio of standard deviation

-------
      to mean values  for all hourly  SO2 predictions is 0.93.  For 6-hour
      average  predictions,  this  ratio  is 0.64 and for 24-hour average predic-
      tions it drops to  0.43.  Over 66%  of the 24-hour average S02  pre.
      dictions are  within ±0.05 part  per  million (ppm) of observed values
      and approximately 90% are within  ±0.1 ppm of actual dosages.
      Hourly  time  series plots  of observed and  actual S02  concentrations
      indicate that the model is sufficiently accurate to  test selective hypo-
      theses  about the  model itself  and  to   evaluate  urban air resource
      management strategies.
AUTHORS
      JOHN  J.  ROBERTS, the  lead engineer for the development of computerized
      mathematical models of atmospheric diffusion  processes for the Chicago Air
      Pollution Systems Analysis Program. CAPSAP is an interdisciplinary, environmen-
      tal science research and development program supported by  the Department of
      Health, Education and Welfare, the Atomic  Energy Commission and  the City of
      Chicago.  It is  directed towards  the  development  of physical and economic
      mathematical models for  use  in air resource  management planning and air
      pollution  systems  analysis. Dr. Roberts  is a major participant in the Argonne
      study of  the design  and  implementation  of episodal and long-range emission
      control  programs.  He  recently organized a  program  of research  to develop
      numerical models of hydrothermal plumes from  power plants and has also been
      associated with  the Argonne-U.S.  Navy magnetohydrodynamic research program.
      Dr.  Roberts received a B.E.E.  from Rensselaer Polytechnic Institute in 1958 and
      then spent  four years  teaching in the  U.S.  Navy Nuclear Power program. He
      received a Ph.D. in nuclear engineering at the University of California, Berkeley,
      in 1966.

      EDWARD J. CROKE is the  director of the CAPSAP. Prior  to joining Argonne
      National Laboratory in 1963, as a systems analyst for the NASA refractory metal
      nuclear rocket  program, he  was with the James Forrestal Research Center at
      Princeton, and  was later propulsion systems project officer for the Agena Satel-
      lite program of  the Air Force Systems Command.
      He holds a B.S.E. in physics from the University of Illinois and an M.S.E. from
      Princeton  in 1960. He has subsequently pursued postgraduate studies in geophy-
      sical sciences, operations research, systems analysis and management sciences at
      the University of Chicago and Northwestern University.

      ALAN S.  KENNEDY received an M.S. in mathematics at Washington University
      in 1962.  He joined  the   Applied  Mathematics Division of Argonne National
      Laboratory in  1963 and has continued his studies toward a Ph.D. in operations
      research at Northwestern University.

      From his study of data  management problems associated with an operational air
      pollution  control department,  he developed a system  of storage, retrieval, and
      analysis of air quality, meteorological, and emission data for CAPSAP.

-------
                               6.   AN  URBAN ATMOSPHERIC
                                            DISPERSION MODEL
               JOHN J. ROBERTS,  EDWARD  S. CROKE
                        and ALLEN S. KENNEDY

                      Argonne National Laboratory
INTRODUCTION
A  computerized,  multiple-source,  atmospheric  dispersion  model  has  been
formulated and  programmed for the  IBM 360-75  system. This "integrated
puff"  model  provides a more realistic  physical simulation of the processes of
smoke  plume dispersion than has hitherto been employed, since it provides
for simulation of near-zero wind speed conditions,  models three-dimensional
wind vector variation and atmospheric diffusion, including upwind-downwind
diffusion, and  permits  variation of  stability and  mixing layer  depth  with
time.

A source-oriented  model is essential for the development of optimal incident
control strategies and  long  range abatement plans. For example,  a prelimi-
nary version  of the integrated puff model was programmed and subsequently
employed to develop a prototype optimal control model for an urban power
plant  network.1 That  prototype model will  be used in combination  with
Chicago's emission inventory data as the primary  analytical tool in the
performance  of a series of air pollution system-analysis studies including:
   1. Cost-effectiveness analysis of alternative control strategies for specific
     source-aggregates such  as power plants, the food processing industry,
     the steel industry, asphalt batching, etc.
  2. Development  of  rapid-response, automated  optimal  incident control
     strategies for the urban power plant network, optimal gas allocation to
     dual  fuel sources, process emission control,  etc, based on operations
     research techniques.
  3. Development  of   long-range  air pollution abatement  plans oriented
     around the computer simulation of the city, with projections  of its
     population, fuel  use,  production and  transportation network growth
*Work  jointly  sponsored  by  the  National Air Pollution  Control Administration, the
 Atomic Energy Commission, and the Chicago Department  of Air Pollution Control.
                                 6-1

-------
      patterns, so that emission control legislation, zoning ordinances, etc.,
      can be tested  on a  simulation of Chicago urban area of the future, and
      cost-effectiveness studies  can  be performed to evaluate the economic
      impact  of  pollution abatement measures.
The model described in this paper has several aspects in common with earlier
efforts2-3-4 (e.g.  Koogler, 1967; Turner, 1964; Clarke, 1964):
   1. An  emission  inventory is  developed;  large emitters are  considered as
      point-sources and smaller residential commercial and industrial emitters
      homogenized as area-sources  on a standard grid.
   2. A  mathematical  algorithm (typically  based on the  representation of
      turbulent diffusion  by a Gaussian distribution) is used to describe the
      advection and diffusion of plumes from each  of the sources or source
      areas.
   3. Contributions from  individual plumes  are  summed  and  sometimes in-
      tegrated over time to give pollution doses at arbitrary locations.
   4. The  model  is validated  by statistical  comparison  with field  measure-
      ments taken by an array of monitors.
   5. While the results of the validation  study may  indicate a refinement of
      certain  physical assumptions  (such  as  the  plume  rise formulation) or
      inclusion of  additional  sources (such  as  distant  power  plants), no
      further  improvement by linear  regression or similar  curve fitting tech-
      niques is  permitted; that is,  the model  is not tuned to a  particular
      receptor or city.  This last point is particularly important if the model
      is to  be used  in the development  of  incident control strategies and
      long-range urban planning.

The essence of an air pollution  dispersion  model as a collection of algorithms,
a set of assumptions and  subsequent calculations for every stage of the analy-
sis, from  the  compilation of the emission inventory to the validation of the
results, is discussed briefly  in the following sections.

DEVELOPMENT OF AN EMISSION INVENTORY
A major task element of the Chicago Air Pollution Systems Analysis program
was  the  development  of  an  inventory of hourly average emissions for the
power plants  and industrial,  residential, commercial, and  institutional  S02
sources that  are the  primary  contributors to  Chicago's sulfur  oxide air
pollution problem.

Figure 6-1  shows the location and distribution of the largest individual coal
and oil burning sources of  sulfur oxides, while Figure 6-2  shows the source
density of  residential emitters. They were too numerous to inventory indivi-
dually.  Figure  6-3  depicts, by  source category, the distribution and  magni-
tude of Chicago's SO2 emitters.
6-2

-------
POLLUTION INCIDENT
  CONTROL PLAN    —
 A  MULTIPLE RESIDENCES      j
 •a  COMMERCIAL ESTABLISHMENTS
 O  INDUSTRIAL PLANTS
 ©  INDUSTRIAL PLANTS WITH DUAL..
      FUEL CAPABILITIES          _   ^_
 |  PUBLIC UTILITY POWER PLANTS  '   ' '  i -
 *  TELEMETERING STATIONS
                                                   '
   Figure 6-1.  Map of major Chicago sources and TAM stations.
                                                          6-3

-------
  TONS OF EMISSIONS PER YEAR
      i	1    0-  -200
            200 -   500
            500-1,000
           1,000-1,500
           1.500 - 2,500
                                                                  :
  Figure 6-2.  Sulfur dioxide from residential space heating, 1968.
6-4

-------
o>
ui
                     GO O


                     LU "o
i?n
en

in
^n
on
in
n

i^
5



9,123
1


5,55
11
:|:|:|x;
1
111
:-:|>:;:






14,974 14,90
85,362 5,832 pi
^^^ M*«B ••:•:•:•:•
•^^HW •••" -1 BlvI'X
Pi n pip






3
4,238
PI





2,2;
II
:•:•:•;•:



4

9

4 275



i
11

1




7

2

826





0,00
1
: S
3-




)

25
'3,12
Fl
^
m
:::::::::
II
III
vXv
I
ii
0






22
                              s
                                                   ,
                                    i^  ujl- uj^  LUd
                                   3
    s
3
                              a 
-------
The Chicago  Air Quality Monitoring System
Chicago  operates  a  network  of eight telemetered  air  monitoring (JAM)
stations that continuously  and automatically record 5-minute, average S02
concentrations  at  15-minute intervals throughout the day. The data  recorded
by these stations during 1966 and 1967 are typical of a city  that is charac-
terized  by a  highly  unhomogeneous S02  source  distribution and density.
Table  6-1  provides a summary of the air quality situation of Chicago as
recorded by the network of TAM stations.

Since the source density, type, and distribution in the immediate vicinity of
a receptor may, under certain meteorological conditions, dominate the am-
bient air quality recorded,  it is of value to examine the  record  of the eight
TAM  stations  from  the  standpoint of  their proximity  to  Chicago's S02
sources. This exercise is particularly  useful, since it provides an indication of
what types of sources  are likely to  have  the  predominant  effect on the
recorded  air quality, and  therefore  serves  as an  indicator of the level of
effort  that should be expended to  develop a  detailed emission inventory.
Moreover,  an  understanding of the distribution of sources relative to each of
the  monitoring  stations  that  provide the  data  required   to validate the
Chicago atmospheric  diffusion model explains certain of the results predicted
by  the model.  Table 6-1  shows the distribution of  1-hour S02 dosages re-
corded at the eight TAM stations.

  Table 6-1.  TOTAL NUMBER OF RECORDED 1-HOUR  SO2 DOSAGES (1966-67)
TAM station
1
2
3
4
5
6
7
8
Total dosage, ppm
0.2
182
618
3210
1990
525
364
934
206
0.3
59
203
1838
966
207
169
332
51
0.4
18
70
975
491
84
94
122
15
0.5
5
30
483
268
45
65
35
6
TAM Station  1
TAM  station 1  is  located near the northwestern limits of Chicago in an area
characterized by residential neighborhoods of relatively  recent construction,
and a limited amount of light industrial  development. Heating plants in this
sector  of the city  are   predominantly natural-gas fired  and  relatively little
coal  or  oil is consumed  by industrial  or  commercial sources.  TAM  1  is thus
sited  in a  relatively  clean  section  of  the city,  insofar  as  sulfur  oxide
6-6

-------
emissions  are concerned.  This conclusion  is  borne out by  the fact that
recorded high ambient  S02  concentrations at JAM 1 are at least one order
of magnitude less frequent than at monitoring stations located nearer to the
urban core area.

TAM Station 2
Station  2, located in the northeast sector of the city approximately 1.5 miles
from the  lakeshore, is  in  a mixed industrial-residential  area that  contains a
large concentration of high-rise gas-fired residential structures.  The prevailing
southwesterly wind flow  tends to transport S02  from  the central  utility-
industrial  cluster into this area, but average dosages tend  to be relatively low
because of the  comparatively  large  transport  distances involved. This  is
evidenced by  the relatively high ratio of low to  high concentrations recorded
at TAM 2  compared to  TAM 3 or  TAM  4 where  localized coal and  oil
burning sources are more  significant. The nearest major S02 source to TAM
2 is a moderate sized coal fired power plant  located approximately 1.5 miles
southwest.

TAM Station 3
This receptor is centrally  located atop a large office building in the Chicago
loop business  district.   It  is  not  only surrounded by a  cluster of  large
commercial and  institutional  space-heating  sources, but it is also situated
directly northeast of a  major  concentration  of industrial  plants  and  a  line
formed  by  the  three largest power plants within  the city limits. Since the
prevailing wind flow is southwesterly, a high incident frequency due to this
industrial-utility  concentration and  to local  space heating  effects is  to  be
expected at TAM 3. As indicated in Table 6-1 this station is, in fact, located
in the most polluted sector of the city.


TAM Station 4
TAM  4 is located  in the  south-central Hyde Park  area of Chicago within 1
mile of Lake Michigan. The  immediate area is characterized by a row of
highrise apartments sited  along the lakefront,  a large concentration of old
coal-  or oil-fired low-rise, six-flat apartment buildings  and a single  major
source  - the  University of Chicago  steam plant. TAM 4 is approximately 5
miles  southeast  of Chicago's central  industrial-utility complex. As indicated
in Table 6-1,  the frequency of high, recorded SO2 concentrations at TAM 4
is  second  only to that  of TAM 3. Statistical analysis of TAM 4  air quality
data has indicated that the air quality in this area  is very strongly correlated
with ambient temperature - a finding that  indicates  the Hyde  Park  area is
largely self-polluted by  local, residential space-heating sources.
                                                                       6-7

-------
JAM  Station 5
TAM  5 is situated approximately 6  miles  inland in the south central  sector
of the city. It is approximately 3 miles south of the central industrial-utility
complex,  atop a  school  building in a neighborhood of low-rise apartments
and  single family  dwellings.  Residential  buildings in  the TAM 5  area are
somewhat newer than in TAM 4, and gas tends to predominate over  coal and
oil  for space-heating  purposes. TAM  5 is nearer  to the central industrial-
utility  cluster than any  of  the other monitoring  stations;  however, only a
few large sources are  located southwest of the site. This fact, combined with
the relatively low  level  of  residential emissions  in its  immediate  vicinity,
results in comparatively modest number of high S02  concentrations  recorded
there.

TAM  Station  6
Station 6, in  the extreme southeast of the city, is sited in a residential area
located  approximately  5  miles west  of  a second large concentration of
Chicago  industrial and  power  plants.  Immediately adjacent  to  the Chicago
source-cluster is the Gary-Hammond industrial complex. Since the receptor is
also located  within one mile of the  southwest city limits of Chicago, it is
exposed  to  emissions from two areas for which an emission inventory is not,
at present, available. According to the data presented in Table 6-1, TAM 6 is
located in one  of the  less polluted  areas of the city.  This conclusion is
consistent with  the fact that the metropolitan region southwest of the city is
relatively  free of  large S02  sources, and that east winds,  which would bring
in  SO2  from the south  Chicago-Gary-Hammond area,  are comparatively
infrequent in the midwest Great Lakes region.

TAM  Station  7
TAM  7 is sited  atop a school building within  0.5 mile of the  western city
limits  and  almost 7 miles directly west  of the  central  business  district. The
immediate  neighborhood is  predominantly residential, although  a few large
industrial  plants  are located within 2  miles of the receptor  in the northeast
and southwest directions. The  station  is approximately 6  miles northwest of
the central industrial-utility cluster.
A detailed emission inventory for the  regions immediately west and  south of
this receptor  is not yet available,  since these  areas lie outside the Chicago
city limits.  It is known, however, that the area west of  TAM 7 is  a  mixed
residential, light-industrial sector in which gas  heating  predominates, while
the highly industrialized city of Cicero, Illinois,  lies immediately  south of
TAM  7, within 2 miles of the receptor.

Although the relative  scarcity of southeasterly winds minimizes the influence

-------
 if the central Chicago  source-cluster on JAM 7, the area experiences signifi-
ilpant S02 concentrations  with moderately high  frequency, making  it the
iiihird or fourth most frequently polluted site.

 TAM Station 8
 _ike the TAM 6  receptor, TAM 8 is  located  in a  residential neighborhood
Sear one of  the southwestern extremities  of Chicago's  irregular  western
border.  Two relatively  large industrial  plants  lie within 1 mile northeast of
'the site; however, the area itself is relatively free of significant S02  sources.
'Since the western city limits  are  within  0.5 mile  and  the southern city
'limits are within  1 mile, emission data  for virtually the entire  southwestern
quadrant, relative to TAM  8, are not available.

As indicated in Table  6-1  the TAM 8 receptor is sited  in one of the least
polluted areas of the city.

Power  Plants
 Electric power for the  Northern Illinois region is provided by  a network of
 15 power generating stations, of which  six are  located in the Chicago metro-
politan  area. These  six coal-fired plants, shown in  Figure 6-1  account for
approximately  65%  of the sulfur  oxides  released  into  the Chicago  atmo-
sphere.  Table 6-2 indicates the relative contribution of each plant in terms of
 its annual fuel consumption.
 Table 6-2.  RELATIVE CONTRIBUTION3 OF NORTHERN ILLINOIS POWER PLANTS
                    TO CHICAGO SO2 CONCENTRATION

Plant designation
Crawford
Ridgeland
Fisk
State Line
Northwest
Calumet
Total
SO2
Amount, tons yr~'
107,954
1 00,540
73,438
70,606
10,566
10,016
373,120
& of Chicago total
18,97
17.67
12.90
12.42
1.86
1,76
65.58
     In terms of annual fuel consumption

Of these installations,  four:  the Fisk, Crawford, Ridgeland, and State  Line
plants  are capable  of  partial or total  conversion from coal  to natural gas
during  periods when the latter fuel  is available at "dump" or "interruptible"
rates.
                                                                       6-9

-------
Because  of the extremely large  emission rates associated  with the urban
power plants and  because of  their complex,  but relatively consistent diurnal
emission cycle,  the acquisition of the most accurate body of emission data
available constituted a high-priority  task  in the emission inventory. Forty.
nately,  the Chicago  utility maintains excellent records of plant operation.
These were provided for  the  period  January 1966  through December 1967.
The data array required was:
   1.  Hourly power generation, in megawatts (Mw), for each generator unit
      of  each  power plant, including an indication of whether the power was
      generated with coal or gas.
   2.  Monthly coal  and  natural gas consumption rates for each boiler-genera-
      tor combination.
   3.  Monthly average coal-sulfur-content analyses for each plant.
   4.  A  set of boiler-turbine-generator efficiency curves for  each generator
      unit. These presented efficiency vs. power  output for each unit.
   5.  A set of stack temperature vs. output curves for each plant.
   6.  A physical  description of each  plant layout,  including the relationship
      between boilers, turbines, generators, and stacks, stack heights, etc.

The  above  information  was combined  in a computer algorithm for calcula-
tion  of the hourly fuel consumption and S02  emission rates for each stack
of each  plant of  the  system. Details  of  the  computational procedure are
provided in the program progress  reports.5

The  resultant body  of power plant emission data  was stored in  the master
data  file of the Air Pollution  Information and Computation System (APICS)
from  which  it could be  retrieved and automatically  input to the dispersion
model.  The  details  of the APICS  system are described by Kennedy and
Anderson,6 and in  the program progress reports.5' 7

Industrial  Sources
There are over 2500 relatively large coal  and oil burning industrial plants in
the City of Chicago, which are responsible for approximately 10 percent of the
average annual sulfur oxides emitted  within the city limits. The  100 largest
plants account for  over  83 percent of the total industrial emissions, and the
largest 50, for over 64 percent.

The inventory of industrial emissions was derived from four sources:
   1.  Department of Air Pollution Control Survey. A Comprehensive annual
      average  inventory compiled in  1963 by the Chicago Department of Air
      Pollution Control  was computer processed to establish industrial SOj
      emissions  on  a square-mile basis. Within any given square-mile sector,
      industrial emissions were assumed to be uniformly distributed.
   2.  Argonne National Laboratory Field Survey. Detailed fuel  consumption
      and  physical  and operating cycle data  for the  50 largest S02  sources
6-10

-------
      were obtained through a field survey system designed by Argonne and
      implemented  by Laboratory  and city engineers. This  involved a pro-
      gram of site visitations and personal or telephone interviews with plant
      personnel.
   3.  Natural Gas Utility Records. The  1966-67 natural gas consumption and
      fuel costs records  for  all dual fuel plants  among  the 50 largest S02
      sources were obtained from the local, natural gas supply utility.
   4.  Fuel Supply Records. The Midwest  Coal  Producer's Institute, an organ-
      ization of wholesale coal  suppliers, provided records of the average
      annual coal  consumption by major  industrial sources during  1966-67.

 Industrial sources that were not sufficiently large to be included among the
 largest 50 emitters were aggregated  into  uniformly distributed square-mile
 area-sources. All plants in this category were assumed to have a stack-height
 of 150 feet.

 Industrial Source Simulation Programs
 The 50  largest sources of SOj were  treated  as individual point sources. A
 detailed analysis of the diurnal, weekly, and seasonal operating pattern of each
 of  these  plants,  in  combination  with  fuel  consumption  records, gas-use
 patterns, and such characteristics as process-vs.  space-heating fuel-use practices
'' was incorporated  into an industrial-plant-emission-simulation computer algo-
 rithm (PLANTSIM) to generate an operating-shift-oriented emission estimate
i for each plant.
I The  data  file  for large  industrial  sources is  designed  to  include enough
I information about each plant to characterize its "expected"  emission pattern
! by operating-shift for each day throughout the year. For this purpose "shifts"
 are defined as follows:
      Shift 1:       12 midnight to 8 am         (0000) to (0800)
      Shift 2:         8 am to 4 pm             (0800) to (1600)
      Shift 3:         4 pm to  12 midnight      (1600) to (2400)
 Up to three daily  emission patterns may be assigned to each plant. These are
 as follows:
      1.    W  — weekday or normal operation.
      2.    S  - Saturday.
      3.    H  — Sunday or special  holidays.
 Finally,  a  monthly weighting  is provided to allow for variation due  to
 seasonal  patterns, etc.  Each plant requires two computer  cards; (1) a source
 identification card, and  (2) a  source emission card. An  example of each is
 shown in Figures 6-4 and 6-5  respectively. A description of each item on the
 cards is provided in Table 6-3  and Table 6-4.
                                                                    6-1-1

-------
                          Figure 6-4.  Sample   of source  identification card.
 MAXIMUM
SPACE HTG
  LOAD
(THERM/hr)
MAXIMUM
PROCESS
  LOAD
(THERM/h<
MONiHLr FUEL usf WEIGHTING
  (% OF MAXIMUM PROCESS)
                             Figure  6-5.  Sample of source  emission card-

-------
                         Table 6-3.  SOURCE I.D. CARD
No.
1
2
3
4
5
6
7
Item
Source no.
Name
Type
Grid
Coordinates
Stack data
Work schedule
Description
Plant sequence number
Plant name (not to exceed 25 characters including
address included if sufficient space available
Standard industrial code
City of Chicago square-mile number

blanks).


x- and y-coordinates relative to standard City of Chicago-
coordinate-system
Four stacks maximum (1 , 2, 3, 4)
HGT - Stack height, (feet)
% — The percent of total emission emitted from
Each day of week assigned one of the following:
W - Weekday
S — Saturday
H — Sunday or holiday
each stack

                     Table 6-4.  SOURCE EMISSION CARD
No.   Item
                        Description
      Source no.
      Max space htg load
      Max process load
      Monthly fuel use
        weighting

      Shift-fuel-use
        weighting

      % Sulfur-coal
      Avg % coal

      % Sulfur oil
      Avg % oil

      % Gas SP HTG

      % Gas process

      Avg gas temp

      Gas price
Plant sequence number
Maxium hourly thermal requirements for space-heating
Maximum hourly requirements for process
Achieves monthly variation in process fuel use. Each month
assigned  weight  (0, 1,...,9) = average  percent (0%, 10%,
...,90%) of maximum process fuel used during month
Within each day  pattern  (W, S,  H), weight (0%, ...99%)
assigned to each shift, = average percent of current monthly
process fuel used during shift
Average percent of sulfur in coal
Average percent of thermal load due to coal (0 implies coal
not used)
Average percent sulfur in oil
Average percent  of thermal load due to oil (0 implies oil
not used)
Percent  of space heating thermal requirements that can be
supplied by gas
Percent  of process thermal requirements that can be sup-
plied by gas
Average ambient temperature  (°F) at which dual fuel plant
receives gas on"interruptible"  supply contract
Gas rate (cents per therm) for  dual fuel plants
                                                                              6-13

-------
The  PLANTSIM  code is completely general. Large, individual space-heating
sources or major institutional  sources such as coal-fired  water-pumping sta-
tions can  be treated in exactly the same manner as industrial plants. In the
case  of heating plants, the process-load portion of the operating cycle input
is zeroed  and emissions are calculated solely on the basis of  temperature. A
special-purpose plant  such as  a pumping  station may be treated as an
industrial  plant that  consumes fuel according to  a characteristic processing
cycle.  Individual  power plants  that have a repeatable diurnal  operating cycle
may also  be simulated by  PLANTSIM; however, power plants that are a part
of a regional  power grid may not have a consistent and  repeatable  emission
pattern, since such  plants  are likely to engage in  power-load  transfer among
plants  — on at least a seasonal, if not a weekly or diurnal  basis.

All   sources  treated  in  PLANTSIM  are  assigned  individual  physical stack
heights based on  the field survey data.

PLANTSIM Computations
When  the temperature, T, is between -10°F and  55°F, a linear relationship
for  the space-heating  thermal   load  Ls is assumed.  This  is  expressed  as
follows:
      i s _ I s     [55   T]      r-1 n < T < 5s t                     (D
      L  ~ LMAX  	—	      (  IU *%  I s= MJ                     UJ
                      65
Then the  total thermal load is

      L = Ls + Lp ,                                                  (2)

where  Lp, the process load, is determined from the appropriate month, day,
and  shift  factors. The amount of load due to coal, Lc, and due to oil, L°,is
then determined, and the  S02 emission due to each  source  is  calculated as
follows:
                            Lc(TH*hr~')  x  10s (Btu    ThT1)
      1.    C(tons/hour)  =
      2.    0(kgal/hr)  =
                             12000 (Btu Ib"1 )  x 2000 (Ib ton"1)

                            SO^ (TH hr"1 )  =   Cltonshf1)  x 3.68 x SC

                               L°(TH) x 10s (Btu TH"1)
                         18000 (Btu Ib"1) x  8000 (Ib  kgaP )

                            S0° (THhr"1)  =  0(kgalhr"')  x  157.0 x S°
 Thus, the total S02  emission is
                         S02  = SO??  + S0°                          (3)
 "Therms
6-14

-------
When the ambient temperature is such that a dual fuel interruptible plant is
likely to  be receiving  natural gas,  the amount  of S02  produced is corres-
pondingly reduced.

In order  to facilitate  data storage  according to a uniform  and consistent
format, each  plant is  assumed to have four stacks. For plants having less
than four  stacks,  zeros are filled-in for  non-existent stacks. The  following
parameters are associated with each  stack:
   1. S02 emission in pounds per hour
   2. Heat emission in  therms per hour.
These parameters are determined  by  weighting  the  total  SO^  and heat
emissions for  the plant by the percentage  emitted from each stack. The heat
emission, H, is assumed to be  15 percent of the thermal input.
Example
Consider the example of the Campbell Soup Company shown in Figures 6-4
and 6-5  and assume that emission data are required  for the following set of
conditions:
           January
           temperature 30°F
           weekday pattern
           first shift
Then,
       s  =
     Ls  =    500 ( —	-)   =  200THhr-'
                      65
     Lp  =   (.70) (.60)5000 =  2100 TH hf1
     L   =   2300 TH  hf1
     Lc  =   2300 TH  hf1
     L°  =   0

              2300 x  10s    -v „„      , _,
     C   =   	 =  10 tons hr1
             12000 x  2000

    S02  =   10 x 36.8 x  2.8  =  1000 Ib  hf1

     H   =   (.15)2300 =  345 TH hf1


For Stack  1

    SC41J =   (.25)1000 =  250 Ib hf1

     H(1)=   (.25)345  =  85 TH  hr"1
                                                                  6-15

-------
For Stack 2

    SOf  =  (.25)1000  =  250 Ib hr'1
       (2)
      H    =  (.25)345  =  85 TH  hr"1

For Stack 3

    SOf  =  (.50)1000  =  500 Ib hr"1
      H(3)  =  (.25)345  =  170 TH hr"1
For Stack 4

    S0(?  =  0
      H(4)  =  0
Residential Space-Heating Sources
Residential  space-heating sources account for approximately  15 percent of
the annual total  emissions for the city. The  inventory of coal-  and oil-fired
residential  space-heating sources  was derived primarily from data supplied by
the market research organization of  the local natural gas supply utility. This
information was in the form of a city-wide, field-sampling survey of space-
heating sources categorized by fuel use and by the number of dwelling units
pter  building  surveyed. For  the purposes of the emission inventory, the
survey data  was  aggregated  into  two groups — moderate sized residential
structures of 19  dwelling units or less and large apartment complexes of 20
or more dwelling units.
Since residential  sources are generally too  numerous to treat individually, the
two  building   size categories  described  above  were  treated  as uniformly
distributed square-mile area sources.  The source density  per square mile  for
each category  was assigned  on  the  basis  of the natural  gas  utility market
research data.

The  operating  cycle  of the  moderate sized  heating plants was based on a
55° F degree-day  proration of a  seasonal average fuel consumption  modified
by  an  empirical  function5  that  represents  the  daily cycle  of automatic
stoking during  waking  periods followed by a hold-fire and cool-down during
sleeping  periods.  This cycle  was  simulated  by a  "Janitor Function"  that
maintains heating sources on  an  automatic stoking cycle between 6 AM and
11 PM  on days  having a minimum  night ambient temperature in excess of
5°F   For nights with  a minimum  temperature  below 5°F,  the automatic
stoking period was assumed  to  begin at 3  AM.  This pattern was derived
through an interview  survey  of  building  superintendents and  heating plant
operators.  A hot  water heating  baseload  equal  to 10 percent  of  the daily
6-16

-------
heating fuel use was assumed  for  all residential  buildings in this category.
Hourly-average emissions were  assumed to be distributed uniformly through-
out the automatic stoking period  and were assumed to be constant during
the hold-fire  period.  All  moderate-size residences were assigned a uniform
physical stack  height of 50 ft.

Large residential  buildings  were  considered to  operate on  an automatic
stoking cycle at all times. Fuel use and SO2  emissions were based on a 55°F
degree-day proration  and a  hot  water heating baseload equal to 10 percent of
the daily heating-fuel  use was assumed.  All  buildings in this size category
were assumed  to have a 200 ft  physical stack height —  an assumption based
on a sensitivity analysis of the effect  of stack height on the ground  level air
quality, and on surveys of actual building  sizes.

The hourly-average emissions for large size residential buildings were assumed
to be distributed uniformly throughout the stoking period.

Commercial  and Institutional Sources
Commercial  and  institutional sources account  for  about 8  percent  of  the
total emissions in  Chicago. Fuel  use  and  SO2  emission  data  for large
commercial and institutional  sources  such  as office buildings, water-pumping
stations, educational  institutions, etc., were derived from a computer  analysis
of a comprehensive inventory developed in 1963 by the  Chicago  Department
of Air Pollution  Control,  combined  with the  results of a field  survey  of
major sources conducted in  1968 by Argonne National  Laboratory and  the
Department of Air Pollution Control.

With a few notable exceptions, the commercial  and  institutional sources are
too small and numerous to  treat individually. They are  therefore aggregated
as area  sources, uniformly distributed  throughout each  square  mile  of  the
city. The square-mile source-density is  based on the 1963 inventory.

Virtually all  sources  in this  category  burn fuel  primarily for space and  hot
water heating and operate on  an  automatic stoking  cycle. Emissions from
these sources  were  therefore calculated in exactly  the  same way as in  the
case of the large residential sources.
Certain  commercial and institutional  plants are  sufficiently large to  warrant
inclusion as  individual point sources. Among  these are the five Chicago
pumping stations  and  a number of space-heating plants that are associated
with extremely  large buildings. An  operational analysis  of the pumping
stations was conducted in order to define their diurnal emission cycle. They
were then  treated  in  essentially  the same  manner as  the  large  industrial
plants.  Individual  large commercial heating plants were  treated  in  a  manner
similar  to the large residential  structures. The validity of the  latter assump-
tion  was established  through a detailed  survey of the operating  cycles of
                                                                     6-17

-------
most of these sources, which included, among others, the large University of
Chicago heating  plant, the Union Station, the Chicago Merchandise Mart, and
the Chicago Housing Authority complexes.

METEOROLOGICAL AND AIR-QUALITY DATA

Chicago's TAM  Networks
In January  1966,  the Chicago Department of Air Pollution Control (DAPC)
began  recording wind speed and direction and sulfur dioxide levels at eight
telemetered air  monitoring  (TAM) stations  (starred  locations on the map,
Figure 6-1). These  continuous measurements are integrated  over 5-minute
periods; 15-minute averages are then telemetered to the DAPC office where
they are recorded. Tape records of these 15-minute observations have been
reduced to  hourly averages. Thus,  although  the  accuracy of data from some
of  the aerovane  sites   is  limited  by  15-  to  30-foot  masts and/or close
proximity  to taller buildings  or smoke  stacks, the procedure  for developing
hourly averages  is excellent. This is in contrast to special airport data where
the aerovane  sites  are  excellent, but  only  brief hourly observations are
recorded.

In the dispersion  model, the  wind speed and wind direction from the TAM
aerovane nearest to the  dose-point is  used to determine the trajectory of all
plumes sensed  at  that  point. This temporary assumption will eventually be
eliminated  by using  all  eight  stations to develop a wind field for the city. In
contrast  to single city-wide  values of  wind speed and direction from the
nearest airport,  the data from a local aerovane is sensitive to  special circula-
tion patterns such  as  lake breezes.

Airport Data
Tapes  containing  hourly standard  weather  data  from Midway Airport and
Glenview Airport  (15 mi  north of Chicago)  have  been obtained from the
National Climatic Center.

Argonne Meteorological Data
The observing station at Argonne National Laboratory  (25 mi southwest of
Chicago) was  specifically designed to measure parameters controlling diffu-
sion: atmospheric  stability, wind speed and direction at five levels up to 150
ft., net and solar  radiation, etc. The hourly averages stored  there on mag-
netic  tape, are the  only  stability records  available  in  the Chicago area.
Unfortunately,  the data are  for  a shallow  layer  taken  in  an open, grass
covered field. Transformations relating these data to the  urban environment
are currently underway.
6-18

-------
Upper-Air Data
Radiosonde observations from  the  U.S.  Weather  Bureau RAWIN station  at
Peoria, Illinois  (140 mi southwest of Chicago)  and Green Bay, Wisconsin
(180 mi north  of  Chicago)  have been obtained from the National  Climatic
Center. These data have  been extrapolated to estimate the height of the
mixing layer in Chacago (see below).

Atmospheric  Stability
Turbulent  diffusion of  pollutant species  is represented  in  the  model  by a
Gaussian  puff  kernel  with  dispersion  coefficients ax.(T), ay.(T), az.(T);
where  T  is  the travel time  and  i,  the atmospheric  stability  index.  The
classification  system proposed  by Turner3  has  been employed using weather
data from  Midway  Airport.

Height of Mixing Layer
The mixing layer defines a zone within which pollutants can be diluted. The
height  of  this layer  is  probably  the  most critical  parameter in determining
ground level  S02  concentrations resulting from  emissions  from tall stacks.
With a high lid  (>  2000 ft )  plumes from power plants tend to travel far and
disperse widely  before  touching ground.  With a sharp temperature inversion
beneath the  physical  stack-height,  plumes will  travel in stable  air, which
greatly inhibits  vertical diffusion, and touch ground far from the local urban
area.  High ground-level concentrations occur  in  the  intermediate  range  of
mixing-layer  heights where the plumes  are trapped beneath  the  lid and
disperse rapidly to the ground  in  unstable air.  Figures 6-6 and  6-7  show
vertical profiles  of temperature and  sulfur dioxide measured  by an instru-
mented helicopter  at a  point 4 miles downwind of a major industrial area in
Chicago.  The sharp inversion  at  2300 ft  (700  m)  has clearly defined a ver-
tical mixing zone within which SO2 levels are nearly uniform.
Except for the  low level rural temperature profiles measured at  Argonne, no
vertical soundings are available for  the Chicago area during the period from
January 1966,  when  the TAM  network became operational,  to  January
1969. This is the  period during  which  meteorological air quality and  emis-
sion data  have been  accumulated at Argonne.  It was therefore  necessary  to
design  a scheme for estimating the height of the mixing layer from historical
weather data. (In July,  1969 the Weather Bureau  in the Chicago Air Quality
Control Region  began  a program of daily balloon soundings at 0600 and
1200 from Midway Airport. This data will be  available for future validation
studies of  the dispersion model and will  also be  used to validate the objec-
tive mixing-layer height estimate described in  this section.)
The computerized  objective mixing-layer depth  estimates  are based  on  an
interpolation  scheme between  two  rawinsonde stations.  Green Bay, Wiscon-
sin, and Peoria, Illinois. The soundings are usually taken at 0600 and  1800
                                                                    6-19

-------
        1,400
        1,200
        1,000
      31
      CD
          800
          600
          400
          200
                                    o$02
                                    • TEMP
                                       •    O   o
                                          •   o
                                           o  •
                                     GROUND-8
                     0.1

                       I
                               0.2
                   0.3
0.4
0.5
                                 S02, ppm
                                I	I
            38
40
 46
48
                               42        44
                            TEMPERATURE, °F
           Figures 6-6 and 6-7 (superimposed) Vertical profile
          of SC>2 concentrations and temperature in Hinsdale,
           Illinois, April  11,  1969 at 10:20  am.  (D.  Nelson,
          Argonne National Laboratory).

Central Standard Time  (CST). The  mandatory and significant levels for each
station are merged  to  form a  single pressure-temperature  array for  each
sounding. The Chicago  rural (Argonne) temperature profile is formed either
by  "shifting"  one of the two to  the Argonne surface temperature or by
interpolation  between the Peoria  and Green Bay  soundings.  If only one
station is in the same air mass as Chicago, the sounding from that station is
transposed, without changing its shape, to the Argonne surface temperature
6-20

-------
at sounding time. If both Peoria  and Green Bay are  in the same air mass as
Chicago (as  determined  by  inspection of the daily  weather map),  then  a
linear interpolation  is performed. This procedure forms a pressure-tempera-
ture array for the 0600 and 1800 rural profiles.

The  next  step  is  to  compute  the  hourly  rural profiles  using  temporal
weighting.  The profile for any  hour is a linear interpolation in time between
the bracketing 0600 and 1800 rural soundings.

The hourly urban mixing depth and mixing depth indices are now computed.
The  height of  the  mixing  layer  is defined  by the  intersection  of  a  dry
adiabat from  the urban  surface temperature  with the constructed Argonne
rural temperature profile.  In the Chicago case  mixing-layer heights calculated
by this objective procedure  are compared with  more subjective  estimates
provided by  a local meteorological  consulting  firm. While actual magnitudes
differed, there was  a good agreement in terms of diurnal trends. Calculated
values agreed very  closely with data from two  helicopter  soundings  but, in
both cases, the  inversions were  above 2000  ft  and therefore not  in  the
critical range where  plume trapping  could occur.

TRANSPORT  OF AIRBORNE POLLUTANTS

Pollution concentrations from a given  source are evaluated  in this model by
forming a convolution sum of two time-series: (1)  piecewise (hourly) con-
stant  stack emissions,  and  (2)  a discrete transfer  function based on  the
mathematical  representation of an  integrated  puff. The latter  is found by
integration of a  Gaussian kernel  representing  advection according to  a time
series of piecewise  constant  wind  vectors and three-dimensional  dispersion
according  to  a  time series  of piecewise constant  classes  of  atmospheric
stability.  (For example,  results indicate  that  one of the more  important
types of  short-term pollution  incidents,  the  morning fumigation, which  is
characterized  by  low winds (up  to 6  mph) and  by  a sharp transition from
stable  stratification  to  strong  vertical mixing,  is simulated successfully.)
Details of the  development  of the kernel  and limitations on  its range of
applicability are discussed  in the following sections.

Dispersion  Kernel
For  a  steady  mean wind, U,  in the  downwind x-direction, with y and z
representing unbounded crosswind and vertical  dimensions, a normal distribu-
tion yields the following Green's Function:
                exp
  Hx-U(t-t'
- < - ^
  I       2°x2
                                      2
                                         ,
                                        +
                                                         ~
     X(t,fJ  =        I       2°x2 _ 2ay2      2az2

      Q(t' )       (27T)3* ax (t - f )  ay (t - f ) az. (t - t' )
                                                                   6-21

-------
 where x(lt"> is  the concentration at time, t due to an instantaneous release,
 Q, at time t' and position (0,0,0)8'9  The dispersion coefficients ax, 
-------
Equation (5), the  contribution from an individual source to  the  concentra-
tion at time mA, must then be summed over all sources.

Equation (5) can be generalized to account for variation  (piecewise constant)
in wind direction, wind speed, and atmospheric stability:
                               m
                         [x-A 2      urum_n + 1 (T-A(n-1) )  ]2
                              j=m-n HI 2

                          [y-A^ _vrvm_n + 1  (T-A(n-1) ) ]'
                 exp-
                                       2ah2(T)
                                       2a22(T)
                                                                     (6)
                                (27T)3/2 ffh2(T) az(T)
wrfere variations in wind velocity are represented  by the piecewise constant
series [Uj]™  and  [Vj]™.  Note  that  this  result  is  contingent  upon  the
assumption of CTX  = ay = ah, a  horizontal  dispersion coefficient. There  are
valid arguments  (to  be reviewed in a later section) why this  assumption is
questionable, except  perhaps, for very  low  wind  speeds. From a computa-
tional standpoint, the puff  appears circular if  ax = ay, and  consequently,
since its x,y shape is independent of wind direction, the generalization of  the
kernel to accommodate variations in wind direction is greatly simplified. The
propagation of a circular puff for varying wind speed and direction is easily
visualized by the schematic of Figure 6-9.
In simulating lake  breezes and unusual  urban circulations, continuity in  the
wind field  and conservation of the mass of  pollutant require introduction of
a  vertical  velocity  component.  This component  as well  as  the horizontal
quantities u and v will be hightly dependent upon space and time variations.
Equation (6) considers only time dependence of the wind, although one  can
use  different sequences [Uj]™   and [Vj]f  depending upon the location of
the  dose point.  Using the lake  breeze  and, in  particular, a dose  point near
the confluences an example, we must  allow for two converging streams  and
a vertical  motion which is highly space dependent. The algorithms presented
in this paper are not yet expressed  in a  manner  sufficiently general to handle
this  space  dependence. However, it seems  relatively straightforward  to  ap-
proximate  a three-dimensional  velocity  field  with  the three components
u(x,y,t), v(x,v,t)  and w(x,y,t) and restate the basic algorithm.

Variable Atmospheric Stability
Postulating variations of a with a travel time T and requiring continuity of  a
at each time step,3  we might describe a  three step process by the paths
shown in Figure 6-10.

                                                                    6-23

-------
 Figure 6-9.   Three-hour propagation of circular puff from an in-
 stantaneous release.  Assumption: 
-------
         10*
         103
       N
       b
         102
          101
           102
103
10*
105
                              T, sec
Figure 6-10.  Three-step variations in atmospheric stability
class, exact method (500 sec,  crcontinuous).
          10*
          103
       bN
          102
          101
            102
103
10*
105
                            T, sec
 Figure 6-11.  Three-step variations  in atmospheric stability
 class by approximate method  (A   500 sec,  T continuous,
 a = a5  + A.a4 + ACTS).
                                                           6-25

-------
Associated with each puff are functions of time, an(t) and crz(t), that could,
in principle, be calculated  according to the procedure of Figure  6-10. This
procedure  requires calculation  of a psuedo-time for travel within each new
stability  class. Although the stability class  does not  change every hour, a
computer program based on hourly  steps must provide a general algorithm
for handling this situation.  Unfortunately, the  process  in Figure 6-10 appears
devoid of any simple algorithm  suitable for economic  programming  on the
computer.  The  problem is that the dispersion coefficient a should be kept
continuous  in spite of the fact that we are forced to  employ a set of discrete
functions.
An approximation to the "continuous a" scheme of Figure 6-10 is proposed
in Figure 6-11 for which the time of  flight is continuous and the value of o(t)
is determined by  adding changes in a. A comparatively simple algorithm can
be written  to  implement  the  procedure shown  in  Figure  6-11. The two
procedures  are compared in Table 6-5 where a tolerable error  is incurred by
the  "approximate"  method of Figure  6-11   as  opposed to  the  "exact"
method  of  Figure 6-10. For this comparison,  we allowed a single change in
stability class at the beginning of period  2.
        Table 6-5.  EXACT METHOD3 VERSUS APPROXIMATE METHOD6
         FOR DETERMINING ATMOSPHERIC DISPERSION COEFFICIENT
                         (A = 5000 sec; a At 10,000 sec)
Atmospheric stability classc
Period 1
4
5
3
5
Period 2
5
4
5
3
Dispersion coefficients
ay(m)d
Exact
2000
1900
2400
2500
Approx.
1900
1700
2500
2100
<7z(m)
Exact
200
200
1000
1100
Approx.
230
220
1100
1120
   Figure 6-10
   Figure 6-11
 d  Piecewise constants (see section— Transport of Airborne Pollutants)
   "m" indicates number of discrete transfer functions (periods)
   to which a is applied (see section— Dispersion Kernal)
6,26

-------
Dispersion  Cofficients
The  integrated  puff  model  is similar  to  earlier dispersion models in that a
Gaussian  distribution with  appropriate dispersion parameters is  the funda-
mental assumption. Although  the puff model is closer to physical reality, it
is still dependent and quite sensitive to these dispersion coefficients.
There are a number of sets of  functions for ay and CTZ with either downwind
distance or time-of-flight as the independent variable, and, for the most  part,
resultant  curves are in mutual agreement.  Some modifications (for example,
by Turner10  in the St. Louis study)  have been made to account for greater
instability over a  city, since most curves are based on experiments over  level
rural terrain.  Recent field  experiments in St. Louis by McElroy and Pooler11
support these modifications; although their estimates  of the  coefficient of
vertical diffusion  may  be  excessive if  applied to smoke plumes, because  their
data was obtained by tracer releases at ground level.

The choice of the function 0X  (j,t)  required for the integrated puff model is a
serious consideration.  Data from  field  experiments  on  instantaneous  and
steady releases12'  11 have been  processed to yield dispersion  coefficients
which fit the plume model, where concentration:
             Q  exp-
                       ' 2ffya     2°z  •
     X  =    	                             (7)
                  27T  CTy  CTZ   U

 Extensive  horizional measurements were employed to determine  0y  statis-
 tically;  and  <7Z  by inferrence. Clearly the  introduction  of an  additional
 dispresion coefficient ax with a corresponding change in the basic equation
 will alter the analysis of the experiments and thus the conclusions regarding
 the magitude of dispersion coefficients.

 Another concern  is the assumption ax=aywhich results in  the advecting/dif-
 fusing circular puff discussed and  shown in Figure  6-9. Although not strictly
 correct, it is convenient to consider CTX  as the superposition  of two compo-
 nents:  (1)  horizontal turbulence due  to eddies  of a scale smaller than the
 puff size,  and (2)  upwind/downwind  elongation  due to vertical shear. Thus,
 while the assumption  ax=ay  may be reasonable  for unstable  and neutral
 conditions  of atmospheric stability,  it  may be inappropriate  in  situations
 where  the effect of wind shear is magnified  by the existence of  a  stable
 atmosphere that  inhibits vertical mixing.13-14  However in these cases, smoke
 plumes  from major emitters  will  probably not touch ground within  the
 urban area.
 The dispersion coefficients used to obtain the  results  presented  here obey
 the sets of functions
                         ory,i(t),ffz,i(t);     i=1,5
                                                                     6=27

-------
from the equation  proposed by Turner10  for use in  his  St. Louis model
(Figures 6-12 and 6-13 for i = 1 to 7). Furthermore, it is assumed that ax =ay.

Plume vs. Puff Calculations
Equation  (5) which, for constant Q, expresses the steady-state concentration
according to  the integrated  puff  formulation; will in  general not agree with
the  plume  model  equation (7)  evaluated  at T  = x/u. However,  on the
centerline (y = 0, z = 0) the two equations yield almost identical results for
u > 1  mph.
The linear plots of az (T)  and ay (T) in Figures 6-14 and 6-15 indicate that
for the first few hours of travel  time, the dispersion coefficients in  Figures
         101
      8  100
      'I
      (T
         10-1
                         I    I  I  r
           10-1
  10°
T, hours
  Figure 6-12.  Sets of lateral dispersion  coefficients, av> j(t), for
  atmospheric stability classes i  = 1 to 7; used in St. Louis model
  by Turner.10
6-28

-------
   6-12 and 6-13 can  be  approximated by the linear relations: ax = cry = aT;
   az = yr. Substitutions of these functions into  Equation (4) for y = z = 0
   yields the following:
            OO


            /dT

           •fo
exp-
        (x-UT)
 (27r)3/2  ax(T)ay(T)az(T)
                                                                   (8)
Figure 6-13. Sets of vertical dispersion coefficients Qx> j  (t), for atmos-
pheric stability classes i - 1 to 7;  used in St. Louis model by Turner."'0
T   plume travel time.
                                                                  6-29

-------
                                  T, hr
Figure  6-14.  Linear plot of  cr2(T) for stability classes 3, 4, and 5.
Plots labeled "P"  are recent data from  St.  Louis."I"1
 where     T =  u
                        r  = \/ 2 a
 and (the error function)   erf(r)  =
*}
dt
Thus the integral can be evaluated as:

      Xt = P F                                                      (9)

where P, the  factor  ciutside  the brackets  in  Equation  (8)  is the plume
equation evaluated  at T, the time of flight to  the dose point x. The terms
within the bracket, represented  by the factor  F, are tabulated below  for
several values of  r.  For a given  value of r, there is a corresponding value of a,
(Figure 6-15)  and thus u for each  stability class i  = 3,4,5. According to Table
6-6 for steady  state on the centerline, the integrated puff calculation is 3
percent greater than  the plume formula when u =  1  mph under unstable
atmospheric conditions.

Table 6-6 also  indicates that the ratio  of  puff to  plume results  becomes
infinite  as u  -> 0. As  the wind speed  approaches zero, the plume  equation
6-30

-------
                         2

                      if
                                                            T, hr
9        Figure 6-15.  Linear plot of 
-------
         Table 6-6.  CALCULATED VALUES OF CENTER LINE3 PLUME  -
                    PUFF CONVERSION FACTOR, F,c FOR
           SELECTED VALUES OF r IN STABILITY CLASSES 3. 4, 5
                  DURING FIRST FEW HOURS OF TRAVEL TIME

r = u/s/2~a
10
1
0.7
0.5
0
Stability class horizontal wind speeds, mph
U3
12
1.2
0.8
0.6
0
U4
8
0.8
0.5
0.4
0
U5
7
0.7
0.4
0.3
0

F
1.00
1.03
1.10
1.22
CO
 3  y = z  = 0
 b  Plume Equation (4) evaluated at T = jj
 c  Equation (9) X^ = pp
 d  See Figure 6-15

with  dispersion  coefficients expressed as functions of travel  time T = (x/u)
goes  to  zero. This is  in contrast to  the  plume formula based on dispersion
coefficients as functions of travel distance  x. This function becomes infinite
as u ->• 0.  In  contrast to these two versions of the plume equation, concen-
trations estimated by  the integrated  puff model in Eq. (5) approach an x"2
distribution as u -> 0.
Effective Stack Height
The  effective  stack height H, is the height above ground which best defines
the centerline of the  plume. The plume  is assumed to  originate as a  point
source at an altitude  H  feet above ground at the location of the stack. If
aerodynamic  downwash  causes  the plume to  break  up  and  mix rapidly
downward in  the vicinity of the  stack, H  may  be set  equal  to the actual
physical  height,  Hs. Observations  of a local  utility indicate this phenomenon
occurs frequently for wind speeds greater than 15  mph. Otherwise a plume
rise AH is added  to the physical stack height Hs, so  that:
                         H  = Hs  +  AH
where AH is estimated as described below:
(10)
Smoke  plume  rise calculations  in  this model  are  based  on  a plume rise
formula  derived  from  observations  at TVA  stations  and  other  power
plants15   (The  momentum term  in the original formulation is omitted in the
following  equation since its contribution is generally negligible compared to
thermal buoyancy.)                    1/2
                                  IxLJc
                         AH   =  —-!_                             (11)
                                    u
6-32

-------
where    AH =  plume rise (ft)
          Qs =  heat emission rate from stack (Btu hr"1)
          U =  wind speed (mph) at the height of the stack
          K =  0.870 (unstable-Class 3), 0.354  (neutral-Class 4),  0.222
                   (stable-Class 5)

Values of  the heat emission  rate may be approximated by a fixed percentage
of the total  heat rate  (heating  value of the fuel times the fuel rate). The
percentage depends  on  the type  and efficiency of the  burning and  heat
exchange equipment  but may be categorized as follows:
          Power plant - 12%
          Large industrial, commercial — 20%
          Small industrial, commercial and residential — 35%
The mean wind  at the physical  stack height, Hs  should never be used in the
plume  rise Equation (11). In  evaluating the dispersion kernel  (Equation 6),
wind at the  effective stack height  H, corresponding  to the plume centerline,
is  used. In each case,  the desired value for the  wind speed will be usually
greater than  that measured at  a local airport or TAM station (typically  at 30
ft  to  100 ft above the  ground).  Equation (12)  provides  a correction term
for altitudes up to  1000 ft. The  exponent P  >s dependent on the stability
class, as shown in Table 6-7.
     U  (at height z2)  =   U (at height zt
(12)
             Table 6-7.  EXPONENT FOR WIND PROFILE LAW
                                                        16
Stability Class
stable (5)
neutral (4) . . .
unstable (3) . ...

P
0 5
0.2
	 0.2

The  experimental  data  in  the  reference  cited,  indicated a  slightly  more
uniform vertical wind speed profile for unstable than for neutral conditions,
but the model is not sensitive to this distinction.

Four restrictions placed on the use of Equation (11) are listed in Table 6-8.

    Table 6-8.  RESTRICTIONS ON PLUME  RISE FORMULA (Equation 11)
Mixing-layer height
Hm
a. Effective stack height H, cannot be
greater than mixing-layer height,
Hm, if stack is beneath mixing layer;
i.e. H  Hm, assume Hm = °° with
plume rise and dispersion accord-
ing to stable atmospheric conditions
Low wind speeds
U
If U estimated to be
< 4 mph, value 4mph
used in Equation (11)
Stability class and
Coefficient K
If the physical stack
height, Hs > 200 feet,
K = 0,354 (neutral
used even though the
stability index indicates
unstable conditions.
                                                                    6-33

-------
The restrictions on mixing-layer height in Table 6-8 are based on the assump-
tion  of an  impenetrable  atmospheric layer  (lid)  at height Hm. The effluent
from  a stack which emits its plume just beneath the lid will probably pene-
trate  into  the  stable  layer rather than  obey  the assumption of perfect
reflection at the lid. Unfortunately, there do not appear to  be any  data that
describe this situation.
The  restriction  in Table  6-8 limiting the minimum wind speed to be used in
Equation  (11)  reflects the  fact  that the  equation  is  an empirical one in
which the parameters n and m in an equation of the form AH  = K Qns Um
were  found  by multiple  linear regression.  A comparison of  Equation  (11)
with  TVA  data17  indicates good agreement down to wind  speeds  of about
6 mph. Below this, the (l/u) factor in the equation causes an over-estimation
of the  plume  rise.  An arbitrary  restriction has  been made, therefore, on a
minimum value of U in Equation  (11).

The  restriction on stability class and coefficient K, occurs because, for plume
rise calculations, the vertical variation  of temperature at heights greater than
200  feet  is best characterized  by a neutral  lapse rate, even  though unstable
conditions may exist in the first few hundred feet above ground.

A single value of  plume  rise is also ascribed  to  each of the three classes of
area-sources (previously discussed)  whenever the wind is less than  a critical
wind  speed. With  winds less than 6 mph,  low-rise  residential buildings are
assumed to  have  an effective stack height of 100 feet; high-rise residential
and  commercial buildings,  300 feet; and   industrial area-sources, 300  feet.
The  model  proves to be  insensitive  to these assumptions as long as the initial
value of az  (oz0) is  greater than or  equal to 100  feet.

Finite Geometry
The  fundamental puff  kernel has been  defined for z from (-00,+°°)  with z=0
on  the  plume  axis.  If  we assume  perfect  reflection  at the ground,  the
concentration C, at  time m, at a  receptor position  (x,y,z), due to a  single
source at  position (x',y',z',)  is:

      C =  x(x-x', y-y', z-z',m)  +  x(*-*',  V-y',  z+z',m)                (13)

where the coordinates  have  been  stated  relative  to the ground and  Equation
(6) determines X-

An objective calculational procedure was developed to estimate the  height of
the  mixing  layer  that  is  assumed to define an impenetrable  barrier for
vertical dispersion.  In  earlier work, the  distribution  under the lid  had  been
calculated by a  single plume until vertical diffusion implied the pollutant had
reached the  lid, at which time abrupt transition  to the equation for uniform
vertical mixing was made.
6-34

-------
In order  to  avoid this  abrupt transition  and simultaneously, to allow for
time-dependent (piecewise, constant,  hourly) values of Hm multiple reflec-
tions between ground and the inversion are represented  by pairs of  image
sources.  For  example, the first image source above the lid  is the reflection
about Hm  of the actual source; the second image source above the lid  is the
reflection about Hm of the  first below-ground image  (Figure 6-16). Depen-
ding upon the lid height and the time  of travel, up to five pairs of images
may be required  to adequately  represent a condition of  uniform vertical
mixing in an  urban area.
                             1	*\S
                                        if—"""
                                 	I  t	 ran
                                                     GROUND
                                          G2
Figure 6-16. Multiple  reflections  of pollutant  between  stable layer
and  ground,  represented  mathematically by   two  sets  of  image
sources.   Image G-|  reflected above  lid  at \_Q.  Image  LI reflected
below  ground  at  62-   S  : first  image  source, reflection of  actual
source.
                                                                6-35

-------
 Since the lid  height may vary  during  the lifetime  of each  1-hour puff,
 algorithms  have been  formulated to allow for specification of a different
 effective lid height for each hour m, look-back time n (Equation (5), and
 source J.  For  example, in  Figure  6-17  the  height of the mixing layer is
 steadily decreasing  during  the three 1-hour periods. Calculations of concen-
 trations due to stack emissions during the first hour  (m=1) would be based
 on lid  height Hmi  for m=1, n=1, and the lid height Hm2 for m=2, n=1 and
 2.  Concentrations due  to stack emissions during  the second hour (m=2)
 would  be based on  lid height Hm2  for  m=2,  n=1, and on Hm3  for m=3l
 n=2. Concentrations due  to stack  emissions during  the third hour (m=3)
 would  be based on  an  infinite lid with  diffusion coefficients corresponding
 to  stable atmospheric conditions.
                     1
                    I           2          I

                    TIME, hours

Figure 6-17.  Variable mixing-layer heights.
Area Sources
Earlier  studies have  emphasized  the  importance of residential coal  use in
determining ambient  SO2  levels.  For example, at Hyde Park (TAM-4) S02
levels of 0.2 to  0.3 ppm  for southwest  winds 6 to 9 mph can be attributed
almost  entirely  to space-heating  by coal-fired boilers  in three-story build-
ings.
    1 9
The  realistic  modeling of residential zones characterized by multiple-sources
too  numerous to  represent by individual plumes is therefore  critical  to the
6-36

-------
success of a source-oriented  dispersion model  for  Chicago.  For the Chicago
grid  system  these so-called  area-sources can  be  visualized  in  one of  two
modes:
   1. For winds less than a  critical speed the individual emissions resemble
     plumes. For three-story buildings,  visual  observations  indicate that a
     representative plume rise  for  wind speeds less than 6  mph is  about
     equal to the building height, i.e., approximately 50 feet.
   2. Above  the critical wind speed, mechanical turbulence, primarily down-
     wash, destroys the plume slightly above the rooftop level.

Similar arguments apply to high-rise, commercial, and industrial area-sources
previously described. In the first case, we might approximate the area-source
as a  uniform  horizontal sheet of small  thickness Az located at z = 100 feet.
In the second  mode,  it is reasonable  to assume  uniform mixing over the
volume of the building.
The  integrated puff model is  ideally suited to the task of synthesizing these
source configurations.  In  contrast to the plume equation, limited to  cross-
wind and vertical coordinates, thereby  representing a two-dimensional  front,
the puff  incorporates the  downwind coordinate and thus  represents a  three-
dimensional cloud of pollutant.

The  area source is approximated  by a  psuedo point source. This is similar to
algorithms employed  in  plume  models; but as mentioned  above,  the  more
sophisticated  puff kernel  results in a volumetric  source rather than a two-
dimensional  wave front. In addition to giving  a more realistic description of
downwind dispersion,  the puff model  can, therefore, directly  evaluate the
effect of the area-source upon the area  itself.

The  volumetric  source is synthesized by specifying three initial dispersion
coefficients: ax0, ay0  and az0. These  are then associated with times tx, ty,
and tz which are the solutions to:

      ffxO = Ox(\,tx), ffyO   =  tfy(i,ty), °zO  =  <7z(Uz)

where a(i,t)  is the family of functions describing the dispersion coefficients
at time,  t, for stability class, i. A virtual source  is thus defined at upwind
distances 6X  = utx and fiy = vty from the center of the area (see Figure 6-18).
The  choices of ax0,  ay0 and az0 as  well  as the  effective source  height, z,
depend upon the representative building  height for each square  mile and the
mode  of initial  transport:   plume  or downwash. The  transport equation
(Equation 6) is then modified with x -5X replacing x, y -6y replacing y,
and T + tx, T + ty, T + tz replacing T in  the appropriate locations.

For a square-area source of side  length,  n the value of ery0 is defined by
     ffy0 =  n/2.4                                                   (15)
                                                                     6-37

-------
                                                     P(x-8x,y-8y)
                                              GROUND
Figure 6-18.  Volumetric synthesis  P, of area source of buildings of
height  z,  at upwind distances of  &x and § y; from center of area
source  (x, y).

This insures that the concentration will be independent of position along a
line on  the downwind side of a row of identical area sources (Figure 6-19).
  Figure 6-19.  Choice of ayg = n/2.4,  insures uniform concentrations
  along line A-A for row of  identical area sources.
 6-38

-------
AN INTEGRATED SYSTEM FOR
AIR RESOURCE DATA  MANAGEMENT AND
DISPERSION MODEL DEVELOPMENT
The preceding  sections of  this paper have described the array of emission
inventory, meteorology and air quality data that was acquired and processed
in  order  to  test and  validate the diffusion  model.  This computerized file
containing nearly  3 years  of hourly-average  data  for a wide variety of pa-
rameters, represented  a singularly valuable resource, but at the same  time,
it presented a formidable  problem in data management. The development of
the "integrated puff"  model  required that  this data  file be exploited in two
ways:
   1. Analysis of  statistically significant  subsets of this  array of data could
     be expected  to yield information concerning the relative significance of
     many of  the  parameters and phenomena that were modeled. Moreover,
     the insights into the micrometeorological characteristics of the Chicago
     metropolitan  area obtained through analysis of this data should  serve
     as a guide to the interpretation and evaluation of the output of the
     model.
   2. Time series  arrays of emission-inventory data,  and raw and processed
     meteorology data were  required as inputs to the puff model, while a
     corresponding array  of  measured sulfur  oxide concentrations had to be
     merged and  compared with puff model  output  in order to analyze and
     validate its theoretical  predictions  against real-world data. The latter
     represented a major and laborious task if it was done manually.
An automatic  data management system was clearly required  in  order to
achieve these objectives. This need was met by the development of a master
Air Pollution Information and  Computation System (APICS)  designed to
automate the entire process of data storage, retrieval, manipulation, analysis,
and display, associated with the dispersion model development effort.

The APICS system consists of the IBM PL-1  data storage and retrieval code
nested  in a  series of operational  subroutines  and computational  codes
tailored to the analysis of  air  pollution data.  With this system,  the data
stored  in the  master  file  can   be  partitioned  in  any  way desired.  Any
combination, array  or subset of emission, meteorology and/or air  quality
data can be formed  and retrieved. The  selected data array  can be accessed
through a FORTRAN  subroutine which allows the  user to manipulate the
array  in  any desired way. For example, the analyst may form a new data
array which consists of products  of  components of the  original array raised
to some  power, etc. Standard subroutines for computing atmospheric stabi-
lity, smoke plume rise, etc., are included  as  options in this system.

The array of data in  the  master file  may be  analyzed  with a multivariate
                                                                  6-39

-------
linear  regression  analysis  code or a discriminant analysis code,  both  part of
the system  and  the results may be displayed as tables, histograms or CAL-
COMP plots.
The air  pollution  data  management  system  described above provides  a
uniquely powerful  tool  for  the development and  testing of the dispersion
model. The emission  simulation model previously  described and the disper-
sion model  algorithms previously described  have been incorporated into the
APICS system, so that the testing and validation process, conducted  with a
3-year computerized  inventory of  hourly-average  meteorology and  sulfur
dioxide air quality  data, is totally automated. Meteorology and emission data
from  the master data file  are input  to the  dispersion model, and  sulfur
dioxide concentrations are calculated; these results are compared with the
corresponding measured air quality data  and displayed in  a  single operation
within the Argonne,  IBM  360-75 computer. The sensitivity  of the model to
variations of  critical  meteorological parameters  and the effects of modifi-
cations in  the structure  of the model itself can  be  tested. When the dis-
persion model  development  effort is completed in  the early part of 1970, a
manual of  operations for the  APICS  system as well as for the dispersion
model itself will  be prepared for general use.

RESULTS, CONCLUSIONS, and  PROJECTIONS

TAM Stations
The month of January,  1967  was chosen  for  an  initial comparison of the
model with recorded  sulfur dioxide data.  Five TAM  stations were selected
representing a wide range of geographical and emission features (Figure 6-1).
TAM  1,  situated  in a low-density  residential area in the northwest corner of
the city, sees little S02  unless southeasterly winds carry plumes from the
power plants and other industries 11 miles away.
Prevailing southwest winds are  responsible for  high concentrations occasion-
ally observed  at TAM  2.  This station  is  also influenced  by  the high-rise
concentration  along  the  lakeshore  during   northeasterly  winds, especially
during lake  breeze conditions.

TAM  3 generally records the  highest  sulfur dioxide levels in the city. It is
situated  in  the  midst  of coal-burning commercial buildings  (for example,
Union Station, several  blocks  to the west emitted 8 x  106  pounds of S02
per year until it recently shifted  to  gas).  Furthermore, southwest winds,
sweeping up the "industrial corridor," along which three power plants and
major  industries  are  located, cause extremely high  pollution concentrations
downtown,  especially when vertical ventilation is  limited  by an inversion.
Peak hourly averages greater  than 1 ppm have been  recorded.
TAM  4
uiiy avciaycD yitaLci LIIQII I  fJ^JIU Mdvt; Deen recorded.

 is situated  along  the lake  at the edge  of  over 20 square miles of
6-40

-------
coal-burning three-story  residences. Studies of the frequency of air pollution
incidents in Chicago from January 1966 to December 1967 indicate that this
station has the highest  number  of  midwinter short-term air pollution inci-
dents  (6 hrs to 24 hrs duration). For example, in January 1966 and January
1967, JAM 4  had a total  of  thirty 6-hour periods for which  S02  levels
exceeded 0.4 ppm. The corresponding  figures for TAMS 1,2, and 3 are  0,1
and 8. Furthermore, plumes from  the  industrial corridor 7 miles northwest
are often sensed at JAM 4.

TAM  5  is similar  to TAM 1 except that northerly winds are responsible for
occasionally high  SO2  levels. Other  TAM  stations  (6 through 8) were  not
included  in  this  initial  validation  effort  because they are situated on  the
borders  of the city, and no  inventory has  yet been  assembled for the areas
of Cook County  immediately  outside  the city limits or for the  Gary-Ham-
mond, Indiana, industrial zone.

Ninety-seven station-days (2328 hourly values) were  studied for  the  month
of January 1967 at TAM stations 1 through 5. A station  malfunction on any
particular  day eliminated that station  from the data set for that day. Mal-
functions,  either  due  to  sensor   or  transmission   difficulties,  were
automatically logged on  the original Chicago air quality tapes.
Computer Requirements

The  model has been  programmed for the IBM -360 computer.  Since each
hourly time step for each source  and each  receptor involves a numerical
integration, some effort has been expended to:
   1.  Limit the number of integrations by employing a preliminary test to
      estimate  whether or  not a particular plume is likely to contribute to a
      given dose point  (for  example, there  is  no reason  to  evaluate  the
      contribution from a stack located 5 miles downwind of the  receptor).
   2.  Approximate integrals, where possible, by the mean value theorem.
   3.  Lump distant area sources together.

As  presently  constructed,  the model  run with hourly  time steps requires
between  0.5 and  0.75 minute  of computer  time  per station-day. Thus  the
run for five TAM stations for the 97 station-days  in  the month  of January
1967  required  less than  2  hours.  Recent analytical work (see Plume vs. Puff
Calculations page  6-28) indicates that dispersion coefficients can  be approx-
imated by the linear relationships  ay =  aT, az = jT  where a  and 7  are
constants. Under  these circumstances, the  integral  (Equation  6)  can   be
evaluated  in closed  form.  It is anticipated that run  time with this  modifi-
cation will be  approximately  equal to that  required  for a  standard  plume
model.
                                                                    6-41

-------
Statistical Results
There  are numerous  statistical  tests that have been applied to source-orien-
ted, pollution-dispersion models. The importance attached  to  any particular
test clearly depends upon the intended use of the model.20 One of the more
standard  presentations involves  evaluating the percentage of calculated values
within  a  given  tolerance  of   the  corresponding  observed values.  This is
equivalent to a scattergram, but less dazzling and more easily interpreted.
Table  6-9 lists these  percentages for each TAM  station and for the citywide
collection.  Hourly  data and 2-hour, 6-hour, and 24-hour averages are evalu-
ated in terms of five criteria.
       Table 6-9. STATISTICAL EVALUATION OF INTEGRATED PUFF MODEL DATA:
         PERCENTAGE OF CALCULATED POLLUTANT CONCENTRATIONS WITHIN
                DESIGNATED TOLERANCE LIMITS OF SAMPLE DATA3
Item
Mean (wobs>
(Mobs-Mcalc) „ inQ%
Mobs
Std dev (obs-calc)
Std dev/mean
% • 0 025 ppm
% ' 0.05 ppm
% ' 0.1 ppm
No data points
Std dev. (obs-calc)
Std dev/mean
% '0.025 ppm
% ' 0.05
% '0.1 ppm
No. data points
Std. dev (obs-calc)
Std. dev/mean
% '0 025 ppm
% '0.05 ppm
% '0.1 ppm
No data points
Std dev (obs-calc)
Std dev/mean
% ' 0.025 ppm
% ' 0 05 ppm
% ' 0 1 ppm
No. data points
Hour
avg.

1

1
1
1
1
2
2
2
2
2
2
6
6
6
6
6
6
24
24
24
24
24
24
TAM stations
1 2
0.06 ppm
10%
0.09 ppm
1.7
58%
69%
84%
288
0.08
1.3
58%
70%
83%
144
0.07
1 1
56%
67%
90%
48
0.04 ppm
0.71
25%
75%
1 00%
12
0.12 ppm
+ 16%
0.10 ppm
0.88
37%
63%
81%
576
0.09
0.78
40%
66%
83%
288
0.08
.65
36%
66%
88%
96
0 05 ppm
0.44
46%
62%
96%
24
3
0.33 ppm
* 14%
0.20 ppm
0.62
12%
23%
47%
432
0.18
0.56
15%
27%
49%
216
0.14
0.42
10%
29%
56%
72
0 10 ppm
0.30
17%
39%
67%
28
4
0.1 5 ppm
3%
0.13 ppm
0.84
26%
53%
79%
408
0.11
0.77
29%
52%
81%
204
0.08
0.58
32%
54%
85%
68
0.07 ppm
5
0.06 ppm
-9%
0.10 ppm
1.6
45%
71%
89%
1-5
0.14 ppm
+ 7.4%
0.1 3 ppm
0.93
35%
57%
77%
624 2328
0.09 ' 0.12
1.5
45%
70%
88%
312
0.07
1.2
46%
72%
91%
104
0.04 ppm
0.45 0.70
59%
77%
88%
17
58%
77%
96%
26
0.83
37%
58%
78%
1164
0.09
0.64
36%
59%
83%
388
0.06 ppm
0.43
43%
66%
90%
97
  • Chicago, January, J967
6-42

-------
Monthly  and  daily  predictions are in excellent agreement with  observed
values.  Marsh and Withers21 indicate that, for their Reading, England, model
as well as for others in the literature, the  ratios of  the  standard deviations
(observed-calculated) to the observed  means  are all  higher than 1.1.  From
Table 6-9 the  corresponding value for 24-hour averages is 0.43. It is 0.64 for
6-hour  averages and 0.93 for 1-hour averages.

Results for  individual JAM stations  vary, depending  primarily upon  the
monthly  mean. For example, almost all values for JAMS 1 and 5 fall within
±0.1  ppm since the means are  only 0.06. For the same reason, however, the
ratios (aobs.caic)/ juobs are high for these stations.

Figure  6-20 (a and b)  shows daily averages  at  JAMS  1 to  5  for January
1967. Six-hour averages are presented  in Table 6-9 because 6 hours  is the
minimum  practical  time  step  for  the implementation  of  incident control
strategies.22  On a citywide basis 59 percent of  the calculated 6-hour values
are within  ±0.05  ppm and  83 percent  within  ±0.1 ppm.  The  results for
6-hour  averages are also presented in  Figure  6-21. Assuming the solid  lines
delineate a region of  successful  predictions, then 59 percent of the calcula-
tions are "correct", the  skill score (based on  chance)  is 0.42; the model
explained 52  percent of the variants in the observed data set. For 24-hour
averages this statistic is 74 percent.

Tables  6-10  and 6-11  illustrate  the  use  of the model in evaluating  the
frequency of  short-term air pollution  incidents. The contingency tables are
based on the following air pollution "war game"  rules:
   1. A threshold S02  level  is defined for a  given averaging period.
   2. If the model  predicts that this  level  will  be exceeded,  then  action  is
     taken.
   3. If action is taken and the observed average is greater than the thresh-
     old  less a  tolerance  (here  .025  ppm), then  incident is said to have
     actually  occurred.
   4. If  no  action is taken  and  the observed average  is  less than  the
     threshold plus a slight tolerance  (here 0.025 ppm), then an incident  is
     said not  to have occurred.
When compared to  chance, the  model is significant at better than the 99.9%
confidence  level and  the  skill  scores for this air pollution war game  range
from 0.64  to  0.80. Skill  scores  are  defined by the ratio (V-VVd-V1)
where  V is the number of correct estimates (here yes*yes + no*no  cases),  T
the total number  of cases,  .and V1  the value of V expected by some means
other than  the model  under evaluation. (For example, instead of chance,  a
simpler plume model  might be used.)  It is  worth noting, also, that in  terms
of incident  control policies;  taking  serious actions, such  as temporarily
curtailing industrial production  is costly. Air pollution  cost-benefit studies
conducted at  Argonne show that as much as $10  million per day can be
                                                                     6-43

-------
              STATIONS
U.tU
0.10

n
u
E
0.
^ 0.20
o
f= 0.10
K 0
z
UJ
o
§ 0.40
£ 0.30
| 0.20
d 0.10
0
°- n
u
o 0.20
LLJ
g 0.10
o:
uj n
> u
0.20
0.10
0
TAIW-1 1 + \ -i- \
I 1
— -+- --0-- "0-- MMMMMMMMM.
-•o- . ~~*~~
--O-- --O--
""TAIW-2

— M M M M M M
o --*
~ TAM-3 ° "°"
	 --O--
- -0- 131 --^: c
— ~*~ ~*~ ^:
—


~TAM-4 ~~~
0
-. M ^ ^ _,

--O-
"TAM-5
-
- --&.. o :*:
--0-- r^r. — o _&. -_
T T


--O--
~~*
-<>.. --O-- -- O- — O--
tl — •— -- J~ * — •— *
M M M M M M
' -0=
—
	


—
--A-- jy, --O-- O ~ *~
5^ -o-- _^_ — •- —

— 1
—
__0._ ^_ ^ ^_
^— "f" . — •— — •—
III 1
                  1   2    3   4    5    6    7    8    9   10   11   12   13   14   15
                                             DAY
Figure 6-20a. Daily pollutant concentration averages,  Chicago, January 1 to 15, 1967.
"M", missing data.

-------
AVERAGE DAILY POLLUTANT CONCENTRATIONS, ppm
U.IU
0.10
0
0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.20
0.10
0
0.20
0.10
0
_TAJH II ' ' _L '
~TAM-2
:3r __„._ '^~ i£r M -—
— -- o-- _ _
--O--
~TAIfl-3
. — M ~°" M MM M
— O—
— O--
"~ TAM-4 =£:
— —*— —0~
" °" M M M M
	 O ,
^TAIW-5
^ M M M
O ..-,_ --n--
	 1 	 1 	 1 	 r — -^r" — i 	 1 	 r "
1 1 1 1 III
. M M M M M MM
--O--
--O-- --^.,
-O--
M — _+_ M_
--O--
—O--
MMMMMMMM
M M
--o-- --o-- —*—
	 r 	 r 	 y 	 1 	 y 	 1 	 1 	 p-1
£
01
               16   17  18  19  20  21   22  23  24  25  26  27  28   29  30  31
                                           DAY

Figure 6-20b. Daily  pollutant concentration averages, Chicago, January 16 to 31, 1967.
"M", missing data.

-------
      ssO.5-
    ~  0.3
    CNl
    O
        0.1





1
1
IT

21
_Ji6
68




3
7
II
15
|io
41
111
9





7
3 1
O

4
4
1

1

3


5
5
3
2
2

4

3


4

1

2
1
2
i

7
4
2

1
2
2
2
0
1



1

1

0

%
2
1






4


2







6
1
3
3
1

1
2

1

                     0.1        0.?        0.3
                             ESTIMATED S02, ppm
0.4       20.5
Figure 6-21.  Six-hour pollutant averages, Chicago, January 1967.
       Table 6-10. FREQUENCY OF 24-HR INCIDENTS FOR STATIONS 1-5
                             JANUARY 1967
Symbol
X
y
z
Threshold
0.1 ppm
0.2 ppm
0.3 ppm
Tolerance
0.025
0.025
0.025
Skill score
(Based on chance)
0.76
0.73
0.75
                             CONTINGENCY



n
c
c
r




Y
e
s

N


Predict
Yes
x 44
V 17
z 9
x 3
y 3

z 3
No
x 9
V 6
z 1
x 41
V 71

z 84
 6-46

-------
      Table 6-11.  FREQUENCY OF 6-HR INCIDENTS FOR STATIONS 1-5
                             JANUARY 1967
Symbol
X
V
z
Threshold
0.1 ppm
0.2 ppm
0.3 ppm
Tolerance
0.025
0.025
0.025
Skill score
(Based on chance)
0.76
0.64
0.80
                             CONTINGENCY
o
c
c
u
r
Predict
Y
e
s
N
o
Yes
x 145
y 62
-L 32
x 14
y 20
z 6
No
x 32
y 28
z 10
x 197
y 278
z 340
involved in lost  wages and production for a city  such  as Chicago. Thus, in
practice, the tolerance  used above would  have to  be enlarged sufficiently to
insure that the model  scores  significantly better  than  shown  in Tables 6-9
and 6-10

Hourly Time Series
Figure 6-22 is a sample of strip charts for January  1967  showing hourly wind
speed, wind direction, temperature (all measured at Midway Airport), atmo-
spheric stability  class,  and the mixing height as estimated  by the objective
method  previously described.  Figure  6-23  is a sample of strip charts also for
January 1967 showing  smoothed  hourly values of observed (solid lines)  and
estimated  (dotted  lines)  sulfur dioxide for TAMS 1-5. For readability, the
hourly  values  have been smoothed  by the  formula:  X(n) =  1/4(X(n-1)  +
2X(n) + X(n+1)  where n  is the hour number.  They are then approximately
equivalent to 2-hour averages. Missing data is denoted by an  "M"

This appears to be the first paper which  presents time series such as these.
They clearly show that the model, which, statistically, looks very  promising,
has its exceptionally  good and equally disappointing moments.  The model as
presently constructed has a number  of strong points: among them a disper-
sion kernel that can respond  in  a  physically realistic  way to transients in
wind direction, wind speed, source  strength, and  mixing-layer height,  and
which synthesizes area sources exceptionally  well; and  industrial inventory
that estimates seasonal  and daily  shift patterns, based  on an emission inven-
tory procedure  geared  to a  boiler-room  superintendent; an objective, mix-
ing-layer  height-estimation technique,  which, although  suffering  from  the
                                                                    6-47

-------
en
CO
                                     DAY                                            DAY
              Figure 6-22.  Stripchart of hourly values of (top to bottom ordinates) atmospheric stability
              classification, estimated mixing-layer height, temperature at Midway airport, and wind speed
              and direction, January 1 to 16, 1967.

-------
                        DAY
                                                                      DAY
Figure 6-22 (continued).  Stripchart of hourly values of (top to bottom ordinates) atmospheric
stability classification, estimated mixing-layer height, temperature at Midway airport,  and wind
speed and direction, January 17 to 31,  1967.

-------
0.40
0.30
0.20
0.10
0
0.30
0.20
0.10
0
0.60
0.50
e 0.40
^ 0.30
° 0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
TAM1





M









1
(
/
1
\^
\
\
i\
\



\





TAM2



M











M








TAIM3


A
/-




i
2 concen-
 trations at TAM stations 1 to 5, January 1 and 2, 1967.
6-50

-------
U.4U
0.30
0.20
0.10
n
TAM1





BOTH



ARE-



0.0




^>

p*\
1

^
V
V
\



X
A

luu
Don
.ill
.10

.bU
Orft
.all
U.4U
09(1
Om
.1U
n
Ojin
.41)
0?n
Oon
.a)
.10

U.4U
0?n
.01)
Don
.&}
n m
n
TAIVI2



,f
\i
~s v




A
\
\
^<


TAM5

—10*^*'
? X




JAMS
\ w /J
^
U




TAM4
rw
^""





"V_

M


A
\v
v-^





•N 	
~^-— '





— > — •




/'ft
I
1
/
/
/
'_/



rs
y-
7




X"
^wx




;>T
/
!
tl
^










v J^i^





~^\
!\
1
1
1
k










A.


M



y
	 '

V



M






^X





K
r^
/
;











L>^"
                   6
12
                 18    24    6
                   TIME, hours
observed,	estimated
12    18
24
                                               M-missing data.
Figure 6-23.(continued). Sample of smoothed hourly values of SC>2
concentrations at TAM  stations 1 to 5, January 3 and 4, 1967.
                                                            6-51

-------
0.40
0.30
0.20
0.10
0
0.30
0.20
0.10
0
0.60
0.50
0.40
I 0.30
Sf 0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
	 r
TAM1


^V



--_fc,_l_'



^^

//I
//
i






M










TAKI2






M








M







r\
1 >
7/'
i
i



TAM3
» Sx-

/^
f \
kX

N
\
\,
/
^

/
7
\
\



\ A
/
^

i
i
i
\
i
TAM4

^A




/


f^


i
i


v v^
A
i_y


I |
LI
i
i
i
\
\
i



\
\
V
"N^_
i r~ i
i
i
•

/
J
/



\
i
\
_L




^ \

TAIV15
\
Ay
v/


"\V\/



\
\ r
^4'


^^«.
~ \



S^>







^^"^^




6 12 18 24 6 12 18 24
TIME, hours
 Figure 6-23 (continued). Sample of smoothed hourly values of SC>2
 concentrations at TAM stations 1 to 5, January 5 and 6,  1967.
6-52

-------
U.4U
Oin
,iv
0.20
Om
.iu

.30
.20
.10

Ocn
.DU
.50
.40
03ft
.oU
n in
U.1U
OAt\
.411
0.30
Don
.iv

Oxn
.40

01 n
.10

TAN11




TAM?

^
V^—

TAM 3


x-J
//


TAM 4


r^

TAM 5



6







S





^A
i





x




.
i


M




v





/^ ^
s/ /










^
2 1







— -3^



/
f
/ \
/ \
' >




^~,
"*




,„.. 	
8 2








• •""


\
V
1'
A ;





b-=^





4 (


M



_ s
f j*-


t\
1
t
1
I
1
'




A
l^





^
\ i







_^ —




\/
^





, — — s^— -






2 1







'""^-v


/
1
1
^T^
-''





:^A.






8 24
                               TIME, hours

            observed, >	estimated.
M-missing.
Figure 6-23 (continued).  Sample of smoothed hourly values of SC>2
concentrations at TAM  stations 1  to 5, January 7 and 8, 1967.
                                                            6-53

-------
      Cvj
      o
      C/5
U.4U
IUU
.LV
.10
0.30
0.20
0.10
0
0.60
0.50
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
TAM1





M



















M






TAM2
//
-._. 	
•*^^
)v
/x \


c^

l\
1-V


. — - — H


:>c^

TAIVI3












M


M














M


M



-JH-'—


3^^5














TAM4


-^=K^P'


x-^
~~+^- '






s^




•-— x
/\
/
/I


\
O~


^c


M
TAM5






/



'^--



~s — —





I

d—
1 1
.' I
^
— — -*^--
I
!\
J \

6 12 18 24 6 12 18 2'
                               TIME, hours

             observed, — •	estimated.
fl-missjng.
Figure 6-23 (continued).  Sample of smoothed hourly values of S

concentrations at TAM stations 1 to 5, January 9  and 10, 1967.
6-54

-------
      CM
     O
U.fU
03H
.JU
09n
,L\I
n in


.30
Don
.i\i
01 n
.ill


.60
Otn
.3D
.40
Don
.oO
Oon
,L\i
01 n
.J.U
0/in
.40
n on
U.JO
Oon
,L\1
.10

0.40
n 3ii
Don
.^0
Oin
.1U
0
TAM1




TAB12

-j


TAM3






TAM4




TAN15
\
\
\
v
->
f


M


A
A^ \
/ V






M





M




~V
^^
i 1






A
S< 	
















s**
,v
2 1






~^^
'"S


















3 2





/s
r\
•*s^
'^. — *-'











r=^





4 (


M



A
•S \
*•




M





/
/^ "^




/^~^
i 1






/N
^ ^










./vy
i



A
A
w
2 1






-^TT-
.x"" ~>










/^X






s^
8 2
                              TIME, hours
            observed,	estimated.
M -missing.
Figure 6-23 (continued). Sample of smoothed hourly values of S02
concentrations at TAM stations 1 to 5, January 11 and 12,  1967.
                                                          6-55

-------
      O
      oo
U.4U
OOA
.30
OOA
01 A
.10
Don
.30
Oon<
01 n



O.bO
Ocn
Ojin
Olfi
Don
Oin

Oin
Oin
n on

0.10
Oin
.40
n 9n
Don
Oin
n
TAKI1




TAM 2
i
J



TAM 3






TAM 4
/
^x-'


TAM 5





M


f\
\
\






M



^
,'\
'\
1
\
>




-^
/\/





y
\ l
VJ7










x\
s




^_^
^~.





l,\
- v.











—^ —
.^^





^^.





' \J

*-^s'











\U
1*~^*




•^ ~^^^E^


M


\
A
K





M




/->L
^L>





tXI







— 	











X*-.'"*




r-








	 .










/N
's:




..^
-+*
                   6    12    18    24    6
                               TIME, hours
              observed, .	estimated.
12    18   24
   M-missing.
 Figure 6-23 (continued). Sample of smoothed hourly values of 802
 concentrations at TAM stations 1  to 5, January 1.3 and 14,  1967.
6-56

-------
u.fu
0.30
0.20
0.10
0
0.30
0.20
0.10
0
0.60
0.50
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
TAIM1





M















M









/ 0.746 i
TAWI2

	 _^


s ~
r~ s__


-*•—•
—*~~-'

— -iiv
/ ^
/
/
*/

\,

\
\
\
\
i
>v^


-^

TAWI3


/
•-x/



^-^
j**>
f




'——•*'



•A
if\
f












M















TAM4

r~*
^4^

/"\
1 V
tr\


/
/ ,
/>


V
V



<-J


yX
K^-







~^S N

TAM5

\
"V



/



J*^



-^



•^-S'



f*








                       12   18    24     6
                              TIME, hours
            observed,	estimated.
12    18    24
   M -missing.
Figure 6-23 (continued).  Sample of smoothed hourly values of S02
concentrations at JAM  stations 1 to 5, January 15 and 16,  1967.
                                                           6-57

-------
0.40
OJU
.20
.10

.30
0.20
.10
OCfl
.bl)
.bU
O/in
A(l
n tt\
U.ol)
090
,/u
.10
Oyin
.41)
.30
0.20
Oin

Oin
Otn
.ou
U.ZO
n in

n

TAIW1




TAIV12

^^

TAIV13






TAIV14

/
s~y- -^

TAIW5



— —-—
6


M




? 	 ~—

/
/
/-/
/




'£
' 1
1




f^
/ ^N
J
1







	 	 ^

\ r-
V
-^ —





^/x>





/


2 1







— . — . — •

/

/ *~'





"^V—
V




^
— • — — x

8 2






/
^^








/
Xy/
	 /


/ \
V'

^-^ —

4 6



„••••

^
X^





M



/^
f \



/
y
A


i





r\
— /








\
\
A x
VxV


\
\
\

V_ >-k ,
2 1


A
J/^-

/A'"'
• v










^V—
\
s





^X.
^
8 21
                              TIME, hours
             observed,	estimated.
I -missing.
Figure 6-23 ( continued).  Sample of smoothed hourly values of SC>2
concentrations at TAM stations 1 to 5, January 17 and 18,  1967.
6-58

-------
U.4U
0.30
0.20
0.10
n
TAM1


=



TAIVI4

-v


TAM5


	 "


/ \
/x~~~-


/VT
/ *
/




/^N
'/-.
f



^
*~


S-x
-•*'


"A
/ ^
\




^^x









.^r




K
V
x/




\
N,






/^

sX^
















I

,\
\
^




M





M





'\























A
^>— ^


















                       12
18    24    6
  TIME, hours
12    18
24
            observed,	estimated.
                   M -missing.
 Figure 6-23 ( continued).  Sample of smoothed hourly values of SC>2
concentrations at TAM stations 1  to 5, January 19 and 20,  1967.
                                                            6-59

-------
U.4U
0.30
0.20
0.10
0
0.30
0.20
0.10
0
0.60
0.50
0.40
0,30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
TAM1


^\








^
\
/ i
' /"^



\
v^<

f\
/'
'/ s

A \

^~


i
\
0.462
TAIV12
V
"~N-

M.


j

s
i














TAIV13








M

















\
***.











\
V




1
J


TAIVI4





M















M










TAM5





M















M









6 12 18 24 6 12 18 21
TIME, hours
            -observed,	estimated.
                                               M -missing.
Figure 6-23 (continued).  Sample of smoothed hourly values of S02
concentrations at TAM stations 1  to 5, January 21 and 22, 1967.
6-60

-------
      o
      oo
u.ou
0.20
0.10
0
0.30
0.20
0.10
0
0.60
0.50
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
n
TAM1







<+


^


V
I
i
i
X"

\
\A


A

TAM2

s
A
r
**"— <^-
- \
V
I,
\ 1
1
A
\ <
\
\
\

\ 1'
\/
I1


\
{ ,
V


^,

TAM3








M























M















TAM4





M















M










TAW 5





M















M









                               TIME, hours


              -observed,	estimated.
                                                       24
I -missing.
Figure 6-23 (continued). Sample of smoothed hourly values of SOg


concentrations at TAM stations 1 to 5, January 23 and 24, 1967-
                                                            6-61

-------
0.30
0.20
.10
1

.30
Oon
.£U
n 1 A
U.1U
1

.bU
Ocn
.JU
Odd

g U.jU
Q-
exi U.^U
O
c/>
Qin
-IU
n
0/in
.4U
Oon
09n
yu
.10
n
TAIVI1



TAIW2


N — <•";

TAM3

^
"\ /
V
\.j


TAKI4




M



A
/ s
^,— ~-


/ V

A
/ \
/
i





M







/•



X\
/
y
/
l
,j













r\
'r-^ <


r"
i \
i \
^
K
I
\













^ ^/
\" — .




v
V ^"^








M




"^-X=
'





^~X
" '^-j




M














— -« —




















k '-
^y>^






U.4U
0,3Q
0.20
0,10
n
TAIW5





M












'V-~



s. 	 .



- , rr>=



f
s
                   i     12    18   24    6    12   18    24
                               TIME, hours

 ^—	;—-observed,---—--estimated.      IVI-missing.
 Figyre 6-23 (continued).  Sample of smoothed hourly values of
 concentrations gt TAM  stations  1 to 5,  Jipuary 25 and 26, 1967
6-62

-------
     n.
     c^
    in
U.4U
0.30
0.20
0.10
0
0.40'
0.30
0.20
0.10
n
TAM4





M















M










TAM5

\
i_
\
\ 	 -
-JMC=^

a
/ v
/
j


L '
^


-\
y^v.



C^~-^-


/-\ /
1


•N
-_
r
;
/
/
/
/^J
                     12    18    24    6
                            TIME, hours
           -observed,-*	estimated.
12
18    24
     I -missing.
rigure 6-23 (continued).  Sample of smoothed hourly values of 862
concentrations at JAM stations 1 to 5, January 27 and 28, 1967.
                                                           6-63

-------
          CXI
         O
UJU
0.20
0.10
n
TAB11



M











M







UJU
0.20
0.10
0
0.60
0.50
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
0.40
0.30
0.20
0.10
0
TAP/12
y
-~"-~— XJ
_,./
r^
1


\
**&**


~"S
"~ *^


^-^

/^
s

^>


__, f
— — ^v

r'
/
/
/
,//
TAM3


— \
\
A. ^ i
\A^





\
\ A
\VW
V\
\




^^ 1

\
s- -



— "^ -

— - —



/^

^s








/
V
/
/
/

TAIV14





M















M









ri
/-i
/ \
/
N/
TAM5


/ 1
A
r


1
,\
V




^-^:



IT^=*-rt^



7*^~-






^
6 12 18 24 6 12 18 2
                                   TIME, hours
             -observed,	estimated
I -missing.
Figure 6-23 (continued). Sample of smoothed hourly values of SC>2
concentrations at TAM stations 1  to 5,  January 29 and 30, 1967.
  6-64

-------
        i.
        Q.
u.ou
.20
01 n
.1U

0.30
Don

.10

.60
.50
.40
.30
.20
Oin
•1U

.40
090
.icu
Oin


.40
Oin
.jU
A on
O.ZU
01 n
.11)
n
TAK11



/A
/ v
/
/
TAM2

TAM3






TAIVI4




TAM5
/--
y
/



M


VN
\N
\ 	




M





M



1
\
I
\
\
"* — d






• — ^^
. — '"























— ~-^
.- 	 '^















/
X





































































































                          12
18    24    6

  TIME, hours
12    18
24
             observed,	estimated.
                 M -missing.
Figure 6-23 (continued).  Sample of smoothed hourly values of SC>2


concentrations at  TAM stations 1 to 5, January 31, 1967.
                                                              6-65

-------
obvious  problems  of  extrapolating vertical temperature profiles over  dis-
tances greater  than 100  miles, correlates well  with subjective estimates of
diurnal trends; and an overall  system  of  data storage, retrieval, and analyses
designed  to  meet  the operational  requirements  of an  urban  air pollution
control administration. Weak  points are also  evident; most significantly, a
poor determination of the wind field, which should be improved by deriving
an  hourly wind map but which, in any  event, is limited in Chicago by the
accuracy  of  the  JAM aerovanes. Under certain circumstances the model is
also extremely sensitive  to  meteorological  parameters-both  observed  and
derived. (For example, when the height of the mixing-layer approaches the
physical stack-height of a major power  plant, the calculation of ground-level
concentrations due to that plant  may  vary  by a factor  of  10 or more,
depending upon whether the  plume is assumed to  travel  above or to be
trapped  beneath the lid.) There are also further refinements that should be
made  in  the  model:  major  point  sources and  area sources  outside the city
limits must be included  in  the emission inventory; the local  effect of the
topography  of city  center on the  mixing-layer  height and  the effective
atmospheric  stability  should be included; plume  rise formulations might be
modified to  synthesize  a gradually ascending plume; the run-time  can be
greatly reduced if the  equations can be integrated analytically.

The  sample  strip  charts give clear evidence of many of these items. The
results, magnitudes, and  trends appear  to  be sufficiently consistent to  test
selected hypotheses, about both the model and the air pollution meteorology
of  the urban  area. The following is a list of topics for which the hourly time
series  of meteorological and  air quality variables give evidence.

The Imponance of Residential and Commercial
Heating Sources and Low-Lying Industrial Sources
The importance of residential  and commercial  heating sources and low-lying
industrial sources can be seen easily by an overall comparison of ambient SOj
levels  at the five stations with emission maps (Figures 6-1 and 6-2). The model
predicts station means for January to within plus or minus 20 percent.

The Importance of  Wind Direction
January 9 and 10 at TAM 4  are days of almost  constant temperature  (28°F),
wind speed (15 mph), and mixing-layer height  (2000 ft). Yet the S02 level
rose steadily from  below 0.1 ppm on January 9, to 0.2 on the 10th, as The
wind  swung  from 220°  (comparatively  low residential  emissions)  to 270°
(significantly higher residential emissions) to 300° (high residential emissions
plus  the  distant  industrial  area).  On  the same  two  days,  TAM  2  saw
approximately  the  reverse trend of upwind source concentration. TAM 3, on
January  27, responded to a  gradual shift in wind direction from NE to  NW.
6-66

-------
Large Point Sources are Assumed to  Touch Ground —
But Do Not
The Crawford  power plant  (III  in  Figure 6-1), as  opposed to others with
different  stack configurations,  is regularly predicted to reach ground  level
whenever its plume centerline  passes over JAM  5  (4  mi south),  but it is
consistently barely detected  (for example, TAM 5, Jan. 5, 10, 27, 29).

Large Point Sources Trapped Beneath a Lid
Cause  Very High Concentrations Downwind
For example:    TAM 1, Jan. 1, 4, 5
               TAM 3, Jan. 15 (note dropping lid  had  little difference
                       on TAM 2)
               TAM 2, Jan. 21, 23

Periods of Very Low Winds
Winds of < 5 mph occur on the  following  occasions:
     Jan  3  (0300-0800)   -   5 (1300-1600)  - 13 (Q600-0900)
         18,  (1000-1300)   -  21 (2000-2200)  - 23 (1800-2400)
         24  (0000-0900)   -  25 (1800-24QO)
Alrpost  uniformly,  the  model yields satisfactory predictions qf  trends and
magnitudes of S02  levels  during these critical  periods of lo\w winds and
potentially high SQ?  levels.

Transition Hours-?Sunri$e and Svnset
From 060Q to  1QQO and from 1500 tp  1900 are diurnal periods of significant
phange  in the, dispersive  capacity  of  the urban atmosphere. During these
periods  the model  is influenced by changes  |n atmospheric stability and
mixing-layer hejght.  The  model appears  capable of simulating  the corres-
ponding diurnal trends in S02  level; slthough, occasionally the sensitivity of
the relationship between  assumed  lid  height 9pd physical  stack height  of
Lipwind  point  sourpes causes large discrepancies between  predicted  and ob-
servecl concentrations (example TAM 4, Jan, 6,).

f/?e Emission Inventory
This should be extended beyond the  pity  limits. Far example, TAM § is
consistently under-predicted when winds are from the SW. The repeptor js 4
mi from  the Cicerq residential/commercial area. Approximately 20 mi tp the
southwest is the Romepyjlle Power Plant  which regularly  emits m0re than 2
x  1Q6  pounds S02  per hpijr. For a southwest wjnd at  10 mph, the plume
cpncentration on the ground at TAM 5 is approximately  0.03 ppm with  nP
||d present. Examples of ynder-predictiops §t TAIV 5 Curing SVV vyinds:
     January 11, 1?, 13  (midday)    January 9, 19 (all  da.y)
                                                                  6-67

-------
 Abrupt Changes in Windspeed and Direction
 On January 23 (1800), January 24 (0700), January 28 (2300), and January 29
 (1300)  abrupt changes in  wind speed and  direction  were  simulated with
 acceptable accuracy.
6-68

-------
REFERENCES
 1. Croke, K. and  A. Kennedy. An Air Quality Control  Program for Large Industrial
   Sources in the  City  of  Chicago.  Presented at 62d  Annual  Meeting of the Air
   Pollution Control Association. New York. June 22-26, 1969.
 2. Koogler, J. B. et al A Multivariable Model for Atmospheric Dispersion Predictions.
   J. Air Pollution Control Assoc. 77:211-214, April 1967.
 3. Turner, D. B. A Diffusion Model for an Urban Area. J. Appl. Meteorol. 3(1):83-91,
   February 1964.
 4. Clarke, J. F. A Simple Diffusion Model for Calculating Point Concentrations from
   Multiple Sources. J. Air Pollution Control Assoc. 74:347-352, September 1964.
 5. Croke, E. J. et al. Chicago Air Pollution System Model; Second Quarterly Progress
   Report. National Center for Air Pollution Control,  Chicago Dept. of  Air Pollution
   Control,  Argonne National  Laboratory. Argonne, III. ANL/ES-CC-002. May 1968.
   160p.
 6. Kennedy, A. and J. Anderson. A Computerized  Information  and  Computation
   System for  Environmental Studies. Presented at 62d  Annual Meeting of the Air
   Pollution Control Association. New York. June 22-26, 1969.
 7. Pasquill, F. The  Estimation of Diffusion from Meteorological Data. In: Atmospheric
   Diffusion. London, D. Van Nostrand Co. Ltd., 1962. p. 179-204.
 8. Slade,  D. Estimates of Dispersion from Pollutant Releases  of  a  Few Seconds to 8
   Hours  in Duration. Air Resources  Laboratories.  Washington, D.  C. Technical Note
   39-ARL-3. 1966.
 9. Croke, E. J. et al. Chicago Air Pollution System Model; Fourth Quarterly Progress
   Report. National Center for Air Pollution Control,  Chicago Dept. of  Air Pollution
   Control, Argonne National Laboratory. Argonne, III. ANL/ES-CC-004. March 1969.
10. Private  communication with  D.  B.  Turner. Bureau  of   Criteria and  Standards,
   National Air Pollution Control Administration. 1968.
11. McElroy J.  L.  and  F. Pooler, Jr. St. Louis  Dispersion   Study,  Vol.11 Analysis.
   National Air Pollution Control Administration.  Arlington,  Va. Publication Number
   AP-53. December 1968. 61 p.
12. Cramer,  H. E.  et al. Meteorological  Prediction  Techniques and  Data System.
   Geophysics Corporation of America. Bedford. Mass.  GCA Technical Report Number
   64-3-G. March 10, 1964. 252 p.
13. Bowden, K. F.  Horizontal Mixing in  the Sea Due to a Shearing Current. J. Fluid
   Mech. 27:83-95. January  1965.
14. Cramer, H. E. and R.  K. Dumbauld. Experimental Designs for Dosage Predictions  in
   CB Field Tests. Geophysics Corporation of  America. Bedford, Mass. GCA Technical
   Report Number 68-17-G., 1968
15. Carson,  J.  E.  and  H. Moses. The  Validity  of  Currently  Popular Plume Rise
   Formulas.  In:  Proceedings of the USAEC Meteorological  Information  Meeting,
   Chalk  River  Nuclear  Laboratories, September   11-14, 1967.  Mawson, C.A. (ed.).
   Atomic Energy of Canada, Ltd. Chalk River Ontario. Report Number AECL-2787.
    1967 p. 1-20.
16. DeMarfais, G. A. Wind-speed Profile at Brookhaven National  Laboratory. J. Mete-
   orol. 76(21:181-190, April 1959.
17. Carpenter, S. G. ef al. Report on Full-Scale Study  of  Plume Rise at Large Electric
   Generating Stations. Presented at 60th Annual Meeting  of the Air Pollution Control
   Association. Cleveland, June 11-16, 1967.
18. Turner, D. B. Workbook of Atmospheric Dispersion Estimates. National Center for
   Air Pollution Control. Cincinnati, Ohio. PHS Publication Number 999-AP-26. 1967.
   84 p.
19. Croke, E. J. ef  al. Chicago Air Pollution  System Model; Third  Quarterly Progress
   Report. National Center for Air Pollution Control,  Chicago Dept. of  Air Pollution
   Control,  Argonne National  Laboratory. Argonne, III. ANL/ES-CC-003. October
   1968. 254 p.
                                                                            6-69

-------
 20. Moses, H.  Mathematical  Urban Air  Pollution  Models.  National Center  for Air
    Pollution Control, Chicago Dept. of  Air  Pollution Control,  Argonne National
    Laboratory. Argonne, III. ANL/ES-RPY-001.  April 1969. 69 p.
 21. Marsh, K. J. and  V.  R. Withers. An  Experimental Study of  the  Dispersion of the
    Emissions from Chimneys in Reading - III. The Investigation  of Dispersion  Calcula-
    tions. Atmos.  Environ. 5(31:281-302,  May 1969.
 22. Croke, E. J. ef a/. Chicago Air Pollution System Model; First  Quarterly  Progress
    Report. National Center for Air Pollution Confrol, Chicago Dept. of Air Pollution
    Control,  Argonne National  Laboratory. Argonne, III. ANL/ES-CC-001. February
    1968. 106 p.
ACKNOWLEDGMENT
The authors of this paper wish to acknowledge with gratitude the significant
contributions of Messrs. J. Norcs and L. Conley of Argonne National Labora-
tory and Mr. J. Lin and the late Mr. R. Votruba of the Chicago Department of
Air Pollution Control.
The advice  and counsel of Mr.  D. B. Turner of the National Air Pollution
Control Administration, Mr. H. Moses of Argonne National Laboratory, and
the late Dr. B. Davidson of New York University were also of great value to the
atmospheric dispersion model development program.
6-70

-------
APPENDIX - GLOSSARY OF SYMBOLS

C               concentration of pollutant at a receptor point
Gn              derivative  of discrete transfer function representing value of
                curve at point of intersection, n
H               effective stack height (centerline of plume)
Hm             height of mixing layer, lid height
Hs              actual, physical  height of stack
AH             plume rise height
i                atmospheric stability class
K               constant, conversion coefficient
LP              process load
LS              space-heating thermal load
m               number fundamental time intervals over which discrete trans-
                fer function is integrated
n               iteration integer; or "look back" time
P               exponent describing stability class
Q(t)             piecewise  constant pollutant release function
r               dimensionless number used to simplify Equation (8)
T               delay time = (t -1')
T               puff travel time
U               steady  mean wind velocity
[Uj] i\[vj] 5",    constant series for variations in horizontal wind velocity
V               number of correct estimates
V'              value of V predicted from a different model
X (t,f)           concentration at time t, due to  instantaneous release, Q at
                time t' and position (0,0,0)
x               horizontal  wind distance,  downwind
V               horizontal  wind distance, crosswind
z               vertical wind distance
Az              approximate, of area-source thickness
                                                                    6-71

-------
  Z               effective source height

  a               parameter used to define a in terms of T, units:
                                                                   time
  6X' 5y           upwind distances

  A               fundamental time increment

  0h              horizontal dispersion coefficient

  ah(t)            dispersion coefficient as function of time

  ax              dispersion coefficient in x-direction
6-72

-------
ABSTRACT
      Analyses of tracer-dispersion data  in Japan, especially those of venial
      concentration profiles, indicate  that  the tracer-cloud height generally
      increases with- the downwind  distance.  The increase is larger over an
      urban area  than a rural area. When the tracer crosses over a hill, it rises
      remarkably. Over a complicated topography the rise of tracer-cloud is
      larger than  over  a flat  terrain, a  circumstance well reflected in the
      surface  concentration  pattern.  When  the  tracer-cloud  height  varies
      greatly  with  the  downwind  distance,  the   surface  concentration
      decreases uniformly downwind in spite of the elevated source.
AUTHOR
      SHIN'ICHI  SAKURABA  is  Chief of  the Applied  Meteorology  Laboratory,
      Meteorological  Research Institute,  Tokyo. He  is currently involved with tracer
      studies of medium range atmospheric dispersions in  littoral, industrial areas ot
      Japan, in connection with  the nation's  air pollution control program.  In other
      meteorological  work he has been concerned with atmospheric turbulence up to
      the 500-meter level and plume rise from a high  stack

-------
       7.   SENSITIVITIES  OF  AIR  QUALITY PREDICTION
                 TO INPUT  ERRORS  AND UNCERTAINTIES


                   SHIN'ICHI SAKURABA*
            Meteorological Research Institute, Tokyo
INTRODUCTION
In earlier dispersion-data analyses, the height of the tracer cloud was usually
assumed to be constant; this frequently resulted in an over-estimate of the
vertical  dispersion  parameter,  especially over an urban  area or a sloped
terrain. When vertical  concentration profile  data are available at one or two
points  downwind,  tracer-cloud  height   can   be   estimated   from  the
surface-concentration analysis. Some dispersion  data  in Japan are analyzed
along this line

RELEASE TIME AND SAMPLING TIME IN DETACHED
PLUME EXPERIMENT
In field  studies of atmospheric dispersion made  in urban  areas of Japan, the
drum-impactor was used at  two points downwind to measure the travel time
of the tracer cloud, the time-concentration curve and eventually the total
time, TD,  required for the passage of a  detached  tracer-plume. In  these
experiments the sampler, which measured the concentration, was actuated
before the dispersing tracer  arrived and was deactivated after the last of the
tracer was  assumed to  have  passed,  so the real sampling time, T, of the
sampler  is usually  equal to TD.  However TD or T is not always equal to the
release time, T, of tracer.

Table 7-1 shows the result of experiments in 1968.
*ln collaboration with M. Moriguchi and I. Yamazi
                              7-1

-------
      Table7-1.  COMPARISON OF RELEASE TIME,r, AND TOTAL TIME Tp
                    OF DETACHED PLUME DISPERSION3

Location of experiment
Toyama . . . 	
Wakayama 	
Oita

Total

Number of
experiments
8
8
13

29


TD=7
5
5
5

15

Frequency
TD>T
3
3
1

7


TD T is
the same as that of TD  < T ,  though the total number of cases is small.
Figures 7-1 and 7-2 show the time-concentration curves at x  = 2.5 km and x
=  5.0 km  respectively, x being the downwind distance.  T  is 27.5 min and
the   source  height,  h,  is  150  meters  (m)   The  ordinate is  1-minute
concentrations in  relative units and the abscissa is time. The  increase of TD
with  the downwind distance is small.  These cases are listed  as  TD = T in
Table 7-1.
Figures 7-3 and 7-4 show the examples of TD > r when T is  26 min and h =
200  m. In  Figure 7-3  the  concentration curve  of surface-source tracer
released  simultaneously with  the 200-meter source is also  shown  by black
circle. In case of h  = 200 m, TD  is 48 min at x = 1.5 km and 72 min at x
=  5.0 km.  The elongation of TD with  the downwind  distance is remarkable,
and  in  these  cases it may happen  that  TD exceeds  the sampling  time T.
Figure 7-5  shows an example  TD < r  when r =  30 min and  h = 150 m.
Now  it is  important to investigate the mechanism of how the time width,
TD ,  of detached plume elongates or contracts with the  downwind distance
 x ,   since  the analysis of  surface  concentration  data  composing most  of
dispersion-experiment data depends exclusively on such mechanisms.
In case of  Figures 7-3 and 7-4  half-hourly pilot balloon observations were
made at the source  x = 0  ( P,  ) and  x = 3.8 km  ( P2 ).  The wind direction
was due west and  the  pibal station; P2,  was  east of the  source. The two
drum-impactor sites D,, and  D2, were also east  of the source as follows:

X
PI
(source)
o
D,
1.5
P2
3.8
D2
5.0 km
7-2

-------
". 100
UJ
o
o
o
                       TOYAMA (JULY 24,1968)
                               D-l

              h = 150 m, TD = 15:16-15:44 = 28 min
              : = 15:00-15:27,5 m jn, ji = 2.3 msec'1, X = 2.5 km
         15:10      15:20      15:30      15:40      15:50     16:00
                            CLOCK TIME
Figure 7-1.  Time-concentration curve at  x= 2.5 km. Trj
                       TOYAMA (JULY 24,1968)
                               D-2
                  h = 150 m, TD= 15:35-16:05 =30 min
                  t = 27.5 min,j^=2.4msec-1,X = 5.0 km
    15:20      15:30      15:40      15:50      16:00     16:10
                           CLOCK TIME

   Figure 7-2.  Same as figure 7-1 out for  X - 5.0 km.
                                                               7-3

-------
                          AUGUST 7,1968 (WAKAYAMA)
                                  X= 1500m
            h = 200 m, TD = 12:12-13:00 = 48 min,t = 12:01-12:27 = 26 min
                    J4 = 2.3, JJ2 = 0.8, ji = 1.6 m sec"1
            h = 0 m,   TD = 12:20-12:52 = 32 min, -c = 12:00-12:30 = 30 min
                        1.25 msec-1
               12:10     12:20     12:30     12:40     12:50     13:00
                              CLOCK TIME

 Figure 7-3.  Time-concentration  curve at  x=  1.5km. TQ>T
                          AUGUST 7,1968 (WAKAYAMA)

                                  X = 5.000 m
                   h=200m    T = 12:42-13:40 =58 min
                              TD_ 12:42-13:54 -72 min
                              f = 12:01-12:27 =26 min
                              }i\  = 1.3,ji2 = 1.5,Ji = 1.5 msec'1
          12:40   12:50   13:00   13:10   13:20   13:30   13:40  13:50
                                CLOCK TIME

       Figure 7-4.   Same  as Figure 7-3 but for  X= 5.0 km.
7-4

-------
                          AUGUST 26,1968 (OITA)

                          X=2.6km
                          TD= 11:07-11:25 (18 min)
                          f= 10:30-11:00 (30 min)
        0
        11:00       11:10      11:20      11:30      11:40

                               CLOCK TIME
    Figure 7-5.  Time-concentration curve at  x= 2.6 km. TQ < t.
The tracer was released at 12:01, ending at 12:27. The wind speeds at 200 m
level at P, and P2 are as follows:

Time
12:00
12:30

Windspeed
(m sec"1 )
2.4
0.9


Time
12:00
12:30
13:00
Windspeed
(m sec'1 )
1.3
1.3
1.3
                                                                  7-5

-------
According to the drum-impactor data, the transport speed of the tracer is;
Speed m sec *
2.3
0.8
2.0
1.1
Part of tracer cloud
front
rear
front
rear
x, km
0 to 1.5

1.5 to 5.0

This, compared with pibal data, suggests that the elongation or contraction
of  TD  occurs by the windspeed difference between the front-part and the
rear-part of tracer cloud.

Next we will  try to compute TD based on pibal data. If the front of tracer
flows with the speed 2.4 m sec"1, it will reach Dj at
         12:01 + 1500 m/2.4 m sec"1 = 12:11
 and P9  at
         12:01 + 3800 m/2.4 m sec"1 = 12:27
 The corresponding arrival time of the rear

         12:27 + 1500 m/0.9  m sec"1 = 12.55

         12:27 + 3800 m/0.9  m sec"1 = 13.37

 The calculated TD is, at  DI

         12:11 - 12:55 = 44  min.

 While the observed TD is

         12:12 - 13:00 = 48  min.
From P2 on, the tracer speed is 1.3 m sec ', so the arrival time at D2 is

         12:27 + 1200 m/1.3 m sec"1  = 12:42      (front)

         13:37 + 1200 m/1.3 m sec"1  = 13.52       (rear)

                TD  = 70 min

The observed TD at D2 is

         12:42 - 13:54 = 72 min.
7-6

-------
The coincidence is good, supporting the view that the elongation or contrac-
tion of a detached plume is due to the windspeed difference at its front and
rear. The sampling time T, by the way, and the estimated or observed TD in
this case are as follows:
X
T
TD
0.6
33
33
1.5
48
[48]
3.0
57
68
5.0km
58 min
[72] min
Here the  numeral  in brackets shows the observed value. Evidently samplers
at 3 km and 5 km  downwind were stopped before the tracer passed by.

AVERAGE CONCENTRATION AND DISPERSION EQUATION
The dispersion equation for a continuous point source is:
— -  C(x,y, z)  =
                           [ffy(x), crz(x), h,y, z]
                                                          (1)
 and the equation of mass continuity is
                  u  C dy dz  =   Q
                                                           (2)
 The notations are customary ones. In dispersion-data analysis we must refer
 to Equation  1. Some kind of average concentration conforming to Equation
 1 must then  be introduced to compensate for having only the exposure data,
 especially when TD  is not equal to T, or when TD is not known.
 Let C be the instantaneous concentration. The exposure, E, and the average
 concentration, C are
E  =
                   C dt
                                                           (3)
        C  (TD)
                       Tn
                              Tr
                          C  dt   = E/Tr
                                                          (4)
 where TQ was already defined in the previous section and t is time.
                                                                  7-7

-------
Case 1.      TD  = T

This corresponds to the case when the sampler has captured  all the tracer
material passing by, which is usually the case with the dispersion experiment.
Then, from the equation of mass continuity,


         ^   tf
         I   dt  I  I u C dy dz    =  Qr                              (5)


Qr shows the total mass of tracer released. Dividing by T,
                   u C  dy dz  = Q r/T                             (6)
where

                                  ,T
                   u C   = 	 /   u
                            T7
Cdt «  u  C  (T)
                                      T

                   C(T)  =   —  /    Cdt
So Equation 6 becomes
                u  C (T) dydz  =  Qr/T=Q'                        (7)
The dispersion equation corresponding to Equation 7 should be

        	           	     T
        u C (T) / Q' =  u C (T) -  =  /
Equation  7 shows  that  the source intensity corresponding to the average
concentration  C(T)  is  not Q, but Q'. Q coincides with Q'  only when T -
r.

Equation 8 may be simplified, by introducing the concentration C,  defined
by
7-8

-------
          •/   ca'-f  '
.T
                C'dt                                (9)
and the corresponding average concentration C (T):

                   1  r
       C  (T)   =   r  I    C' dt                                  (10)
    1   /"T
    rj    -
Namely Equation 8 becomes

     if C  (r)  / Q  =  /                                           (11)
Equation  11  does not contain T explicitly and is convenient for the con-
centration analysis.
Case 2.    TD > T

This is a rather rare case and sampling time does not cover the whole range
of the time concentration curve. Such an example was already shown (Figure
7-4). The best estimate for the equation of mass continuity in this case, is
                          uCdydz =  Qr —                    (12)
                                           TD
which yields:

        u  —  C (T)  I Q  = f                                     (13)
 Equations  11 and  13 will be used later in the surface-concentration analysis.


 Estimation of tracer-cloud height
 based on  surface-concentration analysis

 Since 1967 the vertical measurement of tracer  material  has been made at
 two  points downwind, from  which the center-line height, z ,  of the tracer-
 cloud as well as the vertical dispersion parameter,  az, is derivable. Examina-
 tion  of 45  vertical  concentration   profiles,  made  it clear that  z varies
                                                                   7-9

-------
remarkably  with the downwind distance, and in turn seriously affects the
surface concentration pattern. In the sea breeze  layer, in which most disper-
sion experiments were  conducted, z generally increases with the downwind
distance. The increase is larger over an urban area than a rural area, and over
a  sloped  area  than a  flat area. Next we  will  examine  the  possibility of
estimating, z from the surface concentration.

We will begin with  the dispersion equation:

     -~C(Cx, y, z) =  / [ay(x), az(x), "z(x),  y, z]                  (14)


z is the center-line  height  of  the tracer cloud or the gravity-center height of
concentration and a function  of downwind distance x.
C/Q is:

           C/Q  =   —   C(T)  /Q                                   (15|

as proved in the previous section, and TD/T  is in most cases unity.

To determine z (x) from  Equation  14, the transport speed  u ,  the source
intensity  Q  , the  horizontal  dispersion parameter ay and the vertical disper-
sion parameter  az  must  be  known. The determination of ay is  simple and
needs  no special comment; u  may  be accurately determined by the drum-
impactor  record. At one or  two points downwind we  have the concentra-
tion-profile  data, or z" and  az.  Then Q is determined from the Equation 14,
with the concentration data The constancy  of Q may also be checked if the
concentration-profile data  are  available at  two  points downwind.  It was
found  that  Q   is  almost constant  down to  at  least  5  km, as far as the
elevated-source dispersion is concerned.

Now, from Equation 14 ,

      uCa/Q  =  / [ay(x),az(x), z"(x), 0, 0]                       (16)

Ca = C(x,0,0,) is the axial concentration at the surface. We already have z and
az  at  one  or two  points  downwind.  Then,  from  az  ,  the corresponding
Pasquill  stability could be  determined. If  the Pasquill stability  thus deter-
mined is fairly  constant over the downwind distance concerned, the estima-
tion of  z (x)  from Equation 16 becomes  possible. It was  made clear from
the measurement of az  at  two  points  downwind that in the case of elevated
sources the  Pasquill stability  can be  regarded constant down to about 5 km
distance, even  over an  urban  area.- It is important to note here  that the
Pasquill stability decided from the observed  az  should be adopted.
7-10

-------
In the following sections we will give some examples of  z (x) over an urban
area analyzed along the line stated in this section.

Yokohama experiment
Figure 7-6 is the  sampling  network of  the  Yokohama  experiment.  Two
sources, 60 m and  113  m high, were placed at the Negishi industrial area, at
the southern  extremity of Yokohama City. From the 60-meter source to 1.6
km arc, the terrian is flat. Beyond the  1.6 kilometer arc, there is a tableland
about 50 m  high  bounded  by a cliff (see Figure 7-10). Beyond the 3.5
kilometer arc is the central  part of  Yokohama  City. In  the  experiment
of July 26, 1967,  two fluorescent  particle  tracers  were released from a 60-
meter source and 113-meter source simultaneously. The  tracer  drifted in the
southerly  wind;  hit  the tableland;  crossed over  it; and flowed  over the
congested area of Yokohama City.
         Figure 7-6.  Network of Yokohama experiment.
                                                                 7-11

-------
The concentration-profile sounding was made at two sites, one at x = 2.3 km
on the tableland and the other at x  = 4.8 km  near the coast of Yokohama
Port, the downwind  distance  x  being measured from the 60-meter source
Figure 7-7 shows the concentratron-prof ile at x = 2.3 km The analysis is usually
made, assuming the Gaussian  distribution of concentration in  the vertical
The profile in  Figure 7-7 does not lend itself to such an analysis; however it
is certain that z is larger than 250 m, the  highest level  of sounding.
               h=60m
 JULY 26,1967 (YOKOHAMA)
         10:00-10:30
X = 2.3km      Z = 295m    C7z = 51m(D)
0 08
0.07
n OK
1
8 0.05
f
O
o- °-04
0
0.03
n n?
0.01
0.00

-









— • — 1 — • —











1 •










•
1
•











                 0     40     80    120    160   200    240    280
                                      Z, m

Figure 7-7.  Concentration profile at  X = 2.3 km. source height
meters,
7-12

-------
Figure 7-8  shows two soundings:  one is at x = 4.8 km from the 60-meter
source and the other  is at x = 3.7 km  from the 113-meter  source.  In the
former z is 170 m and az is 56 m,  the Pasquill stability,  E;  , while in the
latter z is  175 m  and az is 54 m, the Pasquill stability, D  to E.  In the
following the Pasquill stability E will be adopted to treat the dispersion from
the meter 60-meter source.
o h = 60 m
• h=113m
       u
       X
      t,
                      JULY 26,1970 (YOKOHAMA)
                              10:00-10:30
                    X = 4.8 Km
                    X = 3.7 Km
= 170m

-------
  Figure 7-9 shows the result of  the surface concentration analysis made with
  the procedure stated in the previous section, The curve is the isoline of z, in
  meters, when the Pasquill stability is E. The white circle shows only how the
X
Z
           cv
        JULY 26, 1967 (YOKOHAMA)
    10:00-10:30   h - 60 m   p = 5.0 m sec'l
(D)     0.8    1.6    (2.3)    3.5    (4.8)
(60)     50     70   (>250)   150    170
                                                           5.0km
                                                           170 Hi
                 0.1
            1.0            10.0
              DISTANCE, km
                                           100.0
                                                  STABILITY, E
Figure 7-9.  Surface axial concentration versus downwind distance
curve, case of Pasquill  stability  E.  Labelled numeral shows tracer-
cloud  height, meters.
  7-14

-------
normalized  surface axial  concentration  varies with the downwind distance
and has no connection with z .  The black circle shows the surface concen-
tration modified such that xy conforms to that in Pasquill  stability E.  This is
because xy and xz are usually  different in Pasquill stability. In the case in
Figure 7-9, the Pasquill stability of xy is  higher than E at x =  0.8 km and
1.6 km, while at x =  3.5  km and  5 km it is about the same; z can be read
from the black circles  in Figure  9.

The  result, depicted together with the  topography in  Figure 7-10  is sum-
marized as follows:
                       JULY 26. 10:00-10:30 (YOKOHAMA)
    N
        100
Figure 7-10.  Tracer-cloud height variation with downwind distance
and topography  in Yokohama experiment.
     x        (0)    0.8      1.6      (2.3)    3.5     (4.8)    5.0m
     "z       (60)    50       70      (>250)  150     (170)   170m
 Here z is estimated from the surface concentration, except for the bracketed
 numbers which are the results of direct analysis of concentration profiles or
 the source height.

 An 0.8-km arc  runs  along  a  pier. The tracer released  from the  60-meter
 height flowed  north,  descended  a  little  over a  small canal ahead  of  the
 0.8-km arc, and  began to rise, passing over the industrial plant area between
 the 0.8-km and 1.6-km arcs. Crossing the tableland beyond the 1.6-km arc it
 rose remarkably,  then descended over the congested area. The difference of
 z between x = 0 and  x = 5 km is 110 m.  Because of such large variations of
 z,  the surface concentration decreases rather uniformly with the downwind
 distance and does not  have the value, characteristic of an elevated source.
                                                                    7-15

-------
Mizushima experiment
Mizushima is an industrial  area facing the Seto Inland Sea,  and  Figure 7-11
shows  one  example  of the  analysis  in  the  Mizushima experiment. The
Pasquill stability is D, which is due to the concentration profile sounding at
x = 5.2 km; and the source height is 90 m.

           x         (0)    1.0      2.0      3.0     5.0     (5.2) km
           I       (90)     65      120     170     130     (130) m

The  1.0-km arc runs along a pier and there is a small canal ahead of this arc,
similar  to  the situation in the Yokohama  experiment. The tracer flowing
north  descended  over  the canal  and  began to  rise significantly over the
congested plant area  between  x = 1.0 km and  x  = 3.0 km.  Beyond the 3.0
km arc there follows the  flat, rural  area with  scattered small towns. There
the tracer descends
Toyama  experiment
The  experiment was conducted at the western suburb of Toyama City, facing
the  Japan Sea,  a  flat rural  area  extending  to  5.0 km arc. The  150-meter
source was sited near the coast and  the tracer  flowed  inland (Figure 7-12).
The  tracer flowed  horizontally down to the 5.0  km arc and rose at the 8.0
km  arc where there are hills.  The  Pasquill stability  was  D.  The surface
concentration curve is typical of the elevated source, since z does not vary.
Kainan City experiment
Kainan, south of Wakayama City, is situated  in a narrow valley surrounded
north and south by mountains about  150 to 400 m high and extends east-
west. The  west end of the  city  is open to the sea.  The 200-meter and
surface-sources were placed at the mouth of the valley. The tracer flowed in
over Kainan City with the west wind.
Figure 7-13  shows the concentration  profile at x = 550 m; for the source
height of  200  m;  z  154 m;  and  az was 33 m; the  Pasquill stability was C.
Figure 7-14 is the  same profile, but for x  =  1.5 km.  Here the ordinate is the
ratio of the  source  to surface  concentration  and the  abscissa is z / z,
parameter  £  being  az / z .  Figure  7-14  shows that  z  is  200 m and  CTZ  is
126 m, when  the  Pasquill  stability is B-C.  Two kinds  of Pasquill stability
were used,  B-C and C.  In  the  other runs  of  the  experiment  the Pasquill
stability was all B-C, so in this case the same stability was also  adopted.
7-16

-------
                   AUGUST 9,1967 (MIZUSHIMA)
                     14:30-15:00  h = 90m
       X
       Z
 (0)
(90)
            0.1
1.0
65
2.0
120
3.0
170
5.0
130
             1.0            10.0
                DISTANCE, km
(5.2) km
(130) m
                                 100.0
Figure 7-11.  Same as Figure 7-9 but for PasquiM  stability  D,
Mizushima experiment.
                                                              7-17

-------
                        JULY 24, 1968 (TOYAMA)
                        15:00-15:27.5 h=150m
                           1.0 0(0-07)     10.0

                              DISTANCE, km
                                                  8.0km
                                                  180m
100.0
  Figure 7-12. Same as Figure 7-11 but for Toyama experiment.
7-18

-------
                     AUGUST 7,1968 (WAKAYAMA)
       L,
       X

      t,
15
10
5
12:00 h 200m Z 154m
<*i 33 m X 550 m


1

0







"0 80 160 240
Z, m
 Figure 7-13.  Concentration profile at X = 550 meters, source
 height  200 meters.    Pasquill stability C.
                     AUGUST 7, 1968 (WAKAYAIY1A)
                              12:00

                       h -- 200 m      Z - 200 m
                       X-1.5 km     ux=126m
Figure 7-14.  Normalized concentration profile at X   1.5  km,
Pasquill  stability B-C, Wakayama experiment.
                                                            7-19

-------
 Figure 7-15 is the final result of the analysis. The tracer descends about 50
 m at the 0.6-km arc,  resumes its initial height at the  1.5-km arc and then
 begins to rise remarkably. At about 150  m  east  of the  source there  is a
 building  about 50  m  high, which  seems to have  affected the lowering of
 tracer to the 0.6-km arc.  This is reflected  in the concentration profile from
 the surface-source released simultaneously with 200-m source (Figure 7-16) z
 being 130 m at x =  1.6 km.
                         AUGUST 7,1968 (WAKAYAMA)
                            12:01-12:27   h = 200 m
                                                      5 km
                                                    >450m
          10-4 F
                           0.5    1.0           5.0    10.0
                                DISTANCE, km
                                                 STABILITY B-C
Figure 7-15.  Same as  Figure 7-9 but for  Rasquill stability F>C,
Kainan City experiment.
7-20

-------
                          AUGUST 7,1968 (WAKAYAMA)
                                   12:00
                            h = 0       Z =130 m
                            X = 1.5km  crz=126m
      2.01  i  i  i
       1.5
   "1.0
   o
      0.5

                                             J  •   ' -•	'	L
                   0.5
1.0       _ 1.5
       Z/Z
2.0
2.5
 Figure 7-16.  Normalized profile of surf ace-source concentration
 at x = 1.5 km, Wakayama experiment.
The weather conditions  such as stability  and wind  speed or profile were
almost  the  same as the  case in  Figure  7-12. The big difference is  in the
terrain. The Pasquill stability is certainly much affected by the topography.

Oita experiment
There are many examples  to show  that the tracer-cloud released from the
surface-source rises  more  or less over an  urban area. One example is  picked
out here from the Oita experiment.

The surface source  was near  the coast and the tracer flowed over Oita City.
The concentration  profile at  x = 2.6 km  is  shown in Figure 7-17;  z is 65 m
and the Pasquill stability  is  D. Figure 7-18 is the result  of the analysis:
                                                                  7-21

-------
                               AUGUST 27,1968 (OITA)
                                       12:00
                              h = 0 m        Z = 65 m
                              X = 2.6 km      crz = 65 m
 Figure 7-17.  Normalized profile  of surface source concentration at
 X   2.6km,  |=crz/Z Pasquill stabil ity D.

           x     (0)      0.6      1.5     (2.6)    3.0      5.0km
           7     (0)      45      65      (65)    80      140 m
The increase of 7 with the downwind distance is conspicuous.

CONCLUSION
Examining dispersion-experiment data in Japan,  we found that  the height of
the tracer cloud generally  increases with the  downwind distance. The height
increase is larger over an  urban area than over a rural area. Therefore,the
earlier analysis,  assuming  constant  height for the tracer clouds, tends to
overestimate   az  ,   especially over an urban  area. The  height variation also
affects the surface concentration. When the tracer crosses over  a hill, its rise
is remarkable. In general, the tracer  rise  is larger  over  a  complicated topo-
graphy. Over a flat terrain the rise is small.

The  source  heights  of  our experiments  were  0 to  200  m.  The present
conclusion is valid in this source-height range
7-22

-------
            X     (0)
            Z     (0)
             1C'2 r
 AUGUST 27,1968 (OITA)
  10:00-10:30   h = 0 m
0.6     1.5     (2.6)     3.0
45     65      (65)     80
                0.1
     1.0            10.0
        DISTANCE, km
5.0 km
140m
   100.0
Figure 7-18.  Surface axial  concentration versus downwind distance
curve,  Pasquill  stability D.
                                                                 7-23

-------
 APPENDIX -  GLOSSARY OF SYMBOLS

A, B, C, D, E  Pasquill stability ratings
C             instantaneous tracer concentration
C             average tracer concentration
D!,D2        drum-impactor sites
E             exposure time
H             source height
P.             pibal station, i
Q             source intensity
QT           total mass of tracer released
t             time
T             real sampling time
TD           total time required for passage of a detached tracer plume
u             transport speed
x             horizontal (downwind) distance from source
z             centerline height, or gravity-center height
I             ffz/^
ffy            horizontal dispersion parameter
°z            vertical dispersion parameter
T             tracer release time
                                                                    7-24

-------
ABSTRACT
      With the advent of computer-oriented simulation models of the physi-
      cal and chemical system that produces varying levels of air quality, it is
      both possible and desirable to assess  the  model's sensitivities to errors
      and  uncertainties  in the  input variables.  From such  analyses, it is
      possible to  derive more explicitly  the levels of accuracy  that must be
      observed  in  the specification  of the  inputs for  source strengths and
      distributions,  wind  velocity,  horizontal and vertical  diffusion rates, and
      pollutant chemical reactions or physical decay and loss rates, if the air
      quality predictions  derived  from  these inputs are to  remain within
      useful  limits of accuracy and uncertainty.

      For  the present report,  a  single  case study is utilized to test the
      sensitivity  of  the  Travelers  Research  Corporation  Regional Model to
      random  and  systematic  errors in  the source-strength  input  and to
      systematic errors in the remaining input  variables. It has been found
      that random errors  in  the source strengths do not produce comparable
      errors  in the air  quality prediction. Beyond  this, and in order of
      decreasing sensitivity,  the  TRC model has been found  to  be highly
      sensitive to systematic  regional  wind-direction errors and moderately
      sensitive to systematic errors in source-strength estimates,  decay or loss
      rates, and vertical diffusion rates.  The model is insensitive to errors in
      lateral  diffusion rates,  at least for  the  multiple-source distribution
      encountered in Connecticut.  These sensitivites  are  quantified  for the
      case studied.

      The case used for the sensitivity analyses also provided an opportunity
      to  determine  the  effect  of  changes in  the  input  variables  on the
      verification  of the  model, since  observations of SC>2  were available
      from the verification program. Significant improvements in the TRC
      model verification,  over that obtained from the original, independently
      chosen, input  variables,  were found when  either the source strengths or
      the decay or  loss rates were adjusted appropriately.  This result suggests
      the  need for  much better  understanding  of  the  decay and  loss of
      airborne SC<2.
AUTHOR
      DR. GLENN Ft. HILST, a native of Kansas, holds degrees from MIT and the
      University of Chicago. He  has  worked in  association with the General Electric
      Company, Argonne National Laboratory, and The Travelers Research Center on
      various facets of air and water quality control and environmental safety. Dr. Hilst
      is presently editor of the Journal of Applied Meteorology.

-------
                          8.    ELEVATION  OF TRACER  CLOUD
                                          OVER  AN  URBAN  AREA'


                       DR.  GLENN  R. HILSTf
                The Travelers Research Corporation
INTRODUCTION
The advantages of a capability to simulate the physical and chemical behav-
ior of the air quality system are so obvious as to require no further emphasis
here. The very existence of this Conference and  the reports we have heard
on  the  state-of-the-art  for  simulation  modeling  (within useful  limits  of
accuracy) also  speak  to these advantages,  even the need for  an ability  to
evaluate alternative air quality  control  strategies before  massive and costly
programs are instituted. This  is the ultimate utility of simulation modeling;
happily, along  the  way, these modeling attempts also  provide us with the
possibility of unusually clear  recognition of where our ignorance, errors, and
uncertainties lie. It  is  on this facet of simulation modeling  of air quality
systems that I should  like to focus.
In particular,  our work at Travelers Research  Corporation (TRC)  now per-
mits us to make a preliminary assessment of (1) just how  faithfully the TRC
Regional Model simulates the real-world air quality  system  in Connecticut
and (2)  the sensitivities of that model's air quality predictions to errors and
uncertainties in  the input  parameters and  variables. I shall discuss these two
topics in reverse order.

At the outset, let  me  emphasize that these  studies  have utilized only the
TRC  Regional  Model;  therefore, the results are  limited  to  that particular
version.  Since the basic methods employed by the model are quite conven-
tional and  common to  most modeling efforts, it is reasonable to assume that
'The work reported here was supported as an in-house program development effort by
 TRC.  The TRC Regional Model was developed under a grant from  the Connecticut
 Research Commission, however, and  the verification data used here were  obtained
 under joint support from the Connecticut Research Commission  and  the National Air
 Pollution Control Administration.
tThe very substantial assistance of Mr.  Charles R. Case, Computer Operations, and the
 advice and counsel of Dr. G. D. Robinson, Mr. N. E. Bowne, and  Dr. R. R. Hippler
 are gratefully acknowledged.
                                  8-1

-------
the sensitivities found  are appropriate to other models that also purport to
predict air  quality  with a  similar time and  space resolution. I  leave this
extension to your own good judgment.

TRC  REGIONAL MODEL
The model whose sensitivities and predictive capabilities are under study here
has been  reported on extensively over  the  past 2  years. Verification of the
model was undertaken in a  1968-69 study  that is  the subject  of a report by
the TRC  staff4  and a preliminary paper by Bowne,5  both of  which are just
now becoming available.
In brief, the TRC model is designed to predict  ground-level  concentrations of
a semiconservative pollutant  emitted  from  multiple sources within a region
of the order of 5,000 square  miles.  The model is capable, in principle, of
resolving  those  concentrations on an hour-by-hour  basis, with a space resolu-
tion of 5,000 feet. Since the design of the model is quite general with regard
to time and  space resolution,  the only  real limitations on  these features are
(1)  computer  capacity and speed, (2)  ignorance  regarding the appropriate
decay  and loss of a pollutant over great distances of travel, and (3) a very
large  degree  of uncertainty regarding the transport and dispersion processes
over very short  distances  of travel. As presently configured, the model fits
nicely into  the  IBM 360/40 and  requires  about  10  minutes of that com-
puter's time  to provide  a  100-point prediction  of  any  single  air quality
pattern.
The TRC model employs the  conventional  Gaussian diffusion  representation
of dilution  of the pollutant  (with  a simple exponential decay)  from each
individual area or point source that could contribute to the concentration at
the point  (or small area) and time of interest. These sources are identified by
computing the prior trajectory of the air resident over that  point at the time
of  interest and by noting the lateral  dispersion appropriate for the  traverse
from the  source to the point of interest. Chronological variations in the wind
field and  atmospheric  dispersion are accounted  for by changing the  input
variables at 2-hour intervals.

The basic equation  used is shown  in Figure 8-1; the  total concentration at
(x,  y,  0)  is obtained by  summing  over  all  appropriate  source locations
(£,??£)• Except  for the transporting wind velocity  field, which defines (x -
£),  (V  — ??),  and (z —f), all  of the input variables  for air quality prediction
are  explicit in this equation. They are as follows:
   Q^,1?,?)    =  the source strength at x = £, y = 77, z  = f •
   u(l,7?,f)    =  the  mean  wind  speed  at  the source  during the time of
                emission (strictly, the harmonic mean).
   u          =  the  transporting wind   speed  that determines the  time of
                flight from x to £,  etc.
8-2

-------
        X(x,y, z |  S,rj,:
Q
            exp -
                                                 X (x - |)
                                                             exp
                                                                          -i
                                                                     o-y(x
                                                                                       exp
                                                                    (x  -
                                                                       V
                                                                        c
                                                                   V
                                                                    D
Class of error
and uncertainty
Source strength
Decay or loss
Di spersion
Positional
Terms of equation
affected
A
B
C, D
A, B, C, D
Nature of error
and uncertainty
Both random and systematic likely
Largely unknown - tends to be systematic
Tend to be systematic
Both random and systematic likely
CO
                 Figure 8-1.  Classification of  input errors and uncertainties for model  sensitivity tests.

-------
   X          = the decay or loss coefficient for  the  pollutant under consi-
                deration.
   ay{x-£)    = the  standard deviation  of  the  cross-wind  spread  of the
                pollutant  eminating from ?,7?,f.
   CTZ(X-£).    = the standard  deviation of the  vertical spread of the pollu-
                tant emanating from ^T^f.

These  are the input variables that we can manipulate in seeking concurrence
between predicted and observed  air quality  levels and to which these pre-
dicted levels will be more or less sensitive. Our first study  is an attempt to
quantify these sensitivities  via the model itself.

POTENTIAL ERRORS AND UNCERTAINTIES IN THE
ESTIMATES OF  INPUT VARIABLES
For  a  typical air quality pattern prediction, the TRC model requires approxi-
mately 200,000 input estimates. The  probability of errors and uncertainties
in these input estimates is  very high.  Figure  8-2 is a pictorial representation
of the different  mistakes  that could  be  made  for a single source situation
and, as noted in the figure, there are eight distinct sources of input error and
uncertainity. They are:
   1. Misplacement of source with respect to  geography and elevation.
   2. Misestimation of source strength.
   3. Misestimation of wind speed at source.
   4. Miscalculation  of trajectory  of  pollutant  cloud  or  plume,  due to er-
      roneous or uncertain wind fields.
   5. Misjudgment of vertical diffusion of cloud or plume, including errone-
      ous distribution functions, as well as parametric errors.
   6. Misjudgment  of  the   lateral diffusion  of cloud  or plume, including
      erroneous distribution functions, as well as parametric errors.
   7. Misestimation of decay or loss of pollutant.
   8. Misplacement of receptor or monitor.

This appears to be an  exhaustive list; at any rate, it is a sufficient one for
our present  purpose.

Figure 8-1 shows that, for  sensitivity-analysis  purposes, these eight sources of
error and uncertainty may  be grouped  into four  convenient classes:
   1. Source strength estimates.
   2. Pollutant decay or loss estimates.
   3. Pollutant diffusion and distribution estimates.
   4. Positional errors and  uncertainties.
Of these, the first three are self-evident  and  essentially independent  of one
another. The fourth, however, in  effect,  affects all of the  other estimates
since,  as shown in Figure  8-1, each of the terms in the  diffusion equation
8-4

-------
      SOURCE MISPLACED
      SOURCE STRENGTH MISESTIMATED
      WIND SPEED AT SOURCE IN ERROR
      TRAJECTORY MISCALCULATED
      VERTICAL DIFFUSION MISESTIMATED
      HORIZONTAL DIFFUSION MISESTIMATED
      LOSSES MISESTIMATED
      RECEPTOR MISPLACED
Figure 8-2.  Sources of input errors and  uncertainties in variables
utilized  in air quality simulation models.
upon which the model is  based  depends upon  the positions of the source
and the receptor. Errors and uncertainties in estimating the relative positions
of the source  and  the  receptor, and in estimating the time of flight and path
between the two,automatically involve errors in the estimates of the source
strength and the decay or  loss and diffusion  of  the pollutant. For example,
if  the transporting  wind direction  is misestimated, the receptor "sees" the
wrong sources, the time of flight  is misjudged, and the diffusion is misesti-
mated. The only time  when no "error" is involved occurs when the sources
and meteorological  conditions  are  absolutely  uniform over the region, a case
of less than trivial interest.
The nature of likely errors and  uncertainties in the  input parameters  is of
considerable interest, particularly the distinction between random  and sys-
tematic errors. In  preparing source  inventories,  we fully expect a random
component of error since we can  seldom even measure the output of a given
source precisely, and we know that most sources do not operate at a single,
steady emission rate.  On  the other  hand, we should  also expect some
systematic  component  of error, with a tendency  for this component to be in
the nature  of  an underestimate. This is  because of the high probability that,
amidst thousands and tens  of thousands of sources, we will tend to overlook
some entirely, but  will not  be  inclined to compensate  with  imaginary
sources.
                                                                   8-5

-------
On the other  hand, quite irrespective of the reality  of  simple exponential
decay  or  losses of  the pollutant,  the error in estimating  this rate of loss is
apt to  be systematic in  nature.  For the more complicated  photochemical
reactions  and  product formation  rates, for which I  have not seen anything
more than rudimentary  simulation models, a mixture of random and system-
atic errors may appear.
For short times of flight  and averaging, we would expect,  from the nature of
atmospheric  diffusion,  to see  sizable  random  errors in the estimates of
vertical and  lateral  diffusion  rates.  Over  longer  distances  and times of
averaging,  however, when these  local variabilities have  been suitably smooth-
ed, any error should tend to become systematic. Since the model requires at
least  1-hour averages through these variations, our primary problem  is pro-
bably with systematic over- or under-estimates  of diffusion.
The  positional errors are of two kinds: (1) geographic misplacement, which
is largely  a  surveying  problem  (for  fixed  sources and receptors)  and  a
fixed-point referencing problem (for mobile sources and  receptors), and (2)
advection  or  trajectory  errors due to erroneous estimates of  the wind field
and  location  within the  wind field. Suppression  of the geographical error
depends upon the precision one is willing to pay for in locating sources and
receptors;  these errors are probably random in nature. The trajectory errors,
however,  will depend upon  the precision  and  spatial   density of airflow
measurements at  various altitudes  in  the  lower  levels of the atmosphere.
Although  both random   and systematic  errors  are  probably present, the
requirements  for  maintaining mass continuity for our representation of these
flow fields operate strongly for systematic errors. The wind fields in adjacent
grid  cells are  not independent of one  another; an error in one estimate  is
reflected  in  neighboring  estimates by this  generally positive correlation,
through the equation of continuity.

Finally, as is evident from the choice of words in the last few  paragraphs, we
recognize  the conventional  distinction between  errors  and uncertainties:
errors  can  be  identified  and  corrected  whereas  uncertainties  cannot be
corrected.  Beyond this, the effects of these  "mistakes" are indistinguishable
for our present purposes, and we  shall  refer to them simply as "errors." The
distinction between random and systematic errors is of much greater  impor-
tance.

THE TESTS FOR SENSITIVITY
The  objective  of  sensitivity tests  is, of course, to measure the error in the
predicted  quantity,  in this case,  pollutant concentration, as  a function of
increasing  errors  in  each  of  the input variables. The  rate of change  of the
predictand error as  a function of  the rate of change  of the predictor error is
the measure of the sensitivity of the former to the latter. Since in air-auality
8-6

-------
pattern predictions (the regional  problem)  we are  dealing  with numerous
predicted  values  that  range through  several  orders of magnitude, it is con-
venient to measure the relative error rather than the absolute error in the air
quality predictions.

Let
     XT   =  the "true" concentration of a pollutant at a given station,
     XE   =  the estimate of concentration after an input parameter error is
             introduced.
Then

  —-	 =  fractional error in X produced by the input error.
    Xt

For  N separate estimates, we have
    XT-XE_   1
and

          2  =
where equations (1)  and (2)  specify the mean and variance of the fractional
error distribution, respectively. We have chosen these distributions and their
mean  and variance  (or  the standard  deviation, [(v'lj2) as the  measures of
predictand errors. We should note that the value of (XT-XE) /XT  must be
less than or equal to one, since  XE  cannot be less than  zero. Assumptions
with regard  to the distribution  functions  for the fractional errors do not
arise in  the present analysis,  but  with this  latter constraint, we expect these
distributions to be skewed toward  negative values of  the  fractional  error of
X.

The technique used for  assessing the fractional error in X was simply: to run
the model for a reasonable set of input parameters, solve for the concentra-
tion at  100 stations uniformly spaced over the Connecticut input parameters
in a stepwise  fashion. The mean and  standard deviations  of the fractional
error in X were then constructed  for each solution referenced on  the original
or "true" solution.

In order to  assure a  systematic  and  controlled variation  of the  input-para-
meter error  (and, incidentally, to conserve computer time), the following
error generator was employed:
                                                                      8-7

-------
 Let
      XT  = the original or "true" value of any input parameter;
      XE  = the erroneous value of the input parameter.

Then take
      XE  = XT (1 + a + b),                                           (3)
where  a   is  the  fractional  error  in X  due to random errors and  b is the
fractional error in X  due to systematic errors. Then

      XT~
        X
                =  ~b  +  a                                          (4)
and  — b + a <  1   In order to give  a  the character of a random variable, it
was  assigned the values   +|a|,o,-a    with equal frequency and applied to
the collection of input variables, such as the source strength Q, on a random
basis.  Likewise,  a systematic error was generated by assigning a single value
to b .
      It follows directly from Equation (4) and the assignments of a that

                        =  -b                                       (5)
and
i.e., the mean and  variance of the fractional input errors are specified sepa-
rately by  a and  b.

For the present sensitivity test, a random error analysis has been done only
for the source strength  field,  Q .  Analyses of systematic errors have been
completed for all  of the input parameters except windspeed. The measures
used are shown in Figure 8-3.

RESULTS OF  SENSITIVITY TESTS
The situation chosen for these  sensitivity tests was based on  a real-world case
for which verification  measurements of SO2  levels over a 21-station network
covering  Connecticut  were available. The exact period was from  0600 to
0800  on  October  30, 1968. (A  2-hour average was used in the verification
program  in  order to  assure  significant sampling of S02, hence this  2-hour
test  period,  rather than  1-hour.)  The  weather  prior to  the  morning of
October 30th was  mostly  clear to partly  cloudy, with  NW  winds  at  4 to 6
meters per second  (m  sec"1}. Surface temperatures during the night were in
the low 30's with some reports of frost.

-------
                     Prediction
                   error measure"
                Input error measure
                                                    Class
                   Random error
Systematic error
                         and
                                                  Source
                                                   strength
Decay
                                                  Dispersion
                                                  Positional
                                                                                 1/2
                      - Q
                                                                                             'Q
  (XT - \E)
                                                                                                 -
00
(b
              Figure 8-3.  Measures of input and air quality prediction errors used  in sensitivity analysis.

-------
 The source inventory for SO2  emissions in  Connecticut  is shown in  Figure
 8-4A. A summary of this inventory shows the following  frequency of emis-
 sions in pounds per day per square mile (Ibs day"1 mf2}.
        rate (Q)
Source cell area
                 Ib day * mi
    100  -
     10  -
      1   -
1000
1000
 100
  10
   1
                                        Number of source cells
                                          per25x106ft

                                                   27
                                                   98
                                                  827
                                                3,248
                                                1,358
As is  clear from  the  map  and summary, the emissions of S02  over this
region are  far from  uniform,  but  they do exhibit  a  pattern of rural land
usage  in the  northwest  and  northeast sections  (with some notable  local
anomalies)  and  an  industrialized  and  urbanized  belt from southwest to
northeast through  the central portion of the region.  There is also a relatively
strong source region  extending  northward along the  Thames  River in the
southeast section of the region.
  Figure 8-4A. Source inventory for SC>2 emissions in Connecticut
  used in TRC regional model.
8-10

-------
                                                                  PUTNAM
                                                                   •
                                                              PLAINFIELD
                                                                       r
                                                                     0.005
                                              0.01
                                                I STAFFORD
                                                '  •
                                 THOMPSONVILLE
        ///Ifl^
 NEW BRITAIN)! * °-
            Ih
        MIDDLETOWN
                                                       NORTH S TON ING TON

                                                                      0.05

                                                                0.01
3REENWICHS
     \»STAMFORD
     V\'   o.'oz
      0.02
     Figure 8-4B.  862 concentrations predicted by TRC model for
     0600-0800, October 30,  1968, from original,  independently
     specified input variables.

  Figure 8-4 B. shows the pattern  of SO2  concentrations predicted  by the
  TRC model from the original choices of input parameters.  (These choices are
  listed  in Table 8-1.) The  values of predicted S02 concentrations range from
  0.003 ppm at North  Stonington  in  the southeast  corner of the  region to
  0.11  ppm  at New Haven  in the south-central part of the region. The  pattern
  of concentrations  shows the major axis of high concentrations extends from
  Hartford to  the coast near Deep  River; New Haven is something of a local
  anomaly.

            Table 8-1.   CHOICES OF  INITIAL  INPUT PARAMETERS
               FOR TRC REGIONAL MODEL USED TO  ESTABLISH
           REFERENCE CONCENTRATIONS FOR SENSITIVITY TESTS
     Input parameter
     Source strength

     Wind field
     Decay coefficient
     Diffusion coefficients
                              Value or method of choice
Connecticut source inventory for SO2, corrected for
diurnal variations and degree days3
320° at 6 m sec'1
0.231, Half-life of 3 hr
Appropriate to a partly-cloudy night with moderate
wind speeds3
    a See Reference 3.
                                                                     8-11

-------
Figure 8-5  shows the frequency  distribution  of  predicted  values of S02
concentration  occasioned  by this  original choice of input parameters. This
distribution  is quite typical, being strongly skewed toward  large  values of
XT. Note however that  there are several'values.of XT  < 0.005 ppm.  Since
XT enters in the denominator of our measures of the mean and variance of
the fractional errors  in  X, these low values can lead to spuriously high
values of these measures. This situation led to no particular difficulty  in the
sensitivity tests, but required  special  attention in the verification  tests dis-
cussed  in  later sections. Examples of  the distribution of (XT-XE)/XT are
shown  in  Figure 8-6,  where they are plotted for  increasing values of the
random error  in  the source strength. As  can be seen  from Figure 8-6, the
distributions of   (XT-XE) / XT  are  reasonably symmetrical near  zero and
are well   behaved.  (The frequency  distributions  of   (XT-XE)/XT  are
tabulated for each sensitivity test in Appendix A.)
            100
             80
             60
  NUWBER/0.005
CLASS INTERVAL
     °F XT   40
             20
                _  /

                             .,-j-"*
CUMULATIVE
                       0.02
0.04      0.06
     XT, ppm
        0.08
OJO    0.12
  Figure 8-5.  Frequency distribution of S02 concentrations pre-
  dicted by TRC model for 0600-0800, October 30, 1968, from
  original, independently specified input variables.
8-12

-------
                                                                                         1/2
               -1.0
                                                                                             0.404
Figure 8-6.  Frequency distributions of fractional error in concentration predictions, (XT - XE)/XT, as
random error of source strengths, (QT - QE)/QT. is increased.

-------
Sensitivity to  Random Errors in  Q.
Since the model prescribes that the concentration X  is directly proportional
to the  source  strength  Q,  there must  be a one-to-one  correspondence
between systematic errors in Q and errors in X   In terms of our measures of
sensitivity,
            [(XT - XE)/XT] =  [(QT - QE)/OT],
and
            (V)2  =  0.

This result, which derives directly from the model, is shown in Figure 8-7.
      XT
    Oh
           i.O
           0.8
           0.6
            0.4
           0.2
                                                         /

                                  */



                                       RANDOM
                                       ERROR
                                           Q
 r   NQ _--*--
- — -
                                1
1.0


0.8


0.6

72" 1/2

0.4


0.2
                                                            •0.2
     •0.2      0       0.2       0.4      0.6       0.8      1.0
                      	rl/2
                       / QT-QE   \
Figure 8-7.  Mean and standard deviation of errors  in concentra-
tion predictions as functions of a random and systematic errors in
source-strength estimates.
8-14

-------
Random errors in  Q produce a different result, namely
     [(XT  -  XE) / XT]  = 0,

  but  (v')2

is non-zero, i.e., while no systematic error in the prediction of concentration
is introduced by  random errors in  Q, the  distribution of the fractional errors
in N  is broadened as the random  errors in Q increase. The calculated values
for this relationship are tabulated  in  Table 8-2 and  plotted in  Figure 8-7.
(The  distributions of  (XT  - XE) / XT  are shown in Figure 8-6.)

  Table 8-2. CALCULATED VALUES OF MEAN AND STANDARD DEVIATION
  OF (XT - XE) /  XT AS FUNCTION OF THE RMSa ERROR IN  Q  (RANDOM)
        (LIMITS ASSESSED AT THE 95 PERCENT CONFIDENCE LEVEL)

QT-QE jl"
aT
0.082
0.163
0.409
0.823

XT-XE
XT
0.001 ± 0.02
-0.004 ± 0.02
-0.016 ± 0.03
-0.033 ± 0.05

(V)2

0.101 < 0.118 < 0.134
0.101 < 0.118 < 0.134
0.146 < 0.170 < 0.194
0.244 < 0.283 < 0.323
   aRoot mean square.
The most notable feature of this particular analysis is  the  relative  insensi-
tivity of the concentration  predictions to large random errors in the specifi-
cation of the source strengths, at least as compared with  systematic errors in
this term.  Evidently, among  multiple sources  these  random errors tend to
cancel each  other effectively. Assuming a Gaussian distribution of

     (XT  - XE) /XT,

which appears reasonable in this case from Figure 8-6, the range of this frac-
tional error  in X increases  only by a factor of 2 while the root  mean square
value of the error in Q increases by a factor of 8. (Extreme errors are listed in
Appendix A.)
This  is  a result  of no small  consequence  when we specify  the accuracy
required for  source  inventories.  If  this  result  is  substantiated  by  more
exhaustive  sensitivity  analyses  of  other  source distributions,  a  very  real
trade-off between required  accuracies in source  inventories,  which are diffi-
cult and expensive at best, and  useful accuracies of concentration predictions
can be proposed. Clearly, accuracy of concentration  prediction cannot be  the
                                                                     8-15

-------
sole criterion for  source  inventory accuracy. The latter must also be used to
determine  abatement  strategies, and  the need for accuracy on this account
may  be overriding. But, with this  provison  it is  clear that concentration
predictions are much  more  vulnerable  to systematic  errors in  the source
specification than they are to random errors in this input variable.

Sensitivity  to Systematic  Errors in
Decay Coefficient,  X.
Although the  physical reality of a  simple exponential decay of S02 with
travel  time has not been rigorously  demonstrated, our experiences with the
TRC  model  show that a  distinct  improvement in S02 concentration predic-
tions   is achieved  with  this corrective  term.  It is,  therefore,  at  least of
practical  interest  to  determine the  sensitivity of  X  to variations in this
empirical input parameter. Our reference here is a half-life of 3 hours (X =
0.231  hf1).
The effect of varying  systematically the decay coefficient, X , on the mean
and standard deviation of  (XT — XE)/XT  is tabulated  in Table 8-3 and
shown graphically in Figure 8-8.

  Table 8-3.   CALCULATED VALUES OF MEAN AND STANDARD DEVIATION
   OF (XT - XE) / XT  AS  FUNCTION OF \, DECAY COEFFICIENT OF SO,
           (LIMITS ASSESSED  AT 95 PERCENT CONFIDENCE LEVEL)
X. ,
hf
0.693 (HLa=1)
0.462 (HL=1.5)
0.231 (HL=3)
0.085 (HL=8)
0 (HL=°^
/XT - XE\
V XT )
0.486 ± 0.04
0.326 ± 0.03
0
-0.439 + 0.08
-0.918 ± 0.17

(V)2
0.187 < 0.217 < 0.247
0.151 < 0.176 < 0.201
0
0.328 < 0.381 < 0.434
0.734 < 0.853 < 0.971
 Half-life.

As   should  be expected,  a  systematic  change  or error in X  produces a
systematic change  in both  the mean  and variance of the  fractional error in
X .  It is also  clear, however, that these errors are more  sensitive  to an
underestimate of X ,  i.e., an overestimate of the half-life,  than to an under-
estimate of the half-life. When referenced on a 3-hour half-life, which is an
arbitrary choice, more  rapid decay leads to only a moderate  rate of increase
in  the mean and range of  concentration  prediction  errors.  Neglect of  the
decay term (X  = 0)  produces the maximum  mean and range of the fractional
error  in X, both of which tend to approach ±1.0.
8-16

-------
  0.6
  0.4
               \
                    o.


 0.2




  0
                             \

                               \

                                \
                                 O
  XT  /                            \       A
-0.2 |-                             \


                                   V
                                      A
  -0.4
  -0.6
  •0.8
  •1.0
                                      A
                                       \
                                                 1.0
                                                     0.8
                                                 0.6
                                                 0.4
                                                    0.7
                                                    0
                  r          ;\

                  ^               \
                         \    /        0
                    I       \ /         !
•0.6       -0.4      ~$2T        0.2        0.4
                (A.-j-_AE), hours'1
 I      I	|	I      I     I	|
 01          1.5          3      8   o=
                  HALF-LIFE, hours
Figure 8-8.  Mean and standard deviation of errors in concentra-

ration predictions as functions of systematic errors in estimate

of the half-life of pollutant.
                                                      8-17

-------
Sensitivity to Systematic Errors in
Lateral  Diffusion Coefficient, ay
As used  in the TRC model,  the lateral diffusion coefficient, cry , specifies the
relative contributions of the sources arranged across the prior trajectory of
the volume of air that is resident over the  point at a time  for which an air
quality assessment  is desired. In principle, this is  no different from the more
conventional practice of carrying a pollutant forward  from each source point
or area  and  speeding  it laterally  according to  the Gaussian distribution
specified by ay.

More  importantly, of  course,  the  lateral  spread of a  pollutant  emanating
from multiple  sources  quickly mixes  pollutant with  pollutant and tends to
produce  a more or  less  uniform cross-wind concentration. It is not  surprising,
therefore,  that we  have  found   the  model's  predictions  of ground level
concentration to be quite insensitive to  errors in the lateral diffusion coeffi-
cient.  These errors are tabulated  in  Table  8-4  and shown graphically in
Figure 8-9. As can  be seen, the mean  error in  (XT  —  XE)/XT is not
significantly different from  zero and the spread of these errors, as measured
by  (v'.
8-7).
comparable to those produced by a random error in Q (Figure
   Table 8^.  CALCULATED VALUES OF MEAN AND STANDARD DEVIATION
        OF (XT - XE) / XT AS FUNCTION OF SYSTEMATIC ERRORS IN
                    LATERAL DIFFUSION COEFFICIENT CTy
          (LIMITS ASSESSED AT 95 PERCENT CONFIDENCE LEVEL)
0 T 0 E
ffyT
-0.2
+3.2
-0.5
+0.5
-0.7
+0.7

/XT
-XE\
V x /
-0.003
0.007
-0.018
0.012
-0.032
0.014
± 0.02
+ 0.03
± 0.03
± 0.03
± 0.04
+ 0.03



0.105 < 0.122 < 0.139
0.130 < 0.151 < 0.172
0.138 < 0.161 < 0.183
0.138 < 0.161 < 0.183
0.170 < 0.197 < 0.224
0.141 < 0.164 < 0.187
Sensitivity to Systematic Errors  in
Vertical  Diffusion Coefficient, az

Vertical diffusion has been  recognized as a  critical component of air quality
prediction for some time, principally through  the employment of measures
of the depth  of the  mixed  layer6 The  inherent sensitivity of concentration
predictions to  errors in  az derives directly from the basic  diffusion equation
(Figure 8-2), since, for ground level concentrations, (z=0),
8-18

-------
XT-XE
•~
, XT ,
1.2
0
1.2
0.4
0.6
)
' V
-!
=.0 	 © 	 ©— 	 © 	 ® ~®~ ® -
-*— *--^ ^-~ *— «--
i i i N/ i i i

,2 1/2
.6
0.4
0.2
0.4
n
                                                          0.6     0.8
                                 OyT-OyE
  Figure 8-9.  Mean and standard deviation of errors in concentra-
  tion predictions as  functions of systematic errors in estimate of
  lateral diffusion  coefficients.
                                                                      (7)
is the fractional  change in X as a function of the fractional change in crz
Equation  (7) indicates that the fractional error in X will  be large when CTZ is
small  relative  to  the  source  height,  f , a  situation that  may extend  to
considerable distances when vertical mixing is slow (stable conditions).

The tests of sensitivity  of (XT—XE) / XT  to fractional changes in crz  are
tabulated  in  Table 8-5 and  shown graphically in Figure 8-10.  The extreme
sensitivity of X to az  when az is small (0zT   crzE -»
from this figure. For substantial overestimates of az
sensitivity behaves very similar  to errors in ay.*
                                                    uzT
                                                         is clearly evident
                                                         -  azE < 0), this
The avoidance of the sensitivity to small values of ay is, in part, due to the fact that area
sources with a minimum dimension of 5,000 feet predominate in the model.
                                                                     8-19

-------
    Table 8-5.  CALCULATED VALUES OF MEAN AND STANDARD DEVIATION
          OF (XT - XE) / XT AS FUNCTION OF SYSTEMATIC ERRORS
                 IN VERTICAL DIFFUSION COEFFICIENT, az
PZT - OZE
ffzT
-0.2
+0.2
-0.5
+0.5
-0.7
+0.7
/XT - XE\
( XT )
0.055 ± 0.03
-0.071 ± 0.03
0.120 ± 0.03
-0.236 ± 0.05
0.151 ± 0.03
-0.429 ± 0.15
. ,,2 1/2
(v r
0.112 < 0.130 < 0.148
0.143 < 0.164 < 0.187
0.133 < 0.155 < 0.177
0.206 < 0.240 < 0.276
0.145 < 0.168 < 0.191
0.669 < 0.776 < 0.885
U.H-
0.2

t v
AT ™ "Q \ 0
T /
-0.2
•0.4


-n.fi
i i i i i i i
~—r> ^ ~
~-©^ 	 /
J"~*~©--^ /
"^©-» / ~
^^^^ i
"^. /
^ / _
I ^ A k
--A 	 'A 	 ! 	 A— — "*" 0~

^X ^ ^
1 1 1 \/ 1 1 1

0.8


0.6
VTll/2
0.4
0.2


0
           •0.8   -0.6   -0.4   -0.2
0.2    0.4   0.6    0.8
 Figure 8-10.  Mean and standard deviation of errors in concentra-
 tion predictions as functions of systematic errors in estimate of
 vertical diffusion coefficients.
Sensitivity to Positional  Errors Introduced by
Systematic  Errors in Wind Direction,©.

As noted  previously, positonal errors, i.e., mislocation of the sources, recep-
tors, and  pollutant  trajectories,  automatically invoke errors  in  all of the
input parameters. For  the present study, a systematic error in the direction
of the general  wind field was  utilized  for the assessment of sensitivity to
these compounded  errors. Variations  in windspeed were  suppressed since,
8-20

-------
between the "stretching" effect at the source and the time-of-flight effect in
the decay of the pollutant,  errors in  windspeed estimates tend to be com-
pensatory.  Rather,  the wind  direction field  over the  region  was simply
rotated  the desired  number of degrees  from  the originally chosen  wind
direction, 320?

As might  be expected,  the errors produced by this  input-error generation
were rather  spectacular. They are tabulated in Table 8-6 and  shown graphi-
cally  in  Figure  8-11.  (One completely  erroneous  station contributed  so
strongly to the variance of  (XT - XE) / XT for  0T-0E = + 20°  that this
point has been deleted.)

   Table 8-6.  CALCULATED VALUES OF  MEAN AND STANDARD DEVIATION
          OF (XT ~'XE) / XT AS FUNCTION OF SYSTEMATIC ERRORS
                         IN WIND DIRECTION, 0
          (LIMITS ASSESSED AT 95 PERCENT CONFIDENCE LEVEL)
0T-0E,
degrees
+10
-10
+20
-20
+40
-40

/XT - XE\
( XT ;
-0.134 ± 0.04
-0.124 ± 0.04
- D.436 ± ?
-0.232 + 0.26
-0.152 ± 0.22
-0.524 ± 0.33

/ ,12 "2
(v'r
0.191 < 0.222 < 0.253
0.193 < 0.218 < 0.248
-
1.14 < 1.32 < 1.51
0.930 < 1.08 < .123
1.45 < 1.68 < 1.92
As can be seen from Figure 8-6, the mean fractional concentration errors are
not spectacular, but the spread of these errors, as measured by (v')2 1/2,  is.
To  relate  this analysis to  a practical  situation, it is  as  though  one had a
single wind measurement  at the center  of  the  region and extrapolated this
measurement, with  its error, to the entire  region. Much  more  sophisticated
analyses are  imaginable and should  be  pursued. But this analysis  points
clearly to the need for accurate representations of the regional wind field.


Summary of Sensitivity Tests
In order to compare  and  summarize these  sensitivity  test results, all  of the
systematic  input errors have been standardized  as a fraction of the original
input  value and the mean  and  standard  deviation of   (XT — XE) /XT have
been plotted for each of these  in  Figure 8-12.  (Errors in A were  normalized
by 0.231,  the value appropriate to a 3-hour  half-life, and errors  in 0 were
normalized  by dividing by  180°, the maximum error that could  be encoun-
tered in this input parameter.)
                                                                    8-21

-------


( X X \
T " E\
V T /



0.4
0.2

0
-0.2


-0.4

i i i i i i i
_\
x
X
\
\
\
\
\ )
\ /-
- lr
\ /
\ /
\ ^~^-^ *
\ s' *** /
I©' x© / /
" \ N / /-
s@d_\ \ / /
/ -+-* \ A /
V ^ \ '
/ \ /
0" \ /
.ncT i i i V i i i
i.a
1.6

14
.4
1.2


1.0
W !
0.8

0.6
0.4


0.2

n
            -40   -30    -20    -10
10    20     30
Figure 8-11.  Mean  and standard deviation of errors in concentra-
tration predictions as functions of systematic errors in estimate of
regional  wind direction.

 In  studying Figure 8-12, it is well to keep  in mind that the reference values
 for the input parameters are arbitrarily chosen. These curves should not be
 constructed as a universal or general representation of error sensitivity.  With
 this caveat, however,  it is clear that the sensitivities of concentration predic-
 tion to systematic input errors in  the expected range of ± 0.5 may be ranked
 as follows:
    1. Wind field directional errors.
      Mean error < ±  50 percent; individual  errors > 200 percent.
   2. Source strength  errors.
      Mean errors < ± 50 percent; individual errors < ± 50 percent.
   3. Decay rate errors.
      Mean error < 40 percent; individual errors < ± 50  percent.
   4. Vertical diffusion errors.
      Mean error < 25 percent; individual errors < ± 50  percent.
 8-22

-------
l.U
0.5/
/XT-XE\ „
I 	 v 	 1
\ XT /

-0.5

-1.0


-1.5

-2.0.
/
i— — __ .^_X /
0^ 	 0— QtT".>. / Ou
tfy""1^^ /jj&JT-^^
/lT X^^N'- V'2l/2
/' © ^ 1.5-
/ Vx
.Q' \ x i.o-
/ °z
f
/ 0.5-
/ i
' \ i i I I i n
e
$
\
T
.


ff
A y ,



e QZ
© ^
/ M
/ • /

j) 	 -ELc,- 0,
\ \ * i """^v^2*^ y v "|
-2.0 -1.5 -1.0 -0.5 0 0.5 1.0 -2.0 -1.5 -1.0 -0.5 0 0.5 1.0
XT-XE XT-XE
XT XT
Figure 8-12. Comparative summary of mean and standard deviation of errors in concentration prediction as
function of fractional systematic errors in each input variable. Sensitivity of model to errors is measured
oo by slope of curves.
CO

-------
   5. Horizontal diffusion errors.
      Mean error < 5 percent; individual errors < 40 percent.

The greatest sensitivites  are  found when errors  become so large that input
parameters  approach  their singular or limiting values (0 for az, and °° for A)
but, assuming that prudence  in the use of models will  obviate this difficulty,
by  far  the  most  serious expected  error is  found when  the wind direction
field  is misestimated. Errors  in the mean of  the concentration prediction,
averaged  over  the  region, approach  ±  25 to  50 percent, and  errors  in
individual  concentration  predictions readily exceed 200 percent of the true
value when wind direction errors reach and exceed 20°.

Surprisingly, air  quality  predictions showed no particular sensitivity to ran-
dom source strength  errors. As noted previously, however, this is probably a
function of the  density of the multiple sources and deserves more rigorous
study from  that point of view.

Finally, this whole study must be recognized as the case study that it is.
Because of the  obvious  importance of a clear and  quantitative measure of
the  accuracies  and  uncertainties  of simulation  predictions,  this  type of
analysis should  be extended   and  generalized  far  beyond this first effort.
Insofar  as these  results  pertain only  to the  TRC Regional  Model,  other
models  should be similarly exercised and  examined for sensitivities  to input
errors and uncertainties.

A CASE STUDY OF SENSITIVITIES OF MODEL
VERIFICATION TO INPUT PARAMETER VALUES
Since direct observations of S02 concentrations are available for 21 stations
within Connecticut for the period  used in the above sensitivity analyses, an
unusually  detailed analysis of the sensitivities of the model verification to
changes in the input parameters arises as a welcome bonus in this study. This
is, of course, a  single  case and must be treated as such. The results are well
worth  noting, however, (For  a more general  analysis of the verification of
the TRC Regional Model, see  the 1969 TRC Staff Report.4)

For this analysis,  we  simply need redefine XT as the observed value of S02
concentration  at any  given station  and XE as the predicted value for that
station for any  given set of input  parameters, including the original set used
as the base reference for the  sensitivity  analyses. The  mean  and  standard
deviation  of (XT-XE) / XT   constitute the test  for goodness of fit or the
verification of the model's predictive capabilities.
The values of XT observed at the  21 verification  stations and  the fractional
error  associated  with  each observation when  compared  with the  predicted
values  using the original input parameters (Table 8-1) are shown  in Figure
8-24

-------
8-13. As can be seen from that figure, the fractional errors in the verification
of  observed concentrations  tend  to  cluster around zero, with the notable
exceptions of  two stations for which  the  observed concentrations are very
low,  0.002  and 0.003  ppm (0.001   ppm  is the  threshhold  level  for  the
measurement system used to monitor S02). If these stations are eliminated,
the  mean value of (XT-XE)/XT  is,  in  fact,  zero,  and there is a great
temptation to do exactly that. The  temptation  has been resisted, however,
primarily  because there are several other observations of S02 in the < 0.005
ppm  region  that did not  produce  anomalous error values; one would be hard
pressed to justify their  retention if  the two bad actors were eliminated on
the grounds of unreliability of observation.  Hence we retained all 21  sta-
tions.
          0.06
   XT,ppm
          0.04
          0.02
• ,
t
                                         -1

                                       XT-XE
                                         XT
Figure 8-13.  Joint values of observed values of 802 concentra-
tions and fractional error in values predicted by  TRC model, using
original, independently specified values of input variables,  for
period 0600-0800, October 30, 1968.
Using  the  original  input  variables  described in Table 8-1, the verification
parameters for the observed versus predicted values of SO2 were
and
                  = - 0.642,
                  =  1.225
                                                                  8-25

-------
hardly  an  inspiring result. The question then asked was, "Could this verifi-
cation be improved by  varying input  parameters for the model?" Normally,
of course,  input parameters must be chosen  independently of the verification
data, but the question is valid for analytical  purposes.

Knowing the general behavior  of the  model and the results of the sensitivity
analysis, we can immediately speculate that  improved  verification of the
mean value of  (XT - XE) / XT can be obtained in any or all of three ways:
(1) reduce the source strengths uniformly by 40 percent (since in the mean,
concentration predictions were overestimated by this amount), (2)  increase
the vertical diffusion rates by an amount that compensates  the overestimate,
and  (3)  decrease the half-life of S02, so as to compensate for the  overesti-
mate of X. In addition, we can test to see if systematic changes in either the
wind direction or the horizontal diffusion rate  reduce the verification errors.

We are  also  vitally concerned  with the individual station verification  (since it
appears  air  quality  assessment and control  will be based on  local values of
concentration,  not the regional  average). It  is imperative, therefore,  that the
range of verification errors.be  reduced  also, a feature that is measured by the
standard deviation  of      (XT —  XE) / XT  , (v')    .  It would be desirable,
of course,  to test all of  the  combinations  and permutations of input para-
meters;   and  this is possible,  given the  time and  resources to perform the
necessary iterative solutions of the model. The present analysis does not go
that far; rather, we have examined only one  input variable at a time.

The effects of random variations in the source strengths on the verification
parameters are shown in  Figure 8-14.  (Values and confidence limits of (Xr
XE) / XT  and   (v')T1/2 are tabulated  for all input variables in Appendix B.)
Clearly,  no  improvement  in the verification  is achieved by this manipulation;
both  the mean  and standard deviation are at their minima  with the original
source inventory.

The  effects  of  systematic  variation   of  the source strengths are shown in
Figure 8-15  and, as expected,  a uniform reduction by 40 percent of all the
source  strengths  reduces  the  mean  error in X  to  very near  zero,  and the
standard deviation of the fractional error in X is reduced  to 0.75  from its
original  value of  1.225. This is a distinct improvement; in fact, it is  the best
result that could be obtained from single-input-variable manipulation. Exami-
nation of Figure 8-16, however, shows that an equivalent improvement could
be  obtained by  reducing the estimated  half-life of S02  from 3-hours to
1-hour.   The mean  error  for  this  case  is also  near  zero  and the  standard
deviation is reduced to 0.860. Figure 8-17 shows that even a very substantial
increase  in the  vertical diffusion coefficients produces only a  slight  improve-
ment in the mean error  of concentration prediction, and no improvement at
all  in the  standard  deviation  of the  fractional  errors in X,  as should have
been  expected from the sensitivity analysis.
8-26

-------
   ,*T_XE
   \   *T
               -0.5
               -1.0
               -1.5
               -2.0
                                                                     2.0
                               1.5
                               0.5
                           0.2
0.4
0.6
0.8
1.0
                                       QT
                                             2" 1/2
Figure 8-14.  0600-0800, October 30, 1968, verification scores for
TRC model as a function of random errors  in source strength speci-
fications.
The effects  of variations  of  the horizontal diffusion coefficient and of the
wind direction  are shown  in  Figures 8-18 and 8-19, respectively. Clearly, the
TRC model  could  not  be  more  indifferent  to the choice of  horizontal
diffusion  coefficients  for  this  case. The  model's sensitivity to the  com-
pounded errors introduced by  changing the wind  direction  is  reiterated  in
Figure 8-19, but  the  original choice of wind direction provides the best fit
for this verification test.

To  summarize this case  study, Figure 8-20 shows  the maximum improve-
ments in  the verification  of the model that were achieved  by varying,  in
turn, the vertical  diffusion,  the decay rate, and  the source strengths.  Of
these, only the latter  two provide a significant improvement  in the verifica-
tion scores.  (See Appendix B for confidence limits on  (XT - XE) /XT and
(v7)7 1/2  for these tests.)
No  sweeping conclusions  can be drawn from this single  case study,  but it
seems unlikely  that the source  inventory for Connecticut is in error by a 40
percent overestimate.  The other verification  tests showed that mean concen-
trations were underestimated as frequently as they were overestimated using
the  same source of inventory.  Rather, this partial result seems to emphasize
the  importance of better understanding of the decay and loss of S02.
                                                                   8-27

-------
00
to
oo
l.U
0.5


0

/XT-XE\ -0.5
V XT j

•1.0

-1.5

-2 0
iii.
/ -


n/ '
v ©/
/ \ p*"
/ \
\
v -
\
\
, \
771-1/2
2-5 0.1


2.0 °-5
/XT-XE\
1.5 I XT ') o
-0.5
1.0
-1.0
0.5
-1.5
n
-1.0 -0.5 0 0.5 1.0
2 0

1 i i i


7TV2
•
•-*--« / •
^©x /
>N. /

^
*~" \
"'""" U \

\
^
\
1 1 1 1
2.5
2.0
1.5

1.0

0.5


n
QT-QE '-0.6 '-0.4 -0.2 o 0.2 0.4
                       QT

Figure 8-15.  0600—0800, October 30,1968, veri-
fication scores  for TRC  model as a function of
systematic changes in source strength specifica-
tions.
                                                                                 (XT-XE)
                                                             Figure 8-16.  0600-0800, October 30,1968, veri-
                                                             fication  scores for  TRC model as a function of
                                                             systematic changes  in decay rate  for SO2-

-------
00
ro
CD
                -0.5
                -1.0
                •1.5
                •2.0
                 -1.0
                         -0.5
                                 °zT
                                          0.5
                                                    2.0
                                                    1.5
                                                    1.0
                                                    0.5
                                          0°
Figure 8-17. 0600-0800, October 30,1968, veri-
fication  scores for TRC model as a function of
systematic  changes  in vertical  diffusion coef-
ficients.
                                                                     •0.5
                                                                /XT-
                                                                V XT
                                                                     -1.0
                                                                     -1.5
                                                                       -0.1
                                                                       	!0__©_.n—0_J0—
                                                                               -0.5
                                                                                                0.5
                                          2.0



                                          1.5

                                           V'2

                                          1.0



                                          0.5
                                                                                                        1.0
                       ayT
Figure 8-18. 0600-0800, October 30,1968, veri-
fication  scores for TRC  model as a function of
systematic  changes  in  lateral diffusion coeffi-
cients.

-------
          0.5
         -1.5
         -2.O
'-0.2


                       \r
                                                    /
                         y©
                     /
                     -1.0
      0
0.1
0.2
                                                            2.0
                                                            1.5
                                 1.0
                                 0.5
                                  8 max
Figure 8-19.  0600-0800,  October 30, 1968, verification scores for
TRC model  as a function of systematic changes in  regional wind
direction.
 The  empirical exponential  decay and  loss term  used in the TRC model
 operates to  reduce  concentrations most at those stations that are relatively
 far removed from  major source areas.  The "effective" distance  from each
 station to the contributory sources for that station, Xe,  may be calculated
 from the model's estimates of X for two different values of X ,
      Xe    =
In
                                         (7)
 a parameter that is expected to change with wind direction. For the present
 case  (NW winds), Xe ranges from 2 to 41 miles in Connecticut (Figure 8-21)
 and clearly portrays both the rural and urban areas and the presence of local
 sources  in  otherwise  nonsource areas.  For  example,  Danbury, Stamford,
 Bridgeport, and  New  Haven are subject  primarily to  local sources of  S02
 with  NW  winds,  while  the  Hartford-Middletown-Deep  River axis  of SOj
 8-30

-------
0.2
fxT-XE\ 0
M
-0.2

•0.4

-0.6
-0.8
1 I I i
X
- Y -
\
rX"° -
/ ^
V
/
/
i
J>
/
i i i i
ORIGINAL ^^ V ^
INPUT % '< \
PARAMETERS %> ^y ^
                                                    Hl.2
                                                    -h.o
                                                    H 0.8
                                                    H U.6
                                                    -U.4
                                                    H0.2
Figure 8-20. Summary of maximum Improvements In verification of
TRC model predictions of SQg concentrations fgr period 0600-0800,
October 30, 1968, by independent  manipulation of vertical diffusion,
decay rate, and  spuree strength  field.
                                                           8-31

-------
  concentrations  (Figure 8-5) represents the regional concentration of  sources
  in  the  Hartford-Middletown area. Similarly, the Norwich-Groton area consti-
  tutes a regional  source of S02 .  On the other hand, the New Milford-Morris
  Dam and  the Colchester regions are  obviously far removed  from the major
  sources of SO2  that affect them.

                          FIXED SAMPLING STATIONS
                                 [___/T~HOMPSON VI LLE
          XFALLS VILLAGE
                    WINSTED
                             SIMSBURY,  /
                                    WINDSOR
                                   S*	^N   BO L
                                  ff~H A RTFORD '
                                     »''V
                         / NEW BRI TAIN
                                  MIDDLETOWN I COLCHESTER NORWICH
                    WATERBURY  MERIDEN
                                                     NORTH STONINGTON
                                                             GROTON
                          INEW HAVEN S  MADISON
        NO RWA L K
 1 O
 GR EENWI CH1 ,

   5-V^STAMFORD
Figure 8-21.  Effective distance to S02 sources for case of NW winds.
  These results are, of course, more  suggestive than conclusive at the present
  level of  analysis. If nothing  else,  however, they  do  point clearly  to the
  powerful analytical  capabilities of a reasonably  faithful simulation  model. As
  noted by one of my colleagues, these are indeed powerful, if  expensive tools.
  8-32

-------
REFERENCES

1.   Hilst, G.  R.  An  Air Pollution  Model of  Connecticut. In:  Proceedings the IBM
    Scientific  Computing Symposium on Water and  Air Resource Management. York-
    town Heights, N. Y. October 1967. IBM.  Data  Processing Division. White Plains.
    1967. 251-274.
2.   Bowne, N. E. Simulation  Model for Air  Pollution Over Connecticut. J. Air Pollu-
    tion Control Assoc.  19:570-574, August 1969.
3.   Hilst, G.  R.  J. E. Yocom, and  N.  E.  Bowne. The Development of a Simulation
    Model  for Air Pollution  over Connecticut, Vol.  I.  Travelers  Research Center, Inc.
    Hartford,  Conn.  Final  Report  to  the Connecticut Research Commission.  TRC
    Report No. 7233-2798.  1967. 66p.
4.   Hilst, G.  R.  Verification by  Observation of the Application of  An Air Pollution
    Model  to  the State of  Connecticut  (in  press). Travelers Research Center,  Inc.
    Hartford,  Conn.  Final  Report  to  the Connecticut Research Commission.  TRC
    Report No. 7242-365. September 1969.
5.   Bowne, N. E. A Mathematical Model of Air Pollution in the State of Connecticut.
    Presented  at  62d Annual  Meeting  of the Air  Pollution Control  Association. New
    York. June 22-26, 1969.
6.   Holzworth, G. C. Estimates  of Mean Maximum  Mixing Depths in the Contiguous
    United States. Mon. Weather Rev.
                                                                           8-33

-------
                            APPENDIX A
   Frequency distribution of errors in predicted concentration,  (XT~XE)/XT,
                 for designated errors or values of variables.
                       (100-station sensitivity tests.)
Range of
(XT XE)/XT
0.860to 0.959
0.760 to 0.859
0.660 to 0.759
0.560 to 0.659
0.460 to 0.559
0.360 to 0.459
0.260 to 0.359
0.160to 0.259
0.060 to 0.159
0.040 to 0.059
-0.1 40 to -0.041
-0.240 to -0.141
-0.340 to -0.241
-0.440 to -0.341
-0.540 to -0.441
-0.640 to -0.541
-0.740 to -0.641
-0.840 to -0.741
-0.940 to -0.841
<-0.940
Avg of<-0.940
Largest (-) value
Avg of
(X-r-XE)/XT-
Std. dev.of
(XT-XE)/XT
Rms error in (Q-j- - Qgl/Q-p
0.082
1
0
0
0
0
0
0
0
8
81
8
1
0
0
0
0
1
0
0
0
0
-
0.001

0.118
0.163
1
0
0
0
0
0
0
2
9
68
13
6
0
0
1
0
0
0
0
0
0
-
-0.004

0.118
0.409
1
0
0
0
0
1
1
6
10
53
9
10
5
4
0
0
0
0
0
0
0
-
-0.016

0.170
0.823
1
1
0
1
0
3
4
5
13
42
4
10
4
4
1
2
2
3
0
0
0
-
-0.033

0.283
Value of half-life, A. (hf1 )
0.693
1
11
17
11
15
14
11
10
10
0
0
0
0
0
0
0
0
0
0
0
0
-
0.486

0.217
0.462
1
0
2
8
13
18
20
20
13
5
0
0
0
0
0
0
0
0
0
0
0
-
0.326

0.176
0.085
1
0
0
0
0
0
0
0
0
4
16
15
13
8
9
10
6
4
1
13
-1.181
1.667
-0.439

0.381
0
1
0
0
0
0
0
0
0
0
4
12
4
6
11
6
7
3
5
2
39
-1.770
3.333
-0.918

0.853
8-34

-------
APPENDIX A (continued)
Range of
(XT-XE)/XT
0.860 to 0.959
0.760 to 0.859
0.660 to 0.759
0.560 to 0.659
0.460 to 0.559
0.360 to 0.459
0.260 to 0.359
0.1 60 to 0.259
0.060 to 0.159
0.040to 0.059
-0.140to -0.041
-0.240 to -0.141
-0.340 to -0.241
-0.440 to -0.341
-0.540 to -0.441
-0.640 to -0.541
-0.740 to -0.641
-0.840 to -0.741
-0.940 to -0.841
<-0.940
Avg of<-0.940
Largest (-) value
Avg of (XT-XE)/XT
Std. dev. of
(XT-XE)/XT
Fractional error in CTZ, (<7zT — CT2E)/ "zT
0.2
1
0
0
0
0
0
0
2
7
38
21
26
3
1
0
0
0
1
0
0
0
-
-0.071

0.164
+0.2
1
0
0
0
0
0
1
8
27
52
4
6
0
0
1
0
0
0
0
0
0
-
0.055

0.130
0.5
1
1
2
1
2
4
9
18
8
25
6
4
8
3
0
3
0
2
0
3
-1.429
1.857
-0.236

0.240
+0.5
1
0
0
0
0
0
19
22
15
38
3
1
1
0
0
0
0
0
0
0
0
-
0.120

0.155
0.7
1
1
0
2
0
1
3
3
7
24
4
9
1
4
1
4
4
2
3
26
1.339
-2.333
-0,429

0.776
+0.7
1
0
0
0
0
11
15
20
13
37
2
0
1
0
0
0
0
0
0
0
0
-
0.151

0.168
                                    8-35

-------
                APPENDIX A (continued)
Range of
(XT- Xe)/XT
0.860 to 0.959
0.760to 0.859
0.660to 0.759
0.560 to 0.659
0.460 to 0.559
0.360 to 0.459
0.260to 0.359
0.1 60 to 0.259
0.060 to 0.159
0.040 to 0.059
-0.1 40 to -0.041
-0.240 to -0.1 41
-0.340 to -0.241
-0.440 to -0.341
-0.540 to -0.441
-0.640 to -0.541
-0.740 to -0.641
-0.840 to -0.741
-0.940 to -0.841
<-0.940
Avg of<-0.940
Largest (-) value
Avg of (XT-Xe)/XT
Std. dev. of
(XT- XE)/XT
Fractional error in cry, ("yT- CTyE'/CTyf
0.2
1
0
0
0
0
1
0
2
8
82
3
1
0
1
0
0
0
0
0
1
-1.000
1.000
0.007
0.151
+0.2
1
0
0
0
0
0
0
1
6
83
3
4
1
0
0
0
1
0
0
0
0
-
0.003
0.122
0.5
1
0
0
0
1
1
0
3
8
78
3
3
1
0
0
0
0
0
0
1
-1 .000
1.000
0.012
0.161
+0.5
1
0
0
0
0
0
0
2
9
75
1
6
2
0
1
3
0
0
0
0
0
-
0.018
0.161
0.7
1
0
0
0
1
1
0
4
8
77
3
3
1
0
0
0
0
0
0
1
1.000
1.000
0.014
0.164
+0.7
1
0
0
0
0
0
0
2
12
65
6
7
3
1
0
1
1
0
0
1
-1.239
1.239
0.032
0.197
8-36

-------
APPENDIX A (continued)
Range of
(XT-XE)/XT
0.860 to 0.959
0.760 to 0.859
0.660to 0.759
0.560 to 0.659
0.460 to 0.559
0.360 to 0.459
0.260 to 0.359
0.160to 0.259
0.060 to 0.159
0.040to 0.059
-0.140 to -0.041
-0.240 to-0. 141
-0.340 to -0.241
-0.440 to -0.341
-0.540 to -0.441
-0.640 to -0.541
-0.740 to-0.641
-0.840 to -0.741
-0.940 to -0.841
<-0.940
Avg of <-0.940
Largest (-) value
Avg of (XT-XE)/XT
Std. dev. of
(XT-XE)/XT
Systematic error in wind direction, 0, (°)
+ 10
1
2
1
0
2
3
6
6
17
32
2
9
5
1
0
1
0
2
4
6
-2.386
4.000
-0.134
0.222
-10
3
1
1
4
1
1
3
6
12
34
6
9
5
1
1
0
2
2
0
8
-2.030
3.615
-0.124
0.218
T20
1
1
0
3
4
8
5
9
9
25
4
8
4
3
2
2
0
3
0
9
-5.317
28.000
-0.436
?
-20
3
4
1
1
3
8
5
7
15
20
1
6
4
2
1
2
3
1
0
10
-3.422
7.667
-0.232
1.32
+40
3
2
4
3
5
6
7
10
2
15
9
12
3
4
0
1
0
4
0
10
-2.410
8.500
-0.152
1.08
-40
6
2
3
5
6
8
5
5
9
7
1
6
5
4
2
0
1
4
2
19
-3.340
7.600
-0.524
1.68
                                    8-37

-------
                            APPENDIX B
          The mean and standard deviation of fractional error in
              verification of SO2 predictions at 21 stations as
              function of indicated changes in input variables.
              (Confidence limits assessed at 95 percent level.)

     Random Variation of Source Strengths, Q.
/QT-QEY"'
( QT )
0
0.082
0.163
0.409
0.823

^XT-XE\
v XT ;
-0.642 ± 0.54
-0.673 ± 0.055
-0.690 ± 0.57
-0.701 ± 0.59
-0.753 ± 0.63
	 ~— 1/2

0.848 < 1 .225 < 1 .60
0.865 < 1.25 < 1.64
0.903 < 1 .30 < 1 .70
0.937 < 1.35 < 1.77
1 .07 < 1 .54 < 2.02
     Systematic Variation of Source Strengths, Q.
/OT_QE\
\ QT /
-0.2
0
+0.2
+0.4
+0.6

/ \
( XT )
-0.961 ± 0.63
-0.642 ± 0.54
-0.323 ± 0.44
0.027 ± 0.32
0.331 ± 0.22
v'2
1 .00 < 1 .44 < 1 .90
0.848 < 1.225< 1.60
0.693 < 1.00 < 1.31
0.501 < 0.723 < 0.947
0.349 < 0.504 < 0.660
     Systematic Variation of the Decay Rate for SO2, X
(\T-*E),*
hr"
-0.462
-0.231
0
0.146
0.231

/XT-XE\
( XT ;
-0.013 ± 0.38
-0.230 ± 0.43
-0.642 ± 0.54
-1.18 ±0.68
-1.70 ±0.82

v'2 "2

0.595 < 0.860 < 1.13
0.648 < 0.987 < 1.29
0.848 < 1 .225 < 1 .60
1.08 < 1.55 <2.03
1,29 < 1.86 < 2.44
     *AT  =  0.231 hr"1  (3-hr half-life)
8-38

-------
                APPENDIX  B (continued)
Systematic Variation of Vertical Diffusion Coefficient, 0Z.
/azT-azE\
V CTZT /
0
-0.2
+ 0.2
-0.5
+0.5
-0.7
+0.7
/ \
V XT /
-0.642 ± 0.54
-0.596 ± 0.54
-0.733 ± 0.54
-0.483± 0.51
-0.911 ± 0.69
-0.428 ± 0.54
-1.06 ± 0.67
v'2
0.848 < 1.225< 1.60
0.858 < 1 .24 < 1 .62
0.851 < 1.23 < 1.61
0.810 < 1.17 < 1.53
1.09 < 1.57 <2.06
0.858 < 1.24 < 1.62
1 .06 < 1 .52 < 1 .99
Systematic Variation of Lateral Diffusion Coefficient, ay
/ayT -ffyE\
I °yr /
0
-0.2
+0.2
-0.5
+0.5
-0.7
+0.7
/XT-XE\
v XT ;
-0.642 ± 0.54
-0.666 ± 0.54
-0.678 ± 0.54
-0.665 ± 0.55
-0.657 ± 0.54
-0.663 ± 0.55
-0.644 ± 0.54

V2
0.848 < 1 .225 < 1 .60
0.858 < 1 .24 < 1 .62
0.858 < 1.24 < 1.62
0.865 < 1 .25 < 1 .64
0.845 < 1.22 < 1.60
0.865 < 1 .25 < 1 .64
0.851 < 1.23 < 1.61
Systematic Variation of Regional Wind Direction, &.
(©T- ©E)
(°)
0
+ 10
-10
+20
-20
+40
-40
/XT-XE\
V XT /
-0.642 ± 0.54
-0.723 ± 0.63
-0.905 ± 0.67
-0.716 ± 0.57
-1.14 ±0.98
-1.02 ±0.86
-
v'2
0.848 < 1 .225 < 1 .60
1 .00 < 1 .44 < 1 .88
1.06 < 1.53 < 2.00
0.903 < 1.30 < 1.70
1.54 <2.22 < 2.77
1.36 < 1.96 < 2.56
-
                                                              8-39

-------
            APPENDIX  C - GLOSSARY OF  SYMBOLS



a            fractional error in X due to random error

b            fractional error in X due to systematic error

Q           source strength field, emission rate

Q(£,i?£)      source strength at x  = £, y = ??, z = f

u            transporting  wind  speed  that  determines time of
             flight  from  x  to £, etc.

(v')2         variance of  fractional error distribution
 (v'l          standard deviation of fractional error distribution

X           concentration

XE          erroneous value of any input parameter

XT          "true" value of any input parameter

X            decay or loss  coefficient for pollutant under consi-
             deration

ay(x -£),
  CTZ(X  — £)   standard  deviation of  crosswind  and vertical  spread
             respectively of pollutant  from £  r?  f

&           wind direction

CTy, az        lateral and vertical diffusion coefficients
8-40

-------
9.  TIME - SPACE MODEL FOR SO2
ABSTRACT

     A multiple-source diffusion model for the simulation and prediction of
     long-term  (climatological)  ground-level sulfur dioxide concentrations in
     urban areas is described. The computer input consists of data from an
     emission source inventory together with  statistics on relevant diffusion
     parameters.

     Because of the capacity of available computers, only a limited number
     of the largest emission sources  (approximately 150) can be treated
     individually.  Smaller industrial emission sources are  treated as residen-
     tial sources.  These are represented by a  large number  of stacks (about
     150)  of  the  same  dimensions,  distributed over  areas of 1  square
     kilometer, for which the mean area emissions have been estimated.

     The  meteorological  input consists of data on wind  direction,  wind-
     speed, and Pasquill-Turner stability classes.1' ^  These parameters are
     assumed  to  be  spatially  homogeneous  throughout the metropolitan
     area.  Low-level  emissions (residential)  are correlated  with low-level
     windspeeds and Pasquill-Gifford diffusion parameters,1'3  whereas high-
     level emissions (industrial) are correlated with extrapolated windspeeds
     and Brook haven diffusion parameters?   The program  also uses corres-
     ponding statistics  for urban  boundary  layer  depths and values for
     parameters affecting absorption at the earth's surface.

     The diffusion model  used is basically Gaussian.  It is modified, however,
     such that  turbulent  diffusion  is restricted  exclusively  to the depth of
     the urban  boundary  layer. This is true for all  sources having effective
     emission heights less than the height of  the upper  limit of the bound-
     ary layer.  The rate of decay of sulfur dioxide is taken into account, as
     well as the experimentally determined absorption at the earth's surface.
     The model calculates fields of steady-state ground-level  concentrations
     that correspond to a given spatial distribution  of emission sources and
     to any possible combination of relevant  meteorological  diffusion para-
     meters. Knowledge  of frequency distributions  of these  meteorological
     diffusion parameters permits the  derivation of frequency distributions
     of ground-level concentrations  for  any location  within  or outside of
     the metropolitan area. The  computerized experiments simulate  fre-
     quency distributions of ground-level concentrations for a great number
     of regularly arranged grid  points  (up to  2500  with a mesh size of 500
     by 500 meters) and for  a  variety  of time periods (months,  heating

-------
      period,  seasons,  year,  etc.).  The frequency distributions are charac-
      terized  by a  limited number of parameters {mean,  percentiles,  etc)
      Each  parameter  is plotted as a system of  isograms  on a map of the
      metropolitan area.

      Experiments to validate the model were conducted during the heating
      period in 1967-68 at four continuously monitoring  stations that had
      been installed at special locations within the limits of the metropolitan
      area  of  Bremen. During  the sampling period, the  assumption  of  a
      sufficiently  homogeneous  wind  field was validated by wind measure-
      ments at the same locations. The calculated frequency distributions of
      half-hourly  mean values of concentrations generally  agreed fairly well
      with those derived from observed values. Comparison, however, shows
      that  the model  does not  simulate ground-level concentration fields in
      the vicinity  of  industrialized areas very well,  because uncontrollable
      low-level emissions from  industrial plants  could  not be taken into
      account in  the diffusion model.
AUTHOR
     HEINZ G. FORTAK is a professor of meteorology and Director of the Institute
     for Theoretical Meteorology at the Free University of Berlin.  His scientific
     interests include: general hydrodynamics, turbulence, dynamic meteorology, and
     oceanography.

-------
                               9.   NUMERICAL  SIMULATION
           OF TEMPORAL AND  SPATIAL DISTRIBUTIONS
            OF  URBAN  AIR  POLLUTION  CONCENTRATION


                     HEINZ G.  FORTAK
             Institut fur Theoretische Meteorologie
                  der Freien Universitat Berlin
INTRODUCTION

Increasing  industrialization in Germany during the 1950's led  to  great in-
terest in  the  problem  of  ascertaining minimum stack heights essential to
pollution  control. In  line, therefore, with its responsibilities in developing air
pollution  standards and  criteria, the Kommission Reinhaltung der  Luft[Mr
Pollution  Control Commision]  within the Association of German Engineers
[Verein  Deutscher lngen/eure(\/D\)]  established  a research group on air
pollution  meteorology  in  1958. The group was  expected  to  develop the
scientific  basis for an approximate solution to the problem. The results of
research during the first period, based mainly on the work of Sutton,5 led to
a simple  nomogram for estimating minimum  stack heights.6' 7 The nomo-
gram  is widely used for legal and  administrative purposes,  although, among
other shortcomings, very little is  known about one of the most important
input parameters  of the model and  therefore of the  nomogram. This para-
meter (in  German,  Immissions-Grundbelastung] characterizes the  temporal
and  spatial distributions (air loadings with time) of ground-level concentra-
tions of pollutants in the area in which the newly  built stack is located. In
most cases the location is in an urban area already  possessing a great number
of emission sources.  It is  extremely difficult to define such  a measure. A
simple number constant in space  and time, as proposed by the VDi,7 will
                               9-1

-------
not  suffice  for  urban  areas.  Instead, such  a parameter  is  dependent on
horizontal space coordinates and, ultimately, on time (e.g., season).

In view of the crucial  dependence  of minimum  stack heights on this para-
meter,  some means of predicting ground-level concentrations  in urban areas
must be found before the stack-height problem can be solved.

Based on  the  work  of  Frenkiel,8 the  multiple-source urban diffusion model
being described in this  paper was developed  by the author in 1962. It was
hoped that this model would be able to solve  problems of the kind mentioned
above.  After financial support  for programming and  computing time became
available, a number  of  simulation  experiments were conducted from  1963
through 1965. A  tentative report was published  early in 1966.9 During that
time  the  first  papers  on  urban air  pollution  modeling  by  Turner2 and
Clarke,10  although they  aimed  at  the solution of the real-time prediction
problem, were of great  help. Especially, the  replacement of Pasquill's stabi-
lity categories1   by Turner's2  proved to be quite useful.

The  purpose of the diffusion model is perhaps understood best by a discus-
sion  of Figure 9-1, which shows the logical structure of some features of the
urban  air  pollution problem. Assuming that  sufficient input data are avail-
able,  a  mathematical multiple-source urban air pollution  model should yield
a set of output data such as that indicated  in  Figure 9-1.

The  most  important  of  these data certainly is a real time short-term predic-
tion  of concentrations for the entire urban area.  Most authors  in the field of
urban   air  pollution  modeling  were  interested primarily in that problem.
Following  Turner,2  they  applied  the well  known  steady-state theory of
transport and dilution to  simulate and predict time series of concentrations.
Quite recently, Marsh and Withers11 demonstrated the inadequacy of such a
procedure. The model of  Davidson12  that applies a  non-steady-state theory
should  provide a means of solving the short-term prediction problem,as well
as the problem of  time  series simulation, more successfully.  If this turns out
to be  true, the very  important feedback  circuit, "warning system," can be
closed;  i.e.,  appropriate  control  measures  can   be  applied  to the source-
emission input in  order to reduce the predicted concentrations below a given
limit.

The  problem  of  minimum stack heights and, more generally, that of city
planning  is connected with  the problem   of simulation and  prediction of
long-term (climatological)  ground-level concentration  fields. Here,time series
of concentrations  are of minor interest;  instead, statistics of  observed or
calculated  concentration fields  for given long periods of time are important.

There is no doubt that  for a given location and a given period of time only
the frequency distribution of ground-level  concentrations forms the basis of
what  could be called  "air pollution climatology." Generally, these frequency
9-2

-------
                                       DISPERSION OF
                                     POLLUTANTS IN THE
                                     REAL ATMOSPHERE
OBSERVED
CONCENTRATIONS
AT A FEW LOCATIONS


STATISTICS OF
OBSERVED
CONCENTRATIONS
                                                                   VALIDATION OF THE MODEL
                                       MATHEMATICAL
                                      MULTIPLE-SOURCE
                                    URBAN AIR POLLUTION
                                          MODEL
  SIMULATION
  OF OBSERVED
CONCENTRATIONS
   STATISTICS OF
SIMULATED OBSERVED
  CONCENTRATIONS
                                                                 SIMULATION
                                                                OF POSSIBLE
                                                               CONCENTRATION
                                                                  FIELDS
                       SIMULATION AND
                       PREDICTION OF
                        LONG-TERM
                      CONCENTRATIONS
      iiw« i cr\m    -
      ENTRATIONS_|
                                                                 REAL-TIME
                                                                PREDICTION OF
                                                               CONCENTRATIONS
                          CONTROL CIRCUIT
                         "WARNING SYSTEM"
                          CONTROL CIRCUIT
                       "MINIMUM STACK HEIGHTS"
                          "PLANNING"
                                            .1
Figure 9-1.  Major elements of the  urban  air pollution problem connected  with mathematical  modelling.

-------
distributions vary from location to location within the urban area and with
time  and  season. Figure 9-2  shows  a  typical example of  the  frequency
distribution in winter of measured  half-hourly mean values of sulfur dioxide
concentrations in Bremen.
     100
          SpOBSERVED GROUND-LEVEL SO^ CONCENTRATIONS:
           gsiTENO. 1 UBERLANDWERK NORD-HANNOVER
            ^PERIOD: 15, 11, 1967-1.6, 1968
        0.0
   Figure 9-2.  Winter time frequency distribution of measured
   half-hourly mean values of sulphur dioxide concentrations
   downtown Bremen.

Although  it is  not believed  that a calculated steady-state field of concentra-
tion  will fit observations at many locations within the urban area very well,
it can  be  expected that statistical evaluation of a great number of such fields
may lead  to  reasonable results in a climatological sense. This  expectation
forms  the basis for  this paper. If the expectation is justified, the feedback
circuit, "minimum  stack  heights, city planning," can  be closed  and, in
addition,  means for the strategic planning of observation sites will  then be
available.
The  following  information  is  used  in  calculating  pollutant concentrations:
period  of  time (month, heating period, seasons,  year, etc.)  divided into equal
9-4

-------
intervals (for example, hours);  data  from  source  emission  inventories; rele-
vant meteorological parameters,  etc. From these data the model then calcula-
tes possible steady-state ground-level  concentrations for each interval of time
and for a large number of grid points within the urban area. From the stored
concentration data, the frequency  distribution of concentrations is obtained
for each grid point.  If the frequency distribution  is characterized by a set of
parameters (mean  percentiles, etc.), these  parameters  are plotted  as a system
of isograms on  a  map of the  urban area.  Such maps then may be used to
define the Immissions-Grundbelastung (ground-level concentrations), to  solve
problems of city  planning as well as problems of strategic planning obser-
vation sites.
From  the very  beginning  the main concern of the investigation was to apply
the model to a  real  situation and to  validate the model by suitable measure-
ments. For several  reasons  the city of Bremen  was  chosen for the first
mathematical experiments.  Local  authorities  of  the city of Bremen  were
willing and  able (in  1963) to collect  information  on source emissions, which
led to a very complete source  emission  inventory. The location of the city,
on flat terrain and only 40  miles from  the North Sea, is favorable in many
respects: the city  is  well ventilated throughout the year and, in addition, the
relevant meteorological fields are approximately  homogeneous horizontally.
In addition, advection  of pollutants from other regions can be neglected. The
input  data, therefore,  were  well defined and relatively simple and allowed
the application of  a simple model.

MATHEMATICAL  MODEL
The steady-state theory of the transport and dilution of  pollutants is based
on a  number of simplifying assumptions: the relevant  meteorological fields
are stationary and are horizontally homogeneous;  dispersion is not limited in
the vertical direction;  mean windspeed exceeds a certain lower limit; and the
earth's surface  is  flat  and not  absorbing. The well-known  formula for the
spatial distribution of a pollutant1 then is:
     Q        . exp
           r   .
X =    exp [71-]
                      r
                       -
                      L
                          0
exp  -   2        exp
                                  \/27r  az           \f~2-n az
                                                                     (1)
As usual, h = hs + Ah  is effective stack height, r = x/U  is travel time, and
T = 1/7  is decay time.
Assuming that the  plume  standard deviations,  ay and az, are functions of
travel  time,  T,   it  can  be shown  that  Equation (1)  is a solution  of the
following13  initial-value and boundary-value problem:


                                              -
                                                                     9-5

-------
                    X  - 7J   6 (y) 5 (h~Z)                             (3)
      Z  = 0  :       =  0                                              (4)
      Z ^ oo :   x  ->  o                                              (5)

The initial condition,  that  is,  Equation (3),  expresses the fact  that a point
source  of strength, Q,  is  located  at  the  effective  stack  height, h. The
boundary conditions show  that there  is reflection of the  pollutant at the
earth's  surface  and  further  that  dispersion  is  not limited  in  the vertical
direction.

It may  be noted that  Equation (1)  has the character of Green's function in
the  special  boundary-value  problem, Equations (2),  (4), and (5).  It seems
legitimate, therefore, to apply Equation  (2) to  problems connected with
boundary conditions different from those described by Equations (4) and (5)
Two important  assumptions are in question;  that  of a non-absorbing ground
and  that  of unlimited  vertical  dispersion.  Denoting  by H  the  height of  a
ceiling  restricting dispersion to a limited  layer  of  the lower   atmosphere
(urban  boundary layer), and denoting by afzylthe absorption coefficient of
the ground, the boundary conditions. Equations (4) and  (5), can be  replaced
by:


                             e ' •"•»>«

      Z = H  :              || = 0                                  (7)

Even if absorption  is not a function of location,  only  mathematical, rather
than experimental,  methods are suitable for solving  Equation  (2)  together
with Equations  (3), (6), and (7).14  In  view  of the uncertainties connected
with absorption  at  the ground, and  with the functional  behavior of plume
standard deviations  for such cases, only experimental calculations were per-
formed  with  boundary-condition Equation (6).

Important for practical  applications,  however, is the assumption that disper-
sion is  confined only  to the urban  boundary  layer; i.e.,  the utilization of
boundary condition  Equation (7) for  the upper ceiling of the layer   (Figure
9-3).

Standard  methods's~'7 allow  the derivation of  a modified version of Eq-
uation (1) so that it now describes dispersion  in a boundary layer of depth H:
                                                                       9-6

-------
     Ah
        hs
                                    -=0
                            U        PLUIVIE AXIS
                                 ar
                                    =o
X  =
Here
 Q
2HU
exp[
                  Figure 9-3.  Model assumptions.
                                        h-z  .  a
                                                                     (8)
                      V277 dy
       03(V,W)   =
is a Jacobian theta-function.
                   L*J
             1    £
             r\A/    i—'
                                     exp  -
                                                                (9)
                             77=-°
It can be shown that Equation  (8) differs only  slightly from Equation (1)
for  H > 3h.  If,  however, the ceiling  approaches the effective stack height,
i.e., if H -> h,  then the ground-level concentration increases drastically.

So far only a single  source has been considered. In an urban area, a large
number of such sources  exist.  In reality, all of them are point sources as far
as emissions are concerned. They may be divided  into three groups. Group 1
consists of  all  industrial  stacks,  including  those of  power stations  and
gasworks;  Group  2 consists of stacks of small industries,  contributing,  say,
less  than  0.02 percent  each  to the  total  output into the  city; Group  3
consists of all  domestic sources burning fuel for space  heating.

Industrial  emission  sources of  Group  1  are treated individually by applying
Equation (8).  Since they are irregularly distributed over the urban area,and
                                                                     9-7

-------
since Equation  (8) applies to a source-oriented coordinate system, the trans-
formation of coordinates shown in Figure 9-4 is performed.
                                                       x + y tan /?
    Figure 9-4.  Transformation of coordinates from a source-
    oriented system to a geographically fixed system.

 If a denotes wind direction  ((3 =  3?r/2  - a) ,  P(x,y) a receptor point, and
 (Sn.^n) the location of a point source, then the individual source distance in
 wind  direction, Xn  ,  and the individual  crosswind distance, Yn, of the
 receptor point,  p(x,y), are given by:
      Xn  =
e-7  =  (x-£n)cos|3+(Y-r;n) sin 0

k(e-y) =-(x-£n) sin 0 + (Y-tjn) cosj3
(9)
where
      7
9-8

-------
Now, source strength, wind speed, travel  time,  plume standard deviations,
and effective stack height are dependent on  (|nT?n)  and  (Xn, Yn) ,  respec-
tively. Individual source strength  is denoted  by Q = Q(^n,r]n) = Qn   Wind
speed relevant  for transport and  dilution of pollutants originating  from the
source at (£n,??n)  is Qiven by  U = U(£n,T?n) . Travel time is indicated by Tn
  Xn/Un,  plume  standard deviations by  ay  = ay  (rn)  = Oyjn  and az  =
uz(Tn)  =  az,n ,  and  local  effective  stack by   h  =  h(£n,?7n)  =  hSil,  +
Ah(!|n,i7n)   Introducing these new parameters and coordinates into  Equation
(8), the ground-level concentration originating from emission source "n,"  i.e.,
Xn(x,y,z) ,  is obtained.

The concentration  fields  from  all individual sources can  be superimposed.
Contributions to  the concentration  at  receptor  point,  P, come  from  all
upwind sources  having coordinates, £n < x + y tan /3, T?n < y +  x/tan )3.    If
N denotes  the total  number of upwind sources, the concentration P is given
by:
                                                                     (10)
     i   fc   ^
     ^   r^1  U"
It is obvious that computing time goes up tremendously with an increasing num-
ber of individual sources. It is impossible, therefore, to treat all domestic sources
of Group 3 individually. In this context one generally talks about area sources.
The source strength, Q, is replaced by a local source strength  density (source
emission per unit area), q(£;r?).  Local relevant mean wind speed, U(|TJ), as
well as local plume rise, AhfijT?),  are connected with the  emission  height
hs(£?7) of /the area  source.  The coordinates, Xn,Yn, with  respect  to  an
individual  source are  replaced by  x,y which obey the  same relations, in
Equation  (9), as Xn and  Yn do.  If the sum in Equation (10) is replaced  by
an integral, the contribution of all upwind area sources to the concentration
at  receptor point, P, is given by:
                                                                     (11)
    1     / d?            /
=  ^rr   J               J
                                                     -i  exp  -
          £^x+ytanj3 17 < y+x/tan )3  U(|T?)
                                                                    ._
                                                                    (T)
                                                                CTy(?)
Superposition of  concentrations,  in  Equations  (10)  and  (11), gives the
steady-state concentration field  at any location  in space, (X,Y,Z), if  indivi-
dual  point sources as well as area sources act together in  that  urban area.
                                                                      9-9

-------
 Apart from the fact that source emission data for area sources, i.e., q(ijr))
 are  not available,  the analytical integration  of  Equation  (11)  cannot  be
 performed.  Numerical  integration replaces the  integral  by  a  sum  which
 represents  the area  source  by a dense,  regularly  spaced system of pojnt
 sources having source strength, q(£j,T/k)AI;AT?. This, in fact, has been done in
 the  model.  The selection of the area element, AJjAr?, depends upon resolu-
 tion and scale.  In  addition, the characteristics of dispersion as well as the
 conditions  of emission (emission  heights)  are  important. In order to find a
 satisfactory  answer to that  question,  a number of mathematical experiments
 was  performed. An  area source, 500 m  by 500  m was represented by a
 successively  increasing number of point sources (Figure 9-5).
 Figure 9-5. Simulation of an area source by the  use of increasing
 number of point sources distributed regularly over an area 500 x
 500 meters.
As  Figures 9-6 and  9-7 show, an area size of A = A|Arj = 56 m x 56 m
should  be  sufficient  for  the representation  of  an area source  by a great
number of point sources  under  the conditions indicated in Figures 9-6 and
9-7.

In  order to  get  the same  degree of approximation  for a wide range of
windspeeds and stability categories, the  500 m by 500  m area source must
be  represented by at  least 100 individual sources (A =  A£A?7  = 50 m x 50
9-10

-------
m) or better, by 144 individual sources (A=A|Ar? «= 42 m x 42 m).  This
corresponds roughly to the mean distance between individual stacks from
space  heating units.
                    S02 CONCENTRATION PROFILES
 Figure 9-6.  Successive approximation of area-source emissions
 under  unstable conditions by the use of  an increasing number of
 point sources.  (Relative crosswind SC>2 concentration profiles
 taken  at Xmax distance from center of area. Uniform emission
 height is 25 meters; stability class, 2; wind speed, 3 meters per
 second.)
                                                            9-11

-------
 s
 o
 cc
                    S02 CONCENTRATION PROFILES

  Figure 9-7.  Successive approximation of area-source emissions
  under neutral conditions by the use of an increasing number of
  point sources.  (Relative crosswind SC>2 concentration profiles
  taken at Xmax distance from center of area.  Uniform emission
  height is 25 meters;  stability class, 2; wind speed, 3 meters per
  second.)
9-12

-------
METEOROLOGICAL DATA  INPUT
A  set of meteorological data  consists of windspeed,  wind direction  (both
taken  at  anemometer level), and  stability  category.  All  three  are  hourly
values taken at the  Bremen  airport.  Observations of windspeed  and wind
direction  at  four  sites in the  city during  winter (1967-68) validated  the
assumption that the airport observations are representative for the urban area
of Bremen,  at least during winter, which is the  most important period with
respect to air pollution in  Bremen. Stability categories were computed using
Turner's scheme.2
It  was discovered  during  simulation  experiments  that  36 wind direction
measurements (wind roses  divided  into 10-degree intervals) are necessary to
provide reasonable ground-level  concentration fields. Windspeed was divided
into seven classes. Including five stability categories, a total number of 1260
combinations exist, of which, however, only about 600 are realized.

Frequency  distributions of  wind  data were  calculated  for  each stability
category  for a variety of periods  (months of the year,  seasons, years, and
five to  ten  years).  Figure  9-8 shows a typical  example of a  long-term
distribution.
Data on plume standard deviations for urban areas were not available during
the years of experimentation.  Therefore, the  well-known  Pasquill-Gifford
values1'3 were  used  for  low emission heights  (space  heating), whereas  the
Brookhaven values 4'18  were applied  in a slightly modified version  to high
industrial stacks (Figures 9-9 and  9-10). The results of the St. Louis disper-
sion study indicate that  utilization of  those  values given  in  Figures 9-9 and
9-10 inevitably lead to a systematic overestimation of ground-level concentra-
tions.19 This trend was,  in  fact, apparent when the results of calculations
were compared with those obtained by observations.
Finally, the problem of  mean windspeed, U, which is relevant for  the
transport and dilution of pollutants, was solved in the usual  manner. Wind
observations were  extrapolated  from   anemometer level  to  physical  stack
height  for each stack  by means  of a  power-law-profile  assumption. This
extrapolation was assumed to be a function of stability and was made by  the
use of parameters taken from the literature (Figures 9-8, 9-9, and 9-10).
EMISSION  SOURCE  INVENTORY
All emission sources, as mentioned earlier, were classified in three groups. All
individual stacks  (Group 1) with  emission  rates  greater  than  1  kilogram


                                                                    9-13

-------
 Figure 9-8.   Frequency distribution  of wind direction and wind-
 speed  for  Bremen,  1954  to  1959.  Stability  class: 4 (neutral).
 Isopleth numbers represent hours per period.

sulfur  dioxide  (S02)  per  hour were treated  individually. The record  con-
tained: geographical  location, physical dimensions  of  the  stack,  output by
volume,  exit  velocity,  exit  temperature,  and,  finally, emission data  that
included  maximum emissions and  mean  winter  and summer emissions.  In
addition, data were obtained whenever possible on daily  variations in emis-
sions and emissions during holidays.
Effective stack heights  were  calculated by applying Stumke's empirical for-
mula,20 similar to the well-known CONCAWE-formula, which is applicable to
all types  of stacks in Bremen.
Group  1  emission sources, consisting of  136 stacks, contributed 75  percent
to the  total emission rate in  Bremen during winter  1965 (Table 9-1). Spatial
distribution per square kilometer is given  in Figure 9-11 for these sources.
9-14

-------
   500
   100
E

kT
bN
   10.
      STABILITY CLASSES:
      2 = UNSTABLE
      3= SLIGHTLY UNSTABLE
      4 = NEUTRAL
      5= SLIGHTLY STABLE
      6= STABLE
 HIGH-LEVEL
 EMISSIONS
      LOW-LEVEL
       EMISSIONS
     30
100
                             CTx,m
1,000
5,000
    Figure 9-9. Crosswind-plume standard deviations.
   10
      STABILITY CLASSES:
      2 = UNSTABLE
      3-SLIGHTLY UNSTABLE
      4= NEUTRAL
      5= SLIGHTLY STABLE
      BISTABLE
      LOW-LEV EL-j
      EMISSIONS
    30
100
                           Ox, m
1,0
5,000
    Figure 9-10.  Vertical-plume standard deviations.
                                                             9-

-------
C~n\ BREMEN 1962
\ ^v./-J" ^) GRID: 1 km x 1
•> 20!
^563 24 | ,^'">
[Z ^"J I ! h ^
|_ 107! 	 . _prrjJ_v:/^V
13?..{?:!!M_L _, . -^4 _|=p.^_. 	 ..-^
V [ \-\--\ : ^ ^~TT-'--I-T~
i'.;i_iL-ii l 	 i 	
\l t""~t f l" I J i
Y r 520 ' t [23!" 1 [ j_
7t' * 96° L i 4 j I"
~\ '. ^'j I8jl8.1»?4_1
[ J38 76ho|
f f- "424-71.
1 i_ r-L'?J_r2
Nl ]56 Jr" \rToTl
1 1 ^^
km
L\



r -
.8.
28




J-4^
\
—
"

^./U
1

:

;
255

17040

-
^


\
1

35









25



—





^
R\
^
-J!
_i j
^
i /
hj
i
/'
V.-
 Figure 9-11.  Spatial distribution of mean winter time industrial
 emissions.
     Table 9-1.  STATISTICS FROM 1965 EMISSION-SOURCE INVENTORY

                             OF BREMEN
Source
Industries and power plants
Small industries
Space heating
Total
Number of
stacks
136
425
-
561
Total emissions,
kg SO2 hf '
Summer
1,423
46
-
3,469
Winter
4,715
116
1,458
6,289
Percent of total
emissions
Summer
99
1
-
100
Winter
75
2
23
100
9-16

-------
A large number of individual stacks from small  industries contributed less
than 0.02 percent  each to the total emission rate, i.e., less than 1 kilogram
862 Per  hour.  They  contributed  only  1  to  2  percent  altogether and,
therefore, were  not  treated  individually.  Instead, the same  technique was
used as that applied to emissions from space heating (Figure 9-12).
                                 BREMEN 1962
                               GRID:  1 km x 1 km
  Figure 9-12.  Spatial distribution of mean winter  time industrial
 emissions from small industries.

Although it contributed less than 25 percent to the total emission rate, space
heating  is the  dominant  factor in air pollution  in  Bremen  because of  low
emission heights.

Emissions from  space heating  were obtained  in  the  following  way.  The
spatial  distribution of  dwelling  units  in Bremen, which  number  about
200,000 was known very well. Further, the total  amount of coal and fuel oil
consumption during heating periods was known. From  the sulfur content of
the fuels, the total  fuel consumption, and  the total  number of  dwelling
units, a mean emission rate of 8 grams S02 per hour  per dwelling  unit was
obtained. This is the amount  that would have been emitted daily  during the
heating  period  if  the daily  mean temperature  had remained  constant. In
several experiments, a  relationship between daily mean temperature and daily
emission was used  to  make emissions from space heating  a function  of time.
                                                                   9-17

-------
A mean emission height of 25 meters was assumed for downtown emissions
and  15  meters for suburban  emissions.  The corresponding effective stack
heights were calculated by a simpler method than that used for tall stacks.

Emissions  from space heating  were treated  as  "area sources" as described
earlier. A  grid produced areas 500 m  by 500  m. The number of dwelling
units in each area was counted. Multiplying by 8 grams S02  per hour (or the
corresponding, temperature-corrected value)  gave the mean emission rate for
that area (Figure 9-13). This amount was then divided by the  number of
individual stacks (from 81 to 144) representing the area source.
                                     BREMEN 1962
                                   GRID:  1 km x 1 km
   Figure 9-13.   Spatial distribution of mean  heating period in-
   dustrial emissions from space heating,  kg SC^/km^-hr.


The  fact that  area sources are represented  by  a  large number of  individual,
equivalent stacks simplifies the calculations a great  deal.  The distribution of
dwelling units  in Bremen  is such that only about 250 areas,  500 m by 500
m, are covered with dwellings. Since the density of these  dwellings  varies
spatially, the emission  rates of the 250 types of stacks, in  general,  differ
from  each other.  Their physical dimensions  and  their  plume  rise values,
however, are equal, or  are grouped  into only two categories (downtown and
suburban).  With the shift  of calculated concentration fields from one  point
source location  to the other within an  area  of 500 m  by 500 m, the
computing time is very  much reduced.
9-18

-------
RESULTS OF SIMULATION  EXPERIMENTS
An inventory  of all  possible steady-state ground-level concentration fields
forms the basis  of simulation experiments. This inventory was obtained by
calculating  concentration fields for all  possible combinations of meteorolo-
gical parameters. It has already been  noted that about 600 such  combina-
tions, representing  given  weather  situations,  can occur  in Bremen. This
number  depends, of  course, on  the classification of windspeed and wind
direction used. If one calculates the concentrations at the points of a grid  (a
grid  distance of  500 meters)  about 600  numbers have to be stored at each
grid  point. Since the  computer  program  was  written for  a grid of 2500
points, a  total number of 1.5 million numbers have to be stored.  If the
source inventory is  assumed to  be time-dependent  (heating period,  non-
heating  period,  etc.),  the number  of  concentration values to  be stored
increases considerably.
The  use of concentration  field   inventory  data  together  with  frequency
distributions of relevant meteorological parameters  permits the derivation of
frequency distributions of concentrations for each  point on  the grid.  Other
statistics, as well, can  be  obtained quite simply,  such  as  S02  wind  roses.
Figures 9-14 to 9-17 demonstrate this clearly. Sulfur dioxide wind roses were
calculated  for  the  four sites  where monitoring stations were later  installed.
These figures,  together with  Figures 9-11  to  9-13, show the possibility of
identifying  large  emission sources  by means of S02 wind roses.  It may seem
that  this is by no  means  an easy  task because  emission sources with quite
different  emission heights  work together with meteorological factors having
complicated frequency distributions. It may be noted that the simulated S02
wind  roses coincide considerably  well  in structure with those  obtained by
measurement.
Among the many experiments performed, the investigation of the influence
of boundary  layer thickness on  ground-level concentration  was the most
interesting.  The upper ceiling was lowered  from 500 m to 25 m. In cases in
which the  ceiling reached the effective  height of  an  individual stack,  this
effective height was reduced with the decreasing height of the ceiling until  it
reached two-thirds  its original value. This stack was then thrown out of the
inventory  on  the assumption that the  plume would  penetrate the ceiling
Figures 9-18 and 9-19 show how the  isogram patterns change and they also
show  the tremendously increased  concentrations that result  if the  depth of
the mixing layer is  approximately equal  to  the effective  height  of space
heating  emissions.  Figure   9-20  demonstrates this behavior  for  a  specific
location  in  downtown Bremen.  It is obvious that this picture looks different
for different  locations. Only mathematical experimentation of the,  kind
applied here can simulate  the complicated behavior of grognd-level concen^
trations.
                                                                   9^19

-------
                                     BREMEN, 1967-68
                                     H! INDUSTRY E2 RESIDENTIAL AREAS
                                     —=2 km      •  MEASURING SITES
     Figure 9-14.  Calculated heating-period-SC>2--wincl rose for
     Site 1 in downtown  Bremen.
                                     BREMEN, 1967-68
                                     iH INDUSTRY E23 RESIDENTIAL AREAS
                                     —=2 km      •  MEASURING SITES
                                     .	.   0.1 mg S02/m3
     Figure 9-15.  Calculated heating-period-S02 wind rose for
     Site 2 in downtown Bremen.
9-20

-------
                              BREMEN, 1967-68
                              EH INDUSTRY m RESIDENTIAL AREAS
                                    m      •  MEASURING SITES
                                      1 mg S02/m:
Figure 9-16.  Calculated heating-period-S02-wind rose for
Site 3 on the outskirts of Bremen.
  swav iviiuaaisaa &Z3  Aaisnawi ^3
                   89-1961 'N3IAI3H9
Figure 9-17.  Calculated heating-period-S02~wind rose for
Site 4, in close proximity to an industralized area.
                                                              9-21

-------
                             BREMEN, 1962
                                 INDUSTRY
                                 2 km
RESIDENTIAL AREAS
MEASURING SITES
    WIND DIRECTION: 300'
    WIND SPEED: 3 m/sec
    STABILITY CLASS: 4
Figure 9-18.  Calculated field of ground-level-SC>2concentration
in  mg/m^ for a special meteorological situation and a boundary
layer thickness of 100 meters.
                             BREMEN, 1962
                                INDUSTRY ^^ RESIDENTIAL AREAS
                            	, 2 kin       •  MEASURING SITES
    WIND DIRECTION: 300'
    WIND SPEED: 3 m/sec
    STABILITY CLASS: 4
Figure 9-19.  Calculated field of ground-level-SC>2 concentration
in mg/m3 for a special meteorological situation and a boundary
layer thickness of 25 meters.
9-22

-------
     1.0 •
  cc
  si
  o
  z
  o
  o
  S1
0.5
     0.0
                 SPACE
                HEATING
                      /INDUSTRY
       101
                                102
                       CEILING HEIGHT, meters
  Figure 9-20.  Variations in ground-level concentrations of S02
  in downtown Bremen as a function of boundary layer thickness.
As mentioned  in  the  introduction,  frequency distributions of ground-level
concentrations were of chief interest  from the beginning of this investigation.
The method  for  obtaining these for desired  periods of time  is straightfor-
ward.  It must  be stated,  however,  that the derived distributions are not
complete because up to now no theory  exists for explaining the dilution of
pollutants under  calm  weather conditions.  These cases, therefore, were ex-
cluded from the statistics as were cases with limited boundary layer depths.
Neither source  of error, however, plays an  important  role  in Bremen. The
frequency of calm conditions as well as the frequency of low-level inversions
was small during all periods of time investigated.

Stored data on fields of steady-state concentrations  (Figure 9-21)  form the
basis of statistics of this kind.
Steady-state concentration fields,  together  with frequency  distributions of
meteorological  parameters, can be used  to  calculate  frequency distributions
of concentrations for  each grid point. These distributions were characterized
by a set of three parameters:
   1. The percentage  of time (in  hours) for which  ground-level  concentra-
    tions  exceeded  a given  value  (0.1  milligram   S02  per  cubic meter).
    Figure 9-22 shows the pattern  of this parameter for the heating  period
    of  1962.  As  seen,  only  20 percent  of the  period  concentrations in
    downtown Bremen were below  0.1  milligram  SO2 per  cubic meter.
                                                                    9-23

-------
                                BREMEN, 1962

                                §Hi INDUSTRY

                               	I 2 km
RESIDENTIAL AREAS

MEASURING SITES
   WIND DIRECTION: SOUTH

   WIND SPEED: 3 m/sec

   STABILITY CLASS: 4
   Figure 9-21.   Typical  possible field of calculated steady-state
   ground-level  862 concentrations in mg/m3

   2.  The mean concentration for the period.  Figure 9-23 shows the pattern
      of  this parameter. Typically, the  pattern of the  mean  concentration
      shows little structure and does not  contain much information.
   3.  An upper percentile; for example,  the 97.5th  percentile. The numbers
      in Figure 9-24 indicate that for only 2.5  percent of the time (in hours)
      concentrations exceeded that value given  by the respective number.

The following Figures, 9-25 thru 9-27 show the corresponding pattern for a
nonheating  period  in  which only  industrial sources are contributing emis-
sions.

It  might  be possible to  define Imissions-Grundbelastung with the help  of
these  maps of  characteristic parameters.  This  investigation  is one step for-
ward  in this direction.
VALIDATION OF THE MODEL
During  the  heating  period  of  1967-68,  four monitoring stations were  in-
stalled at specially  chosen  locations. The locations  were planned as stra-
tegically  as  possible. First of  all,.an  attempt  was made to place all stations
on  a mean  concentration  isogram  (Figure 9-23). Second, an  attempt was
9-24

-------
                         BREMEN 1962
                            INDUSTRY Wflk RESIDENTIAL AREAS
                           j2 km      »  MEASURING SITES
 Figure 9-22.  Percentage of cases (hours) for which ground-
 level  concentrations exceeded 0.1 mg S0/m3.  Winter 1962.
                        BREMEN 1962
                           INDUSTRY W/M RESIDENTIAL AREAS
                          J2 km      •  MEASURING SITES
Figure 9-23.  Mean ground-level-S02 concentration in mg/m3.
                                                         9-25

-------
                            BREMEN 1962

                           liH INDUSTRY HM RESIDENTIAL AREAS

                              .•2 km      »  MEASURING SITES
  Figure 9-24.  97.5th percent! le S02 concentrations in mg/nv
  Winter 1962.
                           BREMEN 1962

                           jmH} INDUSTRY

                              42 km
RESIDENTIAL AREAS

MEASURING SITES
   Figure 9-25.  Percentage of cases (hours) for which ground-
   level concentration exceeded 0.1 mg  S02/m3.  Summer 1962.
9-26

-------
                          BREMEN 1962
                              INDUSTRY (^^ RESIDENTIAL AREAS
                                       •  MEASURING SITES
 Figure 9-26.  Mean ground-level-S02 concentration in mg/m3-
 Summer 1962.
                          BREMEN 1962
                              INDUSTRY
                             j2 km
RESIDENTIAL AREAS
MEASURING SITES
Figure 9-27. 97.5 percent! le of ground-level-SC>2 concentrations
in mg/m3.  Summer 1962.
                                                           9-27

-------
made to locate  stations in areas as diverse as possible. Station 1  was located
in a "normal" downtown area,  surrounded mainly by residences.  Station 2
was  located in the very center of  the city on an  island.  It was  surrounded,
however, on all sides by the water of the river Weser  and of the waterworks.
Station 3 was located on the outskirts of the city, separated from downtown
Bremen by a large  park with tall  trees.  Station  4 was located  in the near
vicinity of a large plant.

The  monitoring stations measured  half-hourly values  of S02 concentrations.
From these values, frequency distributions were derived for every  month of
the  period  and  for  the period as a whole. At the same time, mathematical
simulation  of the same distributions was performed  using the latest  version
of the emission  source  inventory  and  utilizing  meteorological statistics for
the  sampling  period. Figures 9-28 through  9-31  show  comparisons of ob-
served  and  calculated  frequency distributions.  The simulation of Station 1
(downtown "normal"  area)  is quite satisfactory,  as  indicated in Table 9-1.
For  Station 2  (waterworks  on  the Weser island),  the  model obviously
overestimates the concentrations systematically (Figure 9-29). Overestimation
may occur for one or both of two reasons: absorption at the water surfaces,
or an  insufficient  spread  as a  result  of  improperly chosen  urban plume
standard deviations. The same holds for Station 3 (separated from downtown
Bremen by a large park), where the filtering effect of  the park was not taken
into  account.

At Station  4  (in the vicinity of a large  plant), the reverse is observed. The
model   systematically   underestimates the concentrations.  Since  emissions
from low-level sources of space heating are small in the neighborhood  of that
station, low concentration values could be expected. The comparatively high
concentrations  that  actually occur  have  their origin  in  uncontrollable low-
level emissions, which  could  not  be taken  into  account, from  the  nearby
plant.

Table  9-2  summarizes the observed and  calculated mean concentrations for
each monitoring station.

Finally, it  can  be stated  that  it is worthwhile  to  invest  more  effort in
diffusion modeling, for simulation may one day be a very important tool in
city  planning.
9-28

-------
   99.98:

   99.9

   99.5
   CUMULATIVE FREQUENCY POLYGON
     AND FREQUENCY HISTOGRAM

SITE NO. 1, UBERLANDWERK NORD-HANNOVER
      HEATING PERIOD, 1967/68
                                     "r" :>'J?r'90tti PERCENTILET
                                            _	:EZ143iai
                                    =COMPUTEO
                     AVERAGE GROUND-LEVEL CONCENTRATION, mg/m3

  Figure 9-28.  Comparison between observed and computed
  frequency distributions of ground-level  concentrations in
  downtown Bremen.
    99.98
    99.9

    99.5
    0.05
    0.02
    CUMULATIVE FREQUENCY POLYGON
      AND FREQUENCY HISTOGRAM

    SITE NO. 2, WASSERTURM WESERINSEL
        HEATING PERIOD, 1967/68    ,g
     0.01
              0.02
                       0.04   0.06  0.08 0.1
                     AVERAGE GROUND-LEVEL CONCENTRATION, r
                                                            0.6  0.8  1.0
Figure 9-29.  Comparison between observed and computed
frequency distributions of  ground-level concentrations in
downtown Bremen.
                                                                    9-29

-------
                 CUMULATIVE FREQUENCY POLYGON
                   AND FREQUENCY HISTOGRAM
                SITE NO. 3, UMSPANINWERK BLOCKLAND
                    HEATING PERIOD, 1967/68

                                     -tfj?
          0.01       0.02       0.04   0.06  0.08 0.1
                          AVERAGE GROUND-LEVEL CONCENTRATION, 1

     Figure 9-30.  Comparison between observed and computed
     frequency  distributions of ground-level  concentrations on
     outskirts of Bremen.
     qq go ,..	  ...	—	
        ,   CUMULATIVE FREQUENCY POLYGON
     99.9     AND FREQUENCY HISTOGRAM
     99.5
     95
     90
     80

  :-«
  "  60
  a  so
  o  40
  E  30
     20
     10

     4
     2
     1
     0.5
     0.2
     0.05
         SITE NO. 4, PUMPENSTATION RIEDEMANNSTRASSE
                HEATING PERIOD; 1967/68
     0.02

            OBSERVED
                      jz^.' f-^ps  ':   •          __
    0.01       0.02        0.04   0.06  0.08 0.1        0.2       0.4    0.6   0.8 1.0
                      AVERAGE GROUND-LEVEL CONCENTRATION, mg/a?
 Figure 9-31.  Comparison  between observed  and computed
frequency distributions of ground-level concentrations in an
industrial district.
9-30

-------
Table 9-2.  OBSERVED AND CALCULATED MEAN SO2 CONCENTRATIONS
            IN BREMEN; HEATING PERIOD, 1967-1968
                           (mg  ~

Site
November
December
January
February
March
April
May
Total
1
Calc.
0.14
0.10
0.10
0.09
0.08
0.09
0.05
0.09
Obs.
0.11
0.08
0.10
0.08
0.07
0.07
0.06
0.08
2
Calc.
0.15
0.12
0.12
0.12
0.10
0.11
0.07
0.12
Obs.
0.10
0.08
0.13
0.08
0.04
0.05
0.03
0.08
3
Calc.
0.14
0.10
0.09
0.07
0.07
0.06
0.04
0.08
Obs.
0.08
0.06
0.08
0.06
0.05
0.05
0.03
0.06
4
Calc.
0.05
0.03
0.04
0.04
0.04
0.05
0.03
0.04
Obs.
0.10
0.07
0.08
0.08
0.07
0.07
0.05
0.08
                                                            9-31

-------
REFERENCES

 1.  Pasquill,  Frank.    Atmospheric  Diffusion.  London,  D. Van Nostrand Co. Ltd
     1962. 297 p.
 2.  Turner,  D.  B. A Diffusion  Model for an Urban  Area.  J. Appl. Meteorol. 3(1):83-
     91, February 1964.
 3.  Gifford, F  A. The  Problem of Forecasting  Dispersion  in the Lower Atmosphere.
     AEC  Division of Technical Information Extension. Oak Ridge, Tenn. 1961.
 4.  Singer, I. A. and M.  E.  Smith.  Atmospheric  Dispersion at  Brookhaven National
     Laboratory. Int. J. Air Water Pollution. 70:125-135, February 1966.
 5.  Sutton,  0.  G.  Micrometeorology; A Study of  Physical Processes in the Lowest
     Layers of the Earth's Atmosphere. New York, McGraw-Hill, 1953. 333 p.
 6.  Verein Deutsche Ingenieure. VDI-Forschungsheft 483,  Ausgabe B, Band 27, 1961.
 7.  Ausbreitung luftremder Stoffe in der Atmosphare Zusammenhang zwischen Emis-
     sion  und Immission Schornsteinhohen  in ebenem, unbebautem Gelande.  Verein
     Deutscher Ingenieure,  VDI-Kommission Reinhaltung  der Luft.  Germany. VDI
     2289. June  1963. 7p.
 8.  Frenkiel, F  N.  Atmospheric Pollution  in Growing Communities. In: Annual Re-
     port  of  the Board  of Regents of  the Smithsonian Institution,  Publication 4272,
     1956. Washington, D. C. Government Printing Office. 1957. p. 269-299.
 9.  Fortak,  H.  G.  Rechnerische Ermitt/ung der  SO2  — Grundbelastung aus Emission-
     sdaten — Anwendung auf die Verhaltnisse des Stadtgebietes von Bremen. Institute
     for Theoretical Meteorology, The Free University of Berlin.  1966.
10.  Clarke, J. F  A  Simple Diffusion Model for  Calculating Point Concentrations from
     Multiple Sources. J.  Air Pollution Control Assoc.  74347-352, September  1964.
11.  Marsh, K. J. and  V. R. Withers.  An  Experimental Study of the Dispersion of the
     Emissions  from  Chimneys   in Reading  —  III:  The   Investigation of  Dispersion
     Calculations. Atmos. Environ. 5(31:281-302,  May 1969.
12.  Davidson, B. A  Summary of the New  York  Urban Air Pollution  Dynamics Re-
     search Program.  J. Air Pollution Control Assoc. 77:154-158, March 1967.
13.  Frenkiel,  F. N.  Turbulent Diffusion:  Mean  Concentration  Distribution  in a  Flow
     Field  of Homogeneous Turbulence. In: Advances in Applied Mechanics,  von Mises,
     R. and T. von  Karman (eds.), Vol. III. New York,  Academic Press Inc., 1953. p.
     61-107.
14.  Fortak, H. G. Einbeziehung  der Sinkgeschwindigkeit und partiellen Absorption am
     Erdboden  in  die Ausbreitungsrechnung,  speziell  im  Falle nicht-FICKscher Dif-
     fusion. Institute for Theoretical Meteorology, The Free University of Berlin.  1964.
15.  Fortak, H. Zur allgemeinen  Berechnung von Suspensionsverteilungen in turbulenten
     Stromungen. Gerlands Beitr. Geophys. 66(11:65-78,  1957.
16.  Stumke,  H.  Vorschlag einer empirischen Formel fur die Schornsteinuberhobung.
     Staub (Dusseldorf). 25:549-556. December 1963.
17.  Fortak,  H.  G. Ausbreitung  von  Staub und  Gasen um  eine kontinuierliche Punkt-
     quelle in  einer bezuglich  Windgeschwindigkeit  und Austausch geschichteten Atmo-
     sphare. Verein Deutscher  Ingenieur. VDI-Forschungsheft 483.  1961.
18.  Singer, I.  A.  and M. E. Smith. Relation  of Gustiness to Other  Meteorological
     Parameters. J. Meteorol.  70(2) :121-126, April 1963.
19.  McElroy, J. L. and  F  Pooler, Jr. St. Louis  Dispersion Study, Vol. II — Analysis.
     National Air Pollution Control Administration. Arlington  Va. Publication Number
     AP-53. 51 p.
20.  Fortak,  Von Heinz. Sinkstofftransport in Geraden  Kana/en als Randwertproblem.
     Acta  Hydrophysica.  4(11:26-48, 1957.
9-32

-------
ACKNOWLEDGMENT
In  the course of the  research, a  number  of  members of the Institut  fur
Theoretische Meteorologie der Freien  Universitat Berlin contributed to this
work, in  many ways, mainly in programming and organizing. Especially,  the
author wishes to thank the following for the valuable cooperation, without
which  these  experiments  could  not  have  been  performed:  R.  Bleck, J.
Schwirner, H. Buttner, H. Woick, and P. Lenschow.
                                                                 9-33

-------
             APPENDIX -  GLOSSARY OF SYMBOLS

a(r,y)     absorption coefficient of the ground
h         effective stack height
hs        actual stack height
Ah       plume rise
H         height of dispersion ceiling
N         total  number of upwind sources
P(x,y)     receptor point
q(£ 17)     local  source-strength density
Q         source strength
T         decay time, 1/7
U         wind  speed
(Xn,Yn)   source distance  in downwind and crosswind directions, respectively
a j3       wind  direction angle coefficients
A         source area
(In.^n)    location of a point source
TT         3.14
oy,<72     standard deviations of plume spread  in y  and z directions, respec-
          tively
7         travel time
X         concentration
 9-34

-------
ABSTRACT
      A  source-oriented three-dimensional model  of diffusion and transport
      based upon the statistical concept of turbulent diffusion was developed
      for application to an urban area.  A simple algorithm  was found for
      computing  the concentration field from time-dependent continuously
      emitting area sources of different scales.  The model was applied to the
      New  York  City  metropolitan region  by introducing known  specific
      meteorological conditions and the known pollutant (S02) source distri-
      bution as input data.  The  validity of the model has been verified in a
      preliminary manner. The results appear promising, giving concentration
      fields computed on spatial scales down to 0.2 mile.
AUTHORS
      Until his death in 1968, BEN DAVIDSON was Professor of Meteorology at New
      York University,  a consultant to both the U. S.  Weather Bureau and Brookhaven
      National Laboratory, and chairman of the working group on sites for wind power
      installations of the World Meteorological Organization. He has also been on the
      University  of Chicago Review Committee for the Radiological Physics Division of
      Argonne National Laboratory and on  the Air  Pollution  Training Grants Com-
      mittee.  Most of his research was concerned with turbulence diffusion, local wind
      circulations, and urban air pollution models.
      LIAU JANG SHIEH,  a staff member of the  Internationl Business Machines Cor-
      poration Palo A/to Scientific  Center, received his Ph. D. in Meteorology at NYU
      in  1969 for work done with Dr. Davidson. He also holds an M.S. from NYU and
      a B.S. from National Taiwan University, Formosa.
      JAMES P.  FRIEND is  the director of the U. S. Department of Health, Educa-
      tion, and Welfare Air Resources Engineering  Training Program at NYU,  where he
      is Associate Professor of Atmospheric Chemistry. He is a member of the ad hoc
      committee for Air Pollution  Manpower Affairs, NAPCA, and  holds a S.B. in
      chemistry from Massachusetts Institute of Technology, and an M.S.  and Ph. D.
      in chemistry  from Columbia.  His principal  research is in the area of stratospheric
      aerosols and mathematical models of air pollution dynamics.

-------
                               10.    A  MODEL  OF DIFFUSION
                                  IN  URBAN  ATMOSPHERES=
                               S02  IN  GREATER NEW  YORK
               LIAU  J. SHIEH, BEN  DAVIDSON,
                    AND JAMES P. FRIEND

                      New York University
INTRODUCTION
The essential  components of a model of dispersion and transport of pollu-
tants in an urban atmosphere are (1) distribution and strengths of sources of
pollutant  in  space  and  time,  (2)  a set of meteorological conditions as
functions  of space  and time, (3) a  mathematical formula (algorithm) repre-
senting the physical concepts of dispersion and transport, and  (4), in  com-
bination with  the  formula  of  (3), an expression  for  the  removal  (decay,
settling, or absorption on surfaces) of the pollutant.

The model described  below  was  developed  in  the Urban Air Pollution
Dynamics  Research Project at  New York University. It was a task of the
project, not only to develop the model, but to  provide sufficient  data so
that the model could be tested. The pollutant chosen was sulfur  dioxide
(S02)  and the urban region was a  50 by 50 mile  square containing greater
New York City. Commensurate with the requirements  implied by the pre-
vious paragraph, the project provided a survey of SC>2 emissions throughout
the area.  Also, during specific test periods of 2  to  5 days  duration,  it
provided wind, temperature, and S02 concentration measurements at several
points within  the area. Helicopter-borne instruments were  used for vertical
soundings  of temperature and S02 concentrations.

The model of diffusion  and transport in urban atmospheres was fashioned
'Contribution No. 84, Geophysical Sciences Laboratory of the Department of Meteorol-
ogy and Oceanography, New York University. This research was sponsored by the U. S.
Department of Health, Education, and Welfare under Grant No. AP 00328-04.
                               10-1

-------
from the statistical  theory of turbulent diffusion.  In that theory, concentra-
tion profiles in diffusing plumes are represented as  Gaussian distributions. An
attempt was made to maintain generality within the Gaussian framework of
the  theory.  The formulated  theory was thus applicable  in three dimensions
to non-isotropic  and time-dependent  motions of the atmosphere. The model
constructed  on the  basis of the theory was made  to accept  as data (1) any
specified  time and space distributions of area  and point source strengths,(2)
specified  meteorological  parameters  (wind, temperature, thermal  stability,
diffusivity,  and height of inversion layer) as functions of  time and space, and
(3) the (initial)  field of pollutant concentrations over the region  at any given
time. The output of the model consists of fields of  pollutant concentrations
in three dimensions  at 2-hour intervals after the initial time.

DATA
As used here, the term "data"  refers  to quantities entered into the computa-
tional scheme as  input.
A summary  of  the  meteorological  and  SO2  source  and concentration data
obtained  in the Urban  Dynamics project was given by Davidson1 and will be
given in complete detail by Davidson et a/.2

Meteorology
Winds  for the  various  trial periods were obtained from  a fixed network of
stations in the Greater New York region and from analyses of data  obtained
from  pibal ascents at three or four fixed sites plus a  reference ascent at JFK
airport  made on  a  continual  basis throughout each trial period.  Scudder3
analyzed stream-line patterns  derived from various wind data and noted char-
acteristics  ranging from essentially straight streamlines,  through curved in-
flected  ones, to sharp discontinuous ones.  He also found high wave number
patterns. These patterns generally depend upon the strength of the prevailing
synoptic  wind field, time of day,  and  insolation. Evidently the combined
effects of land-sea, land-river, urban-rural, and other  contrasting topographi-
cal features produce complicated mesoscale wind flow patterns.

In view of the observed complexity  of flow over the New York City area  it
is clear that a realistic  model of  diffusion  and transport must be capable of
incorporating space-and time-dependent winds. Turbulent diffusion  processes
are treated  in the model  by  the  statistical theory of diffusion. Transport of
the pollutants is  based on the  average trajectory of  parcels according to the
analyzed  wind field.  This trajectory  method was studied  and verified by
Druyan.4
Bornstein5  frequently found  temperature  inversions  at elevations of 200 to
300  meters over  the metropolitan area in nighttime and early morning hours,
when the inversion was at ground level  in rural areas. This is the heat island
10-2

-------
effect  due to  release of  heat  from  urban  surfaces. The air  beneath  the
inversion  generally  exhibited  vertical temperature gradients  ranging  from
isothermal to adiabatic.
As Davidson1  pointed out,  "the diffusive consequences of the temperature
structure over the urban area are enormous. One would expect good vertical
diffusion  below the elevated inversion, but virtually zero diffusion through
the base of the inversion."
The model includes this thermal  stratification effect by treating the observed
elevated inversion layer as a physical  barrier to diffusion of the  pollutant.
The height of the inversion at all points at all times is entered as data in the
computational  scheme. The meteorological network  and the grid used in
model computations are shown in Figure  10-1.
       74° 45'
74-30'
74° 15'
74-00'
73° 45'
    41'15'
73i»30'
                                                ^ 41-15'
       "74-45'     74-30'       74-15'       74-00'      73-45'      73-30'
                            LONGITUDE, degrees

     Figure 10-1.  Meteorological grid  system and  observational
     network.
                                                                      10-3

-------
SO2  Emissions
A  comprehensive source survey was conducted in  the  50 by 50 mile-square
region shown  in Figure 10-1,  and a portion of it is shown  in Figure 10-2.
Two  types of sources  were delineated: (1) area sources representing emis-
sions  from fuel  combustion used for space heating and production  of hot
water  and  (2)  continuous point sources  representing  industrial  emissions
from  electrical power generating  plants,  large  hospital heating plants, large
industrial  heating plants, and chemical plants.

Messrs.  P. Halpern  and  L. Randall* have  analyzed the diurnal and daily
variations of S02 emission rate for area  sources. Their findings have been
directly incorporated  into the model:

                      Qyear             Qyear
      Qriav  =    07   —^ DD   +  0.3                              (1)
       day             4871             365

where Qday is the daily total output of S02, Qyear tne annual total output,
and DD  is the number of degree days for  the date  of interest. The diurnal
variation of the  area-source strength  was found to  depend upon atmospheric
temperature in the manner shown in Figures 10-3 and 10-4.

The survey  of point  sources was conducted by Messrs. Kaiser, Simon, and
Ingram.*  The most complete data were available for electric power genera-
tion   plants.  For other sources most of  the  important  parameters such as
stack  height,  heat  emission rate, and diurnal variation of  source strength
were  not  available. Hence, assumed  values  of these are used in the  model.
Total  output of  SO2 was, however, obtained for each source.
The  model  treats the  point source emissions in a straightfoward manner
using  the  equations for a continuous elevated  point  source emitting a Gaus-
sian plume.

Initial S02  Concentrations

In all model  computations  to date the initial concentration of S02  has been
taken as zero  everywhere. Experience has shown that under steady wind and
emission conditions the concentrations computed at the end of the first time
step  (At  =  2  hours)  are greater than  or  equal to 90 percent of  the final
steady values.

THEORY AND ALGORITHM
The model  is based on the statistical theory  of turbulent diffusion.  In this
theory it  is hypothesized  that the concentrations in a cloud of trace material
(called pollutant in this paper) emitted from  an  instantaneous point  source
*Members of the research staff, New York University.
10-4

-------
-11  -10  -9  -8
                                          -7   -6   -5   -4   -3   -2   -1   0    1
                                                    MILES EAST OF BATTERY
34567
o
CJI
                                    Figure 10-2.  SC>2 source inventory (tons mi"2 yeaH).

-------
      14
      12
                 i    r
i    i     i    i     r
- MEAN DAILY TEMP. 2(
                          \     I
                        'to29°F.
	MEAN DAILY TEMP. 30 ° to 39° F.
           0200 0400 0600  0800 1000 1200 1400 1600  1800 2000 2200 2400
                              CLOCK TIME

  Figure 10-3.  Diurnal variation of S02 output during days with
  mean temperature ranges of 20° to 29° F and  30° to 39°  F.
16
14
OJ
s 12
CL
|-
0.
§ 8
_i
£ 6
o
>- a
_j *
0 9

—

—



1
— MEAN DAILY TEMP. 40° to 4
r — T - — MEAN DAI LY TEMP. 50 ° to 5
j ! ---MEAN DAILY TEMP.* 60° F
1 L— i
i i
[ |
L_4.™j 	 (___! L_.
"-I.J 1



1 1 1 1 1 1
9"F
9°F-

—
i
i
i

          0200 0400  0600 0800 1000 1200 1400  1600 1800 2000 2200 2400
                            CLOCK TIME

 Figure  10-4.   Diurnal variation of SC>2  output during  days with
 mean temperature 40° to 49°  F, and > 60° F.
10-6

-------
are given by  a  three-dimensional  Gaussian distribution. Empirical data sup-
porting the use of  such an hypothesis  were found  by Sutton,6  Frenkiel,7
Ogura,8 and others.  Application of the theory to the atmosphere was studied
by many investigators,  e.g., Batchelor;9'  10  Hay and  Pasquill;11 Cramer,
Record, and Vaughn;12  Barad and Haugen;13  and Lin  and Reid.14

Models of urban atmospheric diffusion using statistical diffusion theory have
been formulated for several specific urban areas-in one case a whole state,
Connecticut—under  various  simplifying  assumptions  concerning  area-source
representations,  and temporal and spatial uniformity of meteorological con-
ditions. The pertinent features of these  models  including the regions  repre-
sented and  the  assumptions made  were  recently reviewed by  Moses15 and
will not be further discussed here.

In addition to the  diffusion mechanism, the transportation  of pollutant  is
affected  by large-scale flow, which is taken here as mean wind.  It is assumed
that, as the material is released from  a  point  source,  it  is dispersed and the
whole cloud is  simultaneously transported  along a trajectory given by the
mean wind field.
With the  above  assumptions the  equation for  the concentration  distribution
in three dimensions  in a moving  medium from an instantaneous point  source
with non-isotropic diffusion is represented by:
                     Q(t)
C x,y,z,t  = ——	  • exp
            (27r)3/2(Tx(t)ay(t)o-z(t)
  if/x-ut-x'Y  /y-vt-y^
~2\\  ax(t)  /  \Mtj~7
                                                                      (2)
where:
                C =  concentration  (g cm"3)
                Q =  source strength (g sec"1 )
                t  =  time (sec)
           u, v, w =  velocities in x, y, z  directions, respectively (cm sec"1 )

        ax,ffy,crz  =  standard  deviation of diffusion cloud in  three dimen-
                     sions, respectively (cm)
          x', y', z' =  position of release


The  development  of  Equation  (2)  into  a  diffusion  model  applicable  to
continuous point and area sources is divided below into three parts: standard
deviation of cloud, point  sources,  and area  sources.
                                                                     10-7

-------
Standard Deviation of Diffusing Cloud
Numerous schemes  have been  proposed for determining formulas and for
determining the a's  in Equation (2). Most of these are based on G.I. Taylor's16
original work. Among the investigators are Sutton,6  Holland,17 Kellogg,18
Frenkiel   and  Katz,19  Gifford,20  Gifford  and   Pack,21   Inoue,22  and
Hogstrom.23
Pasquill,24' 2S  using  experimental  data, formulated  a set of  rules for deter-
mining  ffy and  az  for  different atmospheric conditions. These have been
widely used in estimating concentrations in  diffusing plumes.

It is generally considered that the diffusion coefficients (i.e., the a's)  relating
to a non-isotropic, non-homogeneous atmosphere depend upon  atmospheric
thermal  stability,  dynamic  wind  field,  aerodynamic  effects of  roughness
elements in the  boundary, and the time of exposure of  the  plume to these
conditions.

In the present model development a modification of the approach presented
by  Singer  and  Smith26  was adopted.  This method was  judged  to  be well
suited  for a computational  method using  Equation (2) as the basis  for an
algorithm. The standard deviations of the diffusion cloud  are given by:

                                                                    (3a)

                                                                    (3b)

                                                                    (3c)

and
where  u  is the  mean  wind  speed (cm  sec"1), o-  and p are dimensionless
parameters depending  on the intensity of turbulence and thermal stratifica-
tion of the atmosphere,  and Uox.°oy .CToz  (miles)  depend on  the character-
istic geometry associated with an individual  source of emission.

The values of a, a0x, a0y, and a0z were determined from parametric studies
using observed data. Integration of Equation (2) was performed for continuous
area sources for a number of combinations  of values of these parameters and
various  wind  speeds.  (The   method of  integration  will be  discussed  later.)
When observed concentrations were compared with those computed from the
integrations it was  possible  to delimit ranges for the values of the different
parameters. Starting with  the assumption that p = 1, it was found that values
of  0.05 «Ca< 0.25 gave  satisfactory estimation of SO2   concentrations for
various  atmospheric conditions.  It was also similarly determined that a0x
ff0y =  0.001 mile  and a0z= 0.01  mile for continuous area sources  in the
10-8
<7X =
ay =
Oz =
a0x + au (t
a0y + au (t
a02 + au (t
_t-)p
-t')p
_fjp

-------
metropolitan  New York  area. Figure 10-5 shows an example of computed
SOs concentrations in the midtown Manhattan  region for a  = 0.2 at two
different wind speeds. The points plotted  are measured S02 concentrations.
     0.3
  I
  =  0.2-
  o
  o
  o
  CM
  o
     0.1
          JAN. 30,1966
                           •1115-1242 (312'• 13 mph)
                           •1339-1434 (315°-16 mph)
                           A1436-1539(315°-16mph)
                 THEORETICAL:  u= 12.5 mph
                                « =  0.2
                                                      -u =15 mph
                                                       a =0.2
                   0.5
                                1.0
         1.5
2.6
                           I
                                   MILES
                                  	I
              I
             B'WAY
                        C.P.W.
5TH AVE    3RD AVE
     E.E.
     Figure 10-5.  Computed SC>2 concentrations in 79th Street
     Manhattan with « =0.2 for two different wind  speeds.

Since an inversion  layer aloft is  considered to  be impenetrable  by  S02
injected  below  it, the value of az  is taken as constant  when an  inversion
stratification is present. With the height of the base of the inversion layer as
H (mi), the  relationship, following Gifford,27 is estimated  as:
                H
           z     2  '

if az as computed by Equation (3c) is greater than H/2.
                                                                   (5)
Continuous Point Sources
Integration of Equation (2) over the time interval  [0,°°]  has been shown  by
Frenkiel28 to give the following expression for the mean concentration from
a continuous point source at ground level:
                                                                  10-9

-------
        C(x,y.z)  _      1      „..„   I  y2
                               -exp-	   +	               (6)
             Q        2-n oy az u       |_2
-------
C(x,x0, y-y0, z-z0,t,t0)   =   /    	_	
                           J     (2n)3'2  0z(t-t')
                          t=to
     exp
 r/z-w.(t-t')-z0V-j
 L    \    Oz(t-t')    /J
              /z+w- (t-t')+zo\i     /      /           1	
              \     az(t-t')    )\    J      J     ax(t-t') • ffy(t-t')
                                    x'=-a    y'=-b
     exp
        [x-x0 -u •  (t-t') -x' -\
             ax(t-t')      J
L         ajt-f)    J
dx'dy'dt'
                                                                     (10)
where:
       C  = concentration (g cm"3)
    Q(t')  = source strength (g cm"2 sec"1)
       t  = time, (sec)
    (t-t')  = time since emission (sec)
   (x',y')  = point of release measured from (x0,y0).
A schematic diagram representing the space referring to any point (x',yf)  and
(t-t'( is given  in Figure 10-6.
                                                                    10-1'

-------
        x=0
        y=0
                                                X-X0-U (t-t')'X'
MEAN POLLUTANT CONCENTRATION CONTRIBUTED FROM
AREA-SOURCE; SUMMATION OF CONTRIBUTIONS OF ALL x',
y ' AND t' FROM INFINITE INDIVIDUAL PUFFS.
   Figure 10-6.  Schematic diagram showing how plume is com-
   puted from area source.

In order  to  simplify the integration  of  Equation (10), the following condi-
tions were set:
     (1)  a = b;
     (2)  t0  = 0,  t =  2 hours;
     (3)  v = w = 0;
     (4)  u  is constant over (0, t);
     (5)  atmospheric stability and turbulent intensity are constant over (0, t)
     (6)  Q(t') = Q, a constant;
     (7)  x0 = Vo = 0.
In other words,  the  integration of Equation  (10) is over a 2-hour  interval
during which emission rates and atmospheric  conditions are steady, and over
an area centered  at  the origin  and  oriented  with one  side parallel to the
wind. These conditions along with  the transformations:
10-12

-------
       X =•
  x- u(t-t') - x'

    ffx(t-t')
                                                                    (11)
and
       Y =
    y-y
  CTV (t-f)
                                                                    (12)
reduce equation (10) to:

     2  hr
--  f
Q  J
      1
(27T)1/2 ffz(t-t')
                 exp
          x- u  •  (t-t') H-  a
                x

                f
                     exp	dX
          x-u •  (t-f) -a
            y + a

           y (t-t'l
     V^   f
exp     —   dY
           y-a
                                           dt'
                                                                    T13)
In order to  prevent accumulation  of  large errors in the double integration, a
polynominal representation of the Gaussian probability function was applied
for the  integration over  X  and Y.  To  perform the integration  over  t',  a
refined  two-part interval was used in a complicated application of Simpson's
Rule. This prevented further undue accumulation of errors.

The integration in Equation  (13) was carried out according to the procedure
                                                                   10-13

-------
above for values of u, a, and a, within their possible limits; for various values
of x, y,  and z;  for p=  1; and for z0 =  100 ft., the  assumed roof-top  level
characterizing metropolitan New York  residential and  small office buildings.
At this point in the development it can be stated:
      -
      UAre
              =  f(x,y, z, z0,a, p, u, a, t, t0).                          (14)
It  is to be noted that (x,  y)  now  represents the distance from the center of
the area source.
Adaptation of Continuous  Area  Source Solution
In order to apply the result of the integration in Equation (10), represented
by  Equation (14), to the model,  it is  necessary  to  adopt the assumptions
equivalent  to  those stated above. The 2-hour interval (0,  t) is  equivalent to
the emission interval, At, over which contributions from the various sources
are summed in the model. Emissions and meteorological conditions over the
entire urban region are assumed to remain steady over the interval, At.

The are no actual data to support a non-zero value for w, and the theoretical
knowledge  of  mesoscale phenomena, which would permit its calculation from
other meteorological  variables,  is limited. It therefore  seems reasonable to
assume w = 0  everywhere in the region.

The  assumption of v = 0 is equivalent to stating that the square area source
elements have one  side parallel to the mean wind. On the basis of computa-
tions on a  simple grid, errors resulting from this assumption are judged to be
small. With the above assumptions, Equation  (14) can  be used as the basis
for modeling multiple area sources in urban  localities. The quantity C/QArea
is  a  tabulated  function  and could  not be conveniently used in a numerical
model  in which  a large  number of area sources are considered. A set of
conditions  (some of which are approximations) that gave simple representa-
tions of  the area source solutions and required reasonable computation time
and memory storage  capacity was found. The representations also preserved
the features of the exact integration, with the conditions:

   1. J7~ =  0; 0  
-------
  3. For (t — t') > At, and p = 1 the area source solution is approximated
     by a  pseudo-Gassian  distribution. The  manner in which  this is  done
     will be discussed below.

  4. j?	=£  f(u). This condition will also be discussed below.
     UArea
Using conditions (1)  and (2) above, the solutions for area source emissions.
Equation (14) becomes:
                = f(x,y, a, p, u,a,t, t0)                              (15)
        Q
         •Area
Examination of the results of the integration in  Equation (13) reveals that
for  (t—t')  >  At the spatial concentration distribution approaches  a  shape
that might  be approximated  by a Gaussian distribution, and  the  maximum
concentration  is far less than when (t-t') < At. An example of this is shown
in  Figure 10-7. It follows then, that the computation of the concentration
field  resulting  from  any area source can  be divided  into two formats.  For
(t-f) < At, i.e., during the interval in which the portion of the plume  under
study  has  been emitted from  the source, one set of grid lines of the area
sources  is  assumed to be  parallel to the mean wind  and the concentration
field  is computed  directly  from the tabulated value of C/QArea based on a
chosen  value of a, and  p  = 1 (Taylor's hypothesis).  For   (t—t') >  At,  i.e.,
for portions of the plume that were in the  atmosphere at the beginning of
interval, At, the field of concentration from an area source is represented in
the following  manner:

The distribution of concentrations is first assumed to  be Gaussian.  Then it is
replaced by a simple  uniform distribution within a rectangular parallelepiped
with dimensions related  to  the  cr's of the Gaussian distribution. This process
is represented  by the equality:
                 Q
       _   (27rr/20xavaz
-00 -00 -00        A  y  £
              g(t)
                                                      dxdydz
          (4)3 (Jxaya2
                                                                    (15)
for   -2ffx
-------
            10
           io-i -
      v  10-2  =
      0  10*3  _
          10-4  -
                                  T
                      T
1    3
              u = 8.00 mi. hrl
              a = 0.2
= 2.0hr        AREA SIZE 1.0 x 1.0 mi
              doz = 0.01
              x0 = 0.5 mi
              0     10     20     30     40     50     60     70
                                   x, miles

  Figure  10-7.  Time  history of a plume from a continuous area-
  source.
           ax   =  aox  +  au(t-t')  +  % a +  1/4 u  (-t')

           CTy   =  CT0y  +  aU (t -t')  +  % a

           CTZ   =  a0z  +  aU(t-t')

           azO   =  CTOy
                                                (17)
With  the  above  conditions  (3)  the dependency of the  tabular function
    Area on P and the time, t and t0, are removed and:
                  rea  =  f (x, y, a, U, a).
                                                 (18)
10-16

-------
Figure 10-8 illustrates the nature of this function for different values of the
parameters.
                                    U = 10.0 mph
                                    x0 = SCALE OF AREA,SIZE/2
                                   -0.1
                                 SIZE-1.0x1.0 mile
                                    a=0.2
                                   SIZE =1.0x1.0 mile
                                              a=0.1
                                              a = 0.15
                                              a = 0.2
                                              a=0.25
                                           SIZE =0.5x0.5 mile
                                  I
              0     0.5     1.0    1.5    2.0    2.5    3.0     3.5
                                   x, miles

    Figure 10-8.  Downwind distribution of pollutant from area-
    sources, .parameterized by values of a.
Finally, by studying the integrated  solution of Equation  (13), it was  found
that:
          Cu
                 = f(x, y,a) a).
(18)
This  is condition (4) above and  is  illustrated  in Figure  10-9. This  was a
significant finding for it  enables the model  to  handle a  number of different
source grid dimensions, a, while including the  effect of various wind speeds.
Using  all of  the conditions discussed above. Equation (10) can be  summed
over all time intervals of duration  At  to obtain the contributions  to the
concentration of all  "puffs"  (each emitted during a different  interval) at the
point  (x—x0, y—y0)  from an  area  source located at (x0,yo). The result is:
                                                                  10-17

-------
o
CO
LINES OF CONSTANT Cu/Q:  y=y0, « = 0.2, (Joz =0.01, x0 =1.0 mi        AREA SIZE 2.0 x 2.0 mi
                      15
                      10
o-> ^
s;
•^ 
^
3 C
3 C
--— ^
X 2
O e«
r o
3 C
3 C
—^
(Si:
3 t
T P
3 C
3 C
—- --
^E)
3 i
3 C
-~^1
f3.:
3 g
3 C
^-^
3ff
^ ^
S £
3 C
—~~
Kut
H 1 1
•
>
+ ut

                        ' 0  1.0 2.0 3.0 4.0 5.0
                               10.0             15.0
                                      x, miles
20.0
25.0
                              Figure 10-9.  Illustrating that Cu/Q is independent of wind  speed.

-------
                  nAt
c(x-x0,y-y0-nAt) =
                  t-nAt+At
         •z    /
         t=nAt   t-nAt
                                        2Q
               D

                                          (277)3/2 CT,(nAt-t')
     exp  -
                az (nAt-
]    •/    /-,
                                                          1
                                          (nAt-t') av(nAt-t')
                                    x =-a  y=-a
     exp
r   j /x-x0-u(nAt-t')-x-Y + /
L   |\  ax (nAt-f)     /    \av
                                        y-y0-y-
                                          (nAt-f)
                         ]'
                                                         dx'dy' dt'
                                                                  (20)
In this equation, N is the number of time intervals from the beginning of all
emission to the  present. When n =  1, the exact solution to Equation (10) is
used, while for all of the remaining terms the center of gravity of the puff is
advected according to the wind,  and the pseudo-Gaussian distribution of
Equations  (16) and (17) are  applied to compute the spatial distribution of
concentrations. Finally, to complete the  model construction. Equation (20)
is summed over  all x0  and y0 to obtain the concentrations contributed to a
point  (x,y) by all continuous area  sources. This summed equation plus an
equivalent  one for point sources plus the tabular function  Equation (19)
represent the  algorithm  for  the  multiple-source  model of dispersion  and
transport.

MODEL CONFIGURATION
To complete the model,  a value of a, remained to  be fixed, a grid represent-
ing area  source emissions to be constructed and  the computation grid to be
specified. Based  upon the parametric studies mentioned in earlier values of a
in the range 0.1 to 0.25 were chosen, depending upon the thermal  stratifica-
tion.
Condition  (4),  in the proceeding  section  permitted simultaneous use  of
several different values of a, the dimension  of the area-source grid.  Figure
10-10 shows the area-source  grid. The values of the dimension a used were
0.5, 1.0, 2.0, and  5.0  miles.  The grid size at  any particular location depends
upon the source strengths and their spatial distributions, and on topographi-
cal and  geographical factors.  Generally the smallest grid  size occurs  in  the
central portion of the region where the annual emissions  per unit area  are
the highest (Figure 10-2).
                                                                 10-19

-------
o
NJ
          Figure 10-10. Area-source grid.  Shaded areas represent zero emissions.  Grid size depends on source
          strengths and spatial distribution, geographical and topographical factors.

-------
Two computation grids (referring  to x,y) with mesh sizes of 0.2 by 0.2 mile
and 1.0 by 1.0 mile were  used. The time interval. At, was 2 hours.

COMPUTATIONS  AND VALIDATION
The computer methodology used in the computations representing the model
is described elsewhere.8
The model was used  to compute surface  concentration patterns of SO2  for
the two test periods March 8-9, 1966, and November 15-16, 1966.
The computations were carried out by a  CDC 6600 computer. The portion
of the program dealing with  winds and  trajectories required  3 minutes of
computer  time for  24 time steps (48 hours) for source  grid system consisting
of 712 points  (565  area sources and  148  point sources). The  diffusion
portion of the program computed  the concentration field at more than 2000
points for each of the 24 time steps in a total  computer time of 19  minutes.

Some  exemplary  results  are  shown  in  Figures  10-11 to 10-18,   maps of
concentration fields of S02 over Manhattan for various times during the test
period, each using  a 0.2 by 0.2 mile computational grid. Isolines with values
less than  0.1 ppm are dashed. Zero time was at 00 hours on March 8 and
November 15.  Figure  10-19 is a map of a larger area computed  for the same
time  as Figure  10-14, but  using a 1.0 by 1.0 mile computation grid.  It is
clear  that the  coarser grid  computation may  miss several important small-
scale features.
                                                                10-21

-------
                                        0800 MARCH 8,1956
    Figure 10-11.  Predicted SO2 concentration field  (ppm), 0800
    March 8, 1966, 0.2 x 0.2 mile computational  grid.
10-22

-------
                                     000 MARCH S, 1966
Figure 10-12. Predicted S02 concentration field  (ppm), 1000
March 8, 1966, 0.2 x 0.2 mile computational grid.
                                                         10-23

-------
                                      1800 MARCH 8, 1966
   Figure 10-13.  Predicted SO2 concentration field (ppm), 1800
   March 8, 1966, 0.2 x 0.2 mile computational grid.
10-24

-------
                                  800 MARCH 9, 1966
Figure 10-14.  Predicted S02 concentration field (ppm), 1800
March 9, 1966, 0.2 x 0.2 mile computational grid.
                                                         10-25

-------
   Figure 10-15.  Predicted SC>2 concentration field (ppm), 1200
   November 15, 1966, 0.2 x 0.2 mile computational grid.
10-26

-------
                                         5, 1966
Figure 10-16.  Predicted S02 concentration field (ppm), 2400
November 15, 1966, 0.2 x 0.2 mile computational grid.
                                                       10-27

-------
   Figure 10-17.  Predicted S02 concentration field (ppm), 1200
   November 16, 1966, 0.2 x 0.2 mile computational grid.
10-28

-------
rigure 10-18.  Predicted S02 concentration field (ppm), 1800
November 16, 1966, 0.2 x 0.2 mile computational grid.
                                                       10-29

-------
CO
o
                               Figure 10-19.  Predicted SC>2 concentration field (ppm), 1800
                                             March 9, 1966, 1.0 mile x 1.0 mile computational
                                             grid.

-------
                           1.0        1.5        2.0
                            DISTANCE FROM HUDSON RIVER, miles
           R.D. N.E.B'NAYAMST. COLAVE. C.P.W.
                                        SlhAVE.  PARKAVE.  3RD 2ND  1ST YORK  E.E.
Figure 10-20.  Observed  and predicted  variation  of SC>2 concentra-
tion  along the  79th Street  crosstown  transect  at  the  indicated
times.

 A comparison of  the  computed  results  with observed  concentrations  is
 shown in  Figures 10-20  to  10-23 that are crosstown transects along 79th
 Street, of the concentration of SOa computed on a 0.2-mile grid. The lines
 represent the  computed quantities while the observations are  plotted as
 points.
                          1.0         1.5        2.1
                            DISTANCE FROM HUDSON RIVER, miles
          R.O. «.E. B'»AY AMST. COL AVE C.P.W
                                       5TH AVE.  PARK AVE. 3RD ZNO 1ST YORK E.E.
Figure 10-21.  Observed and predicted  variation  of S02 concentra-
tion  along  the  79th Street  crosstown  transect at  the  indicated
times.
                                                                    10-31

-------
     BO.IS
     g
            • OBSERVED DATA
             MARCH 1. 1966
             092) • 1025
                            1.0         1.5         2.0
                              DISTANCE FROM HDDSON RIVER, miles
            R.O. ».E.B'««VAMST. COL. AVE.C.P.H.
                                          5TH AVE.  PARK AVE. 3RD 2ND 1ST YORK AVE E.E.
Figure 10-22.  Observed  and predicted  variation  of 862 concentra-
tion  along  the  79th  Street  crosstown  transect at  the  indicated
times.
           0.45

           0.40


           0.35

         E0.30J


         10.25|

         x 0.20
         UJ
         u
         S0.15|
         
-------
While  agreement is not  perfect, the computed  concentrations bear a discern-
ible relationship  to  the  observed concentrations.  Figure  10-24 shows a
transect for the same time as Figure 10-12, but it was computed using a  2.0
mile square area source grid  over the entire  region.  It is evident that  such a
course representation of the area sources causes severe smoothing and dis-
tortion of the computed concentration field.
           I OBSERVED DATA
            MARCH 8,1961
            0927 -1025
                                          PREDICTED AT 1000 HR
                           1.0        1.5         2.0
                            DISTANCE FROM HUDSON RIVER, miles '
           R.D. ».E.B'»AYA'KT.COL. AVE.C.P.*.           5TH AVE PARKAVE.  3RD 2ND 1STYORK AVE.E.E.
 Figure 10-24.   Observed and  predicted variation of SC>2 concen-
 tration along 79th Street crosstown transect,  same period as Fig-
 ure 10-12,  coarse  grid  2.0 miles  x 2.0 miles.
An objective estimate of the relationship between  computed and observed
concentrations for  the two  test periods  is  illustrated  in Figure 10-25, a
scatter  diagram  of  observed versus  computed  values.  The  mean  of the
observed concentrations is 0.19 ppm, and the  mean of the  computed concen-
trations is 0.18 ppm.  There are 121 data points in the diagram. The standard
error of estimate is 0.09 ppm.

The situation depicted in Figures 10-14 and  10-19 for 1800  hours March 9
arose  because of  a  complex wind  field with  a  region of convergence over
Manhattan  Island. The wind field is shown in  Figure 10-26. Because the SO2
concentrations were computed using w = 0 everywhere, the values of concen-
trations shown are higher in the region of convergence than they would be if
provision were  made for non-zero  values  of  w. The concentrations  in the
center  of the convergence  region are about twice as great  as they should  be.

A  final point, illustrating  the  potential usefulness of the model,  is seen  in
Figure  10-27, which shows the  1000-hour,  March 8 SO2  concentration field
resulting from area source emissions only (compare with Figure 10-12).
                                                                   10-33

-------
       0.6
        0.5
    o  0.4
    LU
        0.3
    o
    o
     CNJ
    o
    in
    S  0.2
    if)
    CO
    o
        0.1
                           •        .   X
.X.'«    *     •
     *    .
                   0.1       0.2       0.3       0.4
                     PREDICTED S02 CONCENTRATION, ppm
                         0.5
   Figure 10-25.  Plot of observed against predicted concentra-
   tions of SC>2. Mean observed and predicted concentrations,
   0.19 and 0.18 ppm  respectively.  Standard error of estimate
   is 0.09 ppm.

DISCUSSION

A  mathematical  model  for computing  concentration  in  a  multiple-source
urban region has been developed. The  model is three  dimensional; accepts
input data of meteorological conditions and source strengths that are func-
tions of space  and time; and  computes pollutant concentration fields on a
fine mesh  grid  (0.2  mile-square) over a 50-mile-square region at any given
time. The  framework of the  model is  general enough  so that the vertical
component  of the wind  could  be  included  if it were  known. To do this,
however,  the part of the algorithm giving the integrated expression for a
continuous  area  source  in  tabular form  (Equation  14} would have to be
modified.

The adaptation of the model for the computations of S02 concentrations in
the New  York  urban region involved assumptions of w = 0 and immediate
10-34

-------
uniform  mixing of  area  source  emissions  between rooftop  level  and  the
ground.  No  removal factor was included. SO2  reaching the  ground was
assumed to be reflected. These assumptions are, of course, possible causes of
error  in  the  computed concentration fields.  Other sources of error are:
coarseness of  the  meteorological observational network  (Figure 10-1); poor
data for source strength and inventory, particularly  for manufacturing plant
stacks; and unaccounted for topographical effects on dispersion.

Systems  of the kind  developed in  this work are rarely  amenable to a
posteriori error analysis. The algorithm is based upon semi empirical theory,
not derivable  from known physical  concepts. The sources of error as listed
above are generally  not  readily controlled, and the assessment  of their
individual effects is difficult if not impossible.

The present status  of the model  is one of promising preliminary verification.
Certainly a new effort  in obtaining high-quality test  data for  periods up to 3
or 4 days in duration  would be  desirable. With such data further parametric
studies could lead to improvement of the model.
   Figure 10-26.  Wind  field 1800 March 9, 1966.  Streamlines and
   isotachs represented by solid and broken lines respectively.
   Convergence zone over Manhattan Island, shaded.
                                                               10-35

-------
                                        000  MARCH 8, 1966
    Figure 10-27. S02 concentration field, predicted from area
    sources only.
10-36

-------
REFERENCES


 1.  Davidson, B. A Summary of the New York Urban  Air  Pollution Dynamics Re-
    search Program. J. Air Pollution Control Assoc. 17:154-158, March 1967.
 2.  Davidson, B. et al.  Mathematical  Models  of  Urban Air Pollution Dynamics:  Final
    Report.  Vol. 2. New York University. 1969.
 3.  Scudder, B.  Diagnosing the Mesoscale Wind Field over an  Urban Area by Means of
    Synoptic Data. New York University. M. S. Thesis. 1965.
 4.  Druyan,  L. M. A Comparison of  Low-Level  Trajectories in an Urban Atmosphere.
    J. Appl. Meteorol. 7(4) :583-590, August 1968.
 5.  Bornstein, R. Observations of the Urban  Heat Island Effect in New York City. J.
    Appl. Meteorol. 7(4):575-582, August 1968.
 6.  Sutton,  O. G. Theory of Eddy  Diffusion  in the Atmosphere. Proc. Roy. Soc., Ser.
    A (London).  735:143-165, February 1,  1932.
 7.  Frenkiel, F  N. Application of Statistical  Theory of Turbulent Diffusion to Micro-
    meteorology. J. Meteorol. 9(4):252-259, August  1952.
 8.  Ogura, Y. Theoretical Distribution  Functions of Matters Emitted from a  Fixed
    Point. Meteorol. Soc. Japan J. 32(Series II), January  1954.
 9.  Batchelor, G. K. Diffusion  in  a  Field of Homogeneous Turbulence. I. Eulerian
    Analysis. Australian  J. Sci. Research. 2(4) :437-450, December 1949.
10.  Batchelor, G. K. Diffusion in a  Field of Homogeneous Turbulence II. The Relative
    Motion  of Particles. Proc. Cambridge Phil. Soc. (London).  48:345-362, April 1952.
11.  Hay,  J. S. and F. Pasquill. Diffusion from a Fixed Source at a  Height of a Few
    Hundred Feet in  the Atmosphere.  J. Fluid Mech. 2:299-310, May 1957.
12.  Cramer,  H.  E., F  A. Record, and H. C. Vaughn. The Study of the Diffusion of
    Gases or Aerosols  in the  Lower Atmosphere. Massachusetts  Institute of Tech-
    nology.  Cambridge. AFCRC-TR-58-239. 1958. 70 p.
13.  Barad, M. L. and D. A. Haugen.  A  Preliminary Evaluation  of Button's Hypothesis
    for Diffusion from a Continuous  Point  Source. J.  Meteorol. 76(1):12-20, February
    1959.
14.  Lin, C.  C. and W. H. Reid. Turbulent Flow, Theoretical Aspects. In: Handbuch der
    Physik,  Flugge, S. (ed.),  Vol. 8/2.  Berlin, Springer-Verlag, 1963. p.  438-523.
15.  Moses,  H.  Mathematical Urban  Air Pollution  Models. National  Center  for Air
    Pollution  Control,  Chicago  Dept.  of  Air Pollution  Control, Argonne National
    Laboratory.  Argonne,  III. ANL/ES-RPY-001. April  1961. 69 p.
16.  Taylor,  G.  I. Diffusion by  Continuous  Movements.  Proc. London Math.  Soc.
    20(Series 2):196-212,  August 20,  1921.
17.  Holland, J. Z. and R. F  Myers. A Meteorological Survey  of the Oak Ridge Area.
    U.S. Weather  Bureau, Oak Ridge, Tenn. U.S. Atomic Energy Commission, Tech-
    nical  Information Service, Final Report for  1948-1952. Report  Number ORO-99.
    November 1953. 584 p.
18.  Kellogg,  W.  W.  Diffusion  of  Smoke  in the  Stratosphere. J. Meteorol.  73(3):
    241-250, June 1956.
19.  Frenkiel, F   N. and I.  Katz. Studies of  Small  Scale  Turbulent  Diffusion  in The
    Atmosphere.  J. Meteorol. 73(4) :388-394, August 1956.
20.  Gifford,  F.  A.,  Jr. Atmospheric  Diffusion  from Volume Sources.  J. Meteorol.
    72(3):245-251, June 1955.
    Gifford,  F.  A., Jr.  Relative Atmospheric  Diffusion of Smoke Puffs. J. Meteorol.
    74(51:410-414, October  1957.
21.  Gifford,  F.,  Jr. Statistical Properties of a Fluctuating  Plume Dispersion  Model. In:
    Advances in  Geophysics, Landsberg, H.  E. and J. Van Mieghem (eds.), Vol. 6. New
    York, Academic Press, 1959. p.  117-138.
                                                                          10-37

-------
22.  Inoue,  E.  On the  Shape of Stack Plumes.  National Institute of Agriculture Sci-
     ences, Division of Meteorology, Meteorol.  Res.  Notes. Japan. Vol. 2, No. 5. 1960.
     p. 332-339.
23.  Hogstrom, U. An  Experimental  Study on Atmospheric  Diffusion.  Tellus.  16(2):
     205-251, May 1964.
24.  Pasquill, F. The Estimation of the Dispersion  of Windborne Material.  Meteorol.
     Mag. 90( 1063) :33-49,  February 1961.
25.  Pasquill, F  Atmospheric  Diffusion. London, D.  Van IMostrand Co. Ltd., 1962. 297
     P.
26.  Smith,  M.  E.  and I. A. Singer. An Improved Method of  Estimating  Concentration
     and  Related  Phenomena  from a Point Source Emission.  J. Appl. Meteorol. 5(5):
     631-639, October 1966.
27.  Gifford, F. A., Jr. Atmospheric Dispersion. Nucl. Safety.  7:56-62, March 1960.
28.  Frenkiel,  F.  N.  Turbulent Diffusion:  Mean  Concentration Distribution in a Flow
     Field of Homogeneous Turbulence. In:  Advances in Applied Mechanics, von  Mises,
     R. and  T.  von Karman (eds.), Vol. III. New York, Academic Press  Inc., 1953. p.
     61-107.
29.  Brummage, K. G. et. al. The Calculation of  Atmospheric Dispersion from a Stack.
     Stichting, CONCAWE, The Hague, Netherlands. August 1966. 57 p.
30.  Simon,  C.  Plume Rise and Plume Concentration  Distribution from  Consolidated
     Edison  Plant in  New York City.  New York  University. Technical Report Number
     68-15. 1968.
31.  Halitsky, J. (unpublished report)  Effect of  Turbulence on Gas Diffusion Around
     Obstacles. New York University. 1965.
10-38

-------
APPENDIX - GLOSSARY OF SYMBOLS

a         dimensions of area-source grid
C         concentration
DD       number of degree days for date of interest
g(t)       total source output for a 2-hour period
h         height of stack from which pollutant is emitted
Ah        distance  above  stack  through which  plume rises  by bouyancy
          and exit velocity of stack gases
H         height of base of inversion layer
3f         effective stack height
N         number  of time intervals from  beginning  of  all  emissions to
          present
p         data based parameter to  describe  intensity of turbulence
Q         source strength
Qh        heat emission rate
Qday      daily total output of SO2
Qyear     annual total output of S02
t         time
At        time interval
(t — t')    time since emission
u, v, w    velocities in x, y, z directions respectively
(x'( y', z') position of release
a         data based parameter to  describe  thermal stratification
ox,  oy, oz standard deviations of diffusing cloud in three dimensions respec-
          tively.
                                                                  10-39

-------
ABSTRACT
      This paper presents a statistical approach  to the analysis of multiple-
      station  air  sampling data  in  urban (i.e.,  multi-source}  areas. Measured
      data values are considered to be composed of three contributions: an
      urban scale constituent, a contribution due  to sources in the vicinity of
      the  sampling  station, and a contribution  due  to  measurement and
      analysis  errors. The significance  of correlation coefficients  calculated
      from paired data series and the influence of averaging several individual
      data series prior to calculating correlations is examined.

      The data  base for  the present  study  is  limited,  but the following
      conclusions can be cited:
      — The error contribution to the total variance  in most field data series
      is significant.
      —Correlations among single station data series, and between data from
      one  station and an  external  predictor (such as a meteorological para-
      meter), will be small in most cases. Errors and local phenomena in the
      data from a single station mask much  of  the urban  scale  variability
      that would correlate with most urban scale model predictors.
      —An urban scale average,  drawn from widely separated stations,  is the
      best index of large scale air quality patterns.
      —A maximum  expected correlation coefficient can be  calculated from
      the field data, and this value can serve as a  criterion for the evaluation
      of air quality model projections.
      —Whenever possible,  a double set of instrumentation should be  op-
      erated at one  station  within  a sampling network.  The redundant data
      provide an excellent measure  of the expected error  variance in the field
      observation.
AUTHORS
      JAMES R. MAHONEY is Assistant Professor of Applied Meteorology at Harvard
      University, School  of Public  Health, and staff consultant  at  Environmental
      Research  and  Technology,  Inc.,  Waltham, Massachusetts.  He received a B.S.
      degree in physics from LeMoyne College (New York) in  1959 and a PH.D. degree
      in meteorology from Massachussetts Institute of Technology  (M./.T.) in 1966.

      WILLIAM  O.  MADDAUS is Water Resources Systems Engineer at Engineering-
      Science,  Inc.,  Oakland, California. He received a B.S. degree in civil engineering
      from the University of California at Berkeley  in  1967,  and an M.S. in civil
      engineering degree, together with a C.E. degree from M.I.T. in 1969.

      JOHN C.  GOODRICH is a  teaching assistant in City Planning at Harvard Univer-
      sity, Graduate School of Design, and staff consultant with Environmental Research
      and Technology,  Inc., Waltham,  Massachusetts.  He  received the B.S.  in civil
      engineering degree  from  Princeton University in  1966, and the  Master of Re-
      gional Planning degree from Harvard in 1969.

-------
                     11.   ANALYSIS OF  MULTIPLE-STATION
                                URBAN  AIR  SAMPLING  DATA
        JAMES R.  MAHONEY, WILLIAM O. MADDAUS
                  AND JAMES C. GOODRICH

                        Harvard University
INTRODUCTION
Air  pollution  control depends upon  adequate current data and  projections
describing  concentrations  of the important pollutants within  the  control
region. Average concentrations, temporal and spatial  variability, and  concen-
trations during poor ventilation conditions must all be known.

Recently  a  large investment in  time, effort and  financial expenditure has
been committed to  the collection of multiple-station data in urban areas.
Within the United States, several of the largest  cities have instituted con-
tinuous monitoring for important local pollutants:

-The  Los Angeles  County  Air  Pollution  Control Department instituted a
program  in  1947 Nine stations,  all equipped to monitor carbon monoxide,
oxides of nitrogen,  ozone, hydrocarbons, sulfur dioxide and particulates, are
operated throughout Los Angeles County.

-New York City recently activated a 38-station network, under the direction
of the Air Resources Department. Each station records concentrations and
surface meteorological  data, and ten of the stations transmit data regularly
to a central, computerized  facility.
-Chicago has  an active monitoring  and projection program,  operated with
the cooperation of Argonne National Laboratories.
Many  other cities have operated programs  in  response to local pollution
                                11-1

-------
problems. Pittsburgh, for example, has collected smoke shade data for more
than 40 years. Its monitoring  program was  instituted to support that city's
efforts to achieve a smokeless atmosphere.

In response to the  requirements of the Air  Quality Act of  1967, all of the
states  are  currently  establishing  monitoring  networks  within  federally
designated air quality control regions. All of the data collected is useful only
when it is  properly analyzed and made available for study and  decision-
making operations.  This paper  presents  a  brief analysis of the  uses and
limitations of multiple-station air quality data collected in urban areas. The
illustrative examples cited are  drawn from  an  analysis for a  ten-station
S02   survey  and   a   20-station  particulate  survey  conducted   in  the
metropolitan  Boston region in  1966 by the Massachusetts  Department of
Public  Health.

The  section  headed  "Design Specifications  for a Multiple-Station System"
summarizes the elements required for  adequate specification of  a  sampling
network.  The  section  headed   "Interstation Data   Correlation  Studies"
contains  the results  of  several  interstation  correlation  studies  for the
Boston  data.  These  results  illustrate the   degree of variability  between
simultaneous observations in various parts  of the  same urban region. The
section  captioned  "Correlations  between Air Quality Data  and  Meteoro-
logical  Data" summarizes  a  correlation study of the same  air quality data
related  to  local  meteorological  data.  Such correlation statistics  must be
developed to facilitate projections for special  meteorological  conditions and
for changes  in  local emission  rates. The section  captioned  "The Use Of
Computer Generated Graphic Analyses for Air Quality  Data" illustrates the
use of computer generated graphic displays  for the presentation of regional
air quality data,  and the  concluding section contains recommendations for
the establishment and  analysis of multiple-station air  sampling networks.


DESIGN SPECIFICATIONS FOR A MULTIPLE-STATION
SYSTEM
Before a  monitoring system can be established, several basic questions con-
cerning its purposes must be resolved. The system criteria can be developed
from an  evaluation  of the following topics.  (Note that very similar  specifica-
tions  are required  for the development of diffusion  models intended to
provide urban area concentration data.)

   1.  Basic  uses  of the  data.  Information  concerning  long  term (annual)
      averages, monthly  and  seasonal variability,  day-to-night  variability,
      background levels, and frequencies  and  magnitudes of peak values will
      normally  be   desired.  Also, observed  concentrations  during defined
      periods of poor atmospheric mixing are usually required.
11-2

-------
  2. Spatial resolution.   Is it sufficient to report an average for  the entire
     urban region, or for the "urban center" region? Will it be necessary to
     describe the variability  among  sub-regions of the  urban area, and will
     an  attempt be  made  to  correlate  local  concentrations  with  local
     emission data?

  3. Time  resolution.  Many instruments are capable of near-instantaneous
     observations, and a number of observing systems  are providing data at
     5-minute   intervals.  With  few  exceptions,   hourly   records   of
     contaminant  concentrations provide an adequate description of  mean
     values  and  variability  in  the  urban   atmosphere. Higher  frequency
     data reports usually  contain  significant noise caused by  instrument
     errors  and  transient  changes   in  local  concentrations.   For  many
     purposes   daily   average   values   are  adequate,   and   calibration
     difficulties  frequently   appear   when  the   results  of  24-hour
     observations are compared to corresponding averages  of 24, one-hour
     observations.

  4. Projections  based   upon the  data.  The required projections maybe
     either short-term forecasts for  a single-day or a several-day episode, or
     long-term  planning estimates,  based upon  expected  changes in local
     emission  sources or rates. When  projections are desired, the air quality
     observations  should be supplemented  by surface  and vertical profile
     meteorological data.  The accuracy requirements  for  all data  become
     more stringent when longer  projections are  prepared because  small
     errors in time tendencies or spatial gradients will be magnified by the
     extrapolation process.
  5. Emergency action  decisions based upon the data. Some redundancy in
     measurements is desirable  for  the system if the  output data is  to  be
     employed as the basis for wide-ranging emergency decisions.

  6. Compatability with data from  other regions.  Although every reason-
     able effort  should be  made to  develop uniform data collection and
     reporting  programs,  the physical variability  within  each  urban area
     and  the  rapid development of  observing instruments  make complete
     standardization difficult.

When the purposes of the sampling  network  have been  resolved, the detailed
requirements for the network can be estimated: How many observing points
are needed; how should  they be located relative to  one another and to the
local geography; what sampling  frequency and accuracy are necessary? The
following sections  are  intended  to  provide  some  empirical information in
response to these questions.
                                                                    11-3

-------
INTERSTATION  DATA  CORRELATION STUDIES

When several observing  sites  are used to characterize air quality in an urban
area, the variations in the  resulting data represent a mixture of instrumental
error,  local (or transient)  phenomena,  and large  scale variability  in  con-
taminant  concentrations.  The  best  use  of the  field data  results  when
estimates are available  concerning the  relative magnitudes  of  the observing
errors, local phenomena, and  urban-scale phenomena.

This section describes an approach to such estimates, based  upon correlation
studies  and averaging  techniques applied  to  the  original  field data.  The
analysis proceeds from  the formal  definition of the correlation coefficient,
and it assumes that actual correlation  statistics, calculated  from field data,
represent useful approximations to  the  "true"  statistics  which would be
developed from very large data sets.

The  illustrative  examples  cited  in  this section are  drawn from  the 1966
Boston  Region  Air Quality Study, conducted by the Massachusetts Depart-
ment  of Public  Health.  A  general description  of  this  study and  its results
appears in  a state report;3 only the SO2  and  smoke shade (Coefficient of
haze, Coh)  data are cited here.

Ten S02 observing  locations were operated during the study period. These
ten sites and ten additonal sites were employed for the collection of smoke
shade  data. The  Boston  region base  map in  Figure 11-1  indicates the
locations of all  20 sites.

In the case of  the SO2  data, four-hour air samples were obtained,  and the
West-Gaeke method1  was  employed for all  analyses. The monthly average
diurnal  variability  for the  ten-station average S02  data is  shown in Figure
11-2 for the months of February, March, September and November.  The
observed diurnal pattern has  the expected features:  maximum concentrations
in the early morning  hours, and minima  in  the  late afternoon.  For the
"winter" months, both  the  average  concentrations  and  the  average daily
variability are greater. (December and January data samples  were unavailable
for use in  the  comparative study.) The November  concentrations are prob-
ably higher than normal because of a prolonged stagnation  episode between
November  21, and 25, that year. Concentrations in the mid-summer months
were  all  lower than  the September data illustrated in the figure, and no
diurnal pattern  is evident for  these months.

It should  be  noted that  the  concentrations observed on individual days
frequently  do  not  follow  the  pattern   of  the  monthly averages shown  in
Figure 11-2.  Mesoscale  meteorological  phenomena,  having the time periods
of one day or  less,  mask  the  tendency for day-time mixing and nighttime
stability on many days.
11-4

-------
 BOSTON METROPOLITAN AIR POLLUTION CONTROL DISTRICT (PRIOR TO 1969)
 pi
 U
X - S02 AND SMOKE SHADE
0- SMOKE SHADE ONLY
     SCALE
    IN MILES
  Figure 11-1.  Boston region base map indicating the ocean
  coast and the boundaries of the  Boston Air Quality Control

  Region prior to 1969.
                                                           11-5

-------
 BOSTON TEN-STATION AVERAGE S02 DATA -1966 AVERAGE DAILY VARIABILITY
          10.0
     CVI
     O
     in
                         4-8
8-12        12-16
  TIME OF DAY
16-20
20-24
      Figure 11-2.  Average diurnal variability for the ten-station
      Boston region SC>2 data, 1966.
The  statistical  analysis  of the S02  and smoke shade data is  based  upon
calculated correlation coefficients for data series from individual  stations,
and groups of stations.  The correlation coefficient R, for two data series x,
and yjf having N members each, is
                               S[(Xj -  x)  (Vi -  y)]
            R  = R
                  x,y
                        VE [ (x,  -  x)2 ]   •  2 [ (y, - Y)2
                                (1)
where the  overbar indicates the arithmetic average value for the  series.  The
summation limits are i =  1 through i = N in each case.
11-6

-------
For this discussion  the symbol V  indicates the variance of a series, and C
represents the covariance between two series:
                Vx  - 	fj	                                (2)
                       S[(x,  -  x) (y,  -  y)]
              Cx,y   =              N<3>

Then the correlation coefficient can be defined in terms of the variance and
covariance forms
                *,y     v      V
                       vx     vy
For this discussion the actual value of each field observation  x, (or y,) in a
time series is assumed to arise from an urban-wide contribution, a local scale
contribution, and error due to measurement and analysis:

                x,  = U,  +  Lf  +   El                               (5)

where

            X|   = a  measured  value of S02 concentration or smoke shade
                  at one observing site,
            Uj  = contribution  to X;  which  is  constant throughout  the
                  urban  region,
            Li  = contribution  to x,  arising  from sources  in  the local
                  region of the measuring site,
            E*  = contribution  to X; due to  errors  in measurement  and
                  analysis.

By the same definition  an element of the time series y,  measured at another
site within the urban area  is

             y,  =  U,  +  L?   +  Eyt                                (6)

Note that the urban scale contribution is equal  (by definition) at each site.

With the  aid of  Equation (5)  the variance and  covariance of  the x, and y,
data series can be expressed in terms of the defined urban, local and error
contributions:
                            2
                 S[(x,  -  x)2]
           V*  ~        N
                              VEx + CU|Lx  + CUiEX + CtxiEX            (7)

-------
But  if the  error contributions are  random (i.e. uncorrelated with the urban
plus  local "real" variability, the Equation (7) reduces to

             Vx =  Vu  + VLX + VEX  +  CU|LX                    (8)

A  similar result is obtained for the time series represented by the y, terms:

             vy =  Vu  +  Vi_v  +   VEV   +  CU,LV                  0)

The covariance between  series x, and  y, is

             CX|V
                                              N

                  =  Vy + CLx,Ly + CU|Lx + CUiLy + CLx,Ey         (1Q)

                     + CLy,Ex + CUpEy + Cu>Ex + CEx_Ey

Again, if the error contributions are uncorrelated to the real contributions,
this becomes

             Cx,y = Vy  +  CLX|Ly  +  CU|Lx  +  CU|Ly           (11)

Now  Equations  (4),  (8), (9) and (11)  can be used to write the correlation
coefficient for the x,  and y,  series as
                                                                   (12)

                       _
                       (vu+vLX+vEX+cU|Lx)1/2(vu+vLy+vEy+cU|Ly)I/2
The remainder  of this section is concerned with modifications of the general
form expressed in Equation  (12).

Case 1. Data  from  two sampling instruments  operating at same
         site.
Two sampling instruments operated side-by-side can be  used  to estimate the
magnitude of random measuring errors associated  with the  instruments. In
this case the "urban"  and  "local" contributions to the  measured concen-
trations  are equal  for each  instrument. If the symbol T| represents the true
concentration at the  observing  site, then T; = DI + Li( and  Equation (12)
reduces to the simpler, familiar form

                   __ VT _
              R*'*  ~  (VT + VEX)lfl  (VT  +  VEV)1Q              (13)
 11-8

-------
If the magnitude of the error variance associated  with each instrument is
assumed equal (i.e. Ex = Ey =  E), then

                     _      VT
                Rx,y -  (VT+ VE)                                  (14)

Thus,  for two instruments operating at  the same  location,  the correlation
coefficient  for  the paired  data series represents the  ratio  of the  "true"
variance to the total (true plus error) variance.

As  an example  of this analysis,  two  tape stain  samplers  were  operated
side-by-side on the roof of a 50-foot-high building  at the Harvard School of
Public Health during part of  1968. Coh data obtained from the instruments
operating on identical 2-hour  sampling  cycles during the months of June
through August are summarized in Table 11-1.

If  a value  of R  = 0.75  is  adopted from Table  11-1 as typical  for  the
sampling period, then the  expected ratio of error variance to true variance
for  these instruments can be calculated. Equation  (14) can be rewritten as

             YE =1
             VT      R  '   •

or                                                                  (15)
             VE =  1
             V7   T   ,

based on data of Table 11-1. This significant  contribution due to measuring
errors must always be considered,  particularly when the data  from these
instruments are  used to  make  projections, or to estimate local gradients in
concentrations,

Case 2.  Data from two sampling instruments operated at different
         sites within  urban area.
For  separated sampling sites, the general Equation  (12) must be adopted as
the  appropriate  definition  of  the  correlation  coefficient for paired  data
series. Two  limiting cases of Equation  (12)  may  be considered,  but  they
appear to be  less interesting then the general  case. In one limit, the  "local"
contribution  can be  assumed  to be  perfectly correlated  to the "urban"
contribution;  this case  reduces to Equation (14), discussed previously. In the
other limit, the  "local"  contributions can be assumed  to  be  completely
uncorrelated  to  one another  and to  the urban scale contribution    (i.e.
CU.LX  = CUiLy  =  CLX>Ly  = 0);   in this case the "local contribution can
simply be considered as an additional error or "noise" term, and a form similar
to Equation (14)  is again appropriate.
                                                                    11-9

-------
An interesting approximation  to Equation  (12)  can be developed from the
assumption that the urban and local scale contributions to the measured data
are uncorrelated    (i.e.   Cy Lx  = Cu Ly  = 0).    If the error variances are
also assumed to be equal   (VEx = VEy = VE),  then
         R
                 _
          x'y    (VU+VLX+VE)1'2
For this case the correlation coefficient represents approximately the ratio of
the urban-scale variance plus the covariance of the local contributions to the
total (urban plus local plus error) contributions.

Equation  (16)  is particularly interesting because  the  covariance term  can be
either positive  or  negative. An interpretation of this equation is suggested by
the data  illustrated in Figure 11-3. Monthly correlation coefficients for three
station-pairs  for  February through  November, 1966 are illustrated  in this
figure; the station locations are  all  indicated in the Boston base map, Figure
11-1. Stations  56  and 65  are central city locations, separated by a distance
of only three miles. The covariance  term   CLX Ly   might thus be expected
to be positive  for this pair. Conversely, stations 38  and 87  lie on  opposite
sides of  the  urban area, and on opposite sides  of the  major S02  emission
areas. A  negative  covariance term   CLX Ly   for the local  contributions at
these stations  is  expected  (i.e.  when one  station is upwind of the major
sources, the other is downwind, and  vice versa).

Equation  (16)  and the data of  Figure 11-3 suggest the  following interpreta-
tions:
a.  The  error  variances  for the  station  data series  may be approximately
one-half  the magnitude of the  real (urban scale plus local contributions)
variances.
b.  The  urban  scale  contributions  and  the  local  contributions  may be
approximately equal in the data  series.
c.  For proximate station pairs (such as 56 and 65), and for station pairs with
similar orientations relative to the local maximum emission  regions (such as
56 and 85), the covariance  between the station data series is expected to be
positive.  Calculated  correlation   coefficients for the data observed in condi-
tions a,  b,  and  c  would be in the range of 0.25 to 0.5.  Such results are
obtained  for most of  the  ten monthly calculations for both of these  station
pairs, is illustrated in  Figure 11-3.
d.  For station  pairs on opposite sides of the major emission regions (such as
38 and 87), the covariance  term  CLX Ly  in Equation (16)  is expected to
be negative.  This would  result  in near-zero  values  for Rx  y,  because the
urban contribution variance term, VU( is always  positive.  Figure 11-3 con-
tains  nine  calculated  monthly  correlation  coefficients for  this  pair of
11-10

-------
      S02 STATION CROSS CORRELATION, MONTHLY DATA SERIES 1966
u./
0.6
0.5
i—
LU
S 0.3
§ 0.2
o
1 0.1
< no
LU
g -0.1
o
° -0.2
•0.3
_n ^
— • 	
- .,—
— -^^ ^ ^>^^^^ ^
— xx —— — x 	
/ ~-x • "'
	 / X. /
— 	
                   CODE
   M     J
      MONTH
STATION        ANNUAL
 CODES        AVERAGES

	
56 + 65
56 + 85
38+87
0.343
0.312
0.016
   Figure  11-3.  SC>2 station correlation  coefficients calculated
   for monthly data series.
stations, and the results are all small  compared  to  the other station pairs.
The average of the nine available monthly values is very nearly zero.
It  should be stressed that the illustration described  here  is only suggestive.
Many  more data  samples  must  be investigated  to  validate this statistical
determination  of the scale of urban pollution phenomena.

Case 3.  Use of averaged data to reduce error  variance.
Because of the error contribution to the data record from a single observing
site, average values are frequently generated for  use in model verification and
                                                                 11-11

-------
projection  applications, The averaging process can  be carried out either along
the  time  dimension,  grouping  data  from  a  single  source,  or spatially,
grouping  data  from separate observing sites  operating  simultaneously. In
either  case the  averaging process always reduces the error variance contribu-
tion to the calculated correlation coefficient for paired data series. The real
variance is also reduced by the averaging process.

An example  of  simple time  averaging is  provided by the tape stain data
collected from  the two samplers operating together at  the Harvard School of
Public   Health.  When  the  two-hourly data  values  are grouped  into daily
averages,  the  correlation  coefficients  for  the  three  sequential  periods
described in  Table  11-1  are  0.92,  0.95  and 0.94. The daily average values
from the two samplers contain very little error contribution.

          Table 11-1.  TAPE STAIN DATA COLLECTED AT HSPH, 1968
                            (2-HOUR SAMPLES)
Sampling
period
6/21 - 6/30/68
7/1 -7/10/68
7/1 1 - 8/6/68
Average Con per 103 ft
Sampler A
0.85
0.72
1.00
Sampler B
0.92
0.68
1.01
Number of
data pairs
56
93
269
Correlation
coefficient
0.73
0.73
0.78
When data from two or more observing sites are averaged, the correlation
coefficient for the averaged data  represents an improved  measure of urban
and local  scale contributions, because of the reduction in random errors. The
reduction   in  variance  for   an  averaged  set of  data  can  be  illustrated as
follows. If X,  is defined as an average of two or more data  series, then
               1  M
X           — 	 y   v
          1     R/l  Zj   Aj j
                    _
                       (U,
(17)
where  M  is the number of stations entering  into the averaging process. The
upper  and  lower limits  for  the  variance  Vx of  the  averaged data can be
calculated:  If the local contributions  Lf^  are  perfectly correlated to one
another, the upper limit value is

            vx  = vu  +  VL_X  +  CU]Lx  + jjj   VEX              (18)
If the local  contributions are uncorrelated,  the lower limit value is

            Vx  = ^ +lJr   KX  +  CU|Lx  +  VEx]             (19)
11-12

-------
In  practice the  actual  variance calculated for a  set of  averaged  data lies
between these two limits.

The variances enter  the denominator of  Equation  (12) describing the con-
stitution of the  correlation coefficient, and the limiting cases described above
suggest:

  1. When averages  are  calculated  from  stations  that are  relatively  close
     together, (a)  the local contributions will be more highly correlated, (b)
     relatively  large values  of Vx will  result,  and  (c)  the  correlation
     coefficient RX,Y wi" De relatively small.

  2. When averages  are  calculated from stations widely separated from one
     another, (a)  the local contributions will be less correlated,  (b)  smaller
     values of Vx will result, and (c) the value of RX,Y wi"  be greater.

This pattern  is  illustrated in Table 11-2. The greatest correlation coefficient
results when  the  middle  and outer suburban stations  are  averaged.  These
stations are widely separated, and each is subjected to a variety  of local scale
contributions that do not  affect the others. The inner city stations are  more
closely grouped, and  the local scale contribution to these data is  not filtered
effectively by the averaging process.
   Table 11-2.  CALCULATED CORRELATION COEFFICIENTS FOR GROUPS OF
            TAPE STATION DATA, BOSTON, NOVEMBER 1 - 10,1966
Station Grouping
Station 45 vs station 47 	 	
Station 54 vs inner city average
Inner-city3 average vs middle-suburban average 	
Middle-suburban average vs outer-suburban average 	

Number of
data pairs
120
120
120
120

R
0.11
039
0.65
0.82

  Inner-city stations: 45, 47, 55, 56, 65, 66

 bMiddle suburban stations:  26, 29, 34, 38, 48, 74, 85, 87

 '"Outer suburban stations: 9, 13,51, 92, 99
 Note that even for large groups of stations drawn from throughout the urban
 region  (the  middle and outer suburban data),  the correlation coefficient is
 approximately 0.8. This value might  be taken as an expected upper limit for
 correlations  between  field data  and air quality model data, even when  the
 field measurements and the model projections have been  averaged to  provide
 urban  scale  comparisons.   For  point-by-point  comparisons    of   model
 projections and field  data, significantly lower correlations must be expected.
                                                                     11-13

-------
CORRELATIONS BETWEEN AIR QUALITY DATA AND
METEOROLOGICAL DATA
The time scale of several meteorological phenomena that influence urban air
quality  is less  than 1 day  (e.g. diurnal  changes in stability, front passage,
changes  in "mean" wind direction,  sea  and lake breezes). The relationship
between  meteorological  parameters and air quality  indices must therefore be
examined on a short-time period basis  (e.g.  1-,  2-, 3-hour increments). This
time period limitation often  prevents the use of time averaging as a device to
reduce the influence of errors in the data analysis. Spatial averaging, employ-
ing data from  many stations,  is  , however, helpful  in  the  examination of
meteorological  effects upon air quality.

Up to the present time  the meteorological  information available  for most
urban areas has been gathered  at one primary observing  site  in each area. As
a  result,  the  most appropriate air  quality  index is often  an  urban-wide
average  value.  Figure  11-4  illustrates the  calculated correlations  between
observed S02  concentrations and  wind  speeds  observed at  Logan  Airport,
Boston  during  the 1966 study period.  (Note that  the ordinate scale in  the
figure is reversed.  The calculated  monthly  correlation  between wind  speed
and  SO2  concentration  was always  negative.) As expected, the ten-station
average  data  relate best  with the wind data. For the three individual stations
illustrated, the  correlation is lowest in the case of the coastal .station number
48. Stations 56 and 65 are in urban-center locations.
In view of the error analysis developed in the preceeding section, it is likely
that the correlation between a single meteorological parameter and an index
of  urban air quality may always  be less than  0.5.  This  suggestion results
from three considerations:

   1. A  single meteorological parameter (wind speed, stability, mixing depth)
     only partially describes the dilution capacity of the urban atmosphere.

   2. The errors and the  local  influences reflected in  the air  quality observa-
     tions are never completely filtered out of the data.

   3. The representativeness  of the meterological  data  is unknown in most
     cases. Usually a single  observation, taken  near  the edge of the  urban
     area, serves to characterize the meteorology of the entire area.

Several  other  air  quality-meteorology studies by the authors, similar to  the
analysis  summarized  in  Figure  11-4,  have each  resulted  in  correlation
coefficients   less  than   0.4  for  all  data  periods.   Multiple  parameter
meteorological  indices  might  result  in  larger correlations  because of  the
better  description  they  provide  of  the  local  ventilation  capacity of  the
atmosphere. However, such indices were not considered  in the present study.
11-14

-------
              S02 VERSUS WIND SPEED CORRELATION, 1966
  UJ
  o
  u.
  u.
  LLJ
  O
  O
  ec
  o
  o
•0.7
•0.6

•0.5

•0.4

•0.3

•0.2

•0.1

 0.0
40.1

+0.2
40.3

40.4
	10 STATION AVG. -0.317
	STATION 56      -0.200
	STATION 65      -0.250
  — STATION 48      -0.
         J
                                 J     J
                              MONTH
  Figure 11-4.  862 concentration versus wind speed correlation
  coefficients calculated for monthly data series.
THE USE OF COMPUTER GENERATED GRAPHIC ANALYSES
FOR AIR QUALITY DATA
Spatial variability  in air quality observations is most effectively displayed on
a geographic  base  map. This form of data presentation is desirable  because
the  observations can  be  related  directly to emission patterns, to human
population, and to  land-use distributions.

When the field data are stored in a computer-readable format (e.g., magnetic
tape, punched cards, disks) the use of computer generated graphical  displays
of the data fields is particularly appropriate, with the advantages that:

-Maps can be produced rapidly and at minimum cost.

-The maps permit easy pattern recognition. Relationships between land use,
emissions and air quality  are quickly  evident, and  suspicious data points are
often identified by the contour patterns.
                                                                11-15

-------
—The interpolation scheme for evaluation of variables away from the obser-
ving locations is objective and repeatable.

The Laboratory for  Computer Graphics and Spatial Analysis, of the Harvard
Graduate  School of Design,  has adopted  its SYMAP and SYMVU  programs
for presentation of  air quality data, and examples of these mapping tech-
niques are illustrated in  Figures 11-5 and 11-6. In the two-dimensional-plan
view  maps, illustrated  in  Figure  11-5,  shading  is  proportional  to concen-
tration. The left map  is based upon an inadequate  input data network, and
the widespread high concentration regions are unrealistic. The right map was
produced  with the adoption  of additional  "background" data points, away
from  the important emission  regions.

The three-dimensional view shown in  Figure 11-6 represents an air  quality
(total suspended  particulates) surface  for Southern New England.  The pro-
gram  that  produces this display  can produce a similar display of  the same
data  for all  possible lines of sight  (i.e. all  possible azimuth and  elevation
angles) toward the surface.

These  maps  are  illustrated  here  to underline  the  importance  of pattern
display  in the analysis of air quality field data. The mapping techniques are
important  adjuncts to  the statistical data treatments;  both approaches are
useful for the analysis of urban scale concentration patterns.

RECOMMENDATIONS FOR THE ESTABLISHMENT AND
ANALYSIS OF  MULTIPLE STATION  AIR SAMPLING
NETWORKS
The data  studies reported in  this paper are limited, but they do offer a basis
for several  recommendations concerning sampling networks:

   1.  The  intended  purposes  for  the network  data  should be  carefully
     developed  before  establishment  of  expansion  of  an air  sampling
      network.

   2. At least  one  redundant  set of sampling equipment should be employed
     to evaluate the error contributions  to the  field data series which are
     obtained. The extra sets of sampling equipment might be installed in a
      mobile station  and moved to each permanent  field site on a scheduled
      basis to  provide calibration  checks upon all field equipment. A  mobile
     unit could be used also,  for  special  upwind-downwind studies  near
      major emission areas.

   3. Instead  of equidistant spacing of sampling stations throughout  an urban
     region,  arrangements involving two or  more  stations in  neighboring
      locations  should  be considered. Such  arrangements  may provide more
     useful  data concerning   urban  and  local  scale variability in  pollution
     concentrations.
11-16

-------
                                                      ««"": ..... LEGENDS
                                                      Hi"':;":'"  B  BOSTON
                                                      '.»:::::      p  PROVIDENCE
                                                 ;;;;;;;; ««"'      w  WORCESTER
                                                 "'"*--
                                                                S SPRINGFIELD
                                                                H HARTFORD
                                                                N NEW HA VEN

NORTH       j
 ••••»        T
 0510       |
 MILES       !
                 LONG ISLAND SOUND
SOUTHERN NEW ENGLAND
1966 AIR QUALITY
SUSPENDED PARTt.CU LA TES (UG/M3) - YEARLY AVERAGE

ACTUAL DATA VALUE EXTREMES ARE   25.00    152.00
ARITHMATIC MEAN IS   75.07      STANDA RDA RD O E VIA Tl ON IS   25.77

TOTAL MISSING DATA POINTS IS    4
TOTAL SUPERIMPOSED  DATA POINTS IS   i. THESE OCCUR IN   I LOCATIONS.

VALUE LIMITS FOR EACH  LEVEL
  ('MAXIMUM1 INCLUDED IN HIGHEST LEVEL ONLY)

MINIMUM  25.00  45.80  54.00  57.40  65.40  69.00  79.00  88.60   94.90   110.40
MAXIMUM 45.80  54.00  57.40  65.40  69.00  79.00  88.60  94.90  110.40   152.00

PERCENTAGE LIMITS FOR EACH LEVEL (BASED ON THE TOTAL VALUE RANGE)
  ('MAXIMUM1 INCLUDED IN HIGHEST LEVEL ONLY)

MIMIMUM   0.00  16.38  22.83  25.51  31.81   34.65  42.52  50.08   55.04   67.24
MAXIMUM 16.38  22.83  25.51  31.81  34.65  42.52  50.08  55.04   67.24   100.00
                                                                IjjJKJ
                                                                i
                                                          iiiiii!
         16.38  6.46   2.68    6.30  2.83   7.87   7.56   4.96  12.20
   PERCENTAGE OF THE  VALUE RANGE IN EACH  LEVEL (ABOVE)
    Figure 11-5a.   SYMAP  plot showing Southern New England
    particulate concentrations for 1966.   Insufficient  network-
                                                                           11-17

-------
                                                           LEGENDS
                                                           B BOSTON
                                                           P PROVIDENCE
                                                           W WORCESTER
                                                           5 SPRINGFIELD
                                                           H HARTFORD
                                                           N NEW HA VEN
  SOUTHERN NEW ENGLAND
  1966 AIR QUALITY
  SUSPENDED PARTICULATES (UG/M3) - YEARLY AVERAGE

  DATA  VALUE EXTREMES ARE    25.00   152.00
  TOTAL MISSING DATA POINTS 15  5
  TOTAL SUPERIMPOSED DATA POINTS IS
                                     THESE OCCUR IN  1  LOCATIONS.
  ABSOLUTE VALUE RANGE APPLYING TO EACH LEVEL
    ('MAXIMUM' INCLUDED IN HIGHEST LEVEL ONLY)
                                                             ABOVE
  MINIMUM BELOW  40.00 50.00 60.00 70.00  80.00  90.00 100.00  110.00 1ZO.OO 130.00
  MAXIMUM 40.00   50.00 60.00 70.00 80.00  90.00 100.00 11 0.00  120.00 130.00

  PERCENTAGE OF TOTAL ABSOLUTE VALUE RANGE APPLYING TO EACH LEVEL

                11.11 11.11 11.11 11.11  11.11  11,11  11.11   11.11  11.11

  FREQUENCY  DISTRIBUTION OF DATA POINT VALUES IN EACH LEVEL

  LEVEL    L     123456     7    8     9H
  SYMBOLS  «
     Figure 1l-5b.  SYMAP plot showing Southern New England
     particulate concentrations.  Additional  stations.
11-18

-------
                                    S. NEW ENGLAND AIR QUALITY-PARTICULATES
                                                               AZIMUTH =330
                                                               WIDTH =8.00
                                                               ALTITUDE =45
                                                               HEIGHT =2.00
Figure 11-6.  Three-dimensional perspective view of the
Southern New England particulate surface for 1966.

-------
The  following recommendations are concerned with data analysis techniques
for sampling networks.
   4.  The  best  measure  of  urban scale  concentrations  is an  average  of
      simultaneous data  from stations  throughout the region. This technique
      minimizes both  error variance and variance due to local  sources. Cor-
      relation analysis  involving  single-location field-observation  series and
      model  projections (or meteorological parameters) will  probably  result
      in maximum correlation values of approximately 0.5.

   5. When  a large number of sampling stations are available, it is helpful  to
      calculate   two  urban-wide  averages, from  non-overlapping  sets  of
      stations. The correlation coefficient for the two averages represents the
      upper limit correlation expected from any comparative  study  involving
      the field data. Examples of possible comparative studies include:

        a.  Comparing  time variability  in the concentrations of  two different
           pollutants,
        b.  Relating pollutant  concentrations to meteorological parameters,
        c. Comparing model calculations with  actual field data.

   6. When  daily or longer-period data values are required, time averages  of
      shorter term observations will usually be more reliable than single, long-
      period observations.

   7.  Computer-generated  graphic  displays of air quality  fields should  be
      employed  as data  analysis  tools whenever practical.  The repeatability
      of  the computer-controlled interpolation  scheme  is essential for con-
      sistent interpretation  of data. (As a corollary, it  should be  mentioned
      that  improved  interpolation  schemes,  based upon emission data and
      diffusion  modeling, should be developed.)

The  final recommendation represents  a  restatement of the  theme  for this
paper:

   8. All  air quality observations include significant variability due  to obser-
      ving error and to the influence of  sources near the observing  site. All
      data  analyses, projections, and  model  comparisons should  include  an
      evaluation of the error  and local source  contributions to the variability
      in the measured data. Performance indices  ("skill scores")  for models
      and  projections must  be  judged  relative  to  upper-limit  correlation
      figures developed from the field data.
11-20

-------
REFERENCES


1. Selected Methods for the Measurement of Air  Pollutants. Division of Air Pollution.
  Cincinnati, Ohio. PHS Publication Number 999-AP-11. 1965.
2. Goodrich,  J. C.  Regional Planning  and Air  Pollution  Analysis and  Control:
  Methodology and Case  Studies. Harvard University,  Cambridge, Mass.  M.S. Thesis.
  1969.
3. Investigation and  Study of Air Quality in the Metropolitan (Boston) Air Pollution
  Control District. Massachusetts Dept. of Public Health. 1968.
 ACKNOWLEDGMENT
 Mr. Maddaus carried  out most of the  Boston region correlation calculations,
 and  the data maps are drawn from  a  recent Harvard  Masters Thesis by  Mr.
 Goodrich.2  Other calculations and evaluations  have been provided by Mr. B.
 A.  Egan and  Mr. E.  I. White of the  Harvard  School  of Public Health. The
 Massachusetts  Department of  Public  Health  provided  all  of  the Boston
 Region data discussed in the paper.3

 The preparation  of  this  report  was supported in  part  by the National  Air
 Pollution Control Administration, Consumer  Protection  and  Environmental
 Health Service, Public Health Service, U.S. Department of Health, Education,
 and  Welfare, under Grant Number 1  RO1  AP01013-01.
                                                                     11-21

-------
 APPENDIX - GLOSSARY OF SYMBOLS
 C        covariance between  two series
 E*       error contribution to X,
 i         iteration integer
 L*       constant contribution to X; from measuring site locality
 M        number of measuring stations entering into an averaging process
 N        number of members
 R        correlation coefficient
 U,        constant contribution to X, from urban area
 V        variance of a series
 Xj        measured  pollutant concentration at i
 Vi        point in measured time series
11-22

-------
AUTHOR
      MORRIS NEIBURGER, a professor of Meteorology at the University of California
      at Los Angeles has been a consultant for the Los Angeles County Air Pollution
      Control Department,  the State of California Department of Water Resources, and
      the Public Health Service Community Air Pollution Program. He has been President
      of the American Meteorological Society,  the Society of Sigma Xi and the American
      Geophysical Union. Atmospheric radiation  and dynamics, cloud physics, and air
      pollution have been his primary fields of research.

-------
Banquet Speech
                12.   PROGRESS + PROFITS +  POPULATION
                                                     =  POLLUTION

                       MORRIS NEIBERGER
                University of California, Los Angeles
 Thank you  very  much, Dr. Stern. Or perhaps I  shouldn't thank you, for I
 don't know why  I let  myself get into this situation.  As I  told the audience
 the  only other time I  got roped into giving a banquet speech, "It reminds
 me of Sam Goldwyn's  comment, that anybody who pays good money to see
 a psychiatrist  ought  to have his  head  examined." Anyone who voluntarily
 undertakes to give an  after-dinner talk  at a  technical  conference certainly
 renders his sanity suspect. For, on the one hand, since ladies are present he
 has to be strictly non-technical, since being technical when ladies are  present
 is even worse than  being obsecene.  About the technical  aspects—I  might
 mention that Francois Frenkiel saw  me  looking  at a  manuscript with
 equations today  and  said  in a  voice filled with horror, "You're not going to
 include equations in your talk  tonight." When  I  replied  "It  will  be only
 equations;  I'm not going to include any words, just symbols in equations,"
 Irv Singer said, "That's all right,  as long as you  limit yourself to equations
 my wife can understand."  And when I  asked what kind of equations his wife
 understands, he said,  "None." So, technical talk is out!

 On the other hand, the men in the audience  will be  bored and feel cheated
 if all they get for the price of the dinner ticket is some good food and a few
 jokes and platitudinous remarks.  They feel that at least they should come
 away with some  profound new ideas, stimulating insights  that inspire them
 to attack  the  subject of the conference with  increased  energy, and  predic-
 tions of the future of  America and humanity that give them hope that their
 efforts will not be in vain.

 On top of this,  as I said  in my previous dinner speech, the talk should be
 neither so humorous that  the  laughter is loud  enough to awaken those who
 are using the time for  an  after-dinner  snooze, nor so serious that it arouses
 thoughts that interfere  with the digestive process.
                                 12-1

-------
Obviously, to  succeed  in this talk  is impossible, so you  may well ask: since I
am aware  of  its impossibility, why did  I undertake  it? There is a simple
answer. The only condition  under which Arthur Stern would  invite me to
the conference was if I agreed to be an after-dinner speaker.

So, I  promise that I will  be  only  slightly technical, and not at all obscene,
although it is  with some reluctance that I  forgo the latter.  Knowing that one
should always  start an after-dinner  speech with a little joke, I asked Dr. Claren-
burg,  who  claims to know hundreds of them (though he says they are funny
only  in  Dutch), and  he  recited  a  limerick that was  very apropos of my
subject, as he will see when  I give the title of  my talk.  I would  like very
much  to  tell  you  this  limerick;  unfortunately, there are ladies  present.
Besides, I  have already forgotten it.

About the new  ideas, stimulating insights, and hopeful predictions I'm afraid
that John  Middleton,  in his  keynote speech  for  the conference, anticipated
all  the  ideas I had  intended  to present,  so  I'll  have to  be satisfied to say
things you already know, hoping that I can do so in ways  that may  possibly
be fresh to some of you.
While considering the title I should give to my talk, I was very impressed by
the slogan Walter  Orr  Roberts enunciated a  few  weeks ago, namely, "After
the Moon, the  Earth," and thought of  stealing  it as my topic. I  certainly
concur  with the view  that now that we  have achieved the goal of reaching
the moon we  should divert a  few  billions from the space budget to augment
the few millions now budgeted to  study and  protect our earthly enviroment.
I  have  advocated  making our federal  expenditures  more balanced  in this
regard for a long time. I  have restrained  my kleptomaniac tendency to steal
Robert's subject, however, and following my  pledge to  Dr.  Frenkiel,  have
chosen for the title of  this presentation the equation:

    PROGRESS + PROFITS + POPULATION = POLLUTION

It  is just  eight  years since the conference  on Air over Cities  was held  in
Cincinnati, under  the  auspices of the Division of Air Pollution of the U.S.
Department of Health, Education, and Welfare, Public Health Service. At the
conference I  chaired  a session on dispersion  and deposition  of  pollutants
over cities. In  my introduction to the session I pointed out the deficiencies
in  our knowledge  of diffusion  from multiple sources and area sources. I had
hoped that  the  papers being  presented at the present  conference would  be
evidence of progress toward the  elimination of  these  shortcomings. I fear
that such  evidence  is not clear.  But whether or not the solution of the
problem has been attained, it is certainly true that the intervening 8 years
have seen a tremendous  increase  in the  concern and  attention paid to  air
pollution. Whereas 8 years ago, air pollution  was still regarded as  a minor
problem,  serious  for  only a  few  unfortunate  communities  such as  Los
12-2

-------
Angeles;  now, it  competes with  the Vietnam  war, the racial question, and
the student protest movement for first place in the attention of the country.
Government bodies at all levels  and civic  organizations everywhere devote
large amounts of  time and effort to combating it. Whereas, at  that time, it
was regarded as an  isolated  problem, today it is recognized as an  intrinsic
part of  the complicated impact  civilization and  technology are exerting on
the environment.

In my talk this evening I  wish to present some views about the root causes
of the air pollution problem, and the relationship of this problem to  some
broader  aspects of society.  Just  as  education  is frequently characterized as
being founded in  the three R's,  I suggest that  the roots of air pollution and
other insults  to  the enviroment  lie  in  the three P's: Progress,  Profits, and
Population.  These three  P's  represent  fetishes,  which are  heretical to
question, like motherhood or patriotism. Unquestioning acceptance of them,
however, will in the long run, lead to catastrophe.

As a starting  point, suggesting why I am concerned with the three P's, I shall
give an example from my  home city. In Los Angeles the three P's are closely
correlated: the measure of Progress  has been the increase in Population, the
expansion of industry, and  above all,  the  Profits that  have accrued to the
real estate developers.

Some 15 years ago I  gave a  talk to the members of the research committee
of the Los Angeles Chamber of Commerce. I told them that I did not know
whether  or not the salubrious  climate of Los Angeles would have been able
to destroy itself on its own; but with the energetic assistance of  the Chamber
of  Commerce  it  had been  achieved.  By  advertising  the wonders  of the
Southern  California climate  and  thereby  attracting increasing  numbers of
people and  industries, they  transformed  the clear blue skies, sparkling sun-
shine, bright green  vegetation and  the  year  round invigorating air into a
murky brown smog  that obscured  the  sun,  withered  the plant life, and
irritated  the eyes, noses and  respiratory tracts of the people. I recommended
to  them that, until effective  control measures were  found, instead of trying
to  attract still  more  people  and  industries, they should  sharply apply the
brakes to  the  population  growth of the area,  using every means at their
disposal to dissuade people and industries from moving to Los Angeles, and
persuading those who were already there to move away.

Needless to say, such a heretical  recommendation did not receive a favorable
response from  the Chamber of  Commerce. Its  promotional activities con-
tinued, for its primary motivation was and  is the profits of its members, and
not the protection of its  surroundings. Progress  prevailed and smog  became
more and more wide spread.
I have often  been asked why the peak values of contaminant concentrations,
registered in  those years, have not  been exceeded.  To  this my half-joking
                                                                     12-3

-------
reply has been  that  you can't get automobiles  much  closer together than
bumper to bumper. The tremendous increase in number of cars has  meant
larger and larger areas of high traffic concentrations, but no increase in peak
concentrations of contaminants because the total emission in any given area
has not increased.
Of course, that  is only part of the story. The other part is that we have not
yet  had more severe meteorological  conditions than those experienced in
1954 and 1955.  One of these months, perhaps soon, a prolonged period of
very  light winds and  low  inversion  will  occur,  and  with  it will  come
record-breaking  levels of contaminant concentration, in spite of the measures
that have been  taken so far to reduce emission of pollutants. We can only
hope that it  will not be accompanied by a  disastrous increase in illness and
mortality. If  it is, it will be one more reward for the  prevalence of the three
P's,
A  great deal of attention  has been paid,  recently, to defining air quality
standards in  various  communities and  states, as well  as  nationally.  These
standards represent  the  maximum acceptable  levels of  concentration  of
particular pollutants from the standpoint of  health, safety, and well-being. In
some instances aesthetic considerations enter into the definition  of standards
with  respect  to  particulates in the atmosphere.  But, as all those attending
this  conference  are  aware, these  air quality standards remain  pious  hopes
until  they are utilized to set  maximum allowable emissions from the various
types of sources. The multiple-source  urban diffusion  models  that are the
subject  of this conference, offer, as one of  their applications, the possibility
of determining whether, under adverse meteorological  conditions, the desired
air quality standards could  be met with existing sources of emissions  and if
not, how much  reduction of emission  rates would be required to achieve the
standards. The models  could also  predict  the effects of increased emissions
due  to  expanding population, additional industrial sources, or the speeding
and  redistribution  of the  urban  areas.  In places where the  air quality
standards are already exceeded frequently, the  latter computations  would
seem almost  superfluous.
One principle, which  I think, is readily evident, and which I therefore expect
to be confirmed  by computation with  these models, is the  concept that
emission control  standards must be expressed on a per-unit-area basis  rather
than  a  per-source basis.  This  concept  is  most  readily appreciated   if one
considers  the present practice  of expressing  the allowable emissions in terms
of the percentage of total concentration of effluent emitted from one  source
or the total  emission from  each individual  source. In the former instance, if
the rate of emission of the total effluent is increased, the permitted emission
of  pollution   is   correspondingly  increased.  In  both  cases,  if  in  a given
square-mile area the number  of  sources  is multiplied, the total  pollution
allowed  is correspondingly multiplied. The  pollution  from  one  or two
12-4

-------
sources  in  an industrial plant emitting maximum allowable concentrations
might not exceed  the air  quality standards.  If the same  plant constructs a
dozen  more  stacks, each  emitting  concentrations  of contaminants  indivi-
dually  acceptable,  the total effect might well  be to  produce concentrations
far in excess  of the air quality standards. To prevent such excesses, the limit
must  be expressed in  terms  of total  emission  per  unit  area, whether  the
emission originates from one source or many.

This concept, which  is  so clear  with  respect to industrial stacks, applies
likewise to vehicles. It is not  adequate to set  the emission limit in terms of
individual  cars  without first  estimating  the  consequences  of this limit in
terms of the  total  emission from all cars expected to be operating  in a given
area. As I  pointed out before, and as has been experienced by drivers in all
cities, this number of cars has  a maximum, which is frequently reached when
all lanes on  all streets in  the area are completely filled with bumper-to-
bumper traffic.

The area-limit concept also applies to domestic sources, but if the allowable
emissions are expressed on a per-capita basis there is no maximum  similar to
the automobile  case,  for  modern  apartment  dwellings appear  to have  no
upper-bound  on the density  of population  which can  be concentrated on
small  horizontal areas. While  the vertical spread of  high-rise apartments  has
some  effect of spreading  the  emissions upwards also; the net effect of an
increased  number  of  people  living  in an area, is an  increased amount of
waste produced which  must be disposed of, inevitably  increasing the amount
of pollution that enters the atmosphere.

I  recognize that,  at   present,  the  limits to  allowable  emissions are  not
expressed  on a  per-capita  basis, however democratic this basis may seem.
Surely there  is a sort  of poetic justice in the idea that if I must  share  the
unpleasant  consequences of pollution  with everyone,  I should be  permitted
to contribute an equal  share to everyone else's  discomfort.
Whether or not the emission limits are expressed per-capita, it is self-evident
that  the amount of pollution is closely related to  population. Each addi-
tional person in a community adds to the requirements for electric power,
domestic heating, transportation, and waste disposal. If the pollution result-
ing  is not  proportional to  the  number  of  people, one may suspect that  the
deviation  from  proportionality is  in  favor  of  a  more rapid  increase in
pollution than  in  population.  Small  towns generally have small  pollution
problems,  and large metropolitan areas have  gigantic ones. When regions are
characterized  by  many metropolises their net pollution  seems not merely
additive, but combined in some non-linear fashion.

I  shall return to  the  problem of  population  on  a national and world-wide
basis.  Before  considering it, however, I  should  like to expolore the effects of
progress and profit in  more detail.
                                                                     12-5

-------
Progress  is often measured by the rise in the standard of living of the average
citizen, but it might also be  measured by the increase in power,  which is
associated with the rise in  standard of living, or by the increase in  number of
appliances and machines, which  are used to produce the rise in  standard of
living.  Domestic cooling systems added to  central heating, dryers in additon
to washing machines, color TV in addition to  radios, hi-fi systems and tape
recorders,  multiply the demand  for power by  a  large factor. This increased
demand  leads to construction of more and larger electric generating plants,
most  of  them using fossil fuel and emitting  large amounts of contaminants
into the  air.

The natural demand  for these additional  comforts  of life is augmented  by
intensive  advertising campaigns by the manufacturers of the appliances and
by  the utility companies  that distribute the fuel and electricity. Insofar as
these  industries are motivated by the purpose of rendering the lives of the
citizens more  comfortable, the labors of the  housewives easier,  and every-
one's  leisure  more enjoyable,  one would be  led to  commend  them, while
regretting the degree to which they miss their goal  by  also  contributing  to
discomfort and illness with  the pollution they introduce into  the air.  In
addition  to these high motives,  however,  or,  in some instances,  instead  of
them,  the desire  for  profits  enters  strongly.  Even  when  they  know that
additonal thermal electric generating  plants will augment already objection-
able levels of  pollution, utility companies strive to increase the demand for
electricity, using  all sorts of promotional  gimmicks to get customers to con-
vert to all-electric homes, etc.
Two questions occur to me. One is whether  the damage  to the environment
and  the  discomfort to  the people,  resulting  from  technological advances,
should  not  be  weighed  against  the conveniences  produced  in  deciding
whether  these technological advances really  represent  progress. Perhaps  on
balance it would  be preferable to forgo some of the benefits and be spared
the accompanying  disadvantages.  The second  question is whether, assuming
the validity of asking the first, the decision about expansion  of  electric
generating plants, for  instance, should be based,  as it is for the most part
now,  on  whether  expansion  will yield  a profit  to the power  companies.
Shouldn't the  public's welfare come first?
The automobile  industry  is perhaps the outstanding  instance in  which the
desire  for  profits has  dominated the public  interest. Regretably,  the recent
agreement of  the Department of  Justice to a consent decree in  its  conspiracy
suit against the auto  makers has  prevented  the  evidence in  the case  from
becoming available to  the public. But the agreement of the manufacturers to
cease   and desist,  together  with  the fact that the  court  ruled that the
evidence  be impounded so that it could be  available in damage suits, strongly
suggest that they did  indeed  conspire to prevent and delay the development
of air pollution control devices for motor cars.
12-6

-------
Their  motive has been stated  clearly  by Henry Ford  II, in interviews quoted
in  Fortune magazine and  in newspapers. In these interviews he said that the
industry  has  too large an investment in equipment to make engine blocks,
transmissions,  and other  parts of the internal combustion engine, for it to
give a high priority  to  developing alternative power systems, such as an
electric car. The billions of dollars of machinery owned by the auto industry
act as a  millstone,  holding  the  people  of  our cities submerged in the
pollution exhausted from  internal combustion engines.

The same eagerness for profit has, of course, been responsible for the auto
industry's failure to  develop and introduce safety features  until required to
do so by federal  legislation. Similarly, the profit motive has been responsible
for the industry's disregard  for such factors as the ease and costs of repairs
and the protection of the  car itself from damage in minor collisions.

The auto  industry appears to  judge its progress by the size of engine it can
foist  upon  the  public—500  cubic  inches,  this  year.  With  the  allowable
emissions presently measured  per unit volume of exhaust, it is clear that the
increased cylinder capacity will represent increased amounts of pollution per
car, which when  added  to  the  increased  numbers  of cars will offset the
control  measures. Pollution control considerations obviously  did  not  enter
the decision to put these mammoths on the market and feature them in the
promotional advertising.
The  internal  combustion  engine  is, by its nature, an  incomplete combusion
engine. The automobile industry's answer to the requirements to reduce the
products of that incomplete  combusion, namely hydrocarbons and carbon
monoxide, has been to promote  further combustion in the exhaust system,
where it  produces no useful power.  It has  ignored  the  fact  that,  in that
process, increased amounts of  oxides of nitrogen are emitted.

It seems to me that the solution to  the problem is to abandon the  intrin-
sically inefficient  internal combustion engine, and to develop an essentially
pollution-free  propulsion  system  in its place. It  is inconceivable to me that
the technological ingenuity  that got man to the moon cannot get him from
his home to work in a city without polluting the air.

The high rate  of profit in  the  industry's operations can readily  cover the cost
to industry of developing a smog-free engine. I do not have recent figures at
hand, but the American  Institute of  Management, in  an article appearing in
The Corporate Director,  in  July  of 1956,  reported that General Motors'  net
earnings, at the  1955 rate of profit, "were  sufficient to  recoup the com-
pany's entire net plant investment in  2 years." (Quoted by Ralph  Nader in
The Progressive, September  1968.)
From the standpoint of the profit motive, it is  hard  to understand the past
reluctance  of  the auto industry  to add air pollution  control equipment to
                                                                     12-7

-------
the internal combustion engines. The cost of this equipment has been added
to the price of the car and you may be sure that in arriving at the price, it
has been subject to  the same profit  margin  as all other  parts of the car, so
that pollution control has increased the automobile industry's profits.

I  have suggested  that if the  Federal, State, and municipal governments were
to adopt  laws setting a date  after which no  vehicle injecting any hydrocar-
bons, carbon  monoxide, or oxides of nitrogen  into the atmosphere might be
sold in their  jurisdictions, industry would accept the challenge  and develop
new propulsion systems to meet that requirement.  In  the California State
Legislature, last year,  such a bill was passed by the State Senate  but was
killed in a committee of the Assembly. When it came before the Senate,  the
lobbyists for  industry apparently did not think anyone would  take it  seri-
ously but once it passed in that body, the lobbyists exerted great pressure to
prevent its consideration  in the Assembly. Senator Petris plans to introduce
it again at the next  session, but with  industry alerted I  have no  hope of its
passage, even  though the general public  is more concerned about pollution
than ever before.

That it is necessary  to  develop  completely smog-free vehicles is clear, since,
otherwise,  whatever  degree of reduction of emission  is achieved, the increase
in number of  vehicles will rapidly make up for it and the levels  of pollution
of the atmosphere will  continue to rise. Similarly, unless the emissions from
power  plants  are completely  eliminated, the rising demand for electric power
due to population growth and increased per-capita use will  offset individual
emission limits, and  the deterioration of  the  atmosphere will  continue.  The
same applies,  of course, to other  industrial sources and to domestic  sources.
This situation  is true on a local and  statewide basis  and, also, on a  national
and world-wide basis. The problems of urban  areas  in the more "advanced"
countries are  particularly acute  because of the concentrations of population
there. But  the entire world is undergoing a population explosion, and in the
developing countries there is also a  power  or energy explosion, which will
increase as the people of Asia, Africa, and South America strive to attain the
same  levels of comfort, convenience, and mobility  as Americans and Euro-
peans have. Multiplying the world-wide population explosion by the prospec-
tive demands  for power yields a potential for air pollution that staggers the
imagination. It dramatizes the urgency for the development of  completely
pollution-free  sources of energy  and methods of transportation, and the need
for  contol of population.

I  have  previously pointed out that, even  now, we do not know whether, on
a  world-wide  basis, toxic contaminants are being put into the air faster than
the natural  cleansing processes  of  the atmosphere  remove  them. We know
that pollution  does not stay  in  the air; it is removed eventually.  One way it
is removed, for instance,  is by getting into our eyes and respiratory systems
12-8

-------
and irritating them.  It is also incorporated into the clouds  and falls  out  in
the form  of rain, and  it interacts with vegetation, and  corrodes building
materials. The question is whether  the processes of  removal  are able to keep
up with the  rate of emission. We do know  of  one pollutant, though not a
toxic one, of which there is an accumulation in the atmosphere. Carbon dioxide
has been sampled long enough, and with enough accuracy, to show that the
total  amount is  increasing  steadily year  by  year.  If  the same thing  is
happening to toxic pollutants, such as carbon monoxide or nitrogen dioxide,
eventually,  the  entire atmosphere  will  become so toxic  that  it will  be
necessary for people to go around in gas masks  to survive. Alternatively, one
can visualize  people  living  in air conditioned vehicles, dealing with a compli-
cated  docking procedure similar to space vehicles, in going from building to
automobile and  vice versa. People  could  never  venture out  of doors,  where
they might be subject to  "toxic fresh air."

Regardless  of whether or not we are now putting toxic pollutants into the
air faster than they are being removed; if the number of sources, stationary
and  vehicular,  increases  with  the  population  and  power  explosion, that
situation certainly will be reached, unless sufficently stringent controls  on
emissions per-source  are  put  into effect, and unless the  world population  is
stabilized.
There are,  of course, other urgent  reasons for being concerned about world
population.  To these is  added the  danger that the entire  atmosphere will
become too toxic to  breathe because of the sheer number of people adding
pollution to the air.
What  is the  solution? I  believe that to find it,  it is essential that we make
some  fundamental changes in social attitudes.
In the  short run, of  course, the  present attempts to control sources  of
emission must be strengthened. More and  more  stringent standards of emis-
sion from stationary  sources and from vehicles must  be adopted. Rather than
limiting the requirements by what is technically  feasible, requirements should
be set at levels that challenge technology to meet them.

In the longer run we must revise our attitudes about Progress, Profits, and
Population. The concept  of Progress must be amended to measure it in terms
of the  true  amenities of  life,  including  the enjoyment  of clean air. The
motivation  of industry  and  commerce  to maximize their  profits must be
restrained  by the  requirements of  the  safety  and  welfare of  the  people.
Above all,  we  must take measures to  limit population increase, and ulti-
mately, to stabilize the population of the world.

I  am  not at all hopeful that we will be successful in any of  these tasks. The
concept that  "bigger is better" is so ingrained in American attitudes that any
move  toward smaller  automobiles, fewer appliances,  and  less people per
                                                                     12-9

-------
square-mile will  be regarded as retrogress rather than progress.  Unless we find
some way of having larger  profits accrue to  industry for building  smaller
cars, producing  less electricity, and, in general,  contributing less to pollution;
industry and commerce will  continue to  enlarge their production of energy,
commodities,  and smog. As for  the control of population, since it depends
on  the  slow  process of  education about contraceptive  methods  and on
producing  motivation for  their  use, the prospects are remote indeed.
The  prospects will remain  remote  until the nature of the  problem is recog-
nized. As I said at the  beginning,  I am encouraged by the increase in public
awareness and concern in the past few years. If this increased concern can be
focused on the  roots of the problem, as typified by the three P's, there is a
chance that new attitudes will develop, and solutions will be found.
12-10

-------
AUTHOR
      ARTHUR C. STERN, a mechanical engineer, is Professor of Air Hygiene at the
      University  of North  Carolina  at Chapel Hill. He is  the current President and
      Chairman of the  Board of the Triangle Universities Consortium on Air Pollution,
      which  involves  Duke  and North  Carolina State Universities and the University of
      North  Carolina at Chapel Hill.  From  1935  to 1968 he was associated  with the
      Federal Air Pollution Control Program,  first in Cincinnati, Ohio, as the chief of
      its  Research and Development Program; later, in  Washington D.C., as the Assis-
      tant Director of NAPCA.  He is perhaps best  known as editor of the three-volume
      work. Air Pollution.

-------
Symposium Summary
           13.   UTILIZATION OF  AIR  POLLUTION  MODELS


                       ARTHUR C. STERN
            University of North Carolina at Chapel Hill
It would serve you poorly to have me now recite a series of abstracts of the
papers  you have already  read  and  heard, and  of  the  discussion you  have
heard  and will  be able  to read in final form in the  Proceedings  of  this
symposium. Rather, I  can serve you best by assessing the  impact  of  our
deliberations.
If any  one statement could characterize this symposium, it is that this  was
the one in which a "Gaussian" distribution was qualified by, "if that is not a
dirty word." Those  who  had  incorporated in their models, diffusional pro-
cedures developed to  a  state  of increasing sophistication over the  past  5
years were here on the defensive against those who,  looking ahead,  foresee
newer procedures displacing them over the next 5 years.

This  was the symposium at which  people were  first  able to say  to  each
other, "You have an elegant model, but is it the one  we need? Does  it meet
our objectives for  a  model?" Yet statements of such  objectives were  alluded
to only peripherally by the speakers and discussers.

NEEDS AND  OBJECTIVES IN MODELING
Perhaps, the reason a statement of our needs and objectives in modeling did
not  come  through  loud and  clear is that there is a complete spectrum of
needs and  objectives and that  no single approach can satisfy  all these needs,
equally. We have heard of dimensional  scales, ranging from the global scale,
noted  by our  banquet  speaker, to that of the single city block, discussed by
several  of  you.  Let  us digress to look  at the characteristics  of these scales.
                                 13-1

-------
 They have a unifying feature in that, as the horizontal  scale is changed, there
 are corresponding changes in the vertical and temporal  scales (Table 13-1).
             Table 13-1.  SCALES OF AIR POLLUTION SYSTEMS
System
Global
National
State
Regional
City block
Vertical scale
Atmosphere
Stratosphere
Troposphere
Lowest mile
Heights of buildings
Temporal scale
decades
years
months
days
hours
The time scale in Table 13-1 can be interpreted to mean either the time for
effects  to  manifest  themselves or the  time inherent  in  taking effective,
corrective action.

Purposes Served by Models
Beclouding  a statement of objectives is the  fact that models serve two quite
diverse  purposes.  On one  hand, models serve to  test our understanding of
the  physical nature of the  atmosphere and  the  sources we are  modeling. If
we can't model it, we don't understand  it.  A decade ago, we tended to say
that we understood  these  phenomena, but  lacked the  programming, com-
putational,  and statistical  capabilities  to resolve  the  problem. Today,  the
tables are turned; we have  programming, computational, and statistical capa-
bility far beyond our needs, but lack the  physical inputs and concepts to
fully exploit these capabilities. A model, thus serves to test these concepts
and  inputs. In  this application  models need  to resolve fine detail and to have
fine structure, otherwise they cannot test  concepts  and  inputs.

On  the  other  hand,  models are used  as an aid to decision  making  in air
pollution  control,  and city and  regional planning. In this application we
could not care  less about fine structure and  fine detail.  In fact, fine structure
may  obscure   rather  than  elucidate,  and we may  have  to  seek  ways  to
surpress  the structure in the output in order to make it  effectively useable
for the decision-maker. It is small  wonder  that people  talking about these
two  aspects of model  making may have some problems communicating with
one  another.
Tactical and Strategic  Models
Another thing that did not come through loud and clear in this Symposium
was the differing  tactical and  strategic  roles (Figure 13-1) that  models are
called upon to perform. Only by clearly delineating these differing roles and
13-2

-------

STRATEGY FOR
; .AIRTOLLUTION CONTROL

AIR
QUALITY
STANDARDS
J



SOCIAL
: AND
* POLITICAL
; CONSIDERATIONS
i


:•: AIR
;;i QUALITY
CRITERIA
f 	



•*

EMISSION
STANDARDS
4




EMISSION
ALLOCATION :
I


SOCIAL
AND

POLITICAL '
CONSIDERATIONS:









SOURCES AND THEIR
CONTROL





SOURCES

ALTERNATE
PRODUCTS
AND
PROCESSES
i


COST
EFFECTIVE-
NESS














CONTROL
METHODS


i
COST
FUNCTIONS



DAMAGE
FUNCTIONS




-«!






POLLUTANTS
AND THEIR EFFECTS







POLLUTANT
HALF-
LIFE
i

AIR
QUALITY
t

AIR
POLLUTION
EFFECTS


•—


i
r


POLLUTANTS
EMITTED
^

TRANSPORT
AND
DIFFUSION
	 i




ATMOSPHERIC
CHEMISTRY ^


TACTICS FOR
EPISODE CONTROL


->


EPISODE
CONTROL
TACTICS


SOCIAL
AND
POLITICAL
CONSIDERATIONS


AIR
POLLUTION
POTENTIAL
FORECASTS


CO
CO
Figure 13-1.  A model of air pollution system.

-------
recognizing their differing  dimensional and  temporal scales, can  we plot a
rational path for future diffusion-model development. The tactical model, or
perhaps  better  stated, the  model  for  the  tactics  of  episode  control, is
inherently regional in scale, limited to a vertical scale of the lowest mile and
a  temporal scale  of  hours.  In  contrast to this,  strategic  models are best
developed  on state or national scales, since economic decisions and trade-offs
inherent in the strategy of  air pollution control are statewide, nationwide, or
even (as  in the case  of  fuel oil  and  gasoline  supply)  worldwide  in  their
impact.
If a regional  model  is primarily tactical,  rather  than strategic, there are
implications as to its  diffusional  component to be considered. Here,  I would
like  to  read  into  this record  a paragraph from  a paper  I  presented  in
Australia a few months ago.1

"Some of  our regional air  quality data systems may eventually  prove to  be
white elephants,  or whatever other expression  we use to signify  possessions
entailing great expense out  of  all proportion to  their usefullness to  their
owners. Some of these mistakes, which could have  been avoided if the data
needs of  the jurisdiction  had, themselves,  been subject to proper  systems
analysis, are the  result of a widespread misconception of  the role of an air
quality  monitoring network during  an alert.  In the  United States, there is a
nationwide system  for forecasting periods  of high-meteorological-air-pollu-
tion potential, before  they  occur. The greater the extent and duration of the
stagnation  that  is  forecast,  the  more likely the  readings  at  scattered air
sampling  stations  in  the area,  the  more  representative the data from the
central  station,  and  the  less the necessity to  interrogate a  multiplicity  of
remote  stations. If the central station records the predicted  rise in pollution
levels (particularly  in  hourly-average-time data)  and a forward meteorological
forecast for a continuation of the stagnation  is made, conditions exist for
taking  corrective  measures.  The existence of  short-averaging-time-telemetered
data from a  multiplicity  of  stations would have corroborative value, but
would  not be essential to the tactical decision-making process. During other
than extremely high-pollution situations, the data continually pouring out of
complex continuous sampling networks tend mainly  to  confirm, daily,  what
is already  known about diurnal  and seasonal variation and about the effects
of wind  direction  and velocity.  In  most of these systems, more time and
effort are  spent in keeping them operative, than in using the data acquired."
I  will  have more  to  say,  later,  about the data needs for strategic decision-
making.

Source- and Receptor-Oriented Models
We have  also somehow  or  another,  failed  to make  a distinction  in  our
discussion  at this Symposium between source-and  receptor-oriented  models.
We have talked only  about  source-oriented  models, but  we all  know that
13-4

-------
much  work  is under way  on the receptor-oriented type. In the same  paper
referred to above,1  I  offered a  rudimentary, statistical, air-quality receptor-
oriented, model and, again, will quote:

"Both  Larsen2  and  the author  have  discussed  the  properties of frequency
distributions of air quality  data and of the so-called arrowhead  charts (Figure
13-2)  that display them. However, until now, these data and charts have not
been put to use to analyze  the respective  influences of source and weather
on their make-up. If  we were able  to separate the source factors subject to
human control, from the weather factors beyond  such control, we would be
able to synthesize the distributions of  air quality data that would result from
the application  of specific  control strategies.  We  would  also  be able  to
compare them  with  air  quality  objectives,  expressed  in  like format,  to
determine which strategy comes closest to effecting a match.
second
 1
                                   AVERAGING TIME
                           minute        hour        day       month    yeat
                          1    5 1015 30  1 2 4 8 12 1  Z  4  7 14 1 2 3 6  1   3   10
                             3.412     1.359   0.635 0.<2J       0.121     0.066
                                EXPECTED ANNUAL MAXIMUM CONC IN PPM
                                           GEO. MEAN FOR 1-HR AVE. IS 0.044 PPM
                                           STD. GEO. DEVIATION IS 2.4«
                                           70 PERCENT OF HOURS HAVE DATA AVAIL
           D.0001    0.001
                        0.01     0.1      1      10
                                  AVERAGING TIME, hours
                                                       1,000    10,000
    Figure 13-2.  Arrowhead chart (produced  by computer)  for
    concentration vs.  averaging time and frequency for nitrogen
    oxides in Washington,  D. C. - - December t, 1961 to Decem-
    ber 1, 19642.
 Figure 13-2 might have as its components, figures that look like  Figures 13-3
 and  13-4, respectively,  the weather  factors  and the  source factors.  The
 analysis of Figure 13-4 for  its individual  components would  be  the converse
 of the emission inventory approach, in that the latter seeks  to  arrive at the
 same result through  synthesis,  whereas  the approach  just outlined  seeks to
                                                                      13-5

-------
CO

en
     TJ

     to'
     c

     CD
     c/3
     o
     c
     -^
     o
     (0



     I
     q

     OT
     I
     Q.

     O
     3"
     EC
£2
z
                  FREQUENCY, percent
                     CD ^ -n
                       ~
         10 05
         en z "H
         *"-  c:
            ,
           0-nr
in*   T
> o ^   I

s.2*   I
^  m   *-
                                K)              -

                             TEMPERATURE - °F*

                          AUTOMOBILE  UTILIZATION

                          ELECTRIC POWER  DEMAND

                          MANUFACTURING ACTIVITY

                       REFUSE INCINERATION  ACTIVITY
                                                                 in
                                                              -  s
                                                    o
                                                    o
£2
z
CD

^


m
                                                                   to


                                                                   CD


                                                                   w
                                                                    I
                                                                   CO



                                                                   §
                                                                   CD


                                                                   3-
                                                                   CD
                                                                  3-
                                                                  CD
                                                                  ffi
                                                                  Q.
                                                                   IVJ
                                                                                     f   ?
                                                                         £2
                                                                         z
                                                                         en
                                                                  ^^-     rn
                                                                                          FREQUENCY, percent
                                                                           -n >
                                                                           OT,T
                                                                           1-15
                                                                           >o,z
                               WIND VELOCITY I /mph

                              INDEX OF TURBULENCE-

                          LOW-LEVEL INVERSION DURATION

                           MEAN MAXIMUM MIXING DEPTH-

                                  PRECIPITATION'
                                                                                                                           m
                                                                                                                           n

-------
arrive at it through analysis.  The two approaches should tend to check and
reinforce  each  other, and thus improve  our chances  of determining the
relative influence of various source categories across the averaging-time spec-
trum. This should give us useful  leads to control strategies."

Averaging-Time Considerations
Having just  shown  a model  that  looks at the broad  spectrum of averaging
times,  I would  like  to  point out  something we sometimes  lose sight  of,
namely, that the adverse effects of  air pollution  have  inherently different
averaging times. The range is  as follows:
                                                      Averaging
        Effect                                             time
        Corrosion  ...          .          ....               year
        Soiling   ....                        .    .   month
        Health  . .                           ....           day
        Visibility . .                              .....   hour
        Vegetation damage             ....             . minute
        Odor	minute

 It follows that the  ideal strategic model will  resolve all these averaging times,
 not  just one of them. By the very nature of averaging time, once we resolve
 a  relatively  short  one,  all   longer ones become available to  us. Strategic
 models should, therefore, preferably resolve to averaging  times, measurable in
 minutes.
 In contrast to this, tactical  models,  being primarily  concerned with protec-
 tion of the  health of a regional population from the effects of stagnation,
 need not  resolve to such short averaging times.

 CONCLUSION
 In conclusion, let  me repeat what our banquet  speaker said  last night. We
 have come  a long way  in our ability to model urban  air pollution since our
 conference  in  Cincinnati on The Air Over Cities, a decade ago. We then
 started developing  first-generation  transport and diffusion models for first-
 generation  computers. We are now developing second-generation models for
 third-generation computers. The computers  have  gained a generation  on our
 use  of them for modeling. It behooves us now   to develop third-generation
 models to take full advantage of the generation of computers available to us.
 Yet, in doing so, let us heed the warnings we have heard at this symposium:
 "It  gains  us naught, to apply better programming  to  inadequate  physical
 input data and inadequate physical concepts."
                                                                      13-7

-------
REFERENCES

1.   Larsen,  R.  I. A New Mathematical Model  of Air Pollutant Concentra-
    tion, Averaging Time, and  Frequency. J. Air Pollution Control Assoc.
    79:24-30, January 1969.
2.   Stern,  A.  C.  The  Systems Approach  to  Air  Pollution  Control.  In:
    Proceedings of  the  Clean Air  Conference of  the Clean Air Society of
    Australia and New  Zealand, New South Wales Dept.  of Public Health,
    and New South Wales University, Vol. 2. Sydney, May  1969. p. 2.4.1. —
    2.4.22.
13-8

-------
AUTHORS
      DARRYL  RANDERSON,  Ph.D. is a research  meteorologist with  the  National
      Oceanic and Atmospheric  Administration  (NOAA) Air Resources Laboratory in
      Las  Vegas,  Nevada,  and is responsible for the development of prediction tech-
      niques in forecasting atmospheric pollutant content
      RALPH C.  SKLAREW,  Ph.D.  is a member of the Theoretical Sciences Division
      of Systems,  Science, and Software, has  engaged in applications  of computer
      techniques to modeling high altitude chemistry, and transport and chemical inter-
      action  of air pollutants.  He  holds a Ph.D. in physics from the  University of
      California.
      LESTER  MACHTA, Ph.D. is  the Director of the NOAA Air Resources Labora-
      tories,  holds a Ph.D. in meteorology from the Massachussettes Institute  of Tech-
      nology. He  has  also worked for the  U.S. Weather Bureau in the areas of radio-
      active fallout, atmospheric  trajectories, and air pollution.
      D.  BRUCE  TURNER, on assignment from the  NOAA  Air Resources Labora-
      tories,  serves as  senior meteorologist  for the Bureau  of Critera and Standards at
      NAPCA. As  a research meteorologist he was  involved in developing dispersion
      models for urban areas with  NAPCA's Division of Meteorology. He is a  graduate
      of Carleton  College and the University of Michigan.
      J.  E.  CERMACK,  Professor-in-Charge  of the  Fluid Mechanics Program  at Colo-
      rado State University,  is director of the Fluid Dynamics and Diffusion Labora-
      tory. He earned  a  Ph.D.  from Cornell in engineering mechanics and  was  awarded
      a  NATO  postdoctoral  fellowship  for study  at  Cambridge in  1962.  He  has
      pioneered the development of special wind tunnels for simulation of the lower
      atmosphere and  directs a research program on low-level winds and diffusion over
      urban areas and complex topographies.
      HARRY MOSES, a Meteorology Group Leader at the Argonne National Labora-
      tory, holds an M.S. from the  University of Chicago, a Diplomate in  Meteorology
      from the University of Michigan and has been an Instructor at the University of
      Chicago, where he  was also a member of the Thunderstorm Project,  1945.
      L.  A. CLARENBURG,  Commissioner of Environmental Hygiene of the Rijnmond
      Authority in Rotterdam, has been head  of the physical division of  the chemical
      laboratory of the  Belguim National Defense Research Council. He received his
      Ph.D. in theoretical chemistry in  1960 at the University of Utrecht.
      BURTON FREEMAN,  Vice-President of Systems, Science, and Software,  is one
      of the founders  of that company. He is an authority on large, computer code
      application to radiation  flow  and hydrodynamics. He holds a  Ph.D. in physics
      from Yale.
      DONALD  H.  PACK  Deputy  Director,  NOAA  Air Resources  Laboratories,
      received his meteorology training at New York University, University of  Chicago,
      and  the University of Puerto Rico, and has  worked in forecasting, turbulence,
      and air pollution research.
      ROBERT £ STEWART,  Assistant Professor  of Enviromental Engineering, Uni-
      versity of Florida,  is the author of several papers on atmospheric and oceanic
      diffusion.  He received a Ph.D. from the University of  Waterloo in 1966.

-------
                                              14.   DISCUSSIONS
INTRODUCTION

The  following discussions were  submitted  in  writing to the editor, sub-
sequent to the Symposium. Attendees were given an open invitation to make
comments on the Symposium topic and all of their responses are included in
this chapter. Every author whose work  was  questioned was given  an oppor-
tunity to  read the question and  to write a rebuttal if he felt one was needed.
The chapter is divided into two sections; the first includes the discussions of
Symposium papers and the  second contains brief treatments of some  addi-
tional approaches to multiple-source urban diffusion models.

RESPONSES TO  INDIVIDUAL PAPERS

Lettau paper
                           Frank Pasquill
Regarding the effect  on horizontal  spread of  the  turning  of wind  with
height, I  would like to refer to a matter that is discussed in more detail in a
paper that will shortly appear in  the proceedings of a symposium on Recent
Research  in Air Pollution held by  the Royal Society in  November, 1968.
Evidently, it is necessary to distinguish between the general distortion of a
plume and the ultimate contribution of this distortion to enhanced spread at
a given level.  It turns out that  there  is  a substantial time lag between these
two phenomena. Examination of  field data for stable conditions, available at
the  time  of composing  the  foregoing  paper, indicated that the  effect on
spread at a given  level was unimportant in  relation to the spread produced
directly by  the horizontal component of turbulence within about 5 km from
an elevated  source and about 12 km from a ground source.
                                 14-1

-------
Pasquill paper

                           Kenneth  L.  Calder
It may be appropriate to emphasize that, with very few exceptions, all urban
pollution models developed in the past decade have been derived, with only
minor  modifications, from  Dr. Pasquill's  original  broad estimates of spread
made nearly 10  years  ago  and, frequently, without any serious attempt to
modify  his  estimates for the effects  of  increased urban roughness or the
urban  heating.  It  seems very  difficult to  break with  these  long-standing
affiliations with the Gaussian plume and ay(x),  az(x). In fact, there may be
good reasons for not trying to, particularly since  Dr. Pasquill  now indicates
how some appropriate  modifications in the original estimate might be made
to allow for urban factors. As Dr. G.D.  Robinson has recently  noted, it  is
also  much  easier  to keep  track  of the  conservation of mass in  terms  of
modified Gaussian-plume (or puff) type models  than  in  terms  of theoretical
models derived  from the eddy-diffusity concept.
If some  pessimism may  have  been caused among urban  pollution  modelers
by Marsh and Withers'  recent conclusions based  on an attempt to model the
SO2  concentration distribution in Reading,  England, then  I think that Dr.
Pasquill  has provided  the  necessary  antidote.  Recognition  of the random
errors inherent in  urban pollution  models,  such as those arising from  inac-
curate  specification of  wind direction, is very important. Dr. Pasquill's plea
for more consideration  of the  modeling aspects  for probability distributions
of pollutant concentration, rather  than  just the spatial distribution of average
values appears to  be very timely.


                           Darryl Randerson
In the development of the  Gaussian plume models one  of the fundamental
assumptions  is  that the diffusion is isotropic, so that the "exchange coeffi-
cients" do not vary specially or temporally. Consequently, how do we justify
utilizing  variable   a  values  in  the  Gaussian   plume  models?  Also,  what
justification  is given for varying the atmospheric  stability when another  basic
assumption of the model  is  that the atmosphere is  neutrally stable?

                             Author's Reply
In my contribution to the general discussion I  have included some general
comments on the  principle and  practice  of the 'Gaussian'  assumption in
dispersion models.

However,  I  must confess to being  rather mystified by the emphasis in Mr.
Randerson's  question.  I  have  certainly  not  had the  impression  that the
assumption  was necessarily  prohibited  either by non-isotrop/c  turbulence or
by non-neutral density stratification.
14-2

-------
                          Sin'ichi Sakuruba
In  the  analysis of St. Louis dispersion data I  think the vertical  spread was
overestimated. The tracer material,  even if  released from  near the ground,
tends to be  lifted  bodily, especially  over an urban area.  In the St. Louis case
the overestimation of vertical spread of concentration is unavoidable,  if the
analysis is made assuming a constant height of the tracer-cloud. My  inference
is  based on  the result of dispersion data analysis made in Japan using the
vertical concentration profile  data.

                             Author's Reply
I  am not  sure  whether  Dr.  Sakuruba's comment is offered  primarily as  an
explanation  of the  discrepancy  in  the predicted and  estimated values  of
vertical spread in  daytime conditions or, generally,  as a reason for question-
ing any apparent  agreement.  Clearly the mechanism to which he refers must
be included  in the complex of uncertainty that  applies  to  values of vertical
spread  inferred from surface  concentration, but there is, as yet, no  quantita-
tive indication of the significance of  the effect in the cases analysed.


Calder Paper
                          Ralph C. Sklarew
Simple models, with universal applications, are much sought. Unfortunately,
the relationship between averaging time and spatial resolution limits practical
simplicity. A model so simple that  the required calculations can be carried
out by hand must,  of  necessity, have only gross,  spatial  resolution  and a
correspondingly  long-period  time average.  In  Dr.  Gifford's  model  annual
averages do, in fact, correspond to  a spatial resolution  consistent with hand
calculational efforts. For abatement  strategies and forecasting, where averages
over shorter intervals are needed, finer spatial resolution  is required, resulting
in more lengthily computations.
A  compromise  may  have been anticipated in two comments by Mr. Calder.
The form  of the transfer function,  R, depends upon the path traversed, not
just on the  path length.  Adopting this requirement and using a field or grid
model  would incorporate path  dependence and,  as a secondary benefit,
eliminate  the repetition of  calculations necessary  for  each  source-receptor
pair. The  second comment concerns the possible benefits to be gained from
a  parameter  study with a large, complex  model. A parameter  study with
such a  model would indicate simplifications possible for each specific appli-
cation. This, then, might become the third-generation model: a  large, com-
plex, grid model whose  simplified offspring  are  uniquely adapted to a class
of applications.
                                                                     14-3

-------
Johnson, Ludwig and  Moon Paper

                            Kenneth Calder
Dr. Johnson has indicated that the area-source simulation used in his intra-
-urban model  cannot be applied for very fine resolution of the concentration
pattern, say on  a  scale of less than a city block,  and that for the latter an
entirely different  modeling  concept  is necessary. This  raises an important
general question regarding the ultimate uses of the model and  just how much
knowledge of fine structure is really  necessary,  that  is:  what are the most
important spatial and  temporial scales in  regard to damage effect or inhala-
tion risk for the exposed population. Some clarification in terms of a system
analysis will be  necessary if  there  is to be any guarantee that the urban-dif-
fusion modeling efforts are consistent with, and responsive to the characteris-
tics of the overall system model.  The sensitivity  of the overall model to the
urban diffusion model, when  regarded  as an  input,  should  determine  the
detail and accuracy required for the latter.

                       Response by W.  B. Johnson*
Dr. Calder's remarks are very appropriate, and I fully agree that the temporal
and spatial scales treated  by  a diffusion submodel should be compatible with
the objectives of  any  broader systems model  of  which it is part. The
fundamental,  practical reason  why the  fine  scale must  be  treated in  our
model is that  the  only carbon monoxide (CO) data available for verification
consist of point measurements near streets and  buildings. The  most generally
available data  are from Continuous Air Monitoring Program (CAMP)  stations,
which are located  at streetside,  usually at intersections. It is well known that
these  data strongly  reflect siting effects, so much so  that intercomparisons
between cities are  essentially meaningless.
We are attempting  to design  flexibility  into  our model so   that different
versions  of  it can be  used for  different  purposes. Some  specific foreseeable
applications include the following:
1. Interpretation of  the  significance  of  existing  air  quality   measurements
   (correction for  siting effects)
2  .Estimation  of current urban-wide  air quality  patterns (augmentation of
   existing point measurements)
3. Selection of  optimum control  strategies  (when used  with  a suitable  sys-
   tems model, which  includes economic  considerations)
4. Prediction of effects on future air quality of proposed or forecast changes
   in source distribution and  strength.

The temporal  and spatial scales in .the completed model will be selectable to
fit each of these applications.
*l would like to thank Mr. Wayne Ott of NAPCA and Mr. F. L. Ludwig of Stanford Research
 Institute for sharing  their ideas on this subject with me.
14-4

-------
In  general, simulation  models  should  probably deal with the temporal and
spatial scales of pollutant concentrations that are most significant in terms of
health  and other  effects. The provision of such air quality  criteria is  the
domain of the medical and biological researchers and air quality management
officals. Present knowledge of the time scale factor is presumably reflected
in  current air quality  standards. However, the scale in space has not been
similarly  treated, nor does  the definition of air quality include any spatial
dimension. This situation  is particularly serious when dealing with predomi-
nately ground-level sources, such  as those of CO.  Here,  the  location of a
point  measurement of  air  quality  is  all-important.  The  problem could  be
eased considerably if air quality were  defined and  measured in terms of line
or area integrals. At the very least, there should  be  standards prescribing that
point  measurements  be made  at specific  locations relative  to local sources
and buildings. This would substantially facilitate model verification, as well
as air quality  management.
Roberts, Croke, and  Kennedy Paper

                          Kenneth L Calder
Although this sophisticated model using superposition of instantaneous Gaus-
sian  puffs will  hopefully mark the advent of a new generation of multiple-
source pollution models, there may be some justification for the feeling that
the trend  has yet to be established. That is to  say, an acceptable degree of
agreement between  the model  predictions and observations may  not  be the
only  criterion.  What  is frequenty  surprising  in some  applications  is  the
closeness of the agreement in  spite of the inherent inaccuracies of the model
inputs and the  extreme nature of the  physical idealization  involved.  It is
almost as  though  the  condition for  mass  conservation, common  to  all
models, in itself provided a  very strong   rather than  an extremely weak
constraint on  the concentration  field.  Until it has been shown that  such
extreme complication of the present type of model is absolutely necessary in
order to  fit the  practical observations  with  the required  degree of accuracy,
and that  this complication is  supported  by  sound physical principles, a case
could be made for invoking  Ockham's  razor and continuing  to  use  similar
models.

                           Author's Response
We agree  with  Mr.  Calder that,  while the initial  validation  of  our urban
atmospheric pollution model is encouraging, there is as yet no demonstration
that  success is contingent upon  the  use   of the integrated  puff transport
kernel. Because it  is by  nature a  transient  model,  the  integrated  puff
approach is particularly well-suited to  regions characterized  by significant
diurnal cycling  of windspeed and  direction.  For example, the work of Start
                                                                     14-5

-------
and Markee,* with which  we have recently become acquainted, indicates the
importance  of a puff model in evaluating dosages from  hypothetical radio-
activity releases  in  such  regions.  In this sense,  Chicago  may not be the
optimal city  to  test  the  transient  aspects of  the  model  although diurnal
variations in  mixing-layer depth  and  periods  of  varying windspeed  and
direction  seem to be  modeled  adequately.  As suggested by Mr. Calder, we
are currently  investigating  the sensitivity of our Chicago results to the choice
of transport kernel.
In reference  to  "the extreme complication" of the model,  it is important to
note  that the greatest  demands on computer storage  do not  lie in the
necessity  for  numerical  integration  of the dispersion kernel but are tied to
the extensive efforts to  establish seasonal and diurnal emission patterns and
then  link  these refined  estimates   with  hourly  weather  data and hourly
estimates  of  mixing-layer  depth  to form the  composite  upon which the
integrated puff kernel  operates.

                             Lester Machta
It is my view that uncertainties involved  in predicting weather elements six
or  so  hours  in  advance  raises serious questions about the  usefulness of
forecasts of  air quality  made a similar time in  advance.  The verification of
low-level winds even in the simple terrain of suburban airports has not been
particularly  good; one would imagine that  forecasts of wind  direction and
speed  in  a  city  setting would  be  even  worse.  Since the  high-air-pollution
episode normally  involves a weak windspeed, the problem is  even further
complicated.  Until it  is demonstrated that  forecasts of  air quality can be
provided six or so  hours  in advance I  would  be very  loathe  to advise air
pollution control officials of such a possibility occuring in the near future.
Hilst Paper
                          Kenneth  L.  Calder
Dr. Hilst  indicates that variations in  windspeed were not considered  in the
sensitivity analysis since the  dilution effect  at  the source and the time-of-
flight effect  in  decay of pollutant would tend  to cancel  one another. This
argument  is not clear since we are  considering a function of the form  u"1
exp  |  Axil"1  }   . With the  trajectory analysis that is involved in the TRC
simulation model, it  might be expected that the effects  of errors  in wind-
speed would be very complicated.  If this is so,  then I  wonder about the
validity of the conclusions regarding the decay parameter for S02  that enters
the model in the  combination  Xxu"1. I  personally would feel that juggling
*Start, G. E., and Markee, E. H. Jr., Relative Dose Factors from Long-Period Point Source
Emissions of Atmospheric Pollutants, AECL—2787, Proc. USAEC  Met. Info. Meeting,
Chalk River, Ontario, 1967, pp 59—76.
14-6

-------
with the half-life of SO2 and reducing it from 3 hours to 1 hour in order to
improve  the agreement  with observation  is  a rather risky business  at this
stage of the game, before  a  full error analysis has been  made for windspeed.
Finally-, although  Dr.  Hilst  says that the TRC model  has been extensively
reported over the  past  2  years, it is my  feeling that the available accounts
may not  be sufficiently detailed to allow  independent application  of this
model  to  other  pollution problems. I would like to urge Dr. Hilst to publish,
as soon as possible, a  full technical  report so that others are able to utilize
this sophisticated model.


                            D. Bruce Turner
 In  additon  to the eight errors listed by Dr.  Hilst, I  think there  is a ninth,
which  might be called grid-size error. This is most noticeable  when  calcula-
 tions made at grid-points are  compared  with measurements made at non-
grid-point locations. Specific wind directions will cause individual pollution
 plumes to affect the  measurement station, yet pass between grid-points and
 vice versa. This  can be overcome by models able to calculate concentrations
 for any point. Where this effect was noticed, the model  being  used made
 calculations only at the  center  point of each grid  square. This was because a
 technique had been developed to determine  the concentration at the center
 point  of an area-source square resulting from the emissions coming from
 within that square, but  was not developed for points other than  the  center.
 Even with a model restricted to making computation at the  center  of each
 grid square this grid-size error can  be made  less  noticeable if measurement
 stations used for model  validation are also located as close to the center of
 grid squares as is practicable.


 Sakuraba  Paper
                             Frank Pasquill
 The measurements  of vertical  spread described in Dr. Sakuraba's paper are
 obviously of considerable basic value and will no doubt be widely  studied.
 However, one  should  be careful not to place a literal  interpretation on the
 description  of the  observed vertical spread in terms of the stability categories
 that I  have specified  for use in the absence of more  sophisticated  data  on
 diffusive  conditions. As I noted in  my own  paper in this symposium these
 categories were specified  for  airflow, characteristic  of a land surface with
 small  to  moderate roughness.  On the other  hand, it  seems likely that Dr.
 Sakuraba's  results  also  reflect the  influence  of sudden and  appreciable
 changes of roughness on  an air stream passing from  the  sea  over  a rather
 irregular coastline instead of the effect of stability differences alone.
                                                                      14-7

-------
Fortak Paper
                          Kenneth L. Calder
Dr. Fortak's discussion again emphasizes the point first raised by Dr. Pasquill
at this meeting, that while a calculated spatial field  of  short-period average
pollution  concentration cannot be expected  to agree very closely with that
actually observed, the statistics obtained from  an ensemble  of such  calcu-
lated concentration fields may well be in close agreement with reality.

As a small but important detail I  was interested  to see  that Dr. Fortak has
considered the numerical integration errors associated  with the choice of grid
size for the area-source specification. His conclusion that an area size of the
order of  50  meters by  50 meters  may  be required  for  a  satisfactory
representation of a  large  number  of point sources,  is noteworthy, since, I
believe, this   is  much smaller than the size used  in  some  urban models
currently being recommended for operational use.

Sheih, Davidson, and Friend Paper

                          Kenneth L. Calder
One  point I  found  rather confusing in Drs.  Sheih,  Davidson,  and  Friend's
paper was the initial statement that the model was derived from the statisti-
cal theory for turbulent diffusion,  although  later  in the paper they stated
that  the  model  was semi-empirical  and not derivable from known  physical
concepts.  It would seem that the  latter is a more  true  description since  all
the adjustable parameters  and constants of the model are apparently derived
by fitting observational data for sulfur  dioxide in New York  City. If this is
so, then  agreement  between the  model predictions and actual  observations
may  be  less   impressive than for  some other models where  the parameter
values are estimated independently.

The  considerable effort made in this study to develop an  adequate method
for  numerically  integrating  the  emissions from  a continuous area source is
noteworthy and in strong contrast  to  the crude procedures  used in  many
other models.

Mahoney, Maddaus, and Goodrich  Paper

                             Harry Moses
One  must bear in  mind that the concentrations of a given pollutant such as
SO2, at a given  station, is a function of several variables. Windspeed is one
of these variables. It is possible in a multivariate system to find that several
of the individual independent variables  correlate  poorly  with  the dependent
variable, but when taken together,  show a high multiple correlation.
14-8

-------
It is also  worth pointing out that the windspeed  is a difficult variable to
define. For example, at some distances from an elevated source such as a tall
stack, an  increase  in  windspeed is associated with an  increase in pollutant
concentration on the  ground, because the stack plume rise is reduced  during
high winds. With a ground source  an increase in  windspeed results in  lower
concentrations downwind because of increased turbulent mixing.  One must
take such relations into  account when using windspeed  as a  variable in  a
regression type  model. Thus, it is  not  surprising to find that the zero order
correlation of pollutant  concentration with windspeed or even the reciprocal
of windspeed is unimpressive, even though the windspeed has an ultimately
important influence on the pollutant concentrations.
GENERAL  DISCUSSION

Vertical Circulations and Eddy Diffusion  Coefficients

                           James P. Friend
Until  a dynamic  meteorological model of urban atmospheres can produce
circulations which are physically reasonable, attempts to incorporate circula-
tions  having vertical  motions into  the  present dispersion models are equiva-
lent  to  introducing  new parameters  to an  already  highly  parameterized
system.
It has been suggested that solving a diffusion equation with eddy diffusion
coefficients (K's) would be more scientifically acceptable than application of
Gaussian plume formulas. It seems to  me, with the present state of knowl-
edge,  that this is equivalent to trading one set of parameters for another. We
simply have  no  known  physical  laws by which values of the K's can  be
determined as functions of space, time, and meteorological  conditions. There
has been a  plea  by Dr.  Pasquill and  others  here for more research  in
meteorological physics.  One  can  do no  better than  to second that plea.
Perhaps the resulting knowledge will eliminate  the  need to apply either of
the two systems mentioned above.


Generalization  of Vertical and Crosswind Spread

                            Frank Pasquill
I  am, of course, primarily interested in the meteorological  physics required
in the design of  systems for the  prediction  of air pollutant concentrations
and in the need, which  seems to be  inevitable, for the adoption of some
generalization about the vertical and crosswind spread of material.
Some  models appear to  generalize about the standard deviation of particle
spread in vertical and horizontal  directions,  ay and CTZ, by using practical
                                                                    14-9

-------
observations of the distribution of urban air pollutants. Although not unrea-
sonable,  it should be emphasised  that such  a procedure cannot be accepted
automatically  as justification  for applying  the model to flow and  terrain
conditions other  than those obtaining in the pollutant survey. For the most
useful  generalization  we must seek  independent specifications of  ay  and az
in terms of meteorological  parameters  that  can  either be  measured  at the
time in question or estimated from routine  meteorological data.

The specification  of ay  and  CTZ  may  be explicit  (as from  the statistical
theory) or implicit (as from the gradient-transfer theory, and  solution of the
so-called  classical  equation  of diffusion).  It  seems to be the view of some
participants in this symposium that there is something conceptually superior
in  the  use of  the  implicit  classical diffusion  equation.  I   find this view
surprising in light of the essentially empirical  nature of this approach both in
principle and in application  (i.e. no a priori specification of K is available to
us). On the other hand, I  have not claimed  (as stated in one contribution to
the discussion)  that  the K-approach is entirely  unacceptable. In my own
paper,  however,  I  do argue that the approach is physically plausible and
practically  equivalent or superior to  other approaches only  for the case of
vertical spread from a source that  is effectively at ground level.

Finally on this question of the specification  of cry and az, it may be worth
remembering that, in general, the most  accurate specification appears to be
possible  for short ranges from  continuous sources. On the other hand, the
greatest  uncertainty  still  seems to  apply  to the spread  of  an  individual
isolated puff.
Regarding the incorporation of  estimates of spread in the prediction systems
I  have two other comments to  offer. The first concerns the term Gaussian,
to  which there  has  been frequent reference in  this symposium.  It may be
recalled that the Gaussian  distribution  follows from the classical diffusion
equation only when K is constant and, consequently, the spread is propor-
tional  to the square  root  of the time. This latter relation  has not been
generally  observed.  Furthermore, the empirical  support for the Gaussian
shape comes only from  ensemble averages  of time-mean crosswind distribu-
tions from continuous sources, or of vertical distributions from  effectively
elevated  sources.  It has been noted,  however, that the  precise magnitude of
crosswind spread seems  unlikely  to  be of prime  consequence  in the distri-
bution arising from a  close network of multiple sources!
On the  other  hand,  there  is no real basis for assuming vertical Gaussian
distribution from a ground-level source—instead the theoretical and empirical
indications  point  to  a  shape  between  Gaussian  and simple exponential.
Furthermore, there is  no reliable  guidance  available on the distribution in a
puff. From  all this it may seem rather  irrelevant to attach great significance
to  the assumption of  Gaussian distribution. The Gaussian  shape may be
14-10

-------
convenient to use, but  a  'center-concentrated' plume with a linear fall-off of
concentration from the center  may  turn out to  be just as satisfactory, in
practice,  for  passive  plumes or puffs. (It seems that the distribution in a hot
plume  or puff, initially  dominated  as  it must  be by  the  circulation  and
entrainment  induced  by  relative motion, may, however, have a distribution
much flatter  than Gaussian).

My second comment concerns the suggestion that a 'sequential' puff repre-
sentation  of  a continuous source may  be desirable. While in  principle  this
may seem preferable,  permitting allowance  for variable  source strength  and
widely  variable wind direction and diffusive conditions,  I find  it difficult to
see how  the approach  can be advantageous in practice, unless it  is in  fact
possible to specify the trajectory and spread of  an individual  puff.  On the
contrary,  all experience  points to  the  fact that the spreads  of individual
puffs  in  the same  general   airstream  differ widely, and that reasonable
accuracy in specification  can  be expected only on the basis  of an  ensemble
average. In this context it does not seem to me to  have been made clear how
the envisaged 'puff' predictions are  expected to  be superior to those  that
would  follow from considering the effects of continuous releases over arbi-
trary intervals of appreciable length.

Finally, there have been several references to pollutant movement within the
confines  of  a  street,  and indeed  one  of the papers has considered some
aspects thereof. It seems  to me that there are two opposite extremes in the
types  of  estimates  that  might  be  attempted.  On  one  hand; as  already
in the general discussion, it  may be possible to  make a useful estimate by
considering,  on an  individual basis,  the effective enclosed volume  of air
available to accept the  emission from some position in the street. It may be
useful  to  add to this the refinement of a ventilation coefficient defining the
rate at which this air is changed, so providing a more realistic estimate of the
rate at which the pollutant  concentration  might build up.  At the other
extreme, for a statistical  estimate of the effect over a complex of streets, it
might  be  possible  to  make progress by  developing a two-layer model on the
following basis:
In  the lower (street)  layer one might  reasonably assume that  there is no
advection of the air flow  and  the pollutant but only a vertical exchange  with
the  upper (above  roof) layer. Only  in  this upper layer can one reasonably
expect  the  spread  to  follow  the  'laws'  previously  specified for  plumes.
Essentially, the  first requirement is to specify the distribution of sources in
the upper layer and in the lower layer.  The latter must then be represented
as vertical fluxes at the boundary between the two layers. The concentration
at any position in the  lower  layer will  be determined by  the  concentration
developed  immediately overhead  in  the upper layer,  by  the local rate  of
(vertical)  ventilation of the lower layer and by the local emission rate in the
lower layer.
                                                                     14-11

-------
Physical-Scale Model Advantages

                            J. E. Cermack
A basic deficiency appears in all of the numerical models of urban diffusion
presented in  this Symposium—a deficiency of vital importance, particularly
for the "source-based" models. This deficiency is  the  unrealistic represen-
tation  by  Gaussian  plumes  of diffusion from sources  located within the
urban complex. In particular, the diffusion of carbon  monoxide from street-
level sources  is dominated by local geometry of the city and local meteoro-
logical conditions. The only  feasible approach to  reaching some understand-
ing of this type of diffusion is through the study of  physical models—scale
models  of typical city complexes  placed in flows simulating thermally strati-
fied atmospheres. Models on a scale of  1:200 to  1:1000 such as the 1:400
scale  model  of diffusion  over  Denver  shown  in  a  motion picture at this
Symposium permit detailed diffusion  measurements which can  lead to some
general  conclusion on diffusion characteristics  in city complexes.  Utilizing
even a  modest amount of knowledge  gained  in this  manner,  the  design of
new  numerical models should lead to gross improvement in  their predictive
capability.  Physical  models can  be utilized to great  advantage for "receptor-
based"  numerical models  to help understand the relationship between  source
location and  receptor  response.  This  can be  accomplished  for particular
urban settings under  specified meteorological conditions.

In conclusion,  I wish to  emphasize that physical  modeling of diffusion can
be of great assistance in the operational aspects of air pollution control. This
will be particularly true if the modeling  is accomplished by a team composed
of specialists  in  fluid  dynamics  and  atmospheric sciences  working in  a
meteorological  wind  tunnel of a  type  for  which similarity  of atmospheric
flows has been  verified.*


Second Generation Models—Finite  Difference Approximations

                          Darryl Randerson
Several  of my colleagues  have expressed concern about the limitations in the
Gaussian models  and, consequently,  forsee the advent of "second  genera-
tion"  prediction models.  I would  like to suggest that  these new models will
not be  Gaussian  models  and will be free of  the restraints inherent in such
models. The new generation  models will be both  descriptive and conceptual
and may represent finite-difference approximations  to initial-value problems
that arise in fluid dynamics. Special attention  will be given to the calculation
of time-dependent  fields so that  micrometeorological  phenomena  can  be
simulated   in  time  and  space  and  subsequently  utilized to  predict the
*Cermak, J. E.  and Arya, S. P. S. Problems of Atmospheric Shear Flows and Their Labo-
ratory Simulation. Int. Jl. of Boundary-Layer Meteorology ?:  3-23 January 1970.
14-12

-------
transport of  effluvia  over  non-uniform  terrain.  Initial input for these new
models will be generated from information and knowledge  collected  from
past and future  studies of urban  meteorology. These new models  will be
closely related to meteorological parameters, to chemical behavior of  effluvia
in air  and  sunlight,  to topographical  effects, and to the  generation of a
representative source  term.  Consequently, these models  may become  quite
complex;  however, they should  eventually lead to  a rather  fundamental
model  for  predicting the  transport and  diffusion  of matter  through the
atmosphere.


Tabluation Techniques
                             Harry Moses
A substantial amount of work has been carried out at the Argonne National
Laboratory on a method called the Tabulation  Prediction  Scheme. Since it is
relevant to the problem of developing  and  using  multiple-source urban air
pollution models, I should like to  describe it  briefly  and  show how it may
be applied.

Description of Prediction Tables
This technique involves the  use  of a set of tables that  show the  relation
between air pollutant concentrations and those meteorological variables that
affect  concentration  levels. Hourly readings are used. The meteorological
variables are arranged in an ordered sequence.  For each entry the associated
probability distribution of  pollutant concentration is  presented as shown  in
Table  14-1. For each combination  of meteorological variables, sulfur  dioxide
(S02)  concentrations are given, representing the minimum, the 25,  50, 75,
90, 95, 98, and  99 percentile values,  and the maximum. The interquartile
range,  the difference  in SO2  concentrations  between  75 and 95 percentiles;
the mean; the standard deviation; and the number of cases for each entry are
also presented. These  data represent actual  measurements obtained by the
Chicago Telemetered  Air Monitoring  (TAM)  stations during  a 2-year period.

To arrange combinations of  meteorological variables into an  ordered se-
quence, appropriate group intervals must be selected for each.

Wind  direction values may be grouped into three class intervals; time of day,
also into three; windspeed into five; and so on. A number  is assigned  to each
class interval, let us say, in increasing  order. Any combination  of meteoro-
logical variables  corresponds  to  a equivalent-digit  number.  If letters  are
assigned to each class of each variable, then any combination would corres-
pond  to an equivalent-letter word. Since the combinations of meteorological
variables are  ordered,  it is possible to look  up any  set of meteorological
conditions  just as one would  look  up a  name in a telephone book or a word
in a dictionary.
                                                                    14-13

-------
Table 14-1.  TABULATION PREDICTION TECHNIQUE
Wind dir
degrees
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
0-140
Ceiling,
feet
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
7000-****
Windsp,
kt
0-3
0-3
0-3
0-3
0- 3
0-3
4-7
4- 7
4-7
4-7
4-7
4-7
4- 7
4- 7
Temp,
deg F
70-79
70-79
70-79
80-89
80-89
80- 89
10- 19
10- 19
10- 19
20-29
20- 29
20-29
30-39
30-39
Hour,
CST
0- 3
4-15
16-23
0- 3
4-15
16-23
0- 3
4-15
16-23
0- 3
4-15
16-23
0- 5
4-15
Percentile values of SOjIppm) concentrations
Min
00
0.01
001
0.0
002
0.02
0.01
0.0
0.08
0.02
0.09
005
0 01
0 01

25
0 0
0.01
0 01
0.0
0 02
0.02
0.01
0 0
0.08
0.02
0.10
0.05
001
0 02

50
0 0
0.03
0 01
0.0
002
0.02
0.02
0.0
0.10
0.02
0.15
0.07
003
0.04

75
00
0.04
0 01
0 0
0 03
0 02
0 02
0 0
0.12
0.02
0.20
0 10
0 04
009

90
0 0
0.06
001
0.0
003
0 02
0.02
00
0.12
0.02
0.21
0 27
005
0 13

95
0 0
0.06
001
0.0
0 03
0 02
0.02
0 0
0.12
0.02
0.21
0 27
006
0 15

98
0 0
0.06
0 01
0.0
003
0 02
0.02
0 0
0.12
0.02
0.21
0 27
0 08
0 28

99
0 0
0.06
0 01
0.0
0 03
0 02
0.02
0 0
0.12
0.02
0.21
0 27
0 08
0 29

Max
0 0
0.06
0 01
0 0
0 03
0 02
0 02
00
0.12
0.02
0.21
0 27
0 08
030

75-25
00
0.03
0 0
0 0
0 01
0 0
0 01
0 0
0.04
0.0
0.10
005
003
007

95-75
0 0
0.02
0 0
0.0
0 0
0 0
00
00
0.0
0.0
0.01
0 17
0 02
0 06

Mean
0 0
0.03
0 01
0 0
0 02
0 02
0 02
0 0
0.10
0.02
0.15
0 12
0 02
0 06

St dev
0 0
0.0183
0 0
0.0
0 0050
0 0
0 0047
00
0.0200
0.0
0.0460
0 0864
0 0175
0 0650

rreq
0
6
1
0
2
1
3
0
2
1
4
4
17
24

-------
The probability distribution  of SO2  concentrations, based on  past history,
is given on the same line in  the  table as the combination of meteorological
variables. In practice, the assigned letters or  numbers are converted to actual
ranges as shown in Table 14-1.

Applications of the Tabulation  Prediction Scheme
The Tabulation  Prediction Scheme may  be  used  in at least  three different
ways:
   1. As an adjunct to a source-oriented model.
   2. As a means for monitoring changes in  the source distribution over an
     area.
   3. As a receptor-oriented urban air pollution model.

Use with a Source-Oriented Model
Assuming that the  sources remain constant for the same hour of the day, the
same season,  and the same meteorological conditions, including air tempera-
ture, then the  source-oriented  model  provides  a  single  concentration value
for each point  on  the calculation grid over an urban complex.  The Tabula-
tion  Prediction Scheme shows that, historically, for example, during the  past
two years, not  a single value but a distribution of values has  been observed.
The  Tabulation Prediction Scheme  thus provides a measure of the upper
limit of accuracy  obtainable,  at  least at the  monitoring stations,  by the
source-oriented model. The source-oriented model  may overcalculate in some
parts of a city  and undercalculate in others.  Similarly, systematic errors may
occur  under  some meteorological conditions such as different  wind direc-
tions. The  information provided  by  the tables of the Tabulation Prediction
Scheme could,  thus,  serve as  a  basis for judiciously adjusting the calculations
of the  source-oriented models.

Monitoring  Changes  in the Source Distributions—The tables  of  the Tabula-
tion Prediction  Scheme must be continuously updated if they are to be used
most effectively. One  may accomplish this  by  comparing the distributions,
obtained during a  1-  or 2-month period,   with  the  values  in  the  tables.
Appropriate  tests for statistical significance  may  be used for this purpose.
Automatic updating would, of course, be accomplished on a computer.
As a part of  the updating program one might apply triangulation techniques
to the  data from two  monitoring stations, to detect a  significant change in
the source strength over  a  given area.  In  Figure 14-1, for example, station 1
shows an increase  in a L 1, with respect to  the base-line between stations 1
and 2.  Similarly, station 2 shows an  increase in  the L 2. Such an analysis
indicates the  area where  the  source inventory should be reexamined. When a
network of  monitoring stations  is available, the lines from each station,
indicating  the  direction  in  which a  change in  source  strength  occurred,
                                                                   14-15

-------
 o
 ARG
               NEW SOURCE
                                                            INDIANA
                     10km
   Figure 14-1.  Identification of change in source inventory by

   triangulation techniques and tabulation  prediction scheme.
14-16

-------
delinate the area to be investigated in order to update the source inventory.
This is  yet another  way  in  which  this  technique may be used with  the
source-oriented model.

Use as  a  Receptor-Oriented  Urban  Air  Pollution  Model-In cities  with  a
network of pollutant monitoring stations, the Tabulation Prediction  Scheme
may be used to develop a receptor-oriented model.  For example, for a given
set  of  meteorological conditions,  the  50th  percentile SO2 value  may  be
determined from the tables,  and  lines of equal concentrations (isopleths)
may be drawn. Areas showing values of S02  concentrations above a critical
predetermined value such as 0.3 ppm would  be delineated. Similarly, 75 or
90  percentile  charts  may be  drawn. In  this way, the Tabulation Prediction
Scheme would  serve as an urban air  pollution model that would provide the
basis for action  by  a municipal air pollution control office to request public
utilities or  other large users of  coal to  switch to gas where dual equipment
existed. The Tabulation  Prediction Scheme was used  in this  manner  in a
recent air pollution  incident control test  in Chicago, Illinois.*
A word of  caution  concerning the drawing of isopleths over  a city where
there are an insufficient number of stations is appropriate. The heterogeneity
of  pollutant concentrations  over  an urban  complex  coming  from  large,
widely distributed, discrete sources, must be considered  in drawing isopleths.
For optimal  use of  the Tabulation Prediction  Scheme information on  the
lack of uniformity or "lumpiness" in the pollutant  concentrations should be
obtained.  Traverses  with  mobile  air quality  measuring  equipment would
provide such valuable insight into the spatial distribution of pollutants.

Computerized equipment to provide isopleths of a given variable  over an area
are described in the John, Ludwig, and  Moon,  and  Mahoney, Maddaus,  and
Goodrich papers. These could be adapted for effective use with  the  Scheme.
The Tabulation Prediction  Scheme cables could be stored  in  a computer.
Using forecasts  of  hourly values  of  the pertinent  meteorological variables,
prepared by  a  human forecaster as input, the  computer  can provide maps,
either on a cathode ray  oscilloscope  or on paper, of the 50th, 75th,  90th or
other percentile isopleths of  a pollutant, for  each hour of  the  forecast
period.
Although the  Tabulation Prediction  Scheme  is easy to use, a  considerable
amount of  insight  is necessary to develop an  effective  set of tables.  The
selection of appropriate  meteorological  variables or  even the division of each
variable into optimal  group intervals requires a substantial  amount of calcula-
tion. As an  example:  the  levels of S02  in a residential area are closely
related to temperature; the lower the temperatures, the more coal  is burned,
*Croke, Edward J. and Samuel G. Booras, The Design of an Air Pollution Incident  Con-
trol Plan, APCA No. 69-99. Presented before the 62nd Annual Meeting of the Air Pollu-
tion Control Administration, June 1969. New York, N. Y.
                                                                    14-17

-------
hence, the more SO2 is produced.  Hence,  one is inclined to select tempera-
ture as one of the variables in the  tabulations. The "cooling power," which
is a function of  temperature and windspeed, must  also be considered before
deciding whether this function, or the temperature,  should be used.

In spite of the work  and skill required to construct  an  effective set of tables,
their  usefulness  appears to  be worth their cost. Every technique  available
should be used to provide the air pollution control official responsible to the
public, with the best possible information  to assist  him in his decisions. The
combination  of  the  Tabulation Prediction Scheme  with the source-oriented
model results in  a hybrid  model, which  appears to hold greater promise  than
the purely  source-oriented or receptor-oriented models.


Mathematical Refinements
                            B. E.  Freeman
During the Symposium we  have heard a  comprehensive exposition of  the
formulation and application  of air  pollution models to a number of urban
areas  via the plume and puff approximations.  While many of the participants
in the Symposium have pointed out shortcomings of these models and others
have  called for  a more complete treatment  of  physical phenomena in the
models, there has been  no discussion of mathematical techniques to remedy
these dificiencies. Fortunately,  considerable experience  has been gained  with
suitable methods of  mathematical  approximation which, in  my mind,  hold
out  substantial  hope of  removing  many  of  the objections  to the  existing
models. Clearly,  it  is  not possible  in these  passing remarks to  set out in
detail the methods, which have been developed and  are now being improved.
Consequently, I  shall attempt  to summarize, briefly,  the general character-
istics  of one refined model, its advantages, and its prospects.
Before I  turn to the model itself,  let  me add  another objection  to those
currently used; they  are based  on quasi-analytic  solutions of linear equations
for the evolutions of  pollutant concentration.  Now,  in fact,  many of  the
processes  governing  the  fate  of pollution are  nonlinear; the reactions in-
volving  the components of  photochemical smog, the micrometeorology of
the urban  basin, the modification  of the  eddy diffusion constants by  the
heat island of the city, and  many others. These processes can only  be taken
into account in  a mathematical framework more general than the plume and
puff models.

The tested and  successful method  employed  in many field problems is the
grid or cell model, based on difference approximations to the partial differen-
tial equations governing the phenomena in question.  In this approach, space
and  time  are  divided up  into cells  and the dependent variables are approx-
imated through  averages over the cells.  If the cells are made to diminish in
14-18

-------
size  in a  suitably chosen  manner, the solution of the difference equations
will approach that of the differential equations.

The  difference-equation grid model  is  not a  new technique. In fact, this
method  was applied  some  25  years ago by von Neumann  to  solve  the
hydrodynamical problem of  the  behavior  of a  nuclear explosive-one of the
first applications  of the digital  computer.  Since  that  time the difference-
equation method has become much more  sophisticated.  With advances in the
capacity  and reliability of computing machinery and  in numerical methods,
it  has become  possible to  attack an increasing number  of physical problems
-to  mention only a few:  astrophysicists model the evolution of the interior
of stars and the characteristics of the radiations from the atmospheres of the
stars; physicists describe nuclear explosion  fireballs in  the atmosphere and
duplicate the  hypervelocity  impact of pellets  on  plates. The technique has
been applied to the description of the natural enviroment, too. Of particular
relevance is the progress, in  recent years, on the dynamical modeling  of
global weather.

Now let me mention just  a  few  of the characteristics of the model, that can
be developed   in  modular fashion by  adding  new  physical  phenomena as
warranted by  our knowledge of them.  First, the pollutants themselves are
followed  in space and time. The  governing  equations  are  the conservation
equations for the several pollutant species necessary to  describe urban  pollu-
tion. The equations take  account of  the rate of change  of  the pollution
concentrations  in  all of the  cells of the urban basin due to (.1) advection of
the  pollutant to  neighboring cells  by the wind field, (2) spread of pollutants
through   eddy  diffusion,  (3)  injection  of pollutants  from the sources of
pollution within the urban area, (4) chemical reactions which may take place
within the cell, and (5) advection  of pollutants from outside of the volume
of interest. Each of these processes is modeled  by  adding an  appropriate
numerical term to some or all of the  cells of the basin grid.  The time history
of pollution in the  basin then evolves by time increments that simulate the
actual  temporal  development. In  such a  field description  of the  pollutant
concentrations,  there  is no difficulty  added  in  accounting  for nonlinear
processes over linear ones.

With this model  not  only  is it  possible to  take chemical reactions  into
account,  but by supplying the wind  field, diffusion coefficients, and sources
in the usual way, the multi-source urban pollution application  can also be
solved, as in present models.
In successive  modules it will  be possible to take  account  of  processes
determining the wind field and  the diffusion coefficients  themselves.  The
meteorological  equations,  for  example,  can be solved  by difference approx-
imation,  to predict  the local modification  of the wind field and thermal
stratification by topography, diurnal solar  heating, and heat sources within
                                                                    14-19

-------
the  city. The formulation  is  also compatible  with the  incorporation of
models for  the  growth and  decay of turbulence in order  to  evaluate eddy
diffusivity as a field quantity. While  it is to be expected that the modeling
of the hydrodynamic equations will be quite expensive in computer time, in
contrast  to  the nominal computer  use required by the pollution grid model;
some economics  can be achieved  by using  implicit  methods to increase  the
size of the time  increment  interval.

As in the case of the numerical  weather model, we can only conjecture what
the ultimate value of the grid model for air pollution will be. But to me it is
clear  that substantial improvements in physical  and  mathematical  verisimi-
tude  can be achieved.  As mathematical approximations are  removed,  the
challenge will  lie  in  taking  advantage  of the  growing volume of  urban
pollution data to  improve our understanding of  the  underlying  physical
processes. Only this way can we hope to make air  pollution models quanti-
tatively useful.


Consumer  Applications
                           LA.  Clarenberg
I  would  like to approach  the problem of  multiple-source  diffusion  models
from  the consumers' point of view, that is  from the standpoint of a person
with some responsibility for the air quality  in a region. After listening to the
lectures  given at this symposium  I  have the  feeling that  there  may  be a
discrepancy  between the model builders' aims and the consumers' needs.

The  primary responsibility  of an air pollution control authority is to see that
standards set for  a  pollutant are not violated. Air pollution  standards specify
concentrations and  corresponding  percentages of the time these specified
concentrations may be exceeded. This brings forth a  need  for maps on which
iso-probability lines are drawn,  that  is lines  of equal  probability  that a
specified concentration  has been exceeded,  permitting an immediate evalua-
tion  of the  air quality in  a town or a region. The position of iso-probability
lines is affected only by long-term  (1-year) changes.

The  models we  have been discussing,  thus  far, calculate  the joint contribu-
tion  of a number of fixed sources to the concentration at a fixed site, e.g. a
sampling station,  for a given  set of meteorological conditions. On a map the
location  of  the  sampling  station  with respect to  any one source  can  be
expressed exactly in terms of a  downwind distance (x) and  a lateral distance
(y).  I  would  like to name this a  "space  frame," in which all subsequent
calculations  are  carried  out.  The result is a map covered with  iso-concentr-
ation lines,  pertinent to  that  particular set of conditions. For air  quality
control these lines are of little value, since their positions on the map change
as meteorological conditions  change.  So, one might get the impression that
14-20

-------
model  building  has  become a game-a  word  used  by Dr. Roberts  in this
context-being  played for its own  sake, however,  one that  loses track  of
practical  needs.

Turbulent diffusion is a  process proceeding in  time. Theories, describing the
diffusion of a puff, are worked  out in a "time frame." This means that one
can predict the volume of a puff  being  transported  through the atmosphere
after a lapse  of time, t;  one cannot predict the exact position of the center
of that puff  relative to its origin.  One can, however, predict the probability
of finding the center after  a time, t, in any one  position relative  to the
origin.

If multiple-source urban diffusion  models could be worked  out  in  a  pure
time frame; then the concentration  in any one location could be found  by
probability,  for  a given  set  of meteorological  conditions, at  the  same time
giving accurate prediction. By repeating the calculation for a great  number of
meteorological  conditions  and  assigning the probability of occurrence, the
most  probable concentration distribution, to each set of conditions. From it
is only a small step to produce iso-probability lines.

I have three  reasons for  advocating  this  idea. First,  it seems to me that the
input of  meteorological  data becomes much  easier. Second,  if ever  models
are to  be made for pollutants  other   than  sulfur dioxide—for example,
oxidants— in which  case  one can  neither  locate  nor  specify the sources
exactly,  one would  have  to  (as far as I  can  see)  rely on "time-frame
models."  Last, but  not least,  results would  be in  better  compliance with
practical  needs,  especially  when estimates of reliability are required for the
predictions.


Uses, Limitations and Accuracy of  Models

                          James R. Mahoney
A number of detailed, first  generation,  models for  urban  air contamination
have  been discussed  at this meeting.  In each  case  it would be desirable  to
have a statement of the  specific uses, limitations and accuracy of  the model.
Such  information would facilitate both  applications of existing models and
development  of new models.
The following  specific  observations  result  from   several  of  the technical
discussions during the Symposium:
- The uses and limitations of the diffusion models can be stated in terms of
the time  and space  scales appropriate for  each model. Many of the early
models  have  been applied to a wide spectrum  of  scales,  from near-instan-
tanous analysis  to annual-average calculations in the  time dimension, and
from  a few meters to tens of kilometers in  the  space  dimension. Obviously
                                                                   14-21

-------
the grid size for input data and for calculations should reflect the scale of
the intended, model output data.
— The  greatest deficiency in the  present  models  is  the lack of detailed
meteorological  data for  the  urban  regions being described. It is futile to
refine existing  emission  data  and to reduce grid size in most of the  model
calculations when  the  windspeed  and  direction, atmospheric  stability and
mixing-depth data  are all derived  from a single observation  site  within the
region.  The  new   observation networks in some  major  urban  areas have
surface  meteorological instruments at several sites; the data from such  obser-
vations  should  be  included  in  the  model  refinement  process,  whenever
possible. Significantly improved vertical meteorological  observations are also
required for the development of improved models.
— Current models  lack  descriptions of chemical and photochemical transfor-
mations involving the various pollutants. Realistic, dynamic chemical models
should  be verified  with  detailed and extensive field observations; the first
examples  of such information  are just now becoming available.
— Loss  processes,  particularly absorption and washout of pollutants, are not
sufficiently  understood  for incorporation  into  existing  models, but current
field studies may provide some data on loss mechanisms that can be used in
refined  diffusion models.
— Model  verification schemes, based upon  comparisons with measured  air
quality, must  be analyzed  carefully. Errors and small-scale  phenomena, re-
flected  in air  quality observations  at single sites, may  be as important as
errors in  the model input data. Whenever possible, model results should be
compared to time and/or  space averages  of observed air quality data.
— It is  possible that data from remote  sensing  instruments, presented  in the
form  of  average  concentrations over  extended  horizontal  paths,  may  be
superior to  data from individual  observing sites for  use in  urban diffusion
models. The use  of spatial average concentration  data  would reduce the
errors caused by local sources  near the observing stations.
— Current models  neglect the  initial field  of observed concentrations  in the
preparation  of  short-period  (e.g.   1-day)  forecasts.  Model  forecasts may
improve  if  the changes  in concentration  from a known initial field are
calculated as the predictive  parameters.
— Many current diffusion models are used  to describe regions containing an
urban center and  surrounding suburban and rural areas.  When  the  spatial
gradients  of the emission rates are large only  in one section of  the  model
region  (the  central urban area),  the use of grid telescoping  in that limited
section  can  improve model  accuracy without imposing excessive computation
time and data storage requirements.

— The expense and the importance of urban air-quality  monitoring programs


14-22

-------
in  operation, or  in  development, justify a  significant effort toward  the
development of  sophisticated  advection-diffusion  models  that can  make
optimal use of the data being collected.
Approach to a Universal  Model

                          Robert E. Stewart
I  am somewhat disappointed by the current method of attack for solving the
problem of urban air pollution prediction. I find that, for the most part, one
has to  postulate  "not so  good" assumptions  in order to make the particular
urban diffusion model fit the data. For example, the Gaussian distribution is
a  special  solution  to the semi-empirical diffusion  equation,  provided the
standard  deviations of concentration  are  related to their respective diffusion
coefficients  as given  by  Batchelor.*  This implies  homogeneous  turbulence.
Yet we stimpulate hypothetical,  mathematical expressions for the standard
deviations and, perhaps, even for  the  velocity vector as  well. That is to say,
we turn our  backs  on theory (ever so slightly) to arrive at a not-so-workable
solution. If we can justify this  then  I  suggest that we go all out  and  assume
that the concentration field at a  given time is a function  of the important
variables involved (i.e., vertical-temperature and velocity  gradients, turbulence
intensities, topographical  parameters, stack conditions, pollutant parameters);
and conduct a series of planned experiments, based on  similarity theory, to
arrive at a regression  equation  for  urban areas. If we do  have to revert to
this type of  approach,  let us consult with systems analysts and experimental
statisticians to arrive at a  universally acceptable model.
In any  event,  it would  be beneficial  if a committee or a  specialized group
were  set up  to develop guidelines for future approaches to source-inventory,
receptor-sampling, and  urban diffusion modeling based  on a review of the
existing state of the art.


User-Oriented Design
                               D. H.  Pack
We have heard  of  many advances in simulating the emission of pollutants,
their  transport and  diffusion  through the  atmosphere, and  the resulting
concentrations. Dr. Hilst  has advised us to consider carefully the purpose of
any  model  during  its  design  and not to expect  all  models to answer  all
questions.
It is profitable, I believe, to consider  the use, and the users,  of models in  an
organized way.
*6atchelor, G.K.. Diffusion in a Field,  of Homogeneous Turbulence. I. Eulerian Analysis
Aust. J. Sci. Res. 2: 437-450. 1949.
                                                                    14-23

-------
 Suppose, for example, we set up the following  table for model application:

Meteorologists
Control Agencies
Planners
Analyses

Prediction

Operations Scales of:
Time
Space
Information
Resources
 A  third dimension, scale, can  be added for each  combination of use  and
 user.
 The ultimate urban model could conceivably be used in all combinations but
 we  must consider  the paucity of information  on sources and the deteriora-
 tion of meteorological predictive skill with time.
 It  seems that, for some time to come,  compromises  will be necessary. To
 acknowledge  them, each model  builder should fill  in  the above  boxes (or
 variants thereof) with his estimate of what the model will do and  for whom
 it  will  work  best.  In  the  process,  current  information input,  computer
 availability,  and staff  technical  level should  be considered among  other
 criteria.
 Such an exercise may well  show that many of the models are already very
 successful.  For unsuccessful models this approach  can indicate  where re-
 search effort must be applied.


 General Discussion

                           D.  Bruce Turner
     Here are some miscellaneous comments that I would like to  make:
 - I would  agree with  Mr.  Pack  that the various models  in use have been
 designed for different applications.
 - Each researcher  has  verified  his  model  in  a  slightly different manner
 making  intercomparison  of models difficult.  Several methods of  verification
 should be developed that could be used as standards for verification.
 An additional  problem  in verification is  that  the concentrations calculated
 from the model represent something  quite different from the concentrations
 measured by  the sampling stations. In most models small  sources are com-
 bined and calculations are made  from one, uniform, area source for each grid
 square.  The calculated  concentrations  are thus, more  representative  of an
 area.  But measured  concentrations are made at a single sampling point and
 are greatly influenced by nearby small sources especially those emitting at a
 height nearly  level with  the sampler intake probe   This  difference between
 calculated and observed  concentrations can be  minimized by using  the  mean
 concentration  from  3 or 4  sampling stations  located  hundreds of meters
 apart.  The  derived  area-concentration may then  be  compared  with the
14-24

-------
calculated concentration. An investigator  will seldom  expend his sampling
equipment  in this way, however, due to the cost of establishing and servicing
stations.
-  It appears to  me  that  there  is a  "communication  gap" between  the
researchers developing the models and those wanting to use them. There are
indications that the prospective user is confusing the  researchers' goals for
the future uses of models with what can be done today. This may, in part,
be due to the fact that  the model verifications are  stated in  terms difficult
for the laymen to understand readily.
-  The spatial variation  of air pollution is extremely complex because  of the
great number  of air pollution sources  in urban areas. Only the gross details
of this variation  can be revealed  by air  quality sampling because cost  and
manpower limit the number of samplers that can be applied. It is desirable
to complement  air quality  monitoring with  dispersion modeling to  obtain
isopleths of  concentration.  Since  air  quality is the product of the  micro
enviroment, it is always possible to go to a smaller scale and find many more
details in the  concentration  field.  Therefore, one has to decide on the space
scale suitable for the detail  one wants to observe and  conduct  the emission
inventory, as well  as the model computations, commensurate with that scale.
— The  models that  have  been  presented  at this symposium  haven't con-
sidered photochemistry.  There  is a need to have dispersion models that can
be applied  to reactive  pollutants and their products. Since most  models
presented here  assume  that  the  contribution to  the concentration at a
receptor  from each source can be  determined independent of  emissions from
other sources,  I  doubt if  these models will  be  useful  for  other than
relatively  unreactive substances. It appears that  a different  approach from
normally  distributed  puffs  or plumes is  required  to  model reactive  sub-
stances.
- A statement was made  at this symposium to the  effect  that computer
technology was available  that had  not yet been used. I would  like to disagree
with  that statement because I know of a  number  of  researchers who have
used computers that could be considered in the category with the largest and
fastest computers  in this country, yet they have had to devise  short cuts to
use these computers to make calculations  of air  pollutant concentrations at
reasonable cost and  in a reasonable length  of time. I believe that as advances
in computer  technology become available,  they  will  be  put  to  use by
researchers  formulating  dispersion models.
— There  is certainly room for more simplified models, especially in the area of
community  planning, where estimates  of air quality for future decades is of
interest. Models whose calculations can be made with desk calculators, for ex-
ample, will find wide application when developed.
                                                                     14-25

-------
— A request  was  made  to tailor  dispersion model output to  the  user;
specifically, to determine how frequently a given  pollutant concentration is
exceeded. Dr.  Fortak's model  for  Bremen gives information of this type at
specific points.  I  can  certainly see its utility to control officials and  plan-
ners.
— One  item related to the  success  of dispersion  modeling, which has not
been mentioned at this symposium,  is the computation of plume rise above
the physical stack. Although several  currently used plume rise equations do
rather well when averaged over a number  of trials, we still need improvement
in hour-to-hour calculations of plume rise. Since many  air pollution control
strategies  emphasize the difference between  elevated and low-level sources,
there is a critical  need for models that can be used with  equal reliability on
both low-level  and elevated  sources. This, of course, gets one into all the
difficulties involved  in windspeed  and  direction,  shear,  and  other measure-
ments pointed out by Dr. Lettau.

-------
                                                      15.   APPENDIX
                                                  ATTENDANCE  LIST
Dr. Donald D. Adrian
Civil  Engineering Department
School of Engineering
University of Massachusetts
Amherst, Massachusetts  01003
Dr. Walter D. Bach, Jr.
Research Triangle Institute
Research Triangle Park, N. C.
27709
Dr. Eugene W. Bierly
Program Director of Meteorology
Atmospheric Sciences Section
National Science Foundation
Washington, D. C.  20550

Mr. F. T. Bodurtha
Senior Consultant
Environmental Control
Louviers Building
E. I. Du Pont de Nemours & Co.
Wilmington, Delaware  29898

Dr. W. A. Bowman
Lockheed Missiles & Space Co.
Huntsville Research and
   Engineering Center
P. O. Box 1103 West Station
Huntsville, Alabama  35807

Dr. Al Boyer
Officer-in-Charge of Meteorology
Air Management Branch, Sixth Floor
Ontario Department of  Energy and
   Resources Management
1 St. Clair Avenue West
Toronto 195, Ontario

Mr. Edward W. Burt
Meteorologist
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609

Mr. Kenneth L. Calder,  ESSA
Chief Scientist
Division of Meteorology
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609
Dr. J. E. Cerrnak
Engineering Mechanics
Department of Civil Engineering
Colorado State University
Fort Collins, Colorado  80521

Dr. Jack Chaddock
Chairman
Mechanical Engineering Department
Duke university
Durham, N. C.  27706

Mr. Anton Chaplin
Litton Systems, Inc.
Applied Technology Division
354 Dawson Drive
Camarillo, California  93101

Dr. L. A. Clarenburg
Rijnmond Authority
Schiedam
The Netherlands

Dr. Alan Cole
Earth Sciences  Department
Northern Illinois University
De Kalb, Illinois  60115

Dr. H. E. Cramer
Director
Environmental  Sciences Laboratory
GCA Technology
P.  0. Box 15009
Salt Lake City, Utah  84115

Dr. Gabriel T.  Csanady
Meteorology and Space Science
   Building
The University of Wisconsin
1225 West Dayton Street
Madison, Wisconsin  53706

Mr. Margaret Day
Manager,  Research Systems Analysis
Meteorology Research, Inc.
464 West Woodbury Road, Box 637
Altadena, California  91001

Mr. James L. Dicke
Meteorologist
Office of Manpower Development
Air Pollution Control
   Office
Research Triangle Park, N. C. 27709
                                     15-1

-------
Dr. Authur Dodd
Meteorologist
Army Research Office
Duke University
Durham, North Carolina  27706

Dr. Rudolf J. Engelmann
Acting Chief
Fallout Studies Branch
Division of Biology and Medicine
U. S. Atomic Energy Commission
Washington, D. C.  20545

Dr. R. M. Felder
Department of Chemical Engineering
School of Engineering
North Carolina State University
Raleigh, N. C.  27607

Dr. Bruno Finzi-Contini
Istitutodi Fisica Tecnica
Universita Degli Studi di Trieste
Facolta Di Ingegneria
Via Alfonso Valerio 10, Italy

Dr. Heinz Fortak
I nstitute fur Theoretische
   Meteorologie
Der  Freien Universitat Berlin
 1 Berlin 33
Thielallsee49
 Federal Republic of Germany

Dr. David Fraser
Department of Environmental
   Sciences Engineering
School of Public  Health
University of North Carolina
Chapel Hill, N. C.  27514

Dr. Burton Freeman
Vice-President
Systems, Science, & Software
P. 0. Box 1620
 La Jolla, California  92037

Mr. Francois N.  Frenkiel
Naval Ship Research and
   Development  Center
Washington, D. C.  20007

Dr. James P. Friend
Atmospheric Chemistry
Department of Meteorology and
   Oceanography
School of Engineering and Science
New York University
Geophysical Sciences Laboratory
2455 Sedgwick Avenue
Bronx, New York  10468
Dr. J. Gavis
Department of Environmental
   Engineering
The Johns Hopkins University
Baltimore, Maryland  21218

Dr. F. A. Gifford, Jr.
Director, Air Resources Atmospheric
   Turbulence and Diffusion
   Laboratory
U. S. Department of Commerce
Post  Office Box E
Oak  Ridge, Tennessee  37830

Dr. James Halitsky
Senior Research Scientist
School of Engineering and Science
New York University
University Heights, Bronx, N. Y.  10453

Mr. Harry Hamilton
Engineering and Environmental
   Sciences Division
Research Triangle Institute
P. 0. Box 12194
Research Triangle Park, N. C.  27709

Mr. Steven Hanna
Air Resources Atmospheric
   Turbulence and  Diffusion
   Laboratory
Cheyenne Hall Building
P. O. Box E
Oak  Ridge, Tennessee  37831

Dr. George Herbert
President
Research Triangle Institute
P. O. Box 12194
Research Triangle Park, N. C.  27709

Dr. Glenn R.  Hilst
Executive Vice-President
The Travelers Research Corporation
250  Constitution Plaza
Hartford, Connecticut  06103

Mr. G. C. Holzworth
Chief, Air Pollution Geophysics
   Research Branch
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609

Dr. Walter Hoydysh
School of Engineering & Science
New York University
Bronx, New York  10453
 15-2

-------
Mr. Harvey Jeffries
Department of Environmental Sciences
   and Engineering
University of North  Carolina
Chapet Hill, North Carolina  27514


Mr Robert M. Jimeson
Physical Sciences Administration
Air Pollution Control
   Office
801  N  Randolph St
Arlington, Virginia  22203

Dr. Warren B. Johnson, Jr.
Senior Research Meteorologist
Aerophysics Laboratory
Stanford Research institute
Menlo Park, California  94025

Mr. Edward W. Klappenbach
Senior Meteorologist
City of Chicago Department of
   Air Pollution Control
320 North Clark Street
Chicago,  I Ilinois  60610

Kr. Kenneth Knoerr
School of Forestry
Duke University
Durham,  N. C.  27706

Dr. Richard J. Kopec
Department of Geography
University of North  Carolina
Chapel Hill, N. C.  27514

Dr. Heinz H. Lettau
Department of Meteorology
University of Wisconsin
Meteorology and Space Science Bldg.
1225 West Dayton Street
Madison, Wisconsin  53076

Mr. Denis M. Lohman
Chief, Meteorology Section
Penn. Bureau of Air  Pollution Control
P. 0. Box 90
Harnsburg,  Pennsylvania  17120

Dr. Lester Machta
Director,  Air Resources Laboratories
Environmental Science Services
   Administration
U.  S. Department of Commerce
Sliver Spring, Maryland 20910
 Dr. Jarnes R. Mahoney
 Assistant Professor of Applied
   Meteorology
 Harvard University School of
   Public Health
 Kresge Center for Environmental
   Health
 Department of Industrial Hygiene
 665  Huntington Avenue
 Boston, Massachusetts 02115

 Dr. David B. Marsland
 Department of Chemical Engineering
 School of Engineering
 North Carolina State University
 Raleigh, N. C.  27607

 Mr. Robert A  McCormick
 Director Meteorology Program
 Air Pollution Control
   Office
 3820 Merton Drive
 Raleigh, N  C.  27609

 Mr. Douglas L.  McKay
 Department of  Environmental
   Sciences and Engineering
 University of North Carolina
 Chapel Hill, North Carolina  27514

Dr. John T. Middleton
Commissioner
Air Pollution Control
   Office
801 North Randolph  Street
Arlington, Virginia  22203

 Dr. David Moreau
 Department of City and Regional
   Planning
 University of North Carolina
Chapel Hill, N. C.  27514

 Mr. Paul Morgenstern
 Environmental Sciences and Technology
Waldin Research Corporation
 359 Allston Street
Cambridge, Massachusetts  02139

Dr. William J. Moroz
 Director
Center for Air Environment Studies
Pennsylvania State University
226 Chemical Engineering II
 University Park, Pennsylvania  16802
                                                                             15-3

-------
Mr. Harry Moses
Argonne National Laboratory
9700 South Cass Avenue
Argonne, Illinois  60439

Dr. J. C. Mulligan
Mechanical and Aerospace Engineering
North Carolina State University
Raleigh, N. C.   27607

Dr. R. E. Munn
Meteorological Branch
Department of Transport
315 Bloor Street, West
Toronto, 5, Ontario

Dr. C. R. Murthy
Physical Limnology Section
Department of Energy, Mines and
    Resources
Canada Centre for I nland Waters
Great Lakes Division
P. 0. Box 5050
867 Lakeshore Road
Burlington, Ontario

Dr. Charles O'Melia
Department of Environmental
    Sciences and Engineering
University of North Carolina
Chapel Hill, N. C.  27514

Dr. Morris Neiburger
Department of Meteorology
University of California
Los Angeles, California   90024

Mr. Lawrence E. Niemeyer, ESSA
Assistant Director
Division of Meteorology
Air Pollution Control
    Office
3820 Merton Drive
Raleigh, N. C.   27609

Mr. Donald H.  Pack
Deputy Director
Air Resources  Laboratory
Environmental Science Services
   Administration
Silver Springs,  Maryland   20910

Dr. Frank Pasquill
Meteorological Office
London Road
Bracknell, Berkshire, England
Mr. Francis Pooler
Chief, Boundary Layer Dynamic
   Research Branch
Division of Meteorology
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609

Dr. Darryl Randerson
ESSA-ARL
P. O. Box 14985
Las Vegas, Nevada  89114

Dr. John J. Roberts
Reactor Engineering Division
Argonne National Laboratory
9700 South Cass Avenue
Argonne, Illinois  60439

Dr. G. D. Robinson
Director, Environmental Physics
   and Chemistry
The Travelers Research Corporation
250 Constitution Plaza
Hartford, Connecticut 06103

Dr. Shin'ichi Sakuraba
Meteorological Research Institute
Koenji Kita 4-35-8, Suginami-ku,
Tokyo, Japan

Dr. Walter J. Saucier
Department  of Geosciences
North Carolina State University
Raleigh, N. C.  27607

Dr. Jabbar K. Sherwani
Department  of Environmental
   Sciences and Engineering
School of Public Health
University of North Carolina
Chapel Hill,  N. C.  27514

Dr. L. J. Shieh
Scientific Center, IBM
2670 Hanover Street
Palo Alto, California

Mr. Conrad Simon
Senior Meteorologist
City of New York Department of
   Air Resources
51 Astor Place
New York, N. Y.  10003
 15-4

-------
Mr. Irving A. Singer
Meteorology Group
Brookhaven National Laboratory
Associated Universities, Inc.
Upton, Long Island,
New York  11973

Dr. Ralph C. Sklarew
Systems, Science and Software
P.  0. Box 1620
La Jolla,  California   92037

Mr. David Slade
Fallout Studies Branch
Division of Biology and Medicine
U. S. Atomic Energy Commission
Washington, D. C. 20545

Dr. P. R. Slawson
Department of Mechanical Engineering
Faculty of Engineering
University of Waterloo
Waterloo, Ontario, Canada

Dr. William H. Snyder
Physical Scientist
Division of Meteorology
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609

Dr. F. Y. Sorrell
Department of Engineering Mechanics
School of Engineering
North Carolina State University
Raleigh, N. C.  27607

Dr. Tom Stephens
Southern Research Institute
2000 9th Avenue, South
Birmingham, Alabama  35202

Prof. Arthur C. Stern
Department of Environmental Sciences
   and Engineering
School of Public  Health
University of North Carolina
Chapel Hill, N.  C. 27514

Dr. Robert E. Stewart-
Department of Environmental
   Engineering
College of Engineering
University of Florida
Gainesville, Florida   32601
Dr. Gordon Strom
School of Engineering & Science
New York University
Bronx, New York  10453

Dr. John L. Sullivan
Environmental Engineering
Syracuse University
Syracuse, New York   1 321 0

Dr. Hale Sweeny
Statistics Research Division
Engineering and Environmental
   Sciences Division
Research Triangle Park, N. C.  27709

Dr. Peter A. Taylor
Department of Mathematics
University of Toronto
Toronto 5, Ontario, Canada

Mr. Joseph A. Tikvart
Meteorologist
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609

Mr. Donald B. Turner
Meteorologist
Air Pollution Control
   Office
3820 Merton Drive
Raleigh, N. C.  27609

Dr. Edward Ungar
Divisional Chief of Fluid and Gas
   Dynamics
Battelle Institute
Columbus, Ohio  43201

Dr. Reginald I. Vachon
Mechanical Engineering Department
Auburn University
Auburn, Alabama  36830

Mr. E. A. Ward
TRW, Inc.
Washington Operations
1735 I  Street, N. W.
Washington,  D. C.  20006

Mrs. M. L. Weatherley
Warren Spring Laboratory
Ministry of Technology
Dudley House, Endell Street
LondonW. C.2
England
                                                                             15-5

-------
Dr. Willis L.Webb
Atmospheric Sciences Laboratory
White Sands Missile Range
White Sands, New Mexico

Dr. Allen H.Weber
Department of Geosciences
North Carolina State University
Raleigh, N. C.  27607

Dr. Larry Wendell
Atomic Energy Commission
Box 2108
Idaho Falls, Idaho 83401
Dr. Daniel Werner
Mechanical Engineering
Duke University
Durham,  N. C. 27706

Dr. James J. B. Worth
Associate Director of Engineering
Research Triangle Institute
Research Triangle Park, N. C.  27709
  15-6

-------