RESEARCH ON DIFFUSION IN ATMOSPHERIC BOUNDARY LAYERS:

         A POSITION PAPER ON STATUS AND NEEDS
                          by

                    Gary A. Briggs
                         and
                 Francis S. Binkowski
         Meteorology and Assessment Division
       Atmospheric Sciences Research Laboratory
         U.S.  Environmental Protection Agency
    Research Triangle Park, North Carolina  27711
       ATMOSPHERIC SCIENCES RESEARCH LABORATORY
          OFFICE OF RESEARCH AND DEVELOPMENT
         U.S.  ENVIRONMENTAL PROTECTION AGENCY
    RESEARCH TRIANGLE PARK, NORTH CAROLINA  27711

-------
                                  NOTICE

     The information in this document has been funded by the United States
Environmental  Protection Agency.  It has been subject to the Agency's
peer and administrative review, and it has been approved for publication
as aln EPA document.  Mention of trade names or commercial products does
not constitute endorsement or recommendation for use.
     The authors, Gary A. Rriggs and Francis S. Binkowski, are on assignment
to the Atmospheric Sciences Research Laboratory, U.S. Environmental
Protection Agency, from the National Oceanic and Atmospheric Administration
(NOAA), U.S. Department of Commerce.

                                    ii

-------
                                 ABSTRACT

     The introduction of a new understanding of atmospheric boundary
layers (ABLs) has caused a major change in the view of the diffusion of
pollutants.  The turbulence parameters now standard in ABL work,  have
provided a method for systematically organizing diffusion  parameters.
Concurrently with these advances, alternatives to  the operational  models
have emerged, but existing experimental  data sets  are inadequate  for
model comparisons and evaluations.  The most important knowledge  gap is
the lack of an adequate specification of the relevant meteorology  both  at
the point of release and downwind.  A second major inadequacy  is  experimental
measurements of plume characteristics up to 100 km from the release point.
There is also a great need for formulating new operational  models  based
upon this newly acquired experimental  data and the new alternative
approaches.   Finally,  it is recognized  that a modest but  steady  effort
is necessary.
                                   iii

-------
                            EXECUTIVE SUMMARY

INTRODUCTION

     The purpose of this study is to review the current state of science
concerning diffusion within the atmospheric boundary layer (ABL) and to
identify major issues and research needs.  The study is limited to
considering only the diffusion of non-buoyant, conservative substances
over homogeneous and level  or moderately hilly terrain.  Thus, chemical
transformation, wet and dry deposition,  buoyancy effects, and diffusion
in complex terrain are all  topics which  are excluded.  Certainly these
are very important topics both from the  point of view of the research
scientist and from that of a regulator;  however, each of these topics
deserve a review study of the same scope as attempted here.  The topics
that will be included here are: a review of diffusion experiments and
models, and a discussion of research needs.  We begin, however, with a
highly simplified description of the major meteorological factors affecting
diffusion.  The main report contains the necessary details and references.

METEOROLOGICAL FACTORS CONTROLLING DIFFUSION
     Material released from the ground or from a stack is mixed and
transported by air motion.   This process dilutes the material  and carries
it horizontally away from the source. This mixing process is called
turbulent diffusion, or simply diffusion.  The turbulence responsible
for this mixing may result  from the wind blowing over the ground, generating
turbulent eddies, or it may result from  heated parcels of air or thermals

-------
that are generated in air heated  by contact  with  ground  which  has  been
warmed by the sun.  Thus, we distinguish  between  "mechanical"  and
"convective" turbulence,  because  in addition to having different generative
mechanisms, they have considerably different diffusive properties.

     The current view of  the ABL  divides  it  into  three conceptual  states,
the neutral boundary layer (NBL), the convective  boundary  layer  (CBL),  and
the stable boundary layer (SBL).   The NBL occurs  when there  is a negligible
transfer of heat from the ground  surface  to  the air.  This occurs  with
strong winds and usually  when most of the sky is  covered with  clouds.   Only
mechanical turbulence is  generated in the NBL which  extends  up to  or  even
into the cloud bases.   The SBL grows when the ground is substantially
cooler than the air above it causing a stable profile to develop.   This
means that an upward moving parcel of air, for example,  experiences  a
downward restoring force  because it is colder, therefore denser, than its
surroundings.  Mechanical turbulence is generated as the wind  blows over
the ground, but vertical  motions  are  very much restricted by  the  stability.
Vertical diffusion, consequently, is restricted and  plumes of  effluent
dilute slowly during transport downwind.   The depth  of turbulent mixing
is also very restricted in the SBL, being of the  order of  100  m  or
less.  It is under these  conditions that  some of  the most  serious  air
pollution episodes can occur.  The CBL occurs in  the daytime,  with low  to
moderate wind speeds and  clear to partly  cloudy conditions.  Then  the
ground is warmed and thermals rise from the  surface  to generate  convective
turbulence.  The diffusion process is dominated by the vertical  transfer
of material in the thermals and downdrafts.   The  height  to which

                                    vi

-------
these thermals rise is called the mixing depth.   As a result  of this



efficient type of mixing, the CBL often has a "well mixed"  condition  in



which the vertical profile of a pollutant becomes constant  from a few



meters above the surface to the mixing depth.





     On a day without many clouds, and with neither fronts  nor  shorelines



nearby, the diurnal course of the ABL is as follows: starting about an



hour before sunrise, the ground has cooled to its minimum temperature.



The temperature profile is stable but there is some vertically  restricted



mechanical turbulence from the surface to the top of the SBL.  Any plumes



within this SBL are very well defined.  Any plumes above the  SBL have



virtually no vertical  dilution but may be spread  horizontally by meanders



caused by large scale horizontal  eddies.  As the  sun comes  up,  thermals



begin to rise from the ground but must work against the  stable  temperature



profile until  all  the air in the developing CBL is warmed and mixed.  The



CBL, thus formed,  grows until the diminished surface heating  in  the late



afternoon can no longer overturn stable air above the CBL.  Typically,



the maximum mixing depth is of the order of a kilometer. Any plumes



beneath this height released into the previously  stable  air are mixed



downward to the ground, a process referred to as  fumigation.  Late in the



afternoon, the strength of the thermals diminishes.  Soon a new stable



layer is established at the surface and a new SBL forms. The transitional



times around sunrise and sunset are as yet poorly understood.   Little is



known about the structure of the ABL and even less about diffusion during



these periods.
                                   vn

-------
     In a typical  diurnal  cycle,  the  large  changes  in stability correspond
to large changes in diffusion  rates.   Some  method of indicating the
ambient stability is necessary and  there  are  several indices of stability
currently used.  The index most indicative  of the physical processes at
the surface is the Obukhov length,  L.   By convention L is defined to be
negative with an upward flux of heat  at the surface (unstable conditions).
In principle, this length  is defined  such that at height z = L above the
ground, the buoyant production (or  destruction) and the mechanical
production of turbulent kinetic energy are  comparable.  Hence, above z = L,
the turbulent state of the ABL is dominated by buoyancy effects and
below z = L, the turbulent state of the ARL is dominated by the mechanical
overturning caused by the  wind blowing over the surface.  Profiles of
mean wind speed and other  meteorological  quantities are functions of z/L
near the surface.

     Besides the Obukhov length, other quantities of importance in describing
the turbulence in the ABL  are  the surface heat flux, the surface friction
velocity scale u*, and the mixing depth.  A velocity scale analogous to
u* has been developed for  the  CBL.  Called  the convective velocity scale,
w*, it is essentially defined  by the  surface  heat (or buoyancy) flux the
depth of the CRL; w* has proven to  characterize the turbulent velocities
within the CBL very well.   The depth  of the CRL divided by L serves as a
useful indicator of the relative importance of the  thermals as a major
transfer mechanism; the larger this ratio is, the more important thermals
are to the description of  diffusion.   As  fundamental as these variables
are to our understanding of ABL processes,  u* and the surface heat flux

                                   viii

-------
are not easily measured; both are needed to determine L using its
definition.  An alternative index related to L, called the Richardson
number, Ri, uses a temperature difference and velocity differences between
two heights: in another form known as the bulk Richardson number, only
one wind measurement is used.  There are simple methods to convert from
Ri to Obukhov length.  An estimate of the site roughness length,  z0,  is
necessary with the bulk Ri approach, while the velocity differences
approach strongly demands wind sensors that are properly calibrated.   With
the temperature differences, and at least one wind velocity,  the  same
methods that yield the Obukhov length can be used to obtain u* and the
heat flux.  The only other essential measurement needed for characterizing
ABL turbulence and diffusion is the mixing depth.  This can be obtained
from good profiles of temperature and humidity.  In the case  of the SBL,
a profile of wind speed is also required.  In recent years it has become
possible to estimate this depth by using remote sensors such  as lidars and
acoustic sounders.  Lateral  diffusion at large distances is enhanced  by
wind direction changes with  height, which are best determined from wind
velocity soundings.

EXPERIMENTAL AND THEORETICAL BASES FOR OUR UNDERSTANDING OF DIFFUSION
     The current understanding of diffusion processes is based upon
results from field experiments, laboratory experiments and theoretical
models.  Field experiments may be conveniently grouped into two classes,
those in which the tracer material  is released from very near the surface
and those in which the release is elevated.  The data from previous
                                    IX

-------
experiments has been very useful  in  formulating  our  current  understanding
of the diffusion process but  there are major  inadequacies.   First,  the
tracers used in most early field  studies  were not  conservative.   This means
that as a cloud or plume of tracer was transported downwind,  the  material
stuck to the ground or deposited, thus reducing  the  concentration further
downwind by an uncertain amount.   The theoretical  view of diffusion
prevelant at the time of the  earlier experiments was that the vertical
profile of concentration was  Gaussian.  If this  were true and the tracer
were conservative, the standard deviation of  the concentration, Sigma-z,
is easily obtained from ground samplers alone.  When deposition occurs,
however, the Sigma-z estimate is  biased by an unknown amount  to be an
overestimate.  A second difficulty with these data is that with one
notable exception, the Prairie Grass experiment  done in 1956, the
experiments did not take sufficient  meteorological data for  either
interpretation using contemporary concept or  for validation  of current
diffusion models.  Two recent experiments, one in  Copenhagen, Denmark
and one under EPA sponsorship in  Boulder, CO, had  extensive  meteorological
data.  The Danish experiment  used a  conservative tracer, while the EPA
sponsored experiment used remote  sensing  to get  concentration profiles  at
several locations downwind of the source.  This  experiment,  named Convective
Diffusion Observed with Remote Sensors (CONDORS),  was designed to verify
some new insights into CBL diffusion that were gained from numerical
experiments and laboratory studies.

     Laboratory studies of ABL diffusion  are  gaining wider acceptance as
the scientific community has  learned how to simulate ABL processes more

-------
faithfully.  The major problem faced in a laboratory simulation is to



maintain the proper balance of forces and yet attain the required



reduction in scale.  In strictly neutral  conditions, this is  not too



difficult.  One problem is that if the scale reduction factor is very



large, viscous effects smooth out turbulent eddies in the laboratory



which would not be smoothed in the atmosphere.  This is usually overcome



by making the bottom of the wind tunnel have a rougher texture than strict



scaling would predict.  A major difficulty is properly simulating buoyancy



effects.  Since gravity cannot be easily scaled down on the earth's



surface, density differences must be exaggerated.   This distorts other



aspects of the simulation so that technical trade-offs are necessary.



The most effective use of laboratory work then is  to examine  a particular,



but important aspect of the diffusion problem.  One such application



which illustrates this point is the work done at the EPA Fluid Modeling



Facility that showed that a tall slender building  required a  shorter



stack for good plume dilution than would have been predicted  by using



building height alone.





     Experimental work alone is insufficient for an understanding of



diffusion.  A proper conceptual framework is also  necessary.   Most current



applications of diffusion models are based upon the Gaussian  distribution



in both vertical and horizontal directions in order to characterize the



diffusion.  The fiaussian distributions are based upon empirical  data



collected primarily during near surface releases over relatively flat



terrain.   Unfortunately, most of the model applications are  for situations



that were not considered in the experimental situations which provided





                                    xi

-------
the empirical  data for the formulation of these models.   Finally,  for
diffusion of non-buoyant releases  within  a CBL, the Gaussian  assumption
for the vertical  distribution of material  was  recently shown  to  be
incorrect.  Current practice has adjusted the  necessary  parameters so
that surface concentrations are nearly correct for the CBL  case  even
though the vertical profile is not.   There are however,  several  modeling
methods that have been demonstrated  to be superior for particular  applications,
In the interest of brevity, only three examples will  be  mentioned. The
first is the statistical method which relates  the standard  deviations  of
the concentration distribution directly to the observed  turbulent
intensities of the wind field.  Since the statistics  of  the wind field tend
to be fairly uniform in the horizontal direction, this method has  the
greatest success in characterizing the horizontal  concentration  distribution
for transport over an area of homogeneous land-use.  A second approach is
to simulate the diffusion as a random walk or  "Monte Carlo" process.
Recent work in this area has allowed for vertical  and horizontal
inhomogeneities in the flow field.  A disadvantage to this  approach is
that it generates individual material trajectories and then constructs
the distributions directly.  For realistic cases, a very large number  of
trajectories must be calculated.  Finally, a third method is  numerical
simulation of the diffusion process  by solution of the relevant  governing
partial differential equations in  an Eulerian  framework.  There  has been
great progress made in formulating these models which require a  number of
simplifying assumptions.  For example the laboratory work on  surface
releases in the CBL inspired a set of numerical simulations that in turn
                                   xii

-------
included a prediction about unexpected behavior of an elevated release.
The numerical work predicted that the plume centerline for a non-buoyant
release would descend to the surface, then rise again.  This prediction
inspired further laboratory work that verified the finding.   The CONDORS
experiment was conducted specifically to confirm the numerical and
laboratory findings for diffusion in the CBL.   Here is an example of how
the various methods for studying diffusion interact in the discovery and
confirmation of major new findings in diffusion theory.  All  of the work
mentioned in this example was supported by EPA.

NEEDS FOR FUTURE WORK

     Before listing the future needs it is imperative to state at the
outset that the human and fiscal resources for these endeavors are not
trivial.  A modest, steady commitment of personnel  and funds for an extended
period is necessary, as the work described requires a sustained scientific
effort.

     The first major need is for field data for selected meteorological
scenarios for the development, testing and evaluation of diffusion
meteorology models.  The characterization of those meteorological  variables
important to the diffusion process is a major  impediment to  the implementation
of new and emerging approaches to diffusion characterization.   To meet
these needs requires detailed profiles of the  relevant variables within
the first kilometer or so of the atmosphere.  In addition, sufficient
measurements are needed to characterize the surface energy balance, the
forcing of the wind fields by the large scale  migratory pressure systems,

                                   xiii

-------
the horizontal  temperature field  and  the cloud  conditions.   These  studies
should be conducted over a variety of terrain types  using a  minimum  of
preparation and personnel.  This  would allow the  study  of the meteorological
events of interest at the locations of interest.   A  portable instrumented
tower, a tethersonde (a tethered  balloon carrying an instrument  package),
and an acoustic sounder would be  the basic tools  for such studies.   The
development of such meteorological  field studies  would  greatly facilitate
the transfer of theoretical  results into practical operational models
useful for decision making.

     A second major need is  for the comparison, evaluation and construction
of new operational models.  An interdisciplinary  team with skills  in
meteorology, chemistry, numerical modeling and  statistics is required to
meet this need.  Such a team would evaluate the models, not  only with
respect to experimental data, but also with respect  to  the expected
uncertainty in the model inputs.   These type of comparisons  are  necessary
to make improvements in the operational models.

     A third major need is for diffusion experiments, both  in the  field
and in laboratory settings.   The  laboratory studies  are needed to  test
theoretical results in specific simplified situations that  are free  of
confounding influences.  The field data are needed because  there is  a
wide disparity between the flat homogeneous land  where  the  best  of the
previous field studies have been  conducted and  the complex woodlands and
urban-suburban developments typical of the American  landscape.   One
example of important information  to be gained  from such field studies  is
                                   xiv

-------
the characterization of the concentration  field  at  100  km  downwind  from
the source.  Current models extrapolate to this  distance based  upon
measurements to only 3 km.  Vertical  and lateral  profiles  of  concentration
even at widely spaced plume transects would improve the current  situation
considerably.  Other examples include the  characterization of releases
made during the hours of transition,  that  is near sunrise  and sunset,
as the behavior of the atmosphere during these periods  is  very  poorly
understood.  For any of these examples, a  good diffusion experiment  is of
necessity a good meteorological  experiment, since a full complement  of
meteorological measurements is essential to the  success of the  experiment
and to the continued usefulness  of the data set.
                                    xv

-------
                                 CONTENTS
Abstract	    iii
Executive Summary 	     v
Table and Figures	    xix
Abbreviations and Symbols 	    xx
Acknowledgements  	 xxiii


     1.  Introduction, Purpose and Scope  	   1

     2.  Atmospheric Boundary Layers  	   3
           Important Meteorological Variables and Scales
             Affecting Diffusion   	   5
           The Diurnal Cycle and Transitional Periods 	  13
           Neutral Boundary Layers  	  26
           Convective Boundary Layers 	  29
           Stable Boundary Layers  	  35

     3.  Diffusion Experiments   	  47
           Field Experiments	49
           Laboratory Experiments  	  74

     4.  Diffusion Modeling as Practiced  	  87
           The Gaussian Plume -  Its Utility and Limitations 	  87
           Empirical Diffusion Coefficients and Stability Typing  . .  91
           Evaluations of Sigma Curves and Typing Schemes 	  103

     5.  Diffusion Modeling Alternatives  	  109
           Surface Layer Similarity Models   	  Ill
           Gradient Transfer and Higher Order Closure Models   ....  115
           Convective Scaling Models  	  109
           Large Eddy Simulation Models	130
           Statistical Models 	  134
           Random Perturbation Models   	  143

     6.  What Are the Needs?	149
           Experimental Versus Modeling Needs 	  149
           What Kind of Experiments are Needed?	152
           Meteorological Measurements  	  156
           Evaluation and Concentration Fluctuation Issues  	  160
           Stable Boundary Layer Issues 	  163
           Convective Boundary Layer Issues  	  168
           Calms and Other Inconvenient Events  	  171
           Resource Issues 	  172

     7.  Implementation of Improved Modeling Practices  	  175
           Estimating Needed Parameters from More
             Obtainable Measurements  	  176
           Building Better Operational Models   	  180
                                   xvn

-------
     8.  Assessment   	183
           Atmospheric Boundary Layers  	 184
           Diffusion Experiments  	 186
           Diffusion Modeling as Practiced  	 190
           Alternative Modeling Approaches  	 194
           Needs for Diffusion Experiments	200
           Research Modeling Needs  	 206
           Technical Transfer Needs 	 207

References	210
                                  xvm

-------
                                  TABLES
Number                                                                   Page

  1          Summary of Tracer Experiments in Homogeneous
               Terrains	      52-53
                                 FIGURES
             Temporal  behavior of measurements  in  a CBL at  the
             Boulder Atmospheric Observatory, Colorado, USA
             13 September 1983	      20

             Temporal  behavior of measurements  in  a SBL at
             Cabauw, The Netherlands,  30-31  May 1978	      23
                                  xix

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

ABL          Atmospheric Boundary Layer
b            Oimensionless constant in Eq. 2.5 for h, SBL
C            Concentration (generalized)
CBL          Convective Boundary Layer
D            Wind direction
f            Coriolis parameter
g            Gravitational acceleration
h            Mixed depth height for NBL or SBL
H*           Turbulent flux of buoyant acceleration at the surface:
               H* = (g/V) ^TT1"
i            Subscript, "inversion"
k            von Karman's constant, = 0.4
K            Eddy diffusivity (generalized) (K^, heat) (Km, momentum)
L            Obukhov length: L = -u*3/(kH*)
n            A generalized exponent
N            Brunt-Vaisala frequency: N? = (g/9)d9/az
NBL          Neutral Boundary Layer
o            Subscript refers to surface value
p            Atmospheric pressure
P/G/T        Pasquill/Gifford/Turner
P0           Standard atmospheric pressure (1000 mb)
0            Source strength (rate of release)
r            Correlation coefficient
Re           Reynolds number
Rj           Gradient Richardson number

                                   xx

-------
%           Bulk Richardson number


SBL          Stable Boundary Layer


t            Time


T            Absolute temperature


Te           Eulerian time scale for ABL turbulence


T]           Lagrangian time scale for ABL turbulence


Tv           Virtual absolute temperature


u            Wind speed in the downwind direction (usually defined by

             mean wind direction at source height).


TJ"            Mean wind speed in the downwind direction


Ug           Geostrophic wind speed - the forcing formation for the wind

                                  O    ^"
u*           Friction velocity: u**- = -w'u1 = TO


U            Average "u through mixed depth


v            Wind speed in the lateral  direction


w            Wind speed in the vertical direction


w*           Convective scale velocity: w* = (H*z^)^'^


x            Distance downwind of source


y            Lateral distance from plume center!ine


z            Height above surface


ZT           Mixing depth in CBL, usually capped by an inversion


Z0           Roughness length


Zs           Source height


p            Average slope of a site


9            Potential temperature, T(pQ/p)2/7


A0           Difference in 9


aa           Standard deviation of azimuth angle of wind




                                    xxi

-------
ae           Standard deviation  of elevation  angle of  wind

au, av, aw   Standard deviation  of wind  components u,  v,  or w

av, ay, az   Standard deviation  of concentration  in lateral (y)
             and vertical  (z)  direction

p            Density of air or fluid

T            Horizontal  kinematic  stress,  time  scale

X            Concentration

v            Kinematic viscosity
                                   xxn

-------
                             ACKNOWLEDGEMENTS
     The authors wish to express their deep appreciation to John  S.  Irwin
of the Meteorology and Assessment Division  for his  penetrating  review and
constructive suggestions.  We also express  our profound  gratitude to
Barbara Hinton and Brian Eder for their Herculean  labors in typing this
manuscript and preparing the figures,  respectively.   Finally, since no
task can be finished without supervisory support and  guidance,  we wish to
take this opportunity to thank John F. Clarke and  Francis A. Schiermeier
for their patience and support, especially  during  the final  phase of
delivery of this document.

     Further we wish to thank Evelyn Poole-Kober for  her work in  obtaining
the references and for proofreading the manuscript.
                                  xxi 11

-------
                                SECTION 1






                     INTRODUCTION, PURPOSE AND SCOPE





     The basis of most applied diffusion modeling  today,  excluding  buoyancy



effects, was developed in the late 1950s and  early 1960s.   These  Gaussian-



empirical  models were built on the results of a few field  experiments



that were quite limited in scope and that had problems  with the tracers



employed and instrumental  accuracy,  although  they  represented  the state  of



the art at the time.  Those results  have been extrapolated  to  distances  15



to 125 times as large as the range of measurements and  to more extreme



stabilities; furthermore, they have been applied to much  rougher  or much



more hilly sites and to much higher  sources than used  in  the original



experiments.  Some of these extrapolations were guided  by  statistical



theory, but most were pencil-and-straightedge or freehand  extrapolations.



On this basis, hundreds of multi-million dollar decisions  were made and



continue to be made.  .





     This extrapolative situation, based on such limited  experimental



data, did not come about for lack of progress in diffusion  science  since



1960.  In Section 3, over 20 major field experiments and  numerous labora-



tory experiment are reviewed that were made after  the  two  experiments that



support most applied diffusion modeling today.   In Section  2,  large advances



in knowledge about atmospheric boundary layers  that have  been  made  since



1970 are reviewed; these advances provide much  support  for  better diffusion



modeling.   In Section 5, it is seen  that classical diffusion theories have



been further extended and that many  new theoretical  techniques with





                                    1

-------
considerable validation have been  developed  as  a  result of availability
of faster and cheaper computers.   Hundreds of new papers on diffusion
experiments, theory,  and model  validation appear  every year in conference
volumes or in journals.  Thus,  much  progress has  occurred since  1960 at
the research level.

     There is a perception in the  scientific community that the  best of
the newest work is  not being assimilated  into the regulatory  process in  a
rapid enough manner.   A counterview  is  that  to  effect the transfer of new
technology into the regulatory  process  requires more than a demonstration
of a new idea with  a  single data set; a proof of  superiority  over existing
techniques in wide  variety of cases  is  also  necessary.  Historically,
only a small fraction of the resources  for research has been  committed
for this transfer.

     The purpose of this paper  is  to provide an evaluation of what has
been accomplished by  past diffusion  research and  what most needs to be
done.  The scope is purposely limited to  "basic"  atmospheric  boundary-layer
diffusion, leaving  out the complications  introduced by source buoyancy,
plume deposition, chemical reactivity,  and complex flows due  to  structures,
large terrain features, and land-water  boundaries. This limit is imposed
because basic boundary layer diffusion, in itself, involves complex
phenomena and a large body of literature  (the  few hundred references used
in this report represent only the  most  classic  or most recent research
advances).  Also, a good understanding  of basic diffusion must be the
starting point for advances in  modeling the  more  complex diffusion problems,
Each of these problems can best be addressed separately, by specialists,
                                    2

-------
as some already have.  Most basic and specialized  areas in  diffusion research



were reviewed in Atmospheric Science and Power Production (Randerson,



1984), and complex terrain models were recently assessed by Schiermeier (1984)

-------
                                 SECTION 2

                        ATMOSPHERIC BOUNDARY LAYERS

          What have we learned  about their dispersive properties?

     The atmospheric boundary layer (ABL) refers to that part of the
atmosphere directly affected by turbulent mixing near the surface.  Heat,
water vapor, pollutants, and momentum mix through this layer much more
readily than through the air above it, which is only sporadically turbulent
(in clouds and in "clear air turbulence").  This layer of continuous
turbulent mixing, the ABL,  is virtually always present, although its depth
ranges widely from day to night,  season to season, and place to place (it
is absent only on windless  nights in topographic "bowls" of cold, dense
air).  It is also the part  of the atmosphere in which nearly all human
activity takes place, including release of pollutants.  Fortunately for us,
turbulence is effective at  dispersing and diluting pollutants; without it,
molecular diffusion would diffuse a point-source plume to a height and
width of only 4 m in 24 hr (edges defined by 50% of center-line concentration),
Unfortunately for those people  trying to predict turbulent diffusion, the
rate at which it proceeds varies  quite widely, depending on wind, atmospheric
stability, the height of the ABL, the height of release of pollutants and
surface characteristics in  a rather complex way.

     The ABL may be viewed  as the place where the large scale pressure
systems constituting "weather"  interface with the planet's surface.  The
geostropic wind, ug, is a measure of  forcing by the pressure field, which

-------
changes as the large migratory systems  ("highs"  and  "lows") move  across the
land.  The surface fluxes of heat,  moisture, momentum,  and  pollutants  into
or out of the ABL are all strongly  influenced  by surface  cover  and  local
land-use (a pasture has very different  texture and thermal  properties  from
those of an industrial  area, for example).   It is in  the  ABL that the
influences of large scale forcing and local  surface  fluxes  are  recounciled;
this process accounts for the great variability  in wind,  temperature,  and
diffusion rates which we observe.

     The discussion that follows begins with definitions  and explanations
of all meteorological variables important to diffusion  in ARLs.   Then  the
typical diurnal cycle of meteoroloigcal  change and some of  its  effects on
diffusion is described, with particular attention to  transitional periods.
Finally, current understanding of ABLs, their  turbulence  structure  and
their diffusive abilities are described.  Because stability has the greatest
effect on diffusion rates and because there  are  basic structural differences
in the turbulence of stable boundary layers  (SBLs),  neutral boundary layers
(NBLs), and unstable or "convective" boundary  layers  (CBLs), it is  convenient
to consider these separately.

IMPORTANT METEOROLOGICAL VARIABLES  AND  SCALES  AFFECTING DIFFUSION

     Regardless of the stability, the same basic collection of meteorological
variables are of prime concern for  diffusion modeling.  They will be
introduced in this section.  Overbars indicate a time average,  and  primes
indicate deviations from the time average.   Subscript "o" refers to surface
values.

-------
     There are several  ways  to  measure  stability.   One measure is the
static stability, which in  an  incompressible  fluid  is -dp/dz, where p
is the fluid density and z  is  height  above the surface.  This can be used as
a stability measure in  air,  but the more  easily measured quantity is dT/az,
where T is absolute temperature (air  is nearly an  ideal gas, so at constant
pressure, p « 1/T).  Because air is compressible,  a better stability
measure is 38/dz, where 9 is the "potential"  temperature defined
by 9 = T(p0/p)2/f7 (p is ambient pressure, and p0 is a standard, sea level
value, usually 1000 mb).  This  quantity compensates for heating and cooling
due to pressure changes, so  that if air rises or falls without radiating or
absorbing heat its 9 is constant, i.e., 9 is  a conservative property (this
type of process is called "adiabatic").   If 59/az  = 0, the stability
is "neutral", meaning that  if  air is  displaced vertically, it remains at
the same 9, T, and p as ambient air on  that constant pressure surface (p
is primarily a function of  z);  thus,  it develops no buoyancy.  However,
if 59/dz > 0, upwardly displaced air  will find itself surrounded by
higher 9, higher T, and lower  density air; therefore it will be heavier
than the ambient air.  This  displaced air acquires buoyancy, or negative
buoyancy, which tends to return it to its original  height; hence, the air
is "stable".  If d9/az < 0,  the opposite  happens,  and displaced air
acquires buoyancy acting in  the same  direction as  the displacement.  This
displaced air tends to keep rising or falling, even accelerating, and the
air is called "unstable".  Turbulence develops spontaneously in unstable
air; this is called "convective" turbulence.

-------
     There is a simple relationship between  the  vertical  gradients  of 9 and
T, derivable from the definition of 9, the ideal  gas  equation  of  state,
and the hydrostatic equation, ap/az = - pg,  where g  is  gravitational
acceleration (this equation follows from the fact that  static  pressure is
just the weight per unit area of the air overhead).   For  the earth's
atmosphere, the result is

           as/az = (aT/az + o.ni°c/m) (p0/p)2/7                        (2.1)

     Naturally, the larger the 39/az, the more stable the air  is;
0 < 39/az < 0.01°C/m is considered  slightly  stable;  39/az > 0.01°C/m
is usually considered moderately stable, unless  a9/az > 0.03°C/m,
which is roughly the dividing line  for "very" stable.  However, static
stability is not the only consideration.  A  vertical  gradient  of mean wind
speed, aU/az, tends to encourage overturning and  the  development of
turbulence.  If equal amounts of air from two layers  mix,  one  layer with a
mean wind speed TT = uj and the other with U  = u;?,  the conservation of momentum
principle gives u" = (u^ + U2)/2 for the mixture.   However, the kinetic energy
(KE) of mean motion, U^/2, is less  than the  original  amount, (u^/2 + U2^/2)/2
by the amount (uj> - u^)^/4.  This surplus KE feeds turbulent motion, i.e., it
becomes turbulent KE.  This source  of turbulent  KE is present  whenever aU/az
is not zero, regardless of sign. These considerations  lead to a very
useful stability index, the gradient Richardson  number:
          Ri = (g/9)(ae/az)/(au/az)2                               (2.2)
This number is dimension!ess; it compares the potential energy increase needed
to mix stable air with the surplus  KE that would  be  available  after mixing

-------
due to au/az.  When Ri  > 1,  turbulence is  extinguished  rapidly,  as
there is not enough surplus  KE to drive air mixing  "uphill"  against  the
positive a0/az.  At Ri  = 0.5 there is an approximate energy  balance,
and at Ri < 0.25, turbulence develops spontaneously from  the slightest
perturbation, even if ae/az  is positive.  Thus,  Ri  is a physically
meaningful stability index;  it also highly correlates with diffusion rates
in certain cases, which are  discussed in Section 4.

     Wind speed and its fluctuations are very important quantities  in
diffusion modeling because they have four effects:  transport, axial  dilution
(for continuous releases), shearing of the plume or puff, and turbulence-
induced growth and dilution.  The usual meteorological  convention  is to
define x as the longitudinal horizontal spacial  coordinate  (extending in
the direction of the mean wind), y as the lateral  horizontal  coordinate
(perpendicular to the mean wind), and z as height above the  surface.  For  a
point source, we define x =  0 and y = 0 at the source.  The  corresponding
components of velocity are u, v, and w, respectively.  The direction of  x
is generally defined by the  mean wind direction  at  the  source height, so
that v" = 0 at that height.  However, if the wind direction turns with
height, a7/az * 0; this tends to increase the plume width with distance,
as the plume top is transported in one direction, and the plume bottom is
transported in another.  Turbulent velocities are defined by deviations
from the time average, e.g., v1 = v - 7.  Averaged  over the  same time span,
or "sampling time", we must  have U7", "v"1", and w1" = 0. The mean motion in
the atmosphere closely parallels the ground, so over flat ground we usually
can assume that ^7=0 also.   Turbulence velocity standard deviations are
                                     8

-------
defined by av = (v'v1 j, etc.  These are very important quantities  in
diffusion modeling, as the classical  statistical  theory prediction (Taylor,
1921) for diffusion close to the source is well proven: 
-------
friction effects, the roughness  length,  z0.   It  can  be determined  experimen-
tally from wind speed profiles  near the  ground,  and  it scales  roughly  as
about 1/10 the height of most  "drag elements"  on the surface  (trees, grass,
crops, buildings, etc.); z0 can  be crudely estimated by  inspection of  a
site (Weiringa, 1980).  The velocity scales  Ug and u* are  roughly  proportional
but Ug/u* is influenced by z0  and  by the stability.

     Turbulent velocities and  diffusion  near the surface,  especially their
vertical components, are strongly  correlated with u* and the  flux  of buoyancy
due to heating or cooling of the surface.  This  flux by  (g/Tv)w'T'v =  H*,
where Tv is the virtual absolute temperature,  which  includes  the effect
of water vapor on air density  (Tv  = T/(l - 0.6q), where  q  is  the specific
humidity).  Except over very moist surfaces, the humidity  effect is small
and H* may be determined by using  only the "dry" temperature,  T, in place of
Tv.  H* represents the turbulent flux of positive buoyant  acceleration
caused by heating of the air (or vaporizing  of water) next to  the  ground;
it is proportional to the upward sensible heat flux  plus about 1/7 of  the
latent heat flux (Briggs, 1985a).   When  the surface  is colder  than the
surface air, the heat flux is  downward,  and  H* is negative.  H* is also
equal to the rate per unit mass at which turbulent energy  is  created by
buoyancy; it is the rate of potential energy release in  vertical,  turbulent
motions.  When H* is positive,  99/5z becomes negative due  to  stronger
heating near the surface, and  convective turbulence  is produced.   When H*
is negative, 99/az becomes positive due  to stronger  cooling near the
surface, and H* represents a rate of loss of turbulent energy.  When
turbulence is produced by wind shear, it is  called mechanical.  The

                                     10

-------
rate of mechanical production of turbulent energy is tbu/az, so
u*^ = TO also relates to turbulence production.  Both u* and H* can
be measured using fast-response measurements of u1, w1, and T'  (or Tv').

     Together, u* and H* form an important length scale, the Obukhov length
(Monin, 1959), L = -u*3/(kH*), where k = 0.4, the von Karman constant
(Hogstrom, 1985); this length is defined negative when H* is positive
(unstable conditions).  Near the surface, it is well established that
 aU/az = k-iuVz, so the ratio of buoyant to mechanical  production of
turbulent energy is H*/(u*2dTT/az = -z/L.  Thus, z = |L|  is a dividing
line between dominance of buoyancy effects (above |L|) and mechanical
effects (below |L|).  According to surface-layer similarity theory, dis-
cussed here in Section 5, aU/az = u*/(kz) times a function of z/L and
(g/9)ae/dz = H*/(ku*z) times another function of z/L; these hypotheses
have been well supported by field measurements (e.g., Businger  et al.,
1971).  It follows that in the surface layer (the lowest 10% or so of the
ABL), R-j is a function of z/L.  It is also possible to relate either
Ri or L to the Bulk Richardson number, R^ « AG/TT,  by using surface-layer
similarity equations for IT and 9 (Irwin and Binkowski, 1981).  Both L and
Ri have been shown to be highly correlated with vertical diffusion from
surface sources (e.g., Weber et al., 1977).  This is consistent with the
fact that vertical turbulence quantities, e.g., aw/u*, have been shown to
be functions of z/L.

     The wind speed gradient in the surface layer,  kzau/az = u*m,
where m is a function of z/L, can be integrated to give
                                     11

-------
          u = (u*/k)[ln(z/z0) - <|,m],                                (2.3)

where <|>m is a function of z/L, given by the integral  of (1 - m)d1n(z/L).
This equation provides a way to calculate u for an elevated release using
U measured closer to the surface, if we can measure or estimate L.

     The remaining parameter that is of great importance in describing the
diffusion potential is the height of turbulence mixing in the ABL.  For
CBLs this is usually designated as zj (most CRLs are capped by a temperature
inversion, i.e., positive aT/az, hence the subscript "i").  For SBLs this
is usually designated as h, which will be used here for NBLs also.  First of
all, it is important to know whether an elevated source is within or above the
mixing layer, as there is normally no turbulent diffusion above this layer.
Except in clouds and in occasional patches of "clear air turbulence", the
atmosphere above the ABL is always stable and non-turbulent, regardless of
stability swings at the surface (the upper air is not subject to intense
heating and cooling like the air next to the surface).  Especially on clear
nights with low Ug, when h can be quite small, elevated and even medium-level
sources may be above h.  Second, turbulence intensity and structure is a
function of z/z^ or z/h; the intensity falls off rapidly with height near
z-j or h.  Third, in the CBL it has been found that the size of the largest
turbulent eddies approximates zj and that cru and av are proportional to
the scale velocity w* = (H*z1-)^'  throughout the mixing layer.

     In summary, in the order of discussion, the meteorological variables
and scales most relevant to diffusion are aT/az, 39/az, Ri, TJ", V,
 aw» f» ug» u*» zo» H*» L» and zi or n-
                                     12

-------
THE DIURNAL CYCLE AND TRANSITIONAL PERIODS

     Before discussing neutral, convective,  and stable boundary layers
considered separately, it is worthwhile describing  what is  known  about  the
normal diurnal sequence of stability changes in the ABL.  This  description
excludes overcast, windy conditions, which tend towards neutral  regardless
of the time of day.  In the course of a typical  day, the  ABL  goes through a
wide range of structure, being stable at night  and  convective during most
of the daylight hours.  During the transitional  periods around  sunrise  and
sunset, rapid changes occur.  Rapid changes  also occur during thunderstorms,
at the onset of sea breezes, and with frontal  passages.  Because  these
latter events are accomplished by brisk wind speeds, they are not likely to
be causes of serious pollution episodes.  The steady state  assumptions
customarily made in diffusion modeling can be more  or less  justified for
1-hr averaging during the major phases of the 24-hr cycle,  but  have little
validity during the transitional  periods, as will be seen.  These
transitional periods have received less study than  the major  ABL  states,
but they can be described at least qualitatively.

     A good time to begin this description is the hour before sunrise,
because this is ordinarily a time of steady  state.   Radiative cooling has
created stable stratification in the lowest  300 to  1000 m,  and  there may be
a remnant of the inversion above the previous day's maximum z-j  (radiative
cooling can intensify the inversion if there was much haze  below  that ZT).
There is stronger stability near the surface and a  shallow  mixed  layer  to
h produced either by (1) wind shear-generated turbulence, especially
                                     13

-------
when Ug is large, (2)  slope flows  generated  by  cooling  of the  surface air
and gravity, especially when Ug  is small,  or (3)  convective mixing produced
by more intense radiational  cooling at  the top  of a  fog extending to depth
h, especially when Ug  is small  in  humid conditions on clear nights.

     When sunlight first reaches the surface, at  low angle, most of its
energy goes at first into evaporating dewfall or  fog; this latent heat
produces some increase in buoyancy, as  water vapor is lighter  than air, but
only about 1/7 the buoyancy produced by the  sensible heating of air.
Consequently, there is not much  convective activity  until most of the
liquid water or frost  has evaporated.  This  may take no time at all in a
desert, about an hour  for typical  dewfall  or frost in a humid  climate, or
half a day for a deep  fog.  After  this  evaporation is accomplished, a much
larger fraction of solar energy  ends up as sensible  heat in the air next to
the surface.  The air  becomes lighter and  rises;  small  thermals converge to
produce larger ones, and these impinge  against  the top  of the  mixed layer,
overshoot a little into the stable air  above, mix with  some of it, and
bring it back down as  the mixture  subsides and  feeds into downdrafts.  This
is the beginning of the development of  the daytime CBL. The thermals cause
large undulations in Zj, especially in  the first  hours  of CBL  development.
If there happens to be an elevated layer containing  pollutants, as occurs
when a buoyant plume rises and stratifies  in non-turbulent air at night,
the pollutants, too, are entrained by invading  thermals and are mixed down
to the ground in the "inversion  breakup fumigation"  process.   The depth z-j
continues to grow until early to midafternoon when solar heating and H*
begin to diminish.  Sometimes during the day ZT appears to "jump", as

                                     14

-------
thermals can traverse a near-neutral  layer aloft  quite easily once they
reach nearly the same potential  temperature as the neutral  layer.   The
average 9 is nearly uniform through the bulk of the mixed  layer,  and  it
increases steadily through the day as long as H*  is positive (until about
1 h before sunset).  The heating rate diminishes  through midday and afternoon
as H* decreases and the depth being heated, z-j, maximizes.   The negative
39/az layer near the ground reaches maximum in intensity near noon,
when H* is at a maximum.  As H* decreases  in midafternoon,  this negative
a9/az decreases, so the surface temperature may begin  to decrease  even
though H* is still positive and 9 is still  slowly increasing above 0.1 z-j.

     Exactly what happens in the CBL in the very  late  afternoon is still an
area of speculation.  In clear conditions, H* goes to  zero  about an hour
before sunset, as outgoing thermal  radiation from the  warmed surface  nearly
balances incoming solar radiation.   The CBL may begin  to collapse  before
H* goes to zero as the thermals become weak.  Convective downdrafts outside
the thermals have a slight stable stratification  due to positive
39/at = -wa9/az.  When new thermals are very weak, they cannot  pene-
trate through this stable stratification all the  way to the previous  z-j, so
the mixing depth descends.  It is not known how rapidly this occurs.   The
continuation of positive moisture flux from the surface during  z-j  collapse
may produce a negative humidity gradient,  which leads  to more rapid radiative
cooling at lower heights, and further stable stratification that will
hasten the demise of turbulence aloft.  In dry conditions,  the  CRL collapse
process may go slower.  Without cooling due to radiative flux divergence
(due mostly to humidity), the stratification remains close  to neutral  to z-j
                                     15

-------
and the turbulence gradually "runs  down",  with  a  time  scale ~  ZT/CJW, on
the order of 1 hr in the late day.

     While what remains of the C8L  aloft  is  running  down or collapsing due
to stratification developing from radiation  flux  divergence, the  SRL begins
development before sunset, as soon  as  H*  turns  from  positive to negative.
As the sun sinks, incoming solar radiation falls  to  zero,  while outgoing
long-wave radiation from the surface continues  all night.  Because the surface
emits almost like a black body, this radiation  flux  is « Ts4,  where TS is
the absolute surface temperature.  In  spite  of  nocturnal cooling, absolute
temperature does not change much, so the  radiation balance at  the surface is
rather constant on clear nights (if a  cloud  deck  moves over the site, it
radiates back like a black body at  slightly  smaller  T^, so the outgoing
surface radiation is almost balanced and  the net  radiation is  reduced
drastically).  The net radiative cooling  at  the surface is divided between
cooling the ground slab, condensing or freezing water  in any dew, fog, or
frost, and cooling the air next to  the ground.   If it  is very  humid, as
after a summer shower, fog or dew can  form right  away, but usually this
does not occur until later at night as the surface temperature approaches
the dew point; of course, this does not happen  at all  in dry conditions.
The soil heat flux to the surface is fairly  constant through the  night, but
the surface cooling rate decreases  as  the heat  conducts out of an increasingly
deeper ground slab (Garratt and Brost, 1981).

     Surface parameters as well as  SRL structure evolve rapidly in the
first few hours after transition.  This is evident  from the results obtained
                                     16

-------
with second-order closure ABL models  (Wyngaard,  1975;  Brost  and  Wyngaard,
1978), which appear to give good simulations  of  observations from the  1973
Minnesota experiment (Rao and Snodgrass,  1978; Wyngaard,  1975).  The heat
flux becomes increasingly negative up to  a  point,  about  1  hr after transition,
and then gradually diminishes in magnitude  in spite  of the relative constancy
of the net radiation flux.  At the same time, u*,  L, and  h rapidly reduce
from their values at t = 0 (transition),  but  they  may  approach a rough
steady state in 3 or 4 h, if not interrupted  by  a  turbulent  burst (a sporadic
episode of turbulent mixing due to changes  in the  structure  of Ri).  This
probably happens because the increasing stability  near the surface restricts
turbulent motions, so that u* drops,  and  as turbulence at  the surface
diminishes, the heat transfer from surface  to air  is restricted, which
causes H* to drop.  (Heat must first  pass through  a  "laminar sublayer"
whose thickness is proportional  to v/u*;  this is the same  insulating
layer that our bodies lose at higher  wind speeds,  hence, the "wind chill"
factor.)  The bulk Richardson number  for  the  whole layer gives some insight
into the behavior of h.  Wind shear can support  turbulence working against
stable stratification only as long as Rb  =  (g/9)(9^  -  e^n/u^ is
less than 0.3 or so.  The surface air cools more rapidly than air higher
up, so if ITh remains constant, the increasing (eh  -  QO)» which increases
the "heaviness" of the surface air, pulls h downward.  Of  course, tTn can
also change as h moves downward, or as momentum  is transferred in gravity
waves or due to changes in driving forces like Ugf.  Thus, h  can behave
rather irregularly through a night (e.g., see Fig. 2).   In some of the
Minnesota runs, Ug dropped during the night and  turbulence became immeasurable
                                     17

-------
after about midnight.  In such  cases,  we might  as  well  forget  about turbulent
diffusion until  sunrise, and just  try  to determine in  which direction a
plume of pollutants is heading  and whether  it  is spreading due to horizontal
eddies or meandering.  In this  situation, there is no  substitute for wind
speed and direction measured near  the  point of  release.

     Several efforts have been  made to mathematically  simulate the complete
diurnal  behavior of the ARL, e.g., Yamada and  Mel lor (1975) and Binkowski
(1983).   Roth of these simulations show a rapid decay  of  CRL turbulence  in
the late afternoon, and predict a  maximum in the negative surface heat flux
in the SBL about 2 hr after transition to negative values.

     The temporal evolution of  a fairly typical  CBL, as shown  by actual
measurements, is plotted in Fig. 1.  The measurements  are from the Boulder
Atmospheric Observatory of NOAA's  Wave Propagation Laboratory  during the
CONDORS experiment (Moninger et al., 1983). The day,  13  Sept. 1983, was
cloudless and very convective;  u" dropped as low as 2 m/s  during the midday,
giving Tj/w* = 1  and  |z-j/l |  exceeding 100.  Fig. la shows  the development of
z-j and T at 10 m.  For about an hour after  sunrise there  is evidence of
continuation of a drainage wind from the west  up to a  height of about  100 m
(this flow is peculiar at this  site and probably originates from the east
slope of the Rocky Mountains, which begin 25 km to the west).  This  flow
breaks up soon after the heat flux turns positive  (Fig. Ib) and ZT climbs
more or less steadily until noon,  at which  time it seems  to reach  its  limit
of = 1000 m for that day (z] sometimes leveled off for several hours during
                                     18

-------
(This  page intentionally  left blank)
                19

-------
                                                                        BAO
Figure
                                                                                                                           T
                                                                                                                           E
                                                                                                                       - 22 M
                                                                                                                           p
                                                                                                                           E
                                                                                                                         18 R
                                                                                                                           (C)
                                                   +Z;.LIOAR  • Z: . RAMXNSONOE  0Z;.TOHER  S Z; . SOOAR
                                                     (V-^Z; . INTERPOLATION   -«- TEMPERATURE AT  ION
                                  1.00-

                               E
                               N
                               E  O.7S-
                                                                          TIME - MST


                                                    0  INSOUATZON.  LANOLEYS/MIN     -|-  (8X1 M' T,'  »C-N/S
                                  2.28-


                                  2.00-


                                  1.78-


                                  1.5O-


                                  1.29-


                                  l.OO-


                                  O.7S-


                                  o.ao-

                                H
                                - O.28-
                                S

                                  o.oo-
                                                                          TIME - MST
                                                         , AT  10. 200M
                                                                               AT  10. 200M    * u* -|-u*
                                            00©0®
                                                                                   X
                                                                             X  +X   X
                                                                                                     e    x
                                                                                                       0 +
                                                                                         0        V  0-
                                                                                                      A
                                                                                        ,    2
                                                  •X
. 80 u
    r
    N

 40 °
    D
    I


    C
    T
    I
 3200
    N
                                                                           TIME - MST


                                                            DIRECTION AT   10, 200M     •) 0  WIND SPEED AT  IO, ZOOM
                                                                             20

-------
this experiment, and on some days it appeared to grow in stair-step fashion).
The scatter in the lidar measurements, which were made at various azimuths
and shot elevations, give an idea of the magnitude of undulations in local
z-j, perhaps ± 100 m.  In Fig. Ib we see that the sensible heat flux (« w'T1)
is very much controlled by the insolation, although it lags behind insolation
in the first hours of daylight; this lag may be due to both outgoing thermal
radiation from the surface to the clear sky and channeling of solar energy
into evaporation of wetness in the ground cover (predominantly deep grass).

     Figure Ic shows turbulence velocities at 10 and 200 m together with  w*
and u*. .Although there is much scatter in the av values and the aw
values at 200 m, after ZT passes through 200 m at about 8 AM, they appear
to be essentially controlled by w*.  The 20-min average u* values are much
smaller and fluctuate widely.  Note that w* is nearly constant from 10:10
to 14:10.  In contrast, av and 
-------
hours.  Before CBL development,  the wind  was  from the north  at ~5.6 m/s
at 200 m and from the west at  ~2.1  m/s  at 10  m,  in the presumed drainage  flow.
The wind directions become easterly (from the east)  at midday, probably due
to forcing from the upslope flow on the east  slope of the Rockies  to  the
west, and phenomenon absent at sites far  from mountains.   The wind speed  at
200 m is quite steady until 2-5 approaches that height at  07:20; then  it
drops and fluctuates widely during  the  most convective hours until it
regains some of its steadiness at w'T'  and w* become negligble about  1 hr
before sunset.

     The temporal evolution of a SBL at a flat site (Cabauw, the  Netherlands),
as shown by measurements, is plotted in Fig.  2.   The night,  30-31  May 1978,
was clear and the large scale driving force was  fairly constant,  as is
shown by the geostrophic wind speed and direction in Fig. 2d.   In  spite of
this we see in Fig. 2a large oscillations in  h,  as determined by  the  top  of
the layer with significant downward heat  flux.  Note that the acoustic
sounder or "sodar" representation of h  is in  general agreement, but does
not fluctuate as widely as h determined by w'T1  < 0.001°C-m/s  (this
determination was made by extrapolation in the range 200-270 m).   The
temperature at 10 m falls almost steadily until  45 min after sunrise, and
then rises rapidly and steadily. Various energy fluxes near the  surface
are plotted in Fig. 2b.  The net (incoming minus outgoing) radiative  flux
crosses to negative about 1 1/4 hr  before sunset and then becomes  practically
constant through this clear night;  it crosses to positive about  1 1/4 hr
after sunrise, rapidly increasing.   The tranpirative response  of  the  grass
and trees may account for the continued upward flux of water vapor (positive

                                     22

-------
                                CABAUW
                                        0 _     	^oj> of Tpwer_
(M)100-
                     !®s==©
     IB   17   IB   10   20   21   29   23    0    1    2
  H SOOAR    0 M TOP OF w* T'  < -O.001   •+• TOP OF w' T' » O.OO3    — TEMPeP-ATUHC AT 1
                               * * *    "a"*" X
     10   17   IB   IB   20   21   22   23    O    1    2    3    4    6
                                                                                     Figure  2
                                                  LUX. 2OM  © UH. FRICTION VEuOCITY
  0.5-


  O.4-


  0.3-


  0.2-

M
- 0. 1-
s   :

  0.0-
       »   * •
          A
               *A  .
                   *
                                                               • A
                                                              O  *
OO*  0 O
                        AVA*A*A*
                         * *
                                  -.   AAAV^    °  *  A
                                  '*    *#A##     *AA
     Ifl  17  IB  IB  2O  21  22  23   0    1    2   3

                                  TIME - GMT
                                                          O   8   7
         A,A  6,  AT 20.200 M   f>p 6V AT 20.
    12-

    10-
                          x x xxxx
                                          D

                                       100 I


                                       80  c

                                          T
                                       60  I
                                          0
                                      •40  S
     10   17   IB  IB  2O  21  22  23   0   t   2   3   4   9   S   7   a

                  Sunset           TIME - GMT         Sunrise
                 -{- M2NO DIRECTION AT 20. 200M ^,O WIND SPEED AT 20. 2OOM
                                  23

-------
latent heat flux)  to 1 hr after sunset  and  its  quick  reversal back to  upward
flux minutes after sunrise.   The early  transition  from  positive  to negative
heat flux, about 3 hr before sundown, is  surprising to  us  (the authors);
perhaps the grass  surface is cooling due  to transpiration  more rapidly than
it absorbs solar heat (late  afternoons  in summer,  grass often feels cool  to
bare feet).  The delay in positive sensible heat  flux until  2 1/2 hr after
sunrise mirrors the evening  transition  behavior;  this could  be due to  wind-
induced evaporation of dew.   During the SBL hours,  Fig. 2b shows a striking
correspondence between sensible heat flux and u*,  supporting the constant
T* = w'Tr/u* approximation suggested by van Olden  and Holtslag (1983).

     In Fig. 2c we see that  variations  through  the night in  turbulent
velocities also strongly follow variations  in u*,  especially the velocity
measurements at z = 20 m.  However, aw  at 200 m tends to not respond to
u* and remains at a near constant low value during periods of h  < 200  m
(roughly 1900-2300 and 0200-0700 GMT).   These small 
-------
10 m rises with u*); less stability causes  less  suppression  of  vertical
turbulence, which tends to increase surface friction.   Immediately  after
transition to positive heat flux at 0600 (Fig. 2b)  we  see the rapid develop-
ment of a CRL (Fig. 2a); in its early development,  it  is  capped by  a thick
layer of negative heat flux.

     In Fig. 2d we see relatively small  differences between  20- and 200-m
values of wind speed and wind direction  just before evening  transition  and
increasingly independent behavior afterward. This  characterizes a  principle
difference between CBLs and SBLs.  As h  drops below 200 m, 71200 approaches
and oscillates around ug, the geostrophic wind speed.   (Some of these
oscillations might be due to variations  in  the local Ug or pressure field;
the measured values represent Ug on a large scale.)  In constrast,  ^20  changes
little during this night and is seemingly indifferent  to  changes in other
variables.  After evening transition the wind direction at 20 m begins
turning to the left of the geostrophic wind, which  means  that it turns  toward
low pressure; this is normal behavior near  the surface in a  SBL. This  wind
direction difference begins to diminish  in  the morning at the same  time
that TIQ and u* begin to rise, indicating increased coupling with upper
flow at this time.  The wind direction at 200 m  approaches the  large scale
geostrophic wind closely until 2300, then turns  to  the right nearly 50°, an
event for which we have no explanation.

     Figures 1 and 2 illustrate dramatically the many  complex changes that
occur in ABLs around the times of morning and evening  transition.  They are
also illustrative of many of the understood features of quasi-steady state
                                     25

-------
CBLs and SBLs, so will  be useful  references  for the C8L  and  SBL  subsections
to follow.  In addition to the understood features, measurement  records
like these often reveal events we cannot  readily explain,  such as  the  brief,
marked increases in magnitudes of turbulence variables at  Ifi30 MST in  Fig.  1
and at 0345 GMT in Fig. 2.

NEUTRAL BOUNDARY LAYERS

     A neutral boundary layer is  one in which buoyancy forces are  negligible.
Thus, 56/az = 0 throughout a NBL  and temperature drops 1°C per 100 m
increase in elevation.   For this  condition to hold near  the  surface, H*  must
be zero or very small  because at  small  z/L,  59/5Z = (9/g)H*/(ku*z).
Also, the production (or dissipation) of  turbulent energy  by buoyant forces
must be small compared  to the production  by mechanical forces  (wind shear);
this requires that h <  |L|.  This condition  is met rather  infrequently,
which is perhaps why there has not been as much research on  NBLs as on CBLs
and SBLs.  As L « u*3/H*, a combination of high winds and  negligible heating
or cooling of the ground is required to produce a neutral  boundary layer.
This latter requirement is most likely to be met during  overcast conditions,
when very little solar radiation  reaches  the ground and  most upward long-
wave radiation from the surface is balanced by downward  radiation  from the
cloud base (clouds and most surfaces radiate approximately as  black bodies).
There are also moments of H* = 0  just after sunrise and  just before sunset
in the normal diurnal  cycle when  it is not overcast, as  H* changes sign  from
positive in the day to negative at night.  This condition  is so  transient
that the ARL does not have time to adjust to any kind of neutral structure.
                                     26

-------
     The NBL is driven by the pressure gradient;  the magnitude of the
acceleration in the direction of the gradient is  Ugf, which  follows  from
the definition of ug.  Balancing this when the NBL is in a steady state
(no acceleration) are the Coriolis accelerations  TTf, directed  to the right
of the wind direction in the northern hemisphere, and fractional  drag.
This drag is given by a-t/az, where T is the horizontal  stress; T = u*2  at
the surface and goes to zero at the top of the NBL.   Near the  surface,  U
is small, so the Coriolis acceleration is weak and aT/az nearly
balances the pressure gradient acceleration, Ugf; higher up, the Coriolis
acceleration is stronger and a-t/az becomes weaker, so the wind turns
more to the right and approaches the geostropic wind near z  =  h.   The two
velocity scales ug and u* are more or less proportional, but ratio is
influenced by the surface Rossby number, u*/(fz0).  The ratio  Ug/u*  ranges
from about 20 over rough surfaces to about 50 over very smooth surfaces
(water and snow) in neutral  conditions (Counihan, 1975).

     Turbulence velocities in the NBL are most closely related to u*.   The
vertical velocity standard deviation, given by crw =  (w'w1 J1/2, is about
1.3u* near the surface (Hanna, 1981).  The lateral  (across the wind)  and
longitudinal (along the wind) components of velocity variance, ov and au,
respectively, also scale to u*, but observations  show some variability  in
the ratios crv/u* and au/u* at different sites. This may be  caused by
surface inhomogeneities or topographic influences; av/u* ranges from
about 1.3 to 2.6 and au/u* ranges from about 2.1  to  2.9 near the  surface
(Lumley and Panofsky, 1964).  Observations of these  quantities high  up  in
the NBL are lacking, but Hanna (1981) suggested exponentially  decreasing

                                     27

-------
values using the scaling height  u*/f;  this  is  as  an  approximate  fit  to  a
second-order closure numerical model  of the ABL  (Wyngaard  et  al.,  1974).
Hanna also suggested an explicit form  for the  Lagrangian time scale;  this
is roughly the time it takes diffusing particles  to  "forget"  the turbulent
velocity they experienced at release,  and it is  an  important  time  scale in
diffusion theory (especially statistical  theories.)

     The height of the NBL is of concern, since  it  is  where turbulence
velocities drop off and where the wind speed approaches the geostrophic
value.  It is usually assumed that h  « u*/f, but  this  can  only be  true  if
99/dz = 0 through a deeper layer, so  that h is not  limited by stable capping.
In this case, Deardorff's (1972) large eddy numerical  modeling indicates
that H = Ug at a height 0.2u*/f, but  turbulence  velocities fall  off  gradu-
ally, rather than abruptly.  His results are in  rough  agreement  with Hanna's
(1981) approximations, av/u* = aw/u*  = 1.3  exp(-2zf/u*), except  that
Deardorff's av does not fall off significantly from  its surface  value
until z > n,17u*/f.  Typically,  u* =  0.5 m/s with IT =  5 m/s at z = 10 m, and
f = 10~4 s-1 at midlatitudes, so 0.2u*/f is ~  1000  m.   Obviously,  u*/f  cannot
be the correct form of h near the equator,  because  there f =  0.   NBLs can
also be limited by stable capping at  z = z-j, just like CBLs,  especially when
z-jf/u* is small.  Krishna and Arya (1981) investigated this case with an
eddy diffusivity model.  They concluded that there  is  no significant
entrainment at z-j when Zi > 0.2u*/f,  but there is significant entrainment
when Zj = 0.05u*/f, along with other  changes in  boundary layer structure.
Yokoyama et al., (1977) analyzed aircraft turbulence measurements  in five
early-morning to mid-morning NBLs capped by an inversion,  for which  hf/u*

                                     28

-------
ranged from 0.09 to 0.16 (h was defined by the zero  intercept  of measured
turbulence energy dissipation rate profiles;  h *  zj  within  +150  m and  -50  m).
For these boundary layers, aw was observed to be  proportional  to (1  -  z/h),
a linear dropoff, in contrast to the exponential  dropoff  predicted by  NBL
models with z^f/u* = 0.45 or larger.

CONVECTIVE BOUNDARY LAYERS

     Our understanding of CBLs, which prevail  during the  daytime, has
greatly improved since convective scaling  with z^  and w*  =  (H*z1-)  '^ was
proposed by Deardorff (1970a, b).  Experiments in  CBLs in Minnesota  (Izumi
and Caughey, 1976; Kaimal et al., 1976), England  (Caughey and  Palmer,  1979),
France and Spain (Druilhet et al., 1983),  and other  places,  including  over-
sea measurements, have confirmed the utility  of this scaling for most
turbulence-related quantities, including diffusion rates  (for  a  review of
the development of convective scaling and  its application to diffusion, see
Briggs, 1985a).  The associated temperature scale, 9* = w'T'/w*,  has also
proven useful  for describing the temperature  profile and  temperature
variances in the lower half of CBLs.  For  a comprehensive review of knowledge
of CBLs as of 1981, see Caughey (1981).  Unreferenced information in the
discussion below can be found in Caughey's review.

     An ABL with turbulence driven mainly  by  buoyancy, rather  than by  wind
shear, is considered convective.  Usually, the buoyancy arises from the
surface due to heating by solar radiation  absorbed at the surface.  However,
occasionally an "inverted" CBL develops due to cooling at the  top of a
cloud or a fog layer produced by outgoing  long-wave  radiation; because the

                                     29

-------
magnitude of this heat flux  is  considerably smaller than the typical heat
flux produced at the ground  by  solar  heating, wind speed must be fairly low
for convective effects to  dominate.   Cloud-topped CBLs and convective mixing
in fogs have received very little  attention.

     For the more usual  CBL, with  turbulence driven by heating of the
surface, highly efficient  vertical mixing  is observed to develop; major
updrafts and downdrafts extend  through  nearly the whole layer (Willis and
Deardorff, 1976b).  The following  are some consequences of this vigorous
vertical mixing: (1) horizontal  momentum becomes well-mixed, so that there
normally is negligible wind  shear  from  z/zf = 0.1 to 0.9, (2) potential
temperature becomes well-mixed, but there  is an increasingly negative d0/5z
towards the surface, and (3) passive  material becomes almost uniformly
mixed in the vertical in a time =  4 z-j/w*  after release, typically  about a
half hour.  The CBL is capped by stable air; waves in this air and  undulations
in Zi, are produced by the impingement  of  stronger updrafts against the
stable capping layer.  These strong updrafts also entrain some of the
warmer, stable air aloft and bring it back down into the CBL, creating a
negative heat flux at the top of the  CBL;  this usually ranges from  0.1 to
0.4 times the surface heat flux in magnitude.  Another consequence  of this
entrainment is that significant horizontal momentum is brought into the
CBL; if there is a large difference between the wind in the CBL and the
wind just above z-j, this entrainment  can cause considerable wind shear in
the upper part of the CBL.
                                     30

-------
     In CBLs, velocity variances are most closely related to w*, at least
when Zj > 20|L|.  (The ratio z-j/|L| is considered the index characterizing
the degree of convectiveness.  A z-j/|L| > 50 can be considered "strongly"
convective; however, using a numerical model, Deardorff (1972) found
significant changes from NBL turbulence structure at z-j/|L| = 1.5, which is
considered only weakly convective.)  Panofsky et al. (1977) fit measured
values of both  |L| but < 0.1 zi (provided
that z-j/|L| » 10), and aw = w* times a function of z/z\ at z > 0.1 z-j.
For the lowest 10% of the CBL, Panofsky et al. (1977)  suggest
aw = (2.2u*3 + 2.3 H*z)1/3.  Empirical fits to field observations for the
 W
upper CBL aw have been suggested by Irwin (1979) and Hicks (1981).
                                     31

-------
     One aspect of vertical  velocities in the CBL that  has  considerable
consequences for diffusion is that  they are strongly non-Gaussian  in
distribution, at least through the  bulk of the mid-CBL.   Lamb  (1981)  has
shown this from his numerical modeling, and field measurements also  show
highly skewed w distributions.  For z/z-j  = 1/4 and 1/2,  the mode of  Lamb's
distributions is at w « -0.5w*.  This agrees well  with  the  observed  descent
rate of average plume centerlines from elevated sources, which will  be
discussed later in this report.  Lamb's interpretation  of this, which has
wide support in field observations, is that downdrafts  occupy  a larger
percent of the horizontal  area than do updrafts (thermals); the updraft/
downdraft area ratio is about 0.4/0.6.  Updrafts tend to have  about  1.5
times as large absolute velocity, so the condition w" =  0 is satisfied. The
significant skewness in w1 distributions makes conventional Gaussian  plume
modeling unrealistic for CBLs.

     Hanna (1981) has given relationships for Lagrangian time  scales  in the
CBL; basically they are ZT/W* times functions of z/z-j.   Because updraft and
downdraft velocities average about  0.6w* and 0.4w*, respectively,  the time
it takes for a passive substance to traverse z-j once up and once down is
approximately 4z-j/w*, which is about how long it takes  concentrations to
become almost uniformly mixed.  For turbulence measurements from fixed
sensors  (Eulerian measurements), the peak energy content is observed  at a
frequency = 1.3z-j/IT for the horizontal components and approaches 1.5 Zi/U
for w1 in the upper part of the CBL (Kaimal et al., 1976).   This is  in
rough agreement with Deardorff and  Willis1 (1976b) laboratory  observations
that typical downdraft "cell" sizes were » 1.5 z-j.  One implication  of

                                     3?

-------
these observations is that for good ensemble averaging of any quantity in
the CBL, including concentration, one should average over a period many
times 1.5 z-j/IT, to average out the effects of individual  eddies; 1.5 z-j/IT
typically ranges from 1 to 15 min.

     The height T.\ begins at about the height of the nocturnal  boundary
layer at sunrise and increases rapidly through the morning once heat flux,
or H*, becomes significantly positive (this may be delayed until most dewfall
or surface wetness from precipitation has been evaporated).  As a first
approximation, the growth of z-j can be predicted by using an early morning
temperature sounding and the time-integrated surface heat flux  and by assuming
that all the heat flux is accounted for in warming of the air beneath z-j  from
its initial temperature to a uniform potential temperature, 9s(z-j)
(subscript "s" refers to the sunrise potential temperature profile).  Such
a model  is called an "encroachment" model.  In differential form, it can  be
written  simply as
          Zi(d9s/az)i dzi = (vTHdt,                               (2.4)
where w'T1 is proportional  to the sensible heat flux at the surface and
(59s/az)j is the morning profile value at z = Zj .   According to Tennekes
(1984), the encroachment method handles about 80%  of the Zi prediction
problem; adding an additional 20% heat flux as a crude approximation of the
negative heat flux at z\ adds another 10% accuracy, and further refinements
in the prediction method "tend to get lost in the  unavoidable inaccuracies
of most experiments."
                                     33

-------
     Cumulus clouds often develop just  above the mean  mixing  depth,  especially
in the afternoon on humid days.   These  are caused by the  transport by  thermals
of humid air from near the surface to its  lifting condensation  level  (LCL).
Vertical motion in CBLs is nearly adiabatic, so  air cools nearly  1°C for
every 100 m of rise.  At first,  only the most vigorous thermals containing
the largest absolute humidities  reach the  LCL to form  small scattered
cumuli.  As z-j grows, more and more thermal s reach the LCL and  the percentage
of sky cover by cumuli increases; this  phenomenum has  been discussed and
modeled by Wilde et al., 1985.  One consequence  of this mechanism in very
humid weather is that, at some point during the  day, nearly the whole  sky
becomes covered with cumulus clouds, which limits the  solar radiation
reaching the ground and, in turn, limits further growth of z-j.

     Depending on conditions above z-j,  "cloud venting" (Ching et  al.,  1983)
may occur, causing transport of  air and pollutants out of the mixed  layer
into the overlying stable air.  This venting occurs when  "forced" clouds
(Stull, 1985) are carried by their vertical  momentum past z-j  to a level of
free convection (LFC), at which  point the  latent heat  released  by condensation
of water vapor is sufficient to  cause the  clouds to develop positive buoyancy
relative to the surrounding dry  air.  (The existence of a LFC depends  on
the surface humidity and the mean temperature profile.)  When this happens,
the clouds become "active" (Stull, 1985) and develop vertically into cumu-
lonimlus clouds.  The rate at which CBL air can  be vented aloft by these
clouds has not yet been quantified, but Issac et al.  (1984) have  estimated
that up to 90% per hour of air below cumulus cloud bases  enters the  clouds
in some conditions; much of this air subsides and returns to  the  CBL,

                                     34

-------
but some significant chemical  changes can occur during the wet  phase.





     The CBL tends to "collapse" in the late afternoon as  its driving  force,



the positive surface heat flux, is removed.   Little is known about  how this



occurs.  (This phenomenon was  discussed in the subsection  on "The  Diurnal



Cycle and Transitional  Periods.")





STABLE BOUNDARY LAYERS





     Research on SBLs has lagged behind the  advances  made  for CBLs,  but



much progress has been made in the last seven  years.   Part of the  reason



for delayed progress was the difficulty of developing instrumentation  that



responded adequately to the very low turbulence levels that occur  in SBLs.



In addition, the structure of  SBLs is inherently more variable  than  that of



CBLs and cannot be so easily generalized; it is not known  whether the



relationships that have been developed on the  basis of "ideal"  terrain



experiments will  transfer to more  typical  sites.   Terrain  slope is critical



because even extremely slight  slopes, on the order of 0.1°, lead to  downslope



gravitational forces that are  as important as  the  large scale pressure



gradient in driving the flow (Caughey et al.,  1979).   The  angle between



downslope direction and pressure gradient force also  can have a critical



effect.  Besides these effects, the temperature and stability structure of



SBLs can be considerably affected  by radiational cooling,  and the wind



velocity profile evolves slowly through the  night, rarely  achieving  a



steady state.  These effects cause the Ri  profile  to  evolve also, so that



"turbulent bursts" may occur.   Short of that,  large amplitude wave motion



may develop at the top of a CBL, causing its height to fluctuate (Lu et



al., 1983).



                                     35

-------
     There has been much improvement in  knowledge of  SBL turbulence
structure for flat sites in undisturbed  conditions (no  discernable wave
motion).  High quality experiments  in Minnesota  (Caughey et  al.,  1979)
and in the Netherlands at Cabauw (Driedonks,  van  Dop, and  Kohsiek, 1978)  have
made this possible.  Nieuwstadt (1984a,b)  has shown that many  relationships
that are valid for the surface layer, the  lowest  10%  or so of  the SBL, are
appxoximately valid for the whole SBL if jocal  scaling  is  used.   Instead
of scaling with u*, the surface value of -c1/2,  we scale with the  value
of tl/2 at the height in question.   Likewise, we  use  the local value of
heat flux and a local value of the  Obukhov length, A  =  -t  3/2[k(g/T)wTTr].
The reason that local scaling works is probably because eddy sizes are
limited in stable conditions to a small  fraction  of the total  SBL thickness
(this is not true for CBLs or NBLs).  Thus, each  layer  responds only to the
immediately adjacent layers and not to surface fluxes.  One  of the useful
results of Nieuwstadt's data analysis is that crw  = 1.4T*'2 throughout
the CBL.  For the eddy viscosity of momentum, Km, he  found that
Km = t-^AfjU/A) for the whole layer, in  contrast to the  expression
usually assumed for just the surface layer, Km =  u*Lfg(z/L)  (Km is defined
as T * a~u/az and fj and f2 are functions).

     Nieuwstadt (1984a) also developed a simple prediction for the profiles  of
the local fluxes by using the equations  for 39/at, all/at,  and  aV/at
in horizontally homogeneous boundary layers.   To  do this,  h  assumed
(1) stationarity, with time derivatives  of a0/az, u,  and v set equal to zero,
(2) a constant flux Richardson number, defined by Rf  =  (g/T)w'T'/(i:  aii/az)
(vector values, rather than scalar  values, are used for T  and  aTT/az), and

                                     36

-------
(3) h = 3l/4(kRifU*L/f)*/ ^.  Assumption 1 is not really valid for the mean
quantities like Hand 9, but it seems to give good predictions for turbulence
quantities (i.e., turbulence may adjust quickly to changes in gradients of
mean quantities).  The value of Rf, like Ri, increases from zero at z = 0
to a nearly constant value « 0.2 through most of the SBL.   Assumption 3
agrees with earlier arguments for the functional form of h, to be discussed
later; however, a simple solution for t is obtained only with this particular
choice of constants, for which h = 0.35 (u^/f)1/^.  The solution obtained
is T/u*2 = (1 - z/h)3/2.  This was shown to be in good agreement with the
Minnesota and Cabauw observations.  The result for heat flux,
w'T'/fw'T1)0 = (1 - z/h), follows more readily from just the stationarity
assumption a(99/9z)/3t = 0.  The Cabauw observations fit this well
enough, but it is seen in Nieuwstadt's (1984a) Figure 10 that the Minnesota
observation would better fit a higher power in (1 - z/h)n, with n = 2
instead of 1 (this difference in sites could be due to differences in
radiation flux divergence).  Some consequences of Nieuwstadt's forms for \
and trr are A/L » (1 - z/h)5/4 and aw/u* = 1.4(1 - z/h)3/4.

     Lateral variances for Cabauw were not reported, but it might be noted
that the au/u* values for Minnesota reported by Caughey et al., (1979)
give a fairly good fit to 2.3(1 - z/h)3/4.  Hanna (1981) suggested
a linear decrease with height in aw,  au, and av as empirical  fits to
the Minnesota data, with, for instance, crv/u* = 1.3(1 - z/h).  However,
expressions like these with au and av of the same order as aw can only
represent SBLs in their most ideal state, with only small-scale, nearly
isotropic turbulence and no large horizontal  eddies or meanders.  Thus,

                                     37

-------
they should be interpreted as minimum values;  on some nights,  at  some  sites
 0 and 39/3z > 0.01°C/m  at z < h-f; however, in the
upper part of a SBL and above, ae/az is strongly influenced by radiational
cooling, which depends on the water vapor and  C02 profiles  among  other things.
Also, turbulent exchange tends to reduce  30/dz,  not  increase it,  so
39/az greater than some threshold value makes  no sense as an indicator of
the top of the turbulent boundary layer,  h.  Other doubtful measures of  h  are
the height of significant cooling, h9, and the lowest height of  a maximum
                                     38

-------
in the wind speed profile, hu (it is true that a "jet", a wind speed maximum,
often develops above the SBL, in response to inertia!  forces free of ffictional
drag).  However, there is one mean profile quantity that is a good indicator
of the height of turbulent mixing, and that is the Richardson number, Ri
(Eq. 2.2), or the Bulk Richardson number, Rb = (g/9)(9 - 90)z/Tf2, where 9
and "u are measured at the height z (Rb does not require the precision measure-
ments that Ri does, so it is often the more practical  choice).  By using  data
from 25 nights of the Wangara meteorological  experiment in Australia, Mahrt   -
et al. (1982) showed that the height at which Ri = 0.5 correlated highly
with u* and, by implication, with L, as u* and L were  also highly correlated
(Mahrt and Heald, 1979).  In contrast, the correlations between  hu or hg
and u* were about half as large, about 0.42, and the correlation between  h-j
and u* was even less.  They noted that Ri was normally small  and relatively
constant with height in a layer near the ground, but that this layer
was capped by a sharp increase in Ri.   In the same journal  issue, another
paper using the Wangara nocturnal data appeared by Wetzel  (1982).  He
showed that the height at which Rb = 0.33 correlated well  with u* and L,
with poorer in correlations for hg, hu, and h-j in place of Rb.  He also
found another profile height index that correlated with u* and L almost as
well as the Rb index, namely the top of the linear part of the 9 profile
(these profiles were unusually regular, however, which may be more charac-
teristic of ideal, flat sites like Wangara).  Arya (1981)  showed that u*
and L correlated fairly well (r = 0.63) with h estimated from acoustic
sounder returns at Cabauw, whereas the correlation of  this h  with he was
only 0.28 and with hu was not even statistically significant.
                                     39

-------
     Ideally, we would like to  be able to  estimate  h  on  the  basis of surface
parameters, so that something simpler than a  well  instrumented  tall
meteorological tower would  do the job.  Zilitinkevich (1972)  developed  a
simple SBL model with

                    h * b(u*L/f)l/2                                   (2.5)

This equation has received  much testing in recent years.  While this h
correlates significantly with measured h,  it  is  far short  of perfect.   For
acoustic sounder measurements at Cabauw, Arya (1981)  found a correlation
of r = 0.67, while Nieuwstadt (1984a), who excluded weakly stable
cases, found r = n.77.  Nieuwstadt and Tennekes  (1981)  found better
correlation with a prognostic model;  h never  reached  its equilibrium
value because the relaxation time was too  long (this  time  depended on the
surface cooling rate, a90/at, and became ~ 10 hr late at night).  However,
according to Nieuwstadt (1984a), part of the  success  of  this model was  due to
the fact that the initial  h was specified; because  time  scales  are long,
predicted and observed h only slowly  diverge.  Because there is no ready
way to specify the initial  h, he concludes that  diagnostic models like  Eq.
2.5 are more practical for  now.  However,  Eq. 25 is a steady-state solution,
and the SBL is so slow to adjust to a balance of forces  that a  steady state
may not often be achieved.   A second-order closure  numerical  model developed
by Rrost and Wyngaard (1978) gave predictions that  fit Eq. 2.5  well, but
the steady state, with b =  0.4, was not approached  until t = 5  hr after the
onset of surface cooling.  At t = 2 hr, b  = 0.7  was the  best fit.  The
measured values of h from the Minnesota experiment  runs, which  were centered
                                     40

-------
at about 1 hr after transition to negative H*,  fit Eq.  2.5 with  b  =  0.7  to
0.9 when h was defined as the height at which heat flux extrapolated to
only 5% of its surface value (Caughey et al., 1979).

     Garratt (1982) compared Eq. 2.5 with h data from four different sites,
Koorin and Wangara in Australia (both using the Ri method), Cabauw (acoustic
sounder), and Minnesota (heat flux profile).   By using  a different criterion
for h than did Caughey et al., (1979), he calculated  a  best fit  of b = 0.37
for Minnesota; for Cabauw and Wangara, he calculated  b  = 0.42, and 0.39,
respectively, which supports the generally accepted value of b = 0.4.  However,
for the only low-latitude site, Koorin, at 16°S, the  best fit was  b  = 0.13.
This is three times smaller than the other sites; the f-1/2 term in  Eq.  2.5,
even if incorrect, does not account for this  difference (remember  that f
contains the sine of latitude).  Garratt accounted for  this difference by
running Brost and Wyngaard's model  for each site while  including the signi-
ficant terrain slopes at Koorin (0.11°) and Minnesota (0.08°) and  the
angular difference between the fall  line and  mean surface wind direction;
the results of these model  runs gave very good  agreement with the  best-fit
values of b observed at each site.   This result should  make us cautious
about the general use of a surface parameter  approach like Eq. 2.5 until it
is validated for sites with more typical  slopes; certainly, these  often
considerably exceed 0.1°.

     There are diagnostic alternatives to Eq. 2.5. For sites where  there
are significant slope flows, gravitational  acceleration may be much  stronger
than Coriolis acceleration.  (This  only requires that g(A0/9)p »  uf,
                                     41

-------
where A9 is the potential  temperature difference  between the drainage  flow
and "free" air and p is the terrain  slope,  in  radians.   If A9 =  5°C, 9 =  300°C,
u = 3 m/s, and f-1 = 104 s, p  »  0.1° implies  that  gravitational  force
dominates.)  In such cases, u*/f  should  not be a  significant length parameter.
Instead, it may be found that  h « L, as  Kantha and  Long  (1980) found to be
true in mechanically-stirred stratified  fluid  in  a  laboratory tank.  For  the
simple two-dimensional  slope flow model  that Briggs (1981) fit to Ellison
and Turner's (1959) laboratory slope experiments, h « x, the distance  down
the slope, and the mean flow speed U « (H*x)l/3>  this gives h «  iP/H*  (the
dependency of h and U on p cancel out in this  expression).  If U = u*,
then h « L for slope flow; we might  expect  something like surface-layer
similarity profile assumptions to apply  to  a slope  flow, in which case
constant h/L would imply a nearly constant  U/u* (influenced slightly by a
small increase in In(h/z0) down the  slope), so that there is an  internally
consistent argument for h <* L.  Idealized as this model  may be,  it is
adequate warning that L may be highly variable across rolling or hilly
terrain at night.

     In sorting through these alternatives, one should be aware  of strong
correlations between u*, L, and other SBL parameters; these correlations
also provide possibilities for useful simplifications.   Van Ulden and  Holtslag
(1983) showed that the surface-layer temperature  scale,  T* = w'T'/u*,  is
nearly constant over a wide range of u*  at  Cabauw (see Fig. 2b), becoming
much smaller only at very small u*,  < 0.1 m/s; this constant T*  does depend
somewhat on cloudiness.  Some consequences  of constant T* are H* « u*,
L « u*2, and, if Eq. 2.5 holds, h «  u*3/2.   Venkatram  (1980) showed that

                                     42

-------
L = 1100 u*2 (mks units) gives a good fit to field  experiment  data  at  three
relatively smooth and flat sites; furthermore,  h  =  2400 u*3/2  (mks  units)
gives a good fit to the Minnesota measurements  of h, which ranged from 30  m
to 400 m.  Nieuwstadt (1984a)  carried this simplification one  step  further
to show that h = 28 IP/2 (mks  units), where IT is  measured at  10m, worked
about as well as Eq. 2.5 for Cabauw observations.  However, the constant  in
this relationship probably depends on site parameters like z0; furthermore,
in drainage flows, the relationship may be more like h « Tj"3, as already
discussed.  Briggs (1982a) suggested an easily  measured replacement for L,
Ln = u"3/Rn*, where Rn* is defined like H* but with  net long-wave radiation at
the surface replacing heat flux.  In doing this,  he showed that
H*/Rn* = 0.5(1 + S/L)'1 for the Project Prairie Grass measurements  (Barad,
1958), where S = 15 m, a site-specific length constant.  This  implies  two
regimes.  One is a near-neutral  regime with large L and H* « Rn*, limited
only by the radiation balance; this gives L * u*3 and Eq. 2.5  reduces  to
h « u*2 or h « L2/3.  For a very stable regime, with L « S, H* « L «  u*3/2 and
Eq. 2.5 reduces to h <* u*5/4 « 1.5/6.  jn between  these asymptotes,  there  is a
broad range that approximates  the constant T* solutions, as suggested  by
Venkatram (1980).  Note that this intermediate  regime gives h  <* L3/4,  and
for all cases we have discussed, including slope  flows, a strong correspond-
ence between h and L is predicted; namely, h «  |_n with n ranging from  2/3
(flat, near neutral) to 1 (slope flow).  Therefore, h = function(L) might
be a versatile approximation,  even with L replaced  by an easily measured
quantity like Ln = "u3/Rn*.
                                     43

-------
     Besides slope effects,  in  SBLs  we  must  be  aware  of  radiative cooling
effects on boundary layer structure.  Little work  has been done on this
but Garratt and Brost (1981)  made a  theoretical  investigation by using the
same second-order closure model  that  so successfully  modeled the effects of
slight terrain slopes (Brost  and Wyngaard,  1978; Garratt, 1982).  To do this,
they had to assume some typical  profiles for water vapor, Cf)2, and temperature
in the troposphere plus a value for  the surface emissivity (percent of black
body radiation emitted).  Six different cases were run,  including two for which
radiation flux divergence (radiative cooling) was  left out of the calculation.
The results showed that radiative cooling has large effects on the profiles
of heat flux (w'T1) below O.lh  and on 9 and  Ri  profiles  near and above
z = h.  Radiative cooling also  has significant  effects on the profiles of "u
and 9 in terms of surface-layer similarity  theory  largely because of the
development of an above-surface minimum in  w'T1; this affects H* and L,
which are defined at the surface. Without  radiative  cooling, crw above
z = h does not extinguish nearly as  rapidly after  transition to night.
On the other hand, the inclusion of  radiative cooling in the model has
relatively little effect on TT,  T, 99/dz, and Ri  below z  = h, or on
the evolution of u*, Rh at z  =  h, or b  = h/(u*L/f)1/2 (from Eq. 2.5).  At
z = h, Rb increased from approximately  0.1  at t =  1 hr to 0.25 at t = 2.5
hr regardless of initial conditions  or  inclusion of radiation; b increased
slowly from about 0.32 at t  = 1 hr to 0.4 at t  = 12 hr,  except that it was
of the order of 10% larger in the first few hours  when radiative cooling
was omitted.
                                     44

-------
     At present, there are no models that could  possibly predict  occurrences
of large amplitude waves or large time-scale horizontal  eddies  (meanders)
in the SBL.  It is understood that the possibility  of large gravity  waves,
which may "break" into turbulence or initiate a  turbulent burst,  exists
when Ri is near a critical value (~l/4)  through  a substantial  layer.
However, it is difficult to predict the evolution of Ri  through a night,  as
it depends on far upwind initial  conditions  and  the structure  of  the humidity
profile that develops as the CBL shuts down  and  surface  evaporation  and
plant transpiration continue around the time of  transition.  DeBaas  and
Driedonks (1985) made a case study of a  wave occurrence  measured  at  the
Cabauw meteorological tower; Ri  = 0.2 to 0.3 was measured for  certain wind
directions on the tower at z = 30 to 160 m.   Large  waves were  present, as
could be seen in T versus time traces from the instruments at  z = 120 and
160 m; also, aw more than doubled between 80 and 160 m,  a reverse of its
usual behavior.  The authors developed a linearized wave model which
successfully predicted the frequency range of the waves  that developed,
especially as seen in the turbulence spectra for w1.  However, the wave
magnitude could not be predicted  with this model.   Another episode of large
wave motions near the top of a SBL was described in detail  by  Lu  et  al.
(1983); this episode was apparently caused by the passage of a shallow front.
     Large time-period wind direction shifts and meanders occur in SBLs;
these can cause 1-hr average lateral  diffusion to be many times as large  as
would be predicted using boundary-layer  turbulence  models, which  describe
only three-dimensional  eddies are restricted in  size by  the ground and by
stable stratification.  Large two-dimensional  horizontal  eddies,  with no  such

                                     45

-------
constraints on size, often exist  and  tend  to  be site  specific.   Hanna
(1983) has summarized empirical  information from a  number  of  sites.  At
every site, the magnitude of the  lateral wind direction variance, which
directly relates to initial  lateral dispersion, tends to increase approxi-
mately as u~l in stable conditions.   However, av tends to  be  nearly constant
(within a wide range) with "u at  all sites.  Its average value ranges from
0.3 m/s at Porton (England)  to 1  m/s  at  a  complex terrain  site  in California
Hanna also presents some wind velocity observations for a  site  on the  Snake
River Plain in Idaho, a broad plain bounded by  mountains.   For  this site
av = 0.5 m/s for most 1-hr average values; occasionally values  are much
larger, but they are rarely  less  than 1/2  of  this.   Superimposed on these
fluctuations were wind direction  shifts  with  time periods  ranging from
about 1 to 4 h.  Such large  time-scale oscillations could  be  due to seiche
effects peculiar to that kind of  topography,  with cold air trapped in  a
large basin.
                                     46

-------
                                SECTION 3

                          DIFFUSION EXPERIMENTS
     How far have they done, for what sources,  under what  conditions?

     Good experimental information is essential  for building  confidence
in the abilities of modeling techniques to work under all  proposed
conditions of application.  Some partial  confidence develops  from experience
with model applications, provided that gross  model  failures are  rare,
non-critical, or simply not publicized.  However,  each user is concerned
with only a limited range of source and meteorological  parameters.   Since
there is no collective archive of experiences with model failures and
successes, information of this nature available to a model user  tends to
be anecdotal  and patchy.  It sheds little light on the causes of model
failures and gives little basis for confidence  in  the model for  application
to radically different source types or meteorological  situations.

     It is far preferable to test and compare model  performances against
data from carefully executed diffusion experiments.   Such  experiments
should include trustworthy measurements of the  diffusion pattern of  a
conservative tracer (one that does not lose or  gain  mass flux with distance)
and every source and meteorological  variable  that  might relate to variations
in the diffusion pattern (variables such  as wind direction, stability,
source buoyancy, etc.)  If these measurements are  complete, then the data
set is "universal" in that it provides a  standard  of comparison  for  any
model, past,  present, or future.  Unfortunately, such  complete experiments
are rarely accomplished.  Also, there is  a seemingly limitless number of

                                    47

-------
combinations of source,  terrain  and  surface configurations, effluent
characteristics, and meteorological  conditions  for which dependable
models are desired.  Each field  or laboratory experiment covers only a
very limited range of these combinations.  Usually, only one source type
is utilized, e.g., it may be a passive source of conservative tracer, a
buoyant source with no stack, building,  or terrain downwash conditions
complicating the flow, or a source of  uniform glass beads to study deposi-
tion with known source characteristics.   There  are always constraints on
the range of meteorological variables, due to the location, the limited
goals and resources of the experimenters, or factors  like  the undesir-
ability of allowing the equipment to get wet.   Furthermore, the diffusion
measurements themselves have many limits.  The  sensitivity and number of
sensors used puts limits on the  range  and density of  measurement locations,
and the great majority of field  measurements have been confined to near-
surface concentration values due to  the expense and logistical difficulties
of measurement in the vertical dimension.

     For some of the same reasons that diffusion experiments are limited,
e.g., manageability, this paper  reviews only the state of the science
for continuous, passive, point  sources of (presumedly) conservative
tracers with no complex terrain  or  large surface inhomogeneities, such as
land/water boundaries, within the  range of diffusion  of concern.  These
limitations nevertheless permit  inclusion of most earlier diffusion
experiments and quite a few recent  ones.
                                    48

-------
FIELD EXPERIMENTS

     There have been a few dozen field experiments  on ABL diffusion  that
can be considered of research grade.   While this  seems a substantial  number,
they all have many limitations, and many "data gaps"  continue to  exist.

     A few of the earlier experiments, such as Hbgstrbm's (1964), used
plume photography to determine plume  geometric variables, such as the
vertical standard deviation, cjz, and  the lateral  standard deviation,  cry.
This is a valid and relatively inexpensive method for characterizing
observed plume geometry, for both "instantaneous" or  "relative" plume
diffusion (i.e., diffusion relative to the instantaneous plume centerline)
and for "total" or "time-averaged"  diffusion (Nappo,  1984).   This method
is limited to the range that the plume can be seen  against the background;
depending on the plume opacity and  the background conditions,  this may be
tens of kilometers in very stable conditions or only  a few tenths of  a
kilometer in convective (unstable)  conditions.

     The vast majority of the field experiments use a metered  release of a
specific gas or some type of small  particles as a "tracer".   This is
sampled using aspirated boxes known as "samplers" or, in the  case of  some
fluorescent particles, sticky surfaces which catch  the particles  by
impaction.  A considerable number of  these receptors  are set  out  in  an
array intended to define the plume.  Usually,  they  are set out in near-
surface "arcs" of constant or approximately constant  distance  from the
source with only a few degrees of azimuth  between them.   In a  few of  these
experiments, attempts were made to  define  the vertical  concentration

                                   49

-------
profile of the plume by setting up a vertical  array of samplers.   This  is
always a more difficult and expensive feat,  and  is  often  frustrating.
Because the towers or tethered  balloon cables  supporting  the  vertical
array of samplers are very sparse in the horizontal  plane,  they often
fail to define the plume's vertical  dimension  adequately, or  they  even
miss the plume altogether.  Rarely has such  an array extended through the
whole mixing layer, so its usefulness is confined to near the source
where vertical diffusion is still  within its bounds.  Very  recently,
diffusion experiments have been made that use  remote sensors, especially
lidar, that effectively solve the problem of measuring concentrations in
the vertical dimension; two of these experiments will  be  described at the
end of this section.

     There have been several  good surveys of diffusion field  experiments
up to about 1978, which this discussion will exploit.   Horst  et al.
(1979) selected ten field experiments that they considered  adequate
for testing predictions of 0y and 
-------
somewhat wider than that of the present report,  and  it  included  diffusion
experiments in complex terrain, near shorelines, and within  forests.
Draxler's survey includes 20 "classical" diffusion experiments,  made  over
homogeneous terrain, and 32 experiments in other categories.   Another
survey of some relevance was made by Sklarew and Joncich  (1979);  it
covered field programs plus monitoring networks  at power  plants.   Although
these sources were all buoyant, buoyancy effects dominate lateral  diffusion
only at small to moderate distances.  At tens of kilometers,  plumes from
buoyant sources diffuse much like passive plumes released from the same
effective source height or from any height within a  well-developed convec-
tive mixing layer, which tends to mix all  material  nearly uniformly
throughout its depth.

Classic Tracer Experiments

     It is not possible in a few pages to discuss individually each field
experiment falling within the scope of this paper.   Instead,  information
from Horst, et al . (1979), Draxler (1984), and also  from  tables  published
in Draxler (1976)  and Irwin (1983)  have been used to construct the summary
of experiments shown here in Table 1.  Included  are  four  experiments
that Draxler (1984) considered non-classical  because of rough (but not
"complex") terrain, urban siting, or low wind speed  conditions;  these are
the Julich and Karlshruhe experiments, the St. Louis experiment,  and  the
INEL, Idaho experiment, respectively.  For these experiments, the sites
seem reasonably homogeneous, at least among typical  "real  world"  sites,
as opposed to ideally flat and uniform sites like that  of the Prairie
                                    51

-------
                                                   Table  1.   Summary of Tracer
Site or
Program
Surface Source
Round H111
Prairie Grass
NRTS, Idaho
Ocean Breeze
Dry Gulch
Green Glow
Hanford-30
St. Louis
INEL, Idaho
Hanford-67
Elevated Source
Rrookhaven
Porton, ENR
Han ford
NRTS, Idaho
Agesta, SUED
Suf field, CAN
Julich, GER
Karlsruhe, GER
Hanford-67
Cabauw, NETH
Copenhagen, DEN
Terrain/Roughness
Length (cm)
Low hills, <30m/>10
Flat prairie/0.6
Flat, sagebrush/1.5
Low dunes, <6m
Sloping mesa, ravines
Flat, sagebrush/3
Flat, sagebrush/3
Urban, <30m relief
Flat, sagebrush/1.5
Flat, sagebrush/3
Fields, woods, hills/100
Rolling downs/5
Flat, sagebrush/3
Flat, sagebrush/1.5
Wooded hills <50m/60
Rolling prairie/2
Forest, farm/40 to 80
Forest, farm/110
Flat, sagebrush, 3
Flat meadows, trees/10-2f)
Urban resident ial/60
Stab. a
U-S
U-S
U-S
U-S
U-S
U-S
U-S
U-S
vS
N-S
U-S
sU-sS
s
u
N-vS
U-S
U-S
N
N-S
SU-N
II-N
Release/
Receptor
Heights (m)
0.3/2
0.5/1.5*
-1/-1.5*
2-3/1.5-4.5
2-3/1.5
2.5/1.5*
2.5/1.5*
sfc, bldg"Vsfc
1.5/0.8*
2/1.5*
108/sfc
140-2551/cable
56/1.5*
46/1*
50/photos
7 & 15/0
50 S 100/sfc
100/sfc 20-30
26 S, 56/1.5*
80 4 200/sfc
115/sfc
nur.b
(min)
10
10
60
30
30
30
60
60
60
10-30
60
30
15-60
30
60
30-60
60

10-30
fiO
60
Footnotes:   * Plus vertical profiles,  a Stability range, U = unstable, N - neutral,
S » stable, s ' slightly,  v * very.  b Duration of release or of sampling.  c All
include u" or u profile.  d Deposition may significantly reduce concentrations at
the larger arc distances (Gryning et al. 1983).  e Six 17.5-m towers at x = 100 m.
f Wind and temperature, surface to specified height.  9 Profiles of u and T to 16  m,
aircraft soundings to 3 km, solar radiation, soil temperature profiles and more.
n Significant deposition noted.  1 Five 30-m towers at x = 400 m.  J Significant
deposition loss (Simpson,  1961).  k Reliability questioned by Horst et al. (1979).
References for Table 1:  Round Hill,  Cramer,  Record and Vaughan (1958).  Prairie
Grass, Barad (1958) and Haugen (1959).   NRTS, Islitzer and Dumbauld (1963).  Ocean
Breeze and Dry Gulch, Haugen and Fuquay (1963).   Green Glow and Hanford-30, Fuquay,
Simpson, and Hinds (1964).  St. Louis,  Pooler (1966) and McElroy (1969).  INEL,
Sagendorf and Dickson (1974).  Hanford-67,  Nickola (1977).  Brookhaven, Smith (195fi),
Porton, Hay and Pasquill  (1957).  Hanford,  Hilst  and Simpson  (1958) and Elderkin

                                   52

-------
Experiments  in Homogeneous  Terrain
Tracer
so2d
S02d
Uranine dye"
Diffusion Data
3 arcs
5 arcs, towers6
6 arcs, towers1
Maximum
Range
(km)
0.2
0.8
3.2
Meteorological
Datac
oa, ae, 2m; AT, 12-1. 5m
aa, ae, 2m; profiles1' 16m,
aa, 
-------
Grass experiment (Barad,  1958).   A scan  of  column 2 of  Table 1 shows experi-
ments made in a wide variety of  site  types.   For completeness, the more
recent experiments made in  Cabauw (Nieuwstadt and Van Duuren, 1979) and
in Copenhagen (Gryning and  Lyck,  1984) are  included; both experiments
used elevated releases of S?Q, which  has been favored in recent years as a
very conservative tracer  that can be  accurately measured.

     Table 1 cannot be considered an  adequate, or even  absolutely accurate,
description of each field experiment; it is  very compact and omits many
details.  Yet it does serve to give an overall view of  what has and has
not been studied in the field in  scientific  diffusion experiments.  Within
the surface source and elevated  source categories, the  listings are
approximately chronological. This ordering  shows a trend toward more
mass-conservative tracers such as SF^ in more recent experiments.

     There is evidence of deposition  losses  for almost  all the tracers
used in experiments before the early  1970s.   Certainly, this can be
suspected for any of the  particle or  dye tracers, for surface releases and
for elevated releases after substantial  surface impact  occurs.  Deposition
may not affect lateral diffusion, or  ay, very much, but it can greatly
distort vertical diffusion  at distances  of  about 1 km or more.  For two
of the surface source experiments, it was possible to estimate the decrease
in mass flux of tracer using an  arc of samplers on towers.   In the Green
Glow experiment, which used zinc  sulfide particles, the loss of source
material at x = 3.2 km was estimated  by  Simpson (1961)  to range from 78
to 97%, with 18 to 45% loss even  at x =  0.2 km.  The greatest loss was
for the most stable run.   Similarly,  10  to  54% losses of uranine dye at
                                    54

-------
x = 0.4 km were estimated for the NRTS experiments  (Islitzer  and  Dumbauld,
1963).  Significant deposition was also suspected  for the multiple-tracer
releases of the Hanford-67 series beyond x =  0.2 km (Horst et al .  1979).
Even S02 gas is known to deposit on the ground;  Gryning  et al . (1983)
estimated a 20 to 25% reduction in surface concentration at x = 0.2 km
for the Prairie Grass experiment.  At the farthest  arc,  x = 0.8  km, they
estimated a 25 to 45% reduction, the greatest reduction  being in  stable
conditions.  The significance of these deposition  losses is that  we must
regard all estimated vertical  dispersion results from the surface source
experiments with suspicion or at least as in  need  of careful  interpretation.
The reinterpretation of the Prairie Grass experiment by  Gryning et  al.,
if correct, implies that all  previous determinations of  a2 from this
experiment at x = 0.8 km in stable conditions may  be about a  factor of
1.8 too large, since they were based only on  surface samples  and  the
assumption of a Gaussian, conservative plume.   Hence, for surface  releases,
only the INEL experiment using SFs as a tracer should be free of  this
interpretive problem (this was a rather specialized experiment, run under
extremely stable conditions,  with sampling to only  0.4 km).

     Another tracer that should not be considered  conservative is oil
fog, which was used in several  of the elevated source experiments.   Oil
fog consists of micron-sized  droplets formed  by  condensation  of oil  heated
to a vaporous state and mixed  with air; it is produced much like  steam.
It contains some volatile substances, and these  slowly evaporate  after
the tiny droplets have formed,  as they have a very  large surface  area-to-
volume ratio.  Lidar scans of oil  fog plumes  have  indicated a loss  of

                                    55

-------
total  reflectivity with distance.   Scans from the 1983 CONDORS experiment,
described at the end of this section, indicate a 50% loss  in the total
return signal  (after correction for attenuation and background haze)
from the plume in travel distances of 1.4 to 2.8 km for most cases
(Eberhard, 1985: personal  communication).  The rate of evaporation depends
en the oil type, and the oil chosen for this experiment was  relatively
low in volatile substances;  most oils would evaporate faster.  Presumedly,
this decrease of droplet size, mass flux, and reflectivity with distance
would not much affect plume width  determinations based on  a  certain
percentage of the center!ine concentration, as was used at Brookhaven
(the data base for the Brookhaven  ay and az curves).  However, the
concentration-to-source strength ratio, X/0, of oil fog will  become
increasingly lower than that for a conservative tracer; when az is
inferred from surface measurements by assuming a Gaussian, conservative
plume, as was done at Brookhaven,  increasing overestimates of az with
distance would result.

     The Agesta experiments of Hb'gstrb'm (1964) also used oil  fog, but
plume parameters were determined by photographing puffs of oil fog from
near the releasing point;  each puff was photographed for about 15 min,
until  it became difficult to define, and then another puff was released.
The meandering component of diffusion was determined from  the vertical  and
lateral standard deviations of puff center!ines as the succession of puffs
passed designated travel distances (estimated from travel  time and wind
speed measurements); this determination should not be affected by droplet
evaporation.  However, the relative diffusion estimates, based on the

                                    56

-------
average puff size relative to the puff center,  could  become  increasingly
too small with distance, as evaporation could cause the apparent  edges  of
the oil fog puff to shrink somewhat.   At the furthest distances,  the
opposite effect could occur, as the more rapid  disappearance of the
apparent puff could lead one to believe that it became more  rapidly
diluted by a more rapid increase in ay and  az.

     With the possibility of "disappearing  tracers" in mind, consider the
limitations of the diffusion data in  experiments  listed in Table  1.
First of all, for about half of the experiments,  tracer measurements were
made only within a few meters of the  surface, by  using arcs  of samplers.
In such cases, only the lateral  dispersion  near the surface, such as ay,
and the crosswind-integrated concentration, /Xdy, are measured directly.
The /Xdy measurements are universally meaningful  only if little of the
tracer is lost at the distance of the arc,  i.e.,  if the integrated mass
flux overhead = Q, the rate of release.  If significant deposition has
occurred, this measurement is relevant only for material  with the same
deposition rate under the same conditions.  Very  often,  az  has been
inferred from /Xdy measured on surface arcs by  assuming (1)  a conservative
tracer and (2) a Gaussian form of vertical  concentration distribution
about the release height.  We have reason enough  to believe  that  assumption
1 is very risky after about 0.1 to 0.4 km of plume contact with the
surface for the majority of the tracers (some have yet to be proven or
disproven as approximately conservative tracers).  Assumption 2 has been
shown to be roughly correct in neutral  and  stable conditions, and in
                                    57

-------
unstable conditions for near-surface sources  at distances  less than
0.5 ziTT/w* (Pasquill ,  1974;  Willis  and  Deardorff,  1976a).   At larger
distances and for elevated  sources  in unstable conditions,  this assumption
may be grossly incorrect (Willis  and Deardorff, 1976a,  1978, and 1981).

     The az curves most used in current practice,  Pasquill's and
Brookhaven's (Section  4), were derived  using  surface  sampling data with
assumptions 1 and 2,  above;  therefore,  these  curves should  be considered
very doubtful for the  unstable cases, unless  they  are applied to the same
type (surface or elevated)  of source using  the same assumptions.   Pasquill's
non-neutral az curves  are based mostly  on the Prairie Grass experiment,
so they apply best to  near-surface  sources.  Brookhaven's  curves are
based mostly on their  108-m oil fog releases, so are  more  appropriate  for
elevated sources.  In  both  data sets, it seems likely that  the mass flux
of the tracers decreased significantly  at the larger  distances, violating
assumption 1.  This would have the  effect of  causing  overestimates of  
-------
this deposition would tend to deplete the lower  part  of the  plume more
than the upper part of the plume and thus distort  the profile  shape
compared to that of a conservative tracer plume.  However, surface deposi-
tion should not reduce the maximum height obtained by the tracer at  any
given distance; the settling velocities,  or fall  speeds, of  the particulate
tracers were negligibly small.  Therefore, the az  values computed from
tower profiles for the surface plumes should be  roughly similar to those
of truly conservative tracers, although they will  be  affected  somewhat by
the profile shape distortion.  To take an extreme  example, if  the Gaussian
curve centered at z = 0 is distorted by a factor <= z, the computed second
moment around z = 0 (i.e., az) is increased by a factor 21/2,  or by
41%.  The increase is probably considerably less for  actual  cases affected
by surface deposition, perhaps around 20% at most;  thus, tower-derived az
values may be acceptable in spite of surface deposition. The  Table  1
footnotes on tower distances then suggest that usable az information
exists for surface sources from x = 0.1 to 3.2 km.  However, in every
case the site is flat and smooth; z0 ranges from 0.6  to 3 cm.  There are
further restrictions regarding stability.  For very stable conditions, we
have only the INEL experiment with towers only at  x = 0.2 km (in the most
stable runs the plumes may have reached their maximum vertical growth,
with az only = 4 m, as close in as 0.1 km;  the surface arcs  indicated
no trend in /*dy from x = 0.1 to 0.4 km,  and SFg is considered a conser-
vative tracer).  Moderately stable and neutral cases  are well  covered,
but the towers are too short to encompass the rapid vertical  development
of plumes in unstable conditions.
                                    59

-------
     For elevated sources, surface deposition  should  have  negligible
effect when az < zs, the source height,  because surface  contact  is
minimal up to this point.  Deposition  should also  be  less  serious than  for
surface sources at all  distances,  since  concentrations near  the  ground
are nearly always lower for an elevated  release.   Towers were  used  in two
experiments, but at NRTS they were only  30 m high.  The  source was  at 46 m,
so only a small portion of the plume,  not enough to determine  az, could
be measured.  The towers for the Hanford-67 experiment  (Nickola, 1977)
ranged from 27 to 62 m in height,  so they could have  substantially  defined
the vertical profile only for the 26 m-high release at the farther  arcs
(with the taller towers) in the more stable cases.  Nonetheless, Horst  et
al.  (1979) were able to determine best-fit az values for  a  Gaussian
plume on the basis of the partial  profiles at  x =  0.2 km,  with good
agreement among results for the different tracers  used.  They  considered
this procedure inadequate at the larger  distances,  due to  profile distortion
by deposition.  Two of the elevated-source experiments used  impaction-type
collectors mounted on tethering cables with much  less restrictive heights
than tower networks, but there were no elevated,  lateral plume measurements
(it would be unwieldy to construct an  arc of tethered balloons,  and they
would tend to get tangled in turbulent winds).  The Porton experiments
(Hay and Pasquill, 1957) defined 
-------
     Crosswind-integrated surface concentration  is  directly  related
to az if the profile shape (e.g., Gaussian)  and  the overhead mass  flux
are known.  Most historic az estimates were  made by using  surface
sampling arc /Xdy measurements and assuming  a  Gaussian  plume shape and
a conservative tracer, but one or both of these  assumptions  are  question-
able in the majority of cases, as discussed  earlier.   However, /Xdy
is useful  in its own right, mainly for computing long-term averages  of
surface concentrations.  We can consider it  to be reliably determined
only for those experiments using a conservative  tracer, which would  only
be the SFs experiments in Table 1.  This includes only  the measurements
to x = 0.4 km for a surface release in very  stable  conditions at INEL and
two elevated-source experiments, both relatively recent and  both European.
(It is curious that, in Table 1, all  the surface-release experiments were
made in the U.S.A., while most elevated-source experiments were  made in
Europe.)  At Cabauw, there was only one "arc", really lines  of samplers
following roads.  In Copenhagen, three arcs  were established over  a
useful range of x, 2 to 6 km.  Neither experiment included stable  cases,
nor very convective cases, rare in these northern,  coastal regions.  It
is encouraging that both of these experiments  were  made in "real-world"
locations with surfaces typical of many industrial  and  power production
sites.
     Because ay is not so affected by tracer evaporation or  deposition,
Table 1 indicates that there is a wide variety of useful ay  information,
particularly as derived from sampling arcs.  Many of these determinations
extend to 4 to 6 km downwind, for both source  types.   For  surface  sources,

                                    61

-------
the Hanford experiments listed  in  Table 1  (including  project Green Glow)
extend cjy information to 13 or  26  km,  for  all  stabilities.  However,
keep in mind the fact that  in stable conditions,  ay is  very site dependent;
gravitational  flows of cool, denser air near the  surface develop even on
extremely slight slopes, and these collide and  interact with each other,
causing horizontal  meander. The  investigators  for the  INEL experiment,
which was performed during  very low wind speed  and very stable conditions,
concluded that the best "model" for horizontal  diffusion at short distances
was the frequency distribution  of  observed wind directions (Sagendorf and
Dickson, 1974).  One of the unique aspects of  this experiment was that
the arcs extended over a full 360°; during some of the  one-hour runs, the
plume meandered through the full  360°  of arcs.  Experts have been saying
for years that the only reliable  indicator of  ay  averages of more than
a few minutes duration in stable conditions is  wind variance measurements
at the site at the height of release (Hanna et  a!., 1977); no amount of
experimental information on 
-------
Copenhagen included z-j estimates from soundings.  Of the historic experi-
ments, only the Prairie Grass data tabulation of 1956 included temperature
and humidity soundings to such a depth that Zi could be determined at a
later date, as was done by Nieuwstadt (1980) and Briggs and McDonald
(1978).  Both these references showed that the Prairie Grass unstable
data compared well with predictions that use convective scaling, which
requires z-j, hut scatter was somewhat large owing to the rather short
release time of 10 min; the bulk of the experiments had 30- to 60-min
releases or sampling durations.

     From the point of view of the statistical theories of diffusion, the
most important measurement besides wind speed is wind direction variance.
The lateral direction variance, aa, is needed for correlations with ay,
and the vertical directional variance, ae, is needed for correlations
with az (the subscripts "a" and "e" stand for azimuth and elevation).
Note that aa or "gustiness" of lateral wind direction was measured in
almost every experiment in Table 1.  However, most older wind vanes were
poorly damped and had too much inertia, so they did not yield accurate
cra measurements.  Better instruments were available by the late
1960s.  In about half of the experiments, but dominately the elevated
source experiments, ae was measured; of these, the only ones with
direct tower- or cable-measured vertical  dispersion were Prairie Grass at
0.1 km, NRTS at 0.2, 0.8, 1.6, and 3.2 km (surface source) and 0.4 km
(elevated source), and Porton, with measurements to 0.5 km.  Thus, there
is a rather limited experimental basis at present for comparing vertical
diffusion with statistical modeling predictions.

                                    63

-------
     Surface-layer similarity theory,  favored  by  some modelers,  especially
for vertical  diffusion, requires  some  estimate of the Obukhov  length,  L.
This parameter is very difficult  to  measure directly, as  it  requires high
quality turbulent flux measurements.  This  was done  only  in  the  Cabauw
experiment and in the more recent CONDORS experiment to be discussed
shortly.  However, the Richardson number, Ri,  is  a function  of 7/L  and
can be used to determine L; Ri is defined by temperature  difference and
wind speed difference between two levels.   It  was measured and tabulated
in four of the experiments, but could  be derived  for any  of  the  experiments
having high quality profiles of temperature and wind.   In addition, L  can
be inferred from surface roughness and the  bulk Richardson number by
using known flux/profile relationships from surface-layer similarity
theory.  This number requires only AT  between  the two levels and IT at
one level to be determined; such  measurements  were made for  every experiment
in Table 1, so it is possible in  principal  to  make some estimate of L.
However, precisely calibrated measurements  of  T are  required to  accurately
measure AT; often these measurements are not sufficiently reliable.
Another limitation on this approach  is that the surface layer  may be quite
shallow in moderately to very stable conditions,  so  for the  flux/profile
relationships to be valid, AT and IT must be measured rather  low, preferably
below z = L; this condition was not  met for many  of  the elevated source
experiments.

     It is also desirable to compare diffusion observations  with the more
empirical prediction schemes that use  Pasquill/Turner categories or AT.
While every experiment included either temperature profiles  or AT, any

                                    64

-------
comparison using AT would be confounded by the great variety of height
pairs used to define it; experiments including detailed  temperature
profiles provide the possibility of standardization.  Either cloudiness or
solar insolation measurements are necessary to determine Pasquill  or
Turner categories for each run.   This information  was provided  in  only  five
of the experiments; this seems a serious oversight in the other experiments,
considering the ease of these particular measurements and the popularity
of the Pasquill/Turner method.  It may be possible to assign a  Pasquill
category to these experiments by working backwards through other schemes
that relate L, Ri, R& to these categories (e.g., Colder, 1972).  Such
indirectness should not be necessary.  Unfortunately, too many  past
diffusion experiments seem designed to test only one or  two modeling
approaches rather than all  of them.

Long-Range Measurements

     As might be expected from a simple extrapolation of trends in  Table 1,
as needs for diffusion data extend out to 100 km or more, the experimental
information tapers off rapidly.   This is especially so for vertical
diffusion data, which are virtually nonexistent beyond x = 3 km.   In
convective conditions, which normally prevail  during the daytime,  this
is not a great lack because it is well  established that  tracers and
light particulates become nearly uniformly mixed from z  = 0 to  z-j  by
vigorous vertical turbulence.  Above zj concentrations fall  off rapidly
to near zero at z = 1.1 Z}.  For passive tracers,  this vertical  concentration
distribution is approximately established at a time of travel = 3  z-j/w*,
                                    65

-------
typically 20 to 40 min;  x is  then  typically  2 to  10  km.   In neutral and
stable conditions, theoretical  models  suggest that complete vertical
mixing through the turbulent  boundary  layer  takes a  time  of the order
of f~l, where f is the Coriolis parameter.   This  is  about  3 h at middle
latitudes, far beyond the usual  travel  times associated with vertical
diffusion measurements.

     When one considers  diffusion  over days  of travel, daytime mixing to
the height z-j is so rapid compared to  vertical mixing in  stable conditions
that it is usually adequate to  assume  that vertical  diffusion is "frozen"
from about sunset until  the convective mixing layer  develops through each
layer the following morning,  as long as there is  no  appreciable settling
velocity or scavenging by precipitation. When cumulus clouds develop in
daytime with bottoms near Zj, material  can be drawn  up into them by the
updrafts supporting the  cumuli, a  process known as "cloud  venting" (Ching
et.al., 1983).  This can take material  out of the boundary layer; if the
material has a very long residence time, over a period of many weeks it
will mix through the whole troposphere, where the earth's  weather takes
place.  On these time scales, we inevitably  are faced with processes that
are beyond the scope of this  paper, which is limited to "simple" boundary-
layer diffusion.  To stay within this  scope, we shall limit the travel
times under consideration to  about 1/2 a diurnal  cycle (12 hours), or
travel distances to a few hundred  kilometers.

     For lateral dispersion over such  distances,  somewhat more than a
dozen sets of observations have been  summarized by Pasquill and Smith
                                    66

-------
(1983) and by Draxler (1984).  These observations  are  nowhere near  as
substantial and well-documented as  those in  Table  1  for  short ranges.
Some long-range observations are from a  classified project  (Slade,  1968)
and some are unpublished.   Most of  the published observations have  very
little meteorological information other  than broad designator terms like
stable and neutral.  Because many of these observations  were made by
traversing a plume at a series  of distances  and heights  with a  single
aircraft, and there was inadequate  time  for  repeated traverses  over such
distances, they were really quasi-instantaneous measurements of plume
width or of 0y, rather than a time  average.   Traditional  statistical
theory predicts that  instantaneous  plume widths approach time average
widths at large travel  times compared to the Lagrangian  turbulent time
scale, but this notion has been increasingly questioned  simply  because
horizontal eddies in  the atmosphere at any time scale  up to days can be
found (synoptic "highs" and "lows").   In one of the unpublished experi-
ments (Ferber and List, 1973) referenced by  Draxler  (1984), both instan-
taneous aircraft measurements and time-averaged surface  sampling of a
3-hr release of SFs were made for the same 76-m high release in windy,
neutral conditions; the time-averaged surface-measured ay values were
3 1/2 times as large as the aircraft-measured  ay values  in the  range of
6 to 18 km downwind.   For  1-hr  averages  at x ~ 100 km, the differences
are probably much less, but there appear to  be no  data for testing  this
assumption.

     The ay versus x  data  for the x = 2- to  100-km range summarized by
Draxler (1984) all show similar power laws of  growth, with n in ay  « xn

                                    67

-------
ranging from about 0.7 to 0.9.   This  seems  to  hold  regardless of stability
or of whether ay was from instantaneous  aircraft  traverses or from
surface sampler measurements  of a  3-  or  4-hr tracer  release.  Similarly,
Pasquill  and Smith (1983) fit a diverse  collection  of data in the range
x = 3 to 1000 km with n = 0.875.  These  are significantly larger values
of n than the value 0.5 predicted  by  statistical  or  eddy diffusivity
theories at large x.  Draxler (1984), Pasquill  and  Smith (1983), and many
others have suggested that wind direction  shear may  be  responsible  for
the more rapid growth of ay at large  x.   However, the large-x experiments
so far seem to lack wind profiles  and other supporting  meteorological
measurements for testing models of plume growth,  so  we  cannot progress
much beyond empirical power laws.

Useful Buoyant Plume Observations

     Although this paper does not  cover  buoyancy  effects and will pass
over diffusion data from buoyant sources,  such as those surveyed by
Sklarew and Joncich (1979), diffusion of plumes from buoyant sources is
not always dominated by buoyant effects.  Thus, sometimes observations
from buoyant plumes can be used to test  aspects of   passive diffusion
models, especially when atmospheric diffusion  is  rapid, as in unstable
conditions.  For instance, Carras  and Williams (1983) determined vertical
plume growth versus distance by using time lapse photography at  four
stack sites in Australia during convective conditions.  For the  three
sources of lesser buoyancy, good agreement was found with Lamb's  (1979)
convective scaling predictions for passive diffusion; the most buoyant
                                    63

-------
source had enhanced growth at small  x,  but  its  plume growth  also  fit the
passive plume asymptote at larger distances (>1 km).  For  ay,  Briggs
(1985b) demonstrated that buoyancy has  little effect on  plume  growth in
convective conditions if the dimensionless  parameter F*  <  0.06; F*  can be
defined as (w*/U) times the ratio of the stack  buoyancy  flux to the
buoyancy flux from a surface area n Zj2, owing  to  surface  heating and
evaporation.  Thus, F* is a measure of  the  relative buoyancy contained in
a segment of bent-over plume compared to that in atmospheric thermals.
Except for large sources of buoyancy in low wind speed conditions,  the
F* < 0.06 criterion is usually met.   Briggs (1985b)  also presents evidence
that suggests that buoyant plumes trapped within a  convective  mixing
layer become uniformly mixed in the vertical, just  as  passive  plumes do,
when X = (w*/U)x/zi > 30 F* (or, equivalently,  x >  30  Fb/w*3), where Fb
is the stack buoyancy parameter.  Thus  at these large  distances, the
diffusion of buoyant plumes released into and trapped  with a convective
mixing layer will be essentially the same as  passive source diffusion
from any zs < Zi.  In such cases, plume width measurements at  large
downwind distances are useful  even if the plume is  initially buoyant; a
good example of such measurements are those reported by  Carras and
Williams (1981), which extend from 15 to 1000 km downwind  of an isolated
stack, a considerable extension beyond  the  observation distances in
Table 1.

Remote Sensing Experiments

    Another type of experiment not included in  Table 1 is  remote sensing
experiments, which are very recent.   These  overcome  the  limitations of
                                    69

-------
surface-based sampling by measuring  plume concentrations with reflected
electromagnetic radiation.  Lidars have  actually been  used since the
early 1960s to define the rise  and other geometric  properties of buoyant
plumes.  They send out narrow pulses of  monochromatic  light  from a laser
and then record the time lag and  intensity  of the reflected  return.  Present
operational lidars work best with  plumes composed of ~1 urn-sized particles,
which are easily produced by using an oil fog generator (other, more
conservative tracers are available,  but  these are also more  expensive).
A lidar "scan" consists of a series  of pulses made  at  a fixed azimuth
with a small increment in elevation  angle made between each  pulse.  Up to
10 pulses per second have been  obtained; a  mapping  of  plume  concentrations
at a half-dozen azimuth angles  can be completed  in  2 or 3 min with such
instruments; repetition of such rapid scanning for  30  min or more produces
an adequate ensemble of individual scans for defining  an average plume.
To infer concentrations of tracer, or rather of tracer reflectivity,
requires careful processing of  the return signal to compensate  for range,
background haze, and attenuation  of  the  reflected light returning through
the plume and background haze.   This degree of sophistication is a recent
achievement (Eberhard et al., 1985); earlier experiments used lidar
mainly to define the edges of buoyant plumes.  One  limitation of lidar is
that for eye safety reasons it  should not scan within  15 m or so of the
surface.  Thus, it is best used for  elevated sources or for  surface
releases at distances where they  are diffused through  several hundred
meters or more.  Lidar's great  advantage is that it can detect  sufficiently
dense plumes at a 5-km or more  range.   This allows the mapping of diffusion
                                    70

-------
through a mixing depth of 2 km or more.  Thus,  lidar measurements comple-
ment observations that can be made by using samplers on surface arcs and
towers.

     Lidar has been used in two recent passive  plume experiments, one
carried out in stable conditions and the other  carried out in  unstable
conditions.  Both use oil  fog as the plume tracer for the lidar,  which
was provided and operated  by the National  Oceanic and Atmospheric Admini-
stration's Wave Propagation Laboratory (WPL).   The experiment  made in
stable conditions at Cinder Cone Rutte, Idaho,  was primarily a complex
terrain experiment, but ay and az determinations  were made by  lidar
at small  distances, 0.1 to 0.9 km, upwind  of the  isolated hill  (Strimaitis
et al., 1983).  Source heights ranged from 20 to  60 m, and meteorological
measurements on a nearby 150 m tower included temperature at eight  levels,
three-dimensional wind speed and turbulence intensity at  five  levels,
insolation, and net radiation.  These measurements are sufficient for
testing any presently known modeling approach and have been used  to test
a combination statistical/similarity equation for az (Venkatram et  al.,
1984).

     The CONDORS experiment (CONDORS = Convective Diffusion Observed with
Remote Sensors) made at WPL's Boulder Atmospheric Observatory  was designed
to define three-dimensional  plume concentrations  in highly convective
conditions by using both 1idar-detected oil  fog and doppler radar-detected
"chaff".   This tracer consists of bundles  of aluminum-coated thread which
are chopped into 1.5-cm lengths (half wavelengths)  and ejected in an air
                                    71

-------
jet.  The radar can map the chaff plume  for  tens of  kilometers and can
detect concentrations as low as  10~5  threads/m^ with almost no background
or attenuation problems except  for "ground clutter"  presented by metallic
objects such as power lines.  The radar's lateral  and  vertical resolu-
tion is limited by its 0.8° beam width.  The main  drawback of this remote
sensor is that the lightest chaff presently  available  has a settling velocity
of about 0.3 m/s.  Used in vigorous convection with  vertical turbulence
velocities of about 1 m/s, the chaff  plume does mix  up to z-j, however, its
mode was observed to be displaced downward compared  to the oil fog plume,
with this displacement increasing with distance (Moninger et al., 1983).
In half of the CONDORS runs the chaff and oil fog  were released side-by-side
at heights ranging from 167 to 280 m, using  the 300-m  meteorological
tower.  For other runs, except for one with  both sources on the ground,
the chaff was elevated and the oil  fog was on the  ground.   In addition,
the main part of the experiment in 1983  had  one sampling arc at x = 1.2
km, and SF^ was released from the elevated release point; this made
possible some comparison of X/Q inferred from the  remote sensors and
surface tracer measurements (Eberhard et al., 1985).   Because some of the
chaff deposits and the oil fog slowly evaporates,  a  constant value of 0
cannot be assumed, as for a conservative tracer; instead, the total flux
at each distance, IT //Xdydz, was substituted for Q.   (This  is far superior
to experiments with non-conservative  tracers and only  surface samplers,
because there is then no direct measurement  of the total flux of tracer
versus distance or of the extent of vertical diffusion.) Meteorological
measurements in the CONDORS experiment  included temperature, three components
                                    72

-------
of wind speed measured with sonic  anenometers,  turbulent  variances, and
turbulent fluxes measured at eight levels  on  the  tower.   In  addition,
solar radiation was measured and three or  four  rawinsonde releases were
made during each of the 11 two-hour runs;  z-j  was  determined  from the
rawinsonde temperature and humidity soundings,  the  top of the  haze layer
as seen by lidar, and the top of the chaff as seen  by radar.

New Experimental Capabilities

     It is essential  that the most useful  experiments have been those
with a full complement of meteorological measurements.  Of the historic
experiments, Prairie Grass stands  out, even though  it was one  of the
first, made in 1956.   This data  set has been  used much more  than any
other, largely because almost every needed meteorological  parameter was
measured or could be estimated from the available measurements.  Some of
the recent experiments like Cabauw, Cinder Cone Butte, and CONDORS have
returned to this philosophy of measurement; they, too, should  prove to
be very versatile bases for model  testing  and will  raise  the standards
for future diffusion  experiments.   All  of  the earlier experiments suffered
to some degree from tracer deposition, unreliable turbulence measurements,
and physical/economical  limits on  vertical  sampling.  Recent experiments
have benefited from the use of more conservative  tracers,  much improved
turbulence instrumentation, and  remote sensors.  These capabilities
present many opportunities for dependable  measurements over much greater
ranges of distance, height, and  stability  conditions than  was  previously
possible.
                                    73

-------
LABORATORY EXPERIMENTS

     Fluid modeling experiments  in  wind  tunnels and water channels have
long been used as tools  for  simulating large-scale turbulent flows over
and around obstacles.  Confidence  in  these tools has increased to the
point that they have been  used extensively to  investigate diffusion from
releases on and near buildings and  terrain features (Hosker, 1984).  They
are increasingly used to model diffusion in complex terrain, including
stability effects, primarily because  they are  much cheaper to operate
than large-scale experiments, particularly in  difficult terrain.  In the
last few years, the U.S. EPA has supported several paired field/fluid
modeling experiments on  diffusion  in  complex terrain; these were directed
towards improving modeling of diffusion  in complex terrain and testing
the reliability of fluid modeling  in  such applications  (Schiermeier,
1984).  With some limitations, the  fluid modeling simulations have been
quite successful.

     In addition to being  less expensive than  field experiments, laboratory
modeling offers control  over the meteorological variables, so that both
the flow and surface characteristics  can be idealized.  Therefore, it is
an attractive alternative  for investigation of "pure" turbulent diffusion,
free of the complexities found in  the field due to uneven terrain, uneven
surface heating or cooling,  uneven  surface roughness, and variable winds
that cause part or all of  the plume to miss the samplers.   In bypassing
these complexities, fluid  modeling  is naturally unable  to simulate some
features of the real atmosphere.  The most conspicuous  missing element is
                                   74

-------
plume meander due to large-scale horizontal  flow structures many times
larger than the mixing depth.   Such  eddies,  mini-fronts, or lines of
convergence may be caused  by large-scale  topographic variations, surface
inhomogeneities, wind direction shears, or even cloud  shadows  in the real
atmosphere.  Straight-sided  wind tunnels  and water  channels cannot simulate
wind direction shear effects,  and their walls  prevent  the development of
any horizontal eddies larger than their width.  Thus,  laboratory lateral
diffusion measurements correspond to  the  minimum lateral diffusion that
occurs under the same stability conditions in  the atmosphere during
periods free of these larger scale effects.  Laboratory ay values
probably represent 10- to  20-min averages in the real  atmosphere better
than 60-min averages.

     The great reduction of  scales in  laboratory modeling facilities also
causes a great reduction in  the range of  turbulent  eddy sizes  compared
to the atmosphere.  The smallest eddy sizes  are limited by fluid viscosity
and the turbulent energy dissipation  rate.   In the  atmosphere, the diffusive
ability of very small eddies (~1 mm)  is usually inconsequential, except
for the mixing of chemically reactive species.  The problem with very
large scale reductions is  that viscosity  begins to  smooth and  distort even
the largest eddies, which  are  the most effective agents of diffusion, so
that none of the full-scale  turbulent  diffusion is  well simulated.  In
the atmosphere, only the motion of eddies smaller than about 1 cm are
effectively smoothed and dissipated  by the action of molecular viscosity.
The degree of scale reduction  that avoids viscosity-induced distortions
is most limited when modeling  plume  buoyancy or ambient stratification

                                    75

-------
(stable or unstable atmospheres).   This  is because gravity then plays a
role, and gravity cannot  be adjusted.  Gravitational acceleration has the
dimensions of velocity squared divided by length; if lengths are reduced
by a factor S, velocities such as  wind speed must be reduced by a factor
S1/2 to maintain the ratio of inertia! forces to gravitational forces (this
assumes that full-scale density  differences are maintained).  The importance
of viscosity is measured  by the  inverse  of the Reynolds number, Re, which
is a length scale times a velocity scale divided by the (kinematic)
viscosity.  Thus, Re must be reduced by  the factor S3/2.

     To illustrate, suppose we wish to model a 1 km width full-scale terrain
segment in a 2-m wide wind tunnel, so S  = 500.  A full-scale wind of 10 m/s
blowing over roughness elements  of size  1.5 m has Re = 10^.   If we model
neutral flow, we can maintain the  10 m/s flow speed and reduce the roughness
size by the factor of 500, to 3  mm; this results in Re = 106 * 500 = 2000,
which is large enough for good turbulence simulation.  However, if we
wish to model a stratified atmosphere, we must also reduce the wind speed
by a factor of (500)1/2 to 0.45  m/s, which reduces the Re to 90; this is
submarginal , so that viscosity smoothes  the flow around the roughness
elements and they appear  smoother  to the wind than they should be.
A non-practical solution  to this problem is to put the whole wind tunnel
in a giant centrifuge in  order to  increase the effective gravitational
acceleration by substituting centripetal acceleration.  More  practical
solutions are (1) to use  less scale reduction, thus modeling  less area,
(2) to use increased density or  temperature differences to increase buoyant
accelerations, which is equivalent to increasing gravity, or  (3) to

                                   76

-------
experiment with exaggerated roughness  elements to  achieve  a good simulation
of full-scale wind speed and turbulence  profiles.   All  of  these strategies
have their limits, and the bottom  line is  that there  are limits to the
scale reduction possible to successfully simulate  atmospheric turbulence or
turbulent diffusion.  Several  other  scaling  limitations of fluid modeling
have been discussed by Snyder (1972  and  1981).

     Laboratory diffusion experiments  have been made  in wind tunnels,
water channels, and water tanks.   The  latter have  been  used to study
diffusion by purely convective turbulence.  Since  tank  walls prohibit
crossflow, there can be no turbulence  generated by boundary-layer wind
shear.  Water channels are relatively  easy to stratify  and are most often
used as towing channels for buoyant  sources  or for elevated releases
upwind of a terrain feature; that  is,  the  model is towed along the length
of the channel.  This simulates the  effect of a uniform crossflow velocity.
This technique allows easy maintenance of  strong,  stable stratification,
which is not easily maintained in  a  return flow mode  (model stationary,
water moving through channel  and  recirculating).   It  is used to study
only the effects of crossflow and  stable stratification, however, it does
not permit simulation of ambient  shear-generated turbulence.  Such simu-
lation requires a flow of fluid across a long fetch because the turbulent
boundary layer that develops increases in  depth slowly  and is only a
small fraction of the fetch.   The  boundary layer depth  usually determines
the scale-down factor.  Only wind  tunnels  have been used to study diffusion
in shear-generated turbulence. Large  facilities with long fetches are
cheaper to build and to operate using  air  as the operating fluid rather

                                    77

-------
than water because water is  so  heavy;  it  is also easier to measure flow
velocities and turbulence in air.   Large  wind tunnels used to simulate
atmospheric boundary layers  are generally about 2 m high, 2 to 4 m wide,
and 30 m long.

Neutral Boundary Layers

     Simulation of diffusion in NBLs  is relatively easy to perform in
wind tunnels because no  heating or  cooling is required and large flow
speeds can be used.  This helps keep  model Reynold's numbers high enough
to create well-developed turbulence that  closely resembles full-scale
turbulence.  The main requirement for good boundary layer simulation is a
long fetch over a rough  surface.  The boundary layer develops to a reason-
able depth, e.g., 1 m, more  quickly if turbulence is initiated through
the whole depth of the tunnel at the  upwind end by using an array of drag
elements known as "vortex generators". Surface-release diffusion experi-
ments in neutral flow were performed  by Robins (1978) using such a confi-
guration.  The wind tunnel was  unusually  wide, 9.1 m, and 2.7 m high;
this aspect ratio should allow  freer  development of horizontal eddies and
lateral diffusion.  Experiments were  performed with two different floor
roughnesses, one rather  smooth  to simulate "rural" diffusion (z0 = 3 cm
full-scale) and one rough to simulate "urban" diffusion (z0 = 130 cm
full-scale).  The depth  of the  developed  boundary layers at the release
point was 0.6 m and 2 m, with scale-reduction factors of 1000 and 300,
respectively, for a full-scale  boundary layer thickness of 600 m.  Measured
turbulence intensities,  when scaled with  the friction velocity, u*, were
                                    78

-------
within 5 to 10% of best estimates from field measurements.  The mean
plume measurements revealed nearly Gaussian lateral  plume profiles and
quasi-Gaussian profiles in the vertical, fitting X <* exp(-azn)  better
with n = 1.7 than with n = 2, the Gaussian value.  The growth of the
vertical depth of the plumes fit quite well  a prediction based  on surface-
layer similarity theory, 6 In(0.176/z0) = 0.22x, where 6 is the
height at which the concentration falls to half its  surface value.  The
lateral spread averaged approximately 20% less than  a prediction based on
statistical theory.  Overall, Robins'  experiment showed that quite good
simulation of turbulence and diffusion, especially for vertical  diffusion,
in a neutral ABL can be achieved in a large wind tunnel.  For lateral
diffusion, growth due to wind directional  shear cannot be simulated in a
parallel-walled tunnel.  In the atmosphere,  wind shear effects  may
contribute substantially to 
-------
of wind and wind shear, but  the  tank  has  proven  to  provide a  reasonably
good simulation of purely convective  turbulence;  turbulent velocity
variances, nondimensional ized  by the  convective  velocity scale w*  and
plotted against z/z-j,  were in  good  agreement  with aircraft measurements
for the vertical component,  aw»  and were  about 30%  lower than the
atmospheric measurements for horizontal components  (Willis and Deardorff,
1974).  This may be due partly to the horizontal  motion-inhibiting effect
of the tank walls.

     Diffusion experiments were  made  in this  tank by  instantly releasing
over 1000 tiny (~1 mm), neutrally-buoyant oil droplets  from a line
source crossing the tank at  various fixed source heights, zs.  These were
photographed looking down the  plume axis, and inventories of  particle
positions were made at a series  of  times, t,  after  release.   To  interpret
these results for vertical and lateral diffusion  in terms of  distance
downwind of a continuous, point  source in a crossflow,  it is  simply
assumed that x = ~ut.  This assumption appears reasonable, considering the
fact that "u is nearly constant with height in a  very  convective  ABL,
except close to the surface.  Since the tank  diffusion  is, in effect,
integrated along the longitudinal coordinate, x,  in the photographs, this
technique fails to account for longitudinal diffusion;  Deardorff and
Willis (1975) estimate this phenomenon to be  significant for  point sources
only when U < 1.5w*, rather low  wind  speeds  (a typical  summer midday
value of w* is 2 m/s).
     The results of the Willis and  Deardorff  convective tank  diffusion
experiments were surprising and  somewhat  troubling  to the diffusion

                                   80

-------
modeling community because they revealed features  of vertical  diffusion
that were at odds with assumptions common to all Gaussian  models,  so
widely accepted and used.  The near-surface releases, with zs/z-j = 1/15,
diffused in a Gaussian manner at first,  up until t = 0.5 z-j/w*, but
shortly thereafter the locus of maximum  concentration lifted  off the  surface
and lofted into the upper half of the mixed layer.  This was  accompanied
by a rapid decrease in surface concentrations  until  nearly uniform vertical
mixing from z = 0 to zj was achieved  at  t = 3  z-j/w*, the limit of  the
experiment.  (In convective ABLs, updrafts and downdrafts  extend through
nearly the whole layer with mean up and  down speeds  of 0.6w*  and 0.4w*,
respectively (Lamb, 1982); therefore, the time needed for  one complete
"stir" is approximately Zj/(0.6w*) +  Zi/(0.4w*)  =  4zj/w*.)  The elevated
releases, at zs/Zi = 1/4 and 1/2, exhibited non-Gaussian vertical  behaviors
from the outset;  the locus of maximum concentrations descended from  the
release heights until  they reached the surface at  t  « 2 zs/w*.  Naturally,
this produced higher surface concentrations than predicted  by  Gaussian
models, which assume that the centerline of the plume remains  at z =  zs.
At larger t, the locus of maximum concentration lifted off the surface and
lofted into the upper part of the mixing layer,  just as had occurred  for
the surface releases.

     These results were unexpected and stirred some  controversy, particularly
because convective tank experiments are  a relatively new tool  for  simulating
atmospheric dispersion.  Some modelers questioned  whether  the  observed
diffusion patterns occur in the real  atmosphere.   Unfortunately, the
answers could not immediately be found in existing field data  due  to  many

                                    81

-------
deficiencies already discussed,  especially  the  lack  of  vertical  concentra-


tion profiles to any substantive height.  The same phenomena  were  observed


in the numerical modeling results of Lamb (1979), which provided some


confirmation.  (Lamb's work followed the surface-release tank experiments


and actually preceded and inspired the elevated-release tank  experiments.


For a history of developments in convective diffusion,  see  Bn'ggs,  1985a.)


Papers by Bn'ggs and McDonald (1978) and Nieuwstadt  (1980)  confirmed the


fact that surface concentration  results from the Prairie Grass field


experiments were consistent with the results of Willis  and  Deardorff


(1976a) and of Lamb (1979), when convectively scaled, except  that  ay


values in the tank were about 30% smaller than  the  field values (this


seems consistent with the smaller horizontal velocity variances observed


in the tank).  Preliminary results from the CONDORS  field experiment,


which was largely motivated by the results  of Lamb's and Willis and


Deardorff's experiments, did show the descending plume  behavior for a


zs/z-j = 1/2 case, with a remarkably good comparison  with the  convectively-


scaled tank results for crosswind-integrated concentrations (Moninger  et


al., 1983).  (Convectively-scaled means that heights are scaled with z-j,


time after release or x/TF is scaled with ZT/W*, and  concentrations

                   _  O
are scaled with 0/(uzj ).)  Further good comparisons of the tank experiment


results with some power plant plume data, the Cabauw field  experiment,


and CONDORS data were shown in Rriggs (1985a).   The  convective tank


experiment results are now being taken quite seriously, and many researchers


have been developing mathematical diffusion models  that fit these  results


(such models are surveyed in Briggs, 1985b).
                                    82

-------
     More recently, Poreh and Cermak (1984)  demonstrated  that  convective
ABL simulations producing results  similar to those of the tank experiments
could be performed in a large boundary-layer wind  tunnel.  This  has the
advantage of including some of the crossflow and wind speed  shear  effects
found in the real  atmosphere.  They ran  these experiments using  both  a
smooth floor consisting of a 12 m  long heated aluminum plate and a rough
floor, with twisted-link chains stretched across the  plate.  The simulation
had some shortcomings, e.g., the fetch of heated plate upwind  of the
source was only about 1 TT(zi/w*),  which  is not sufficient for  full convective
turbulence development (Briggs, 1985a),  and  the convectively-scaled vertical
velocity variances were only about 2/3 of values observed in the atmosphere.
Nonetheless, plume behaviors similar to  those observed in the  tank experiments
were observed in the tunnel.  The  lateral  diffusion rate  was larger,  even
somewhat larger than that observed for the Prairie Grass  field experiments
(Nieuwstadt, 1980), in spite of the lateral  width  restriction  of about
2.4 z-j in the tunnel.  An improved convective diffusion experiment was
recently completed in the same tunnel;  stronger stable capping  and
increased dimensionless upwind fetch over the heated  plate were  achieved
(Poreh, 1985: personal  communication).

Stable Boundary Layers

     Some basic diffusion experiments in SBLs layers  have been done in
the same wind tunnel  facility at Colorado State University.  In  fact, a
recent series of baseline diffusion measurements were made in  this tunnel
for every combination of (1) stable, neutral, or convective  boundary
layer, (2) smooth or rough floor and (3) surface source or elevated
                                    83

-------
source (one or two different  heights)  (Cermak  et  al.,  1983).   The  smooth
floor configuration still  developed  turbulence (a tripping  device  in
these experiments helped maintain  turbulence throughout  the test section
of the tunnel).  The effective roughness  length of a  smooth surface  is «  v/u*,
where v is the kinematic viscosity of  air (0.15 cm2/s  at room  temperature);
in these experiments it corresponded to approximately  1  cm  at  full scale
(similar to the Prairie Grass field  site).   The rough  floor configuration
gave z0 = 16 cm, full  scale,  typical of mixed  surface  cover.   To achieve
the SBL, incoming air was  preheated  to 78°C before entering the test
section, which had its floor  temperature  cooled to about 0°C.  This  large
temperature difference helped to compensate for scale-down  effects by
increasing buoyancy forces.  Unlike  earlier experiments  that achieved
only slightly stable boundary layers,  with  diffusion  slightly  reduced
from the neutral case (Chaudhry and  Meroney, 1973), the  recent experiments
achieved substantial stability.  Vertical diffusion at the  larger  distances
was reduced by factors up  to  5.5 compared to neutral  conditions with the
same tunnel configuration.  The estimated values  of the  Obukhov length,
16 cm and 25 cm for the smooth and rough  floors,  respectively, are frac-
tions of the boundary layer depth  =  1  m), that correspond to moderately
stable conditions in the atmosphere  (Pasquill  category "F").   Measurements
from this series of experiments were included  in  a comparison  of wind
tunnel ay measurements versus field  data  and statistical-theory predic-
tions (Li and Meroney, 1984).

     A recent diffusion experiment in  a boundary-layer wind tunnel  in  Japan
(Ogawa et al., 1985) achieved even stronger SBLs, with Obukhov lengths

                                    84

-------
as small as 1/13 times the boundary layer depth.   Vertical  diffusion  was
sharply reduced from its neutral  growth at the greater  stabilities,  results
comparable to those of Cermak et  al. (1983).   A somewhat  surprising  result
was that lateral diffusion was reduced from its neutral growth  by  about
20% in the moderately stable SBLs,  while it was actually  somewhat  larger
than its neutral growth in the strongly stable SBLs.  Ogawa et  al.
attributed this growth to the development of  horizontal meanders in  the
very stable flow, a phenomenon that field observations  of lateral  turbulence
velocities tend to support (Hanna,  1983).  The Ogawa  et al. (1985)
observations also showed significant departures from  Gaussian vertical
profile shapes for the neutral and  very stable cases.  These experiments
were conducted only for a surface source and  a smooth floor configuration.

Potential  Further Uses of Laboratory Simulations

     The results reviewed in this section support  the idea  that fluid
modeling,  in the laboratory, is a valid way to improve  our  understanding
and modeling of diffusion.  However, it is obvious that this tool  has not
been fully exploited.  Considering  the fact that it is much cheaper  to
run diffusion experiments in the  laboratory,  under controlled conditions,
than in the field, it makes sense to use laboratory facilities  as much as
possible.   These laboratory experiments should be  supplemented  by  selected
field experiments to provide confidence-building and  to measure features
that cannot be produced in the laboratory, like crosswind shear-induced
diffusion.  For neutral conditions, elevated  releases in  a  well-developed
boundary layer could be run by using several  contrasting  floor  roughness.
In a wide  tunnel it would be informative to run experiments with a patchwork
                                    85

-------
pattern of contrasting surface  roughnesses  (on a  scale small compared to
the tunnel width);  this would be much  more  like the  real world and might
reveal  effects of the surface inhomogeneities on  lateral diffusion (surface
drag inhomogeneities  would  cause vertical torques that might produce
larger and more vigorous horizontal  eddies).  For convective conditions,
a similar exploration could be  made of the  effects of surface heat flux
inhomogeneities, which would provide forcing for  the turbulent mixing
much more like the real world than  a uniformly heated aluminum plate
(most places of interest have a complex pattern of different types of
vegetation, different radiation balances, and different degrees of moistness
and thus, large contrasts in surface heat flux).  For stable conditions,
it now appears possible to  simulate fairly  strong stable ABLs.  There is
a need for much more work on the effect of  height of release relative to
the boundary layer height for all  stabilities.
                                    86

-------
                                SECTION 4



                     DIFFUSION MODELING AS PRACTICED



         How well does it reflect the current  state of knowledge?



THE GAUSSIAN PLUME - ITS UTILITY AND LIMITATIONS



     The vast majority of "production line"  diffusion  models  are Gaussian,


i.e., they make the basic assumption that concentration distributions


relative to the center of a puff or plume have a Gaussian  shape.   In a


recent summary of scientific reviews of eight  models (Fox  et  al.,  1983),


most of the reviewers "felt that they were reviewing a single,  physically


incomplete, Gaussian model  with slight variations,  rather  than  eight


substantially different models".  They also  felt that  these models do not


represent the state of the science.  The reason that Gaussian plumes have


come to dominate the model  market is a matter  of precedence and convenience.


Early diffusion models were naturally influenced by much earlier treat-


ments of molecular diffusion.  Molecular diffusion  in  a fluid is caused


by countless collisions and rebounds among the molecules,  which can be


treated statistically as a "random walk" process.  When concentration and


diffusivity vary on scales that are large compared  to  the  mean molecular path


length, the diffusion process can be described by the  "parabolic equation",


                    5      ax
          dx/dt = 	  (K — ) ,                                 (4.1)
                  sx-j     ax-j




where x-j (i = 1, 2 or 3) represents the three  orthoganal spacial coordi-


nates.  In this context, K, the diffusivity, is directly related to the



                                    87

-------
product of mean molecular path  length  and  speed.   In  the  absence of shear
and when K is constant, the solution of this  equation  for a  tracer released
at a point at t = 0 is Gaussian;  the standard deviations  about the center
are
          a  = (2Kt)l/2 .

     This approach was adapted  to modeling diffusion  in the  atmosphere by
replacing molecular diffusivity with "eddy diffusivity",  which in analogy
to random molecular motions is  supposed to be proportional to the product
of mean turbulent eddy size (or spectral  peak)  and  turbulent velocity
variance.  Whenever the atmosphere is  turbulent,  this  K is orders of
magnitude larger than the molecular diffusivity,  so the latter is thought
to be negligible.  The classical  treatment for  diffusion  from a ground
source using the parobolic equation with u" «  zm and Kz «  zn  was first
published by Sutton (1953), where Kz is the vertical  component of K.  The
resulting distribution in the vertical  is  Gaussian  only for  the special
case m = n.  In general, x * exp  (-constant • zP) with p  = 2 + m-n.   If
n = 1, as it is observed to be in the  neutral surface layer, and if wind
shear is neglected (m = 0), an  exponential  distribution results.  Values
of p fitting observed surface-source concentration  distributions have ranged
at least from 1.15 to 3 (Pasquill, 1974, p. 205;  Ogawa et al., 1985).
However, Pasquill demonstrated  that the practical effect  on  the ratios
of various profile constants is very small  for  1.25 <  p < 2.5, which  suggests
that there is no harm done in assuming a Gaussian distribution (p = 2)
for these cases.  Large errors  result  if a Gaussian distribution in the
vertical is assumed when the locus of  maximum concentration  from a ground

                                   88

-------
source becomes elevated; no value of p fits this case.  This phenomenum
has been observed in convective conditions, and has undoubtedly resulted
in some erroneous diffusion modeling as a consequence of assuming  Gaussian
plume distributions.

     There are many in which a Gaussian plume shape is an acceptable
approximation.  For a sufficient sampling duration, lateral  diffusion
near the source is Gaussian if the distribution of lateral  turbulent
velocities is Gaussian.  This is normally the case, except  when the  sampling
period encompasses a substantial wind shift.  The same applies  to  vertical
diffusion from an elevated source, almost to the point that  significant
concentrations reach the ground, but this is not valid in the convective
mixing layer.  For this case, it has been amply demonstrated that  vertical
turbulent velocities have a skewed distribution favoring downdrafts.  The
simple Gaussian assumption is wrong here and underpredicts  maximum ground
concentrations (Briggs, 1985a).  The vertical  velocity distribution  can be
fit with a double-Gaussian shape, however, and this has been assumed in
some recent dispersion models.  A near-Gaussian shape is often  observed
in lateral  concentration distributions at distances of many  kilometers.
This is to be expected when the plume width far exceeds the  width  of the
largest horizontal eddies because then horizontal  diffusion  becomes  akin
to a random walk process.  At moderate to large distances in unstable
conditions, after a travel  distance of about 3(U/w*)zj, passive material
becomes almost uniformly distributed from the surface to the top of  the
mixed layer.  Gaussian models can approximate this by assuming  reflection
at the surface and at z = Zj (see Pasquill, 1976b).

                                    89

-------
     Some further discussion of when  plumes  are  and  are  not  Gaussian  is
given by Pasquill (1978, pp. 6-9).   He notes that  power  station  plumes
have been observed to be Gaussian only in  their  lower  half.   Weil  (1983)
speculates that the Gaussian assumption may  perform  satisfactorily in
convective conditions for vertical  distributions of  strongly buoyant
plumes, citing numerical simulations  of Lamb (1982); however, the  assump-
tion works poorly for passive or weakly buoyant  plumes.   Strengths and
weaknesses of Gaussian plume modeling have been  discussed by Smith (1981).
The conditions of applicability he  lists are somewhat  more restrictive
than the ones stated here.  Strengths that he lists  are  ease of  computation,
familiarity, and adaptability; this means  that ay  and  az can be  altered
to better fit new observations without fundamentally changing the  model.
Some of the limitations of Gaussian diffusion models that he disccusses
are the following:  they do not account for  wind direction shear,  they
cannot describe near-calm situations, and  they do  not  represent  diffusion
during stable conditions very well, when vertical  diffusion  may  be controlled
by source properties (for buoyant  plumes)  and lateral  diffusion  is due
mostly to wind direction shear and  meandering.  Gaussian models  have  also
been extrapolated to long range diffusion  predictions  over curved  trajec-
tories and changing meteorological  conditions, which has no  basis.

     To a considerable extent the  shortcomings of  Gaussian models  can be
overcome by adjustment of ay and az values to account  for complicating
factors which distort plumes from  this ideal. At  least  this is  usually
possible for making practical estimates of surface concentrations.  Lateral
surface concentration distributions can almost always  be fit reasonably

                                    90

-------
well with a Gaussian curve.  However, manipulations of 
-------
are certain to serve as metersticks  for progress  in  developing  improved
models.  A fairly comprehensive discussion  of existing  sigma curves  and
stability typing schemes, including  some that will he skipped here,  has
been given by Weber (1976).   The discussion here  is  limited to  four  generic
schemes.  Gifford (1975, 1976)  also  offers  an overview  and considers
schemes for exceptional flows,  including near calm,  very  stable condi-
tions, urban diffusion, diffusion over water, diffusion downwind of
buildings, and diffusion in  rugged terrain.

Brookhaven Curves and qa Typing

     Stability categorization methods based on wind  direction fluctuations
go back at least 50 years (Giblett et al.,  1932).  The  meteorology group  at
Brookhaven National Laboratory  (BNL) developed a  system of five stability
categories based on inspection  of one-hour  segments  of  strip chart recordings
of lateral wind direction fluctuations (Singer and Smith, 1966); this  is
a quick way to estimate aa.   They related these categories to measurements
of tracers released over about  1-hr periods at the top  of a  108-m tower
at BNL, which is in rolling, forested terrain of  relatively large roughness.
The tracers were uranine dye, oil fog, and  argon  41. The lateral plume
dispersion, ay, was measured at the surface directly, but az was inferred
indirectly by assuming conservation of the  tracer fluxes  and a  Gaussian
plume concentration distribution centered at the  release  height in the
vertical  (these were shown in Sec. 3 to be  somewhat  risky assumptions).
It was also assumed that ay and az have the same  power  law dependences
on x, and optimum fit power laws were determined  for each category.
                                    92

-------
     This typing scheme has a very direct physical  basis  for expecting
good correlation with lateral diffusion because near the  source passive
material goes as the wind blows, i.e., it tends to  travel  in straight
lines at first, so that the wind direction distribution also describes
the concentration distribution.  In terms of Taylor's (1921) statistical
theory, ay = aax at small x.  As material travels  far enough to be
influenced by a number of turbulent eddies of various scales, ay/(aax)
tends to get smaller, but this dropoff is gradual  and the rate of dropoff
is not very different in different stabilities.  Hence, aa or the wind
direction tracer measured at the height of release  over the desired  sampling
period correlates well with ay at any distance, provided  that the
stability has not undergone large changes since the time  of release.

     To address vertical diffusion, statistical theory gives us az = aex
near an elevated source, where ae is the standard deviation of vertical
wind angle.  This measurement is somewhat more difficult  to obtain than aa
and is often not made.  However, aa can substitute  for ae as a measure
of vertical diffusion if atmospheric turbulence is  nearly isotropic, so
that ay and az are of the same order.  This is true in NBLs and CBLs
but not in SRLs, because there horizontal motion can far  exceed the  scale
of vertical motions, which are restricted by vertical  stability.   Cramer
(1957) noted that aa correlated well  with ay measurements  made day
and night and with daytime values of az, but the correlation with
nighttime az was nil.  Briggs (1985a) argued on the basis  of convective
scaling theory that aa should also correlate well with vertical  diffusion
from a ground source while it is in the "free convective  regime", with

                                    93

-------
az <«: H*1/^3/2.  This follows from the approximate relationships t  =* x/u
and aa a av/u = 0.6w*/u; substituting the definition w* =  (H*z-j)1//3,
the free convective az « (aax)3/2z.j~1/2.  Because a-,3'2 ranges  far  more
widely than z-j'1/2, the correlation between az and aa is strong.

Pasquill/Gifford/Turner Curves and Pasquill Typing

     Pasquill's typing scheme, based on wind speed, insolation, and
cloudiness, has been very widely adapted, sometimes with minor  modifications,
The scheme is more or less intuitive and was suggested as  a way to  relate
diffusion rates to easily obtainable measurements.  It was in use at the
British Meteorological Office in 1958 and became widely used after
Gifford's (1961) adaptation of the scheme for Gaussian plume model ing.
Turner (1961) elaborated on the scheme somewhat by replacing subjec-
tive solar insolation categories with solar elevation and  cloud cover
estimates.  Smith (1972) proposed a scheme using wind speed, insolation
and cloudiness, but with a continuously variable numerical stability
index.  His az curves were based on an eddy diffusivity model and
included corrections for surface roughness; 
-------
the worst hours of the year; this approach would tend to overestimate
typical  1-hr concentrations, however.)   They are based primarily on  appli-
cation of the statistical  theory of Hay and Pasquill  (1959)  to wind
direction fluctuations observed at a 16 m height over relatively smooth
terrain with z0 = 3 cm.  There were some very limited tracer measurements
supporting the 
-------
1977 or Briggs, 1982b);  this correlation also applies  to the Richardson
numbers, Ri and Rfo, since they relate strongly and monotonically to L.
Golder (1972) used five sets of meteorological  data to develop nomo-
grams relating R-j and L to Pasquill  categories and surface roughness.
Briggs (1982a) showed that L relates well  to "psuedo"  L's defined by
Ls = -u"3/R| and Ln = -u"3/Rj!;, where R* and R* are g/(cppT) times
insolation and net radiation, respectively.  It was pointed out that Ls
and Ln simply amount to quantified forms of Pasquill's categories.
The moistness of the surface affects the ratios H*/RS* and H*/Rn*, so  also
affects the relationship between Pasquill  categories,  Ls, or Ln and L;
however, this is secondary compared  to the role of Tf3  in establishing  this
relationship.

     Pasquill's categorization method is less effective for ay because
scales other than the height above the ground can have more influence on
lateral turbulent motions.  In convective conditions,  z-j affects lateral
diffusion even near the ground, since av = 0.6 w*.  At small x,
 ay/x = aa = av/U <= (u*/u) |z.j/L| '•', and z^ has as much influence
on ay as L.   In stable conditions, large horizontal meanders are very
terrain sensitive and are not controlled at all by L,  which is the scale
limiting vertical turbulent motions.  However, the Pasquill's categorization
method has validity for 3-min sampling times, as represented by Pasquill's
cfy curves, since large eddies have little effect at this time scale;
the small eddies are more isotropic, even in stable conditions.
                                    96

-------
     A very dubious, but frequent, practice is the application  of Pasquill/



Gifford/Turner (P/G/T) az values to pollutants from tall  stacks.   This



was a major point of criticism in recent rural  model  reviews  by a committee



of atmospheric diffusion specialists (Fox et al .,  1983).   They  felt  that



the system is based too much on surface data and  that the classification



system has a strong bias towards the neutral stability category.   We now



know that, in convective conditions, the locus of maximum concentrations



from near-surface sources lift off the ground at  x = 0.5  ("u/w*)z-j, and



thereafter the surface concentrations reduce very rapidly; this has  a



strong bearing on the interpretation of Pasquill's oz curves  for  "A"



and "B" conditions (Briggs, 1985a).  These curves are based on  observed



jurface concentrations from Prairie Grass surface releases, and they



depend on the assumption that the vertical  concentration  distribution is



Gaussian with its maximum at the surface.  Combined with  the  conservative



tracer assumption, this gives az <* (/xdy)~*.  However, if the locus  of



maximum X elevates, causing /Xdy to reduce rapidly at the surface,



the above assumptions lead to an exaggerated growth in the calculated



cfz.  Reanalyses of the Prairie Grass data using convective scaling



suggest that this is what happened, since the surface concentrations, at



least, behaved very much like those in laboratory tank experiments (Briggs



and McDonald, 1978; Nieuwstadt, 1980; Briggs, 1985a).  The Pasquill  "A" az



curve in x expands much faster than "very unstable" curves derived from



elevated sources, such as the BNL curves.  Little harm is done  if the



P/G/T curves are applied to low sources using a Gaussian  model, since



that is how they were derived, but large distortions can  result if they
                                  97

-------
are applied to very elevated sources.  For instance at about 1.8 km in
"A" conditions, a plume is predicted to be four times as high as it is
wide.  This seems counterintuitive given the current understanding of
diffusion under convective conditions (Lamb, 1979).

Hogstrom's Sigmas and (d9/az)/Uf2 Typing

     Remarkably little attention has been given in diffusion literature
to Hb'gstrb'm's experimental study of diffusion from elevated sources made
two decades ago (Hdgstrbm, 1964).  The data are very useful and his analysis
was detailed, scientific, and full  of significant conclusions, especially
regarding diffusion from elevated sources in stable conditions.

     The experiments were unique in that lateral  and vertical diffusion
were measured optically using photographs of puffs shot from near the
release point, rather than from the usual lateral camera position.
Each puff consisted of a 30 s oil fog release.  The puffs were assumed to
advect at the rate of the mean wind speed measured at source height.
Total diffusion was measured by releasing a succession of puffs over
about 1 h, each one released as the former one faded.  For each experiment,
the standard deviations of the puff center!ines around their mean position,
(jyC and ozc, were determined from the photographs.  In addition, the
mean relative diffusion around each center-line, ayr and azr was estimated
by assuming a bi-Gaussian X distribution and that the visible edge of the
puff represents a constant value of /Xdx (Roberts, 1923); cases of
uneven or variable background were omitted.  Total diffusion for each
experiment was calculated by assuming vectorial addition of center!ine and

                                    98

-------
relative diffusion, e.g., -     jr i


were performed at Agesta with zs = 50 m and at Studsvik with zs = 24 m



and 87 m (both sites are in Sweden).  Conditions were mostly neutral to



very stable, with only a few experiments made in slightly unstable condi-



tions (plumes disappear more rapidly in unstable conditions).





     Five different stability indices were tested for their ability to



order data consisting of the most stability-dependent measure,



ayr ' Ozr at x = 2000 m, for the largest group of experiments, Studsvik



at 87 m.  Plots were made for each index and a line of best fit was drawn



by eye; correlation coefficients were computed using this function.  The



best correlation, 0.95, was obtained with s = 105(m3/C°-s2)(d9/dz)/Uf2,



where Uf is the speed of the "free wind" determined from routine, twice-



daily rawinsondes released nearly 100 km away from this site averaged



through a layer near 500 m.  Uf actually worked better than u measured at



at the site itself.  The lowest correlations were with d"u~/5z instead



of IT, and with d9/3z alone.  In all cases, ae/az was determined only



from the 30-m and 122-m levels of the meteorological mast, so was the same



as A9/(92m).  Thus, Hogstrom's stability index is a form of the Bulk



Richardson number, with s <= A9/Uf2; however, it represents stability near



the height of release better than surface layer stability, so is not



directly related to L or Pasquill categories.





     The components of relative diffusion for the Studsvik 87 m releases



were highly correlated with s.  At distances between 1 and 4 km, a2r



approximated its neutral-condition value divided by (1 + 0.022s).  Similarly,
                                    99

-------
cfyr a^ 1 km approximated its neutral  value divided by (1 + 0.012s).
A very significant conclusion about aye, which can be regarded as
the "meandering" component of diffusion, is that it showed no dependence
on stability.  It grew roughly linearly with distance in the range 1/4 to
4 km.  Another interesting conclusion made by Hogstrom is that the wind
direction shear effect on aye was also almost independent of stability;
the larger shear with increasing stability was compensated by smaller crz.
(He estimated this effects contibution to 
-------
     It is of interest to note the relative importance of azr and 
-------
approximated if the buoyancy flux  F^  =  300  rrr/s^  (typical  for these
sources) and u = 6 m/sec.   Although ambient turbulence or  stable stratifi-
cation are bound to be important  in this  distance range, it seems likely
that growth induced by buoyant  rise is  also a  significant  component of
the TVA 
-------
AT/Az tends to be a function of heat flux only, while u has much
stronger influence on dispersion; thus, no correlation at all  is seen
between AT/Az and measured az or f^dy in unstable conditions
(Weber et al., 1977; Briggs and McDonald, 1978).  During stable conditions,
AT/Az, Weber et al. noted that u*, H*, and L are all  well  correlated
with each other in the surface layer and AT/AZ correlates fairly
well with near-surface vertical diffusion, provided it is measured  through
a substantial layer roughly coinciding with the plume layer.  (The  best
correlations with az were obtained with (22 - zj)  > 5m and zz/zj >  2,
where Z2 and zj are the upper and lower heights of temperature measurement,
respectively).

EVALUATIONS OF SIGMA CURVES AND TYPING SCHEMES

     We find surprisingly few comparisons of diffusion data with rival
schemes for ay, az, and stability classification.   There have  been
dozens of papers demonstrating that different stability schemes disagree
with each other, but only a few papers address the question of which one
is best.  Weil (1977) compared the performance of  different Gaussian
model schemes applied to three power plants, which involves the additional
uncertainties associated with buoyant plume rise.   The Brookhaven ay
and az values worked best, but they were chosen by an algorithm of
Weil's own devising based on wind speed and A9 (this  most  closely relates
to bulk Richardson number classification).  The TVA scheme was the  poorest
performer.  Pasquill's scheme was intermediate in  performance, even though
its basis was in surface source data;  its performance was  greatly improved
by displacing the stability category towards the unstable direction by one

                                    103

-------
category.  Similar results were obtained at all  three power plants,  except
for a few exceptional  cases at Chalk Point; this was the only plant  having
cra measurements, so it seems that  aa classification schemes were
not adequately tested.

     Quantitative tests of the correlation between various  stability
indices and vertical  diffusion from surface sources were made by Weber et
al., (1977).  Only linear correlations were tried, and az was mostly
inferred from surface-arc /Xdy measurements and  the Gaussian plume
assumption, using the Prairie Grass, Green Glow, and NRTS data sets.  The
indices tested were AT, net radiation, aa, ae, IT, the gradient
Richardson number (Ri), and z/L.  For unstable cases, AT and net
radiation did not correlate with az at all.  The wind direction variances
had mixed results with mostly rather weak correlations, but the wind
vanes used in these early experiments may have been poor.  The best
correlations were with Ri and z/L, which are closely related to each
other.  For stable conditions, ae gave the most  consistently good
correlations with az, but Ri, z/L, and AT also gave good corneldtions.
At the farthest distance, 3.2 km,  all correlations fell to 0.5 or less.
The Prairie Grass data set provided the best correlations;  these data
were further used by Rriggs and McDonald (1978), who compared (/Xdy)~l
with various indices and nondimensionalizations.  The single best stability
index was L, but U^/A9 worked almost as well and it is much easier to measure;
1J2, by itself, applied separately to stable and  unstable data categories,
worked slightly less well but still provided good correlation with /Xdy;
this supports Pasquill's scheme for ground-level sources, since it depends

                                   104

-------
most strongly on u.  Measured between 1 and 8 m, A0 also provided good


correlation in stable conditions, but none at all  in unstable conditions.


Rood ordering of the data was obtained with nondimensional  plots involving

   _o
L, uVAS, or, for unstable cases, convective scaling with z^ and w*;


L-scaling had a slight advantage in stable conditions, as did convective


scaling in unstable conditions.




     Comparisons of ay and az schemes against data from a large number


of both surface- and elevated-source experiments were made by Irwin


(1983).  He tested the P/G/T scheme, with classification by Turner's


(1970) method, along with five schemes based on wind direction fluctuations,


with 
-------
must remember that Pasquill's 
-------
of the atmosphere itself; for these 60-min averages in neutral  to



moderately unstable conditions, then, 50% of the /Xdy fell  within about



a factor of 1.?. of the mean and 75% of the observations fell  within about



a factor of 1.5 of the mean prediction; even the "perfect model"  nay



produce little improvement in this scatter.  Three models had insignificant



bias: Hay and Pasquill (1959), Draxler (1976), and the P/G/T method



(Turner, 1970).  The az enhancement due to roughness and urban-enhanced



heat flux recommended by Pasquill  (1976a) resulted in underestimates of /Xdy  .



for this suburban area about a factor of 1.4.  The largest  mean underestimate,



of about a factor of ?., was by a model  taken from Hanna et  al .  (1982);



although Gryning and Lyck did not state the model  specifically, it probably



is the Briggs, "urban" ay and ax curves, which are based on curves



from the St. Louis diffusion experiment (McElroy and Pooler,  1968).  This



suggests that urban cry and az curves give too fast a growth rate  for



suburban locations.  R)r ratios of predicted-to-observed maximum



concentrations for Copenhagen, all  four methods using aa to predict ay



gave the least scatter; three of these also gave insignificant  bias: Hay



and Pasquill  (1959), Hraxler (1976), and Pasquill  (1976).  For  overall



performance in predicting ay, /^dy, and *max, it appears that the Hay



and Pasquill  and the nraxl er methods were clear winners in  this contest.





     To summarize these evaluations, it appears that models based on



ay = aax F(t), with a moderate degree of sophistication in  the



specification of f(t), like Draxler (1976), give the best results so far



for lateral  dispersion.  The nraxl er (1976) formula for az, based on ae



measurements, also appears to be the most dependable method for predicting





                                   107

-------
vertical  dispersion, although  several  similar methods  are  competitive
with it.   This supports the recommendations  of  scientists  at the various
workshops since 1977 that favor characterization  of  
-------
                                SECTION 5

                     DIFFUSION MODELING ALTERNATIVES

     The "conventional" models discussed in Section 4 all  assume a Gaussian
plume and all are mostly empirical.  However, we find that most of the
stability classification schemes have some physical  basis.  The Brookhaven
03 classification relates directly to ay at short distances,  with
the relationship degrading very slowly with distance.  It  relates  well
to az when conditions are nearly isotropic, i.e., lateral  and vertical
turbulence velocities and length scales are about the same;  this condition
holds best for elevated sources in neutal  and unstable conditions.
Pasquill's classification based on IT and insolation or cloudiness  is
crudely related to the Obukhov length, L,  the most important  length scale
affecting vertical turbulence in the lower part of the ABL.   Thus, it
correlates best with az from surface or low sources.  It correlates
fairly well  with short-term averages (3 min)  of ay,  since  the small
eddies affecting this time scale are more isotropic, but considerably
underestimates typical  1-hr averages of ay.  Hogstrom's (ae/az/Uf2)
classification is a bulk Richardson number, which is strongly related  to
L if ae/az is measured near the surface.  Measured near 100  m,  it
is a good index of the relative vertical  stability at heights of elevated
releases.  For neutral  to stable conditions it correlated  very  well
with az, both-relative and total, and with ayr, the relative  part  of
lateral diffusion (this was accomplished by smaller eddies which are more
isotropic);  however, it did not correlate with the total (time-averaged)
                                    10Q

-------
cry.  TVA's index AT/Az and  variations  on  it  like  AT,  A9,  or  a0/az
are very weak indicators of the atmosphere's diffusive capacity, because
wind speed has a stronger effect than  the static  stability.

     Thus, these classification schemes,  except for AT/AZ, each have
some areas of success as diffusion  indicators, and none  is "universal".
The Gaussian plume assumption has been a  useful workhorse, but we  have
identified some diffusion conditions during  which it  leads us astray, as
the atmosphere takes on more complex behavior.  The worst aspect of the
models reviewed in the previous section is that each  was  built on  a
very small set of diffusion measurements, measurements that  had quite a
few shortcomings; since the curves  are empirical, there  is simply  no
basis for extrapolating them to much larger  x or  using them  for all types
of sources and terrains.  Here is where more scientific  models become
practical tools.  If they can be validated for enough parameter combinations
within their range of applicability, there is some basis  for confidence
in extrapolations and interpolations.

     Over homogeneous terrain and under near steady conditions, typical
Gaussian models predict peak concentrations  within about  a factor  of 2.
There is widespread belief that this kind of performance can be
improved upon by models which incorporate adequate physics,  if they
can be supported by adequate meteorological  measurements. This section
describes six general methods of approach that have had  proven success  in
some areas of testing; none of them appear to he  the  universal solution
for all diffusion problems, but a combination of  approaches  should be able
to provide the large step in modeling  improvement that  is greatly  needed.

                                    110

-------
SURFACE LAYER SIMILARITY MODELS

     In the early 1960's, there were a number of attempts  to  rationalize
atmospheric dispersion rates in terms of fundamental  boundary layer
parameters that were discussed in Sec. 2.   The inspiration for this  work
was the so-called "similarity theory" of Monin and  Obukhov (1954), which
was becoming a successful tool for describing profiles  in  the surface
layer of wind speed, temperature, humidity,  and vertical turbulence
quantities.  This theory assumes that vertical  turbulence  near the surface
is a function of height and velocity and length scales  based  on surface
fluxes of heat (or buoyancy) and momentum  (these scales were  defined in
Sec. 2).  Thus aw is assumed to be u* times  a function  of  normalized
height, z/L, on dimensional  grounds.  Vertical  gradients of u", T, water
vapor, etc., are assumed equal to their fluxes at the surface divided  by
(u*z), times some function of z/L.  These  functions must be determined
empirically and may be different for different quantities.  The usefulness
of this approach, essentially a dimensional  analysis, was  shown decisively
by the results of the 1968 Kansas field experiment  (Businger  et al.,
1971; Kaimal et al., 1972).

     If the theory describes the structure of turbulence near the ground,
then it should also be capable of describing turbulent  diffusion of
passive tracers released at  the surface; this was tried in  1957 by Kazansky
and Monin for describing the shape of smoke  plumes  (see Monin, 1959).
Rifford (1962) reviewed progress with similarity theory up to that time
and developed predictions of diffusion from  the surface starting with  the
                                    111

-------
equation
          cfz/dt = cu*<|)(I/L),                                    (5.1)

where T is the mean height of the concentration, c is a dimension!ess
constant, and  is a universal  function of z/L.  Other early examples of
the application of Monin-Obukhov similarity theory to diffusion are the
papers by Cermak (1963), which used mostly wind tunnel diffusion observa-
tions, and by Panofsky and Prasad (1965).  The latter authors noted that
lateral dispersion does not fit the theory as well as vertical dispersion;
this observation has been reconfirmed by many others since then, and
seems to follow from the fact that lateral turbulence is influenced by z-j
at day and by topographically influenced meanders at night.  Thus,  surface
layer similarity theory omits significant factors affecting lateral
dispersion.

     The subject lay in a relatively dormant state, stymied in part by
the difficulty of measuring L, until after analyses of the 1968 Kansas
wind and temperature profile data were published.  Chaudhry and Meroney
(1973) made use of the Rusinger et al., (1971) relationship between
Richardson number, R^ = (g/9) (ae/az)/(a"u7az)2, and L, to make improved
estimates of L for the Prairie Grass data.  They also advanced
the theoretical argument that $ = ^~^, where ^ is the dimension!ess
temperature gradient given by ku*ae~/az =  (w1 e'/z)h, and 9 is potential
temperature.  The argument was made using an eddy diffusivity approach.
The simplest way to do this is to recast  Gifford's equation as
                                    112

-------
          dz2/dt = 2bu*z = 2K ,                                  (5.2)

where K is an eddy diffusivity for passive material  diffusing from the
surface.  Chaudry and Meroney argued that this diffusion mechanism is
similar to that for heat flux from the surface, so K « Kn, the eddy
diffusivity for heat; this is defined by w'e'  = Khd9/dz.  Substitution
of the profile law gives K^ = ku^z/^ and, therefore, <{> = n  •
These authors were able to demonstrate that this hypothesis leads to a
much better fit with the Prairie Grass data than does Monin's (1959)
hypothesis for <|>, which is too weakly dependent on L.  They used the
observations at x = 100 m, the only arc having samplers on towers; this
permitted direct determination of "z (no Gaussian assumption required).

     Horst (1979) carried this analysis further using all  the arc distances
of the Prairie Grass experiment, ranging from 50 to  800 m, but he had to
assume a vertical distribution function to relate 7 to /xdy at the
surface, since only near-surface samplers were used.  Based on Prairie
Grass profiles at x = 100 m and other evidence, he chose a modified Gaussian
profile assumption: X-l « exp(z/b"z")r with r = 1.5.  He then showed that
K = Kn leads to a satisfactory fit with observations, much better than
K = Km, which had been suggested by Pasquill.  He also compared  the predic-
tion with /Xdy observed at x = 3.2 km at the National Reactor Testing
Station (Islitzer and Dumbauld, 1963), finding more  scatter but  definite
skill using K = Kn.  Van Ulden (1978) also used K =  Kn in a similar analysis
for arc distances x = 50, 200, and 800 m; further, he made an allowance
for the non-zero height of the Prairie Grass samplers, z = 1.5 m.  In a

                                    113

-------
recent re-analysis of the Prairie Grass data, Gryning, van 111 den and
Larsen (1983) again demonstrate a good fit of the observations using
surface layer similarity and K = K^, with poorer results using Km.  They
find improved fits when they include an estimate of deposition in their
model, as S02 is now known to deposit on vegetation.

     Two recent papers attempt to simplify application of surface layer
similarity to vertical diffusion from the surface.  Both of these utilize
the Prairie Grass data.  Venkatram (1982) argues that both the neutral
and very stable asymptotes for /Xdy at the surface are independent of
u".  (To simplify discussion, Twill be used in place of az and cloud
height h, since they are all closely related.)  For surface, concentrations
from surfaces sources f^dy « 0/(TT"z).  As T « u*t « (u*/u)x for the neutral
case, the U dependence cancels out: f-^dy « 0/(u*x).  For extreme stability,
with ~z~ » L, Venkatram assumed u" « u*z/L and a constant eddy diffusivity,
K « u*L.  Substituting this into 7« (Kt)1/2, with t * x/u" and the above
IT, we get T« L2/3xl/3 an(^ j-^y « 0/(u*L^/^x^/3).  (jne existence of this
asymptote in the ARL can be questioned in light of new knowledge about  SBLs,
because h/L never gets large enough.)  Venkatram used profile determinations
of L and u* to nondimensionalize the data, showing that the two asymptotes
fit the neutral-stable Prairie Grass data rather well, except for some
widely scattered points at larger x/L values.  Briggs (1982b)  derived the
same asymptotic relationships, arguing that there is a very slight dependence
on ~z/z0 at intermediate values of x/L, which can be neglected.  He also
extended this approach, using cruder approximations, into the convective
regime, and suggested analytical expressions for Q/(u*x f-^dy)  versus

                                    114

-------
x/L.  The numerical  constants in his  analysis  differ  from  Venkatram's



asymptotes by about  20%, partly because Briggs adjusted  for  the  non-equality



of the source and receptor heights in the experiment.  Venkatram (1980)



also demonstrated a  strong correspondence between  u*, L  and 7j  in stable



conditions, which could be used to simplify application  of these results.





     It can be concluded that surface layer similarity models  can  give



very good results for vertical  diffusion from  near-surface sources at



distances up to 1 or 2 km.  The theory does not contain  enough information



about the variation  of turbulence with height  to give good results for



sources at more than 1/10 the ABL height.  Except  for averaging  times of



less than 10 min, for which diffusion is accomplished mostly by  small,



isotropic eddies, this approach can only predict the  minimal (rather than



typical)  lateral  diffusion.  This is  because it contains no  physics relating



to large horizontal  eddies.








GRADIENT TRANSFER AMD HIGHER ORDER CLOSURE MODELS





     One of the earliest and most frequently used  approaches to  prediction



of diffusion in the  atmosphere is the gradient transfer  assumption, or



"eddy diffusivity" theory, or "K theory".  It  assumes that the local



turbulent flux of a  material  is directly proportional  to the mean  local



gradient of that  material, and is aligned into that gradient (towards



lower concentration).  The constant of proportionality,  K, is  called the



eddy diffusivity  and is supposed to be a property  of  the flow, rather



than of the substance.  Thus, in a steady state situation, K should be
                                   115

-------
only a function of position  (x, y,  z)  and  boundary conditions.   For just
the vertical  component of turbulent flux of  substance C, we can  write the
gradient transfer assumption quite  simply
               = -K aC/az   .                                       (5.3)
If there is no mean advection,  then  the  rate of change of concentration
at each point is given by

          aiU/at = - owrc"I7oz =  o(KoC/az)/dz.                     (5.4)

     This approach is tempting  in  its  simplicity,  and builds on  a long
history of mathematical  treatments of  molecular diffusion.   In that
context, K is just 1/3 of the mean molecular speed times the mean free
path length between molecular collisions.   The most direct analog for
turbulent diffusion is K proportional  to the mean  eddy velocity  times a
mean eddy diameter, or turbulence  length scale (such as aw times the
Lagrangian time scale).   However,  this analogy is  often inappropriate,
and there are only certain situations  in atmospheric diffusion for  which
K theory is approximately valid.   Pasquill  (1974,  section 3.1) has  discussed
this at some length.  The question of  validity has to do with the difference
between diffusion and dispersion,  terms  that usually are used as if they
were interchangable.  "Diffuse" in its Latin root  means to pour  or  to
spread out, and seems to connote  smearing.  Thus,  local flux proportional
to local gradient seems a good  description  of "diffusion".   "Disperse" in
its Latin root means to scatter,  which better describes what big eddies
do to small plumes - wave them  this  way  and that.  The relative  size of
the eddies and the plume or puff  is  the  key.  The  idea that  local mean

                                   116

-------
concentration gradients have anything  to  do  with  "dispersion" of small
plumes by large eddies seems absurd,  if we try  to visualize this process.
Yet K theory has been indiscriminately applied  to dispersive situations.
On the other hand, it is intuitively  a good  assumption to apply to "diffu-
sion" of large plumes by small  eddies. Pushed  to its limit, we might
expect this assumption to hold  as long as mean  concentration gradients do
not change drastically on the scale of the most effective eddies.

     There are three cases of atmospheric diffusion which might be
describable using ordinary K theory,  as described above.  The diffusion
of material released near the surface  is  one.   This application actually
occupies the fuzzy region between validity and  non-validity, because the
most effective eddies doing the upward mixing are those of the order of
the plume thickness or slightly larger; it can  be viewed as an approach
that might work, especially when the  eddy structure maintains similarity
relative to the height above the ground.  This  condition is most closely
met in the lower part of the NRL.  Here,  K « u*z  has proven validity, as
shown by many of the papers referenced in the last subsection.  The SBL
is also a reasonable candidate  for K theory, because the eddy size becomes
gradually restricted by the scale L as z/L becomes larger and vertical
diffusion becomes more of a diffusive  process (eddies small compared to
scale of variations in mean gradients).   The unstable case is just the
opposite; the higher the material  mixes,  the larger the dominate eddy
size until the dispersant is drawn into thermals  which lift the material
all the way to z-j.  Here, the eddy diffusivity  assumption is inappropriate,
At best, it might apply as long as the maximum mean concentration remains

                                   117

-------
on the ground and the vertical  distribution  is  quasi-Gaussian; there  is

growing evidence that this condition holds only for  t  <  0.5  z-j/w*, or


-------
only after the plume thickness  is  of the  order  of  L  or A, the "local"
Obukhov length defined by Nieuwstadt (1984a,b).  The stable  stratification
limits vertical mixing motions  to  a  depth comparable to  L or A, except
near the ground where z is a more  limiting factor.   Observations of
layers of near-constant a¥/az and  aU/az in stable, dry boundary layers
above z = L indicate that Km =  Kn  =  0.08u*L, where Km and K^ are eddy
diffusivities for momentum and  heat, respectively.   We would expect K for
the diffusion of passive matter to be about the  same. Ordinary K theory
has no validity for elevated plumes  in neutral  and unstable  conditions,
because the plume thickness is  then  smaller than the most active eddies,
which are dispersive rather than diffusive.

     The third possible case for limited  validity for K  theory is that of
lateral diffusion in convective conditions at distances  beyond several
(U/w*)zj.  By this time the plume  width is several z-j, larger than the
dominant cell size (about 1.5 z-j).  The asymptote  for oy suggested by
Deardorff and Willis (1975), based on their laboratory tank  experiments,
is 0.53 (w*zit)1/2, which would result from K =  0.14 w*z^ (note that, in
this case, convective scaling with w* and z-j applies, rather than surface
layer scaling with u* and L).  However, in the  atmosphere, where wind
shear plays a role and there are no  horizontal  restrictions  like the tank
walls, there are horizontal  eddies or rolls of much  larger scale than zj
in the line of mean wind direction.   Some indication of  this is given
in the report on the Minnesota  convective boundary layer experiment
(Kaimal, et. al., 1976).  Very  frequently, wind  direction shifts are
observed to occur on time scales of  about one hour on convective days

                                  119

-------
(personal  observation).   Then  continuous  plumes are  still "dispersing" on
these time scales, and a  ay «  x*'^  or  « t*'*  regime  may be  short-lived,
about 1 hr at most.   Lateral dispersion observations in neutral and
unstable conditions  have  mostly yielded best  fit  power laws between x^-?
and xO*9,  in contrast to  the x^«5 asymptote given by K theory.  This suggests
that plumes are usually influenced  by  larger  and  larger horizontal eddies,
contrary to K theory assumptions.

     One unrealistic consequence of any diffusivity  model is that non-
zero concentrations  are predicted out  to  infinity, implying an  infinite
velocity of propagation.   An example is the Gaussian distribution, which
results when K is assumed constant.  Although these  far-flung concentra-
tions are quite insignificant  in magnitude, the physical unreality of
this aspect of such  models is  discomforting.  An  alternative approach,
called the telegraph equation, has  been developed by Monin  and  Yaglom
(1965).  Essentially, it  restricts  the rate of change of flux by adding
the term t dw'C'/at  to w'C1 in the  flux-gradient  equation,  where T is
a limiting time scale.  The result  is  that there  is  a limited velocity of
propagation proportional  to (K/-c)l/2,  and the flanks of the concentration
distribution have a  cutoff. Closer to the center of the distribution
there is a slight change, but  not enough  to have  any practical  consequence
(see Pasquill 1974,  Fig.  3.2).

     The most serious 1 imitation'of simple eddy diffusivity approaches is
that non-local effects, such as  those brought about  by large turbulent
eddies, are neglected.  The local turbulent flux  of  concentration  is
                                   120

-------
assumed to be a function of only the local  gradient  of  concentration.



When large eddies bring in fluid containing zero  concentration, which



therefore comes from regions of zero gradient,  this  assumption of depend-



ence only on local  gradient is obviously  wrong.   Several  attempts have



been made to construct models which avoid this  deficiency.   One involves



a more sophisticated version of K-theory  called spectral  diffusivity



(Christensen and Prahm, 197fi; Rerkowicz and Prahm, 1980).   The concentra-



tion field is described in terms of Fourier components, with  c(k) being



the amplitude of the Fourier mode of concentration with wavenumber  k



(waves per unit length).  The eddy diffusivity  is assumed to  be a function



of k; thus, K(k) can account for the variability  of  turbulent energy in



each range of eddy size.  The eddy diffusivity  equation for a Fourier



mode becomes ac(k)/5t = -k^K(k)c(k).  In  a  sense, this model  is



non-local , since each Fourier component represents concentration spanning



the whole dominion under consideration.  A  particular form  that has been



proposed for K is KQ/(1 + Bk^'^).  This gives a constant  K, with the same



result as ordinary diffusivity theory,  for  very large wavelengths (small



k), i.e., for variations in concentration on a  scale large  compared to



energetic turbulent eddies.  For small  wavelengths K <*  e^/^k-^/^,



where e is the turbulent energy dissipation rate; this  is the asymptote



for the "inertia!  subrange" of turbulent  eddies driven  by the larger



eddies.  Prahm, Rerkowicz and Christensen (1979)  applied  this theory to



time averaged plume concentrations, adding  the  effect of  "meandering" of



the plume center! ine by means of a "spectral phase diffusivity coefficient'



that is made to fit statistical  theory  predictions for  large  and small
                                   121

-------
asymptotes of k and t.  The  resulting  plume  growth  is, not  surprisingly,
in agreement with the main results  of  statistical theory, but the  shape
of a narrow plume is different  from Gaussian.   There  is no  experimental
confirmation of this shape as yet,  and its practical  significance  seems
doubtful.  The theory has primarily been  applied to two dimensional global
transport models, with eddy  sizes  ranging up to several thousand kilometers.

     Another type of non-local  K theory was  suggested by Voloshchuk
(1976).  In this model, K is both  a function of z and of a  displacement
distance, 1.  If K(z,l) has  a non-zero value only for 1 = 0, the equation
reduces to the telegraph equation  for  the local case, like  that developed
by Monin and Yaglom (1965).  While this method  appears to have potential
for accounting for non-local concentration gradients, the actual specifi-
cation of K(z,l) for various stabilities  might  be quite difficult, con-
sidering the effort that has been  spent on trying to  specify K for a
single variable, z.

     A new mechanism for turbulent transfer  has been developed by  Stull
(1984) and applied in one form  by  Stull and  Hasegawa  (1984).  This mechanism
is called transilient turbulence theory,  from the Latin verb "transilire"
meaning to jump over or leap across.  .Stull  introduces the  mechanism  in  a
finite difference framework, then  goes to a  continuous representation.
In either form, the mechanism  adjusts  the concentration at  a point by
considering the concentrations  at  all  neighboring points which can be
influenced by the flow.  For example,  consider  only vertical transfer,
for which the concentration  at  a point is influenced  by all of the points
                                   122

-------
in the vertical.  The formal  expressions for the  continuous  case  are
similar to the equations for radiative transfer in  an  absorbing,  but
nonscattering medium.  Fiedler (1984)  independently developed  a similar
concept, but started from a different  perspective,  namely  the  spectral
diffusivity concept of Berkowicz and Prahm (1980).   Stull's  approach  was
to visualize turbulence as consisting  of a spectrum of eddy  sizes each
contributing to the turbulent transfer.  The interesting feature  of the
mechanism is that the coefficients defining the transfer process  can  be
specified ahead of time, or can be made responsive  to  changing flow
conditions.  The coefficients can be made to represent any type turbulent
flow.  The same model with different coefficients can  represent diffusion
in a CRL or in a SRL.  Further, Stull  points out  that  the  transilient
approach can also be used in combination with a second order closure
model to close the higher moments.  This new development has much promise.

     K theory is considered "first-order" closure,  meaning that closure
assumptions are applied to the second  moments of  turbulent quantities,
such as w'C1 and w'u1, that relate them to first  order quantities like (T
and "u.  About 10 years ago, work accelerated on "second-order" closure
models for turbulent boundary layers and diffusion, carrying this approach
to a new level.  These models begin with the exact  equations for  second
moment terms like concentration flux or momentum  flux, which are  needed
to solve for a"C/at or 5~u7dt.  As in all  turbulence  models, the exact
equations contain terms with new unknowns, and always  more unknowns than
equations, so closure assumptions are  still  necessary.  In this case, the
new unknowns are third moments,  such  as  u1w'Cr ,  and  assumptions must  be

                                   123

-------
made relating them to terms  already  being  computed,  i.e., second moments
or mean quantities.   These assumptions may be just as questionable as
assumptions made about K in  first  order  theory, and  the third moment
terms may be modeled quite crudely;  the  advantage is that second moment
terms like turbulent fluxes  do not depend  on these assumptions alone, but
also depend on exact terms;  thus,  they are less susceptible to error on
account of the failure of a  closure  assumption.  Nevertheless, the output
of second-order closure models depends on  the performance of the closure
assumptions, and these should  be scrutinized for proper physical charac-
teristics and tested against data  in a variety of flow situations.  One
shortcoming that second order  closure models so far  share with K theory
is that the closures are local; thus, the  effect of  large eddies on the
third moment terms cannot be handled properly.

     One of the advantages of  second-order closure modeling of diffusion
is that buoyancy-driven vertical mixing  is included; this is done with
the term (g/e")9'C' in the equation for aw'C'/5t.  Whereas gradient
transfer theory fails to model the lifting of the mode of maximum
concentration from a surface source  in convective conditions, Lewellen
and Teske (1976) demonstrated  a second-order closure model that does
this.  The qualitative features of /Xdy  that were observed by Willis
and Deardorff (1976) in the  laboratory are reproduced quite well, but the
surface concentrations do not  reduce quite as fast in this numerical
model.  Pasquill (1978) also reports some  success with second-order
models in the case of vertical diffusion from an elevated source, another
case which cannot be handled by conventional K theory.  Second-order

                                  124

-------
closure models have also been used to study the  SBL,  as  reported  in  Sec.  2



Wyngaard (1975) and Rao and Snodgrass (1978)  were  able to  reproduce  in



their models the approximate behavior with  time  of quantities  like heat



flux, friction velocity, and boundary layer height, h, which is of vital



relevance to diffusion modeling.   It  evolves  rapidly  after sunset, and a



low or medium level source could  be below or  above h; this determines



whether or not vertical  diffusion occurs  at all.





CONVECTIVE SCALING MODELS





     A very active area of research at present is  the application of



convective scaling to daytime diffusion models.  The  idea  was  first



applied to atmospheric boundary layer turbulence by Deardorff  (1970b),



who first tried this scaling in the context of a large eddy simulation



model of neutral  and unstable boundary layers; this model  will be discussed



in the next subsection.   Subsequent successes using this scaling  in the



context of CBL turbulence measurements were reviewed  in  Sec. 2.



For  |zj/L|  equal  to 4.5 and larger, Deardorff (1972)  found that w* scales



the magnitude of turbulent velocities better than  u*.  Typical daytime



values of convective and surface  layer scaling quantities  are: z-j ~ 1000 m,



u* ~ 0.4 m/s, and H* ~ 0.004 m2/s3.  This gives  L	40  m  and  |z.j/L| ~ 25,



which is quite convective according to Deardorff's results.  The criterion



|z-j/L| > 4.5 would be met even with l/5th the above values of  z-j or H*,



or with these values unchanged and u* = 0.7 m/s; this typically corresponds



to wind speeds of about  6 m/s near the ground; U = 6 m/s with  moderate



insolation corresponds with stability class "C-D"  in  Pasquill's table,
                                   125

-------
between "neutral"  and "slightly unstable".   We can conclude that convective
scaling would be appropriate for most  of the daytime on most days at most
sites on land.  One condition that  produces  sustained maximum average
ground concentrations near large power plants is  a sunny summer day with
low wind speed;  these are very convective conditions, with  |zj/L| much
larger than 100.  This is a case which should certainly benefit from the
convective scaling models recently  under development.

     Application of the convective  scaling idea to diffusion began with
the now-famous convective tank experiments of Willis and Deardorff (1974,
1976, 1978, 1981,  1983).  These experiments  and their controversial results
were discussed in  Section 3 under "Laboratory Experiments"„  These results
were unexpected and cast much doubt on the ability of conventional Gaussian
models to accurately model  daytime  diffusion.  Naturally, there was much
interest in verifying these results outside  the modeling tank.  The
numerical experiments of Lamb, using Deardorff's  (1972) model, was a
first attempt.  The results were qualitatively quite similar to those of
the tank experiments, with nearly the  same descent rate of the maximum
concentrations from elevated releases, but with slightly less decrease in
surface concentrations at the larger distances.   Utilizing field data
that were already available, both Briggs and McOonald  (1978) and Nieuwstadt
(1980) demonstrated that the "Prairie  Grass" surface values of /Xdy
were quite similar to those for the lowest release in the tank when
nondimensionalized using convective scaling. The field experiments
showed more scatter; this was attributed to  the short release time, only
10 minutes, which is inadequate to  sample many eddies  (a typical "eddy

                                   126

-------
passage time" is 1.5 z-j/u, which can range from 2 to 30 minutes or more).
The Prairie Grass ay/Zi values were about 40% larger than those in
the tank; part of this may be due to the absence of horizontal  motion
restriction in the atmosphere.

     It is ironic that field experiments done in the two decades following
Prairie Grass did not include sufficient measurements to determine z-j, a
parameter that turns out to be indispensible for modeling diffusion in
convective conditions (the 1956 experiment included aircraft soundings of
temperature and humidity, as well as rawinsondes, so z-j was easy to
determine in retrospect).  A recent field experiment designed to test
convective scaling on diffusion in the atmosphere, especially for compari-
son with the tank experiments, was carried out at the Roulder Atmospheric
Observatory in August and September of 1982 and 1983; this was  a coopera-
tive experiment between the U.S. EPA and NOAA's Wave Propagation Laboratory.
It is described in Section 3.  Some preliminary results have been reported
by Moninger et. al., (1983), Eberhard et al. (1985) and by Briggs (1985a).
In the first paper, a striking resemblence was seen for nondimensionalized
j"*dy between a 29-min period with z^/z^ - 0.45 and the tank experiment
for zs/Zi = 0.49.  However, it is not yet known how much the 0.3 m/s fall
speed of the chaff tracer distorts the distributions (this concern motivated
the simultaneous side-by-side releases with oil fog).  In Rriggs (1985a),
five periods of chaff distribution were analyzed in comparisons with the
tank and numerical  experiments and three other field experiments including:
the most convective cases from Project Prairie Grass, the five  most
convective 30-min periods of SF^ releases at Cabauw, the Netherlands,

-------
which really were only marginally convective (van Duran  and  Nieuwstadt,
1980), and, finally, some SO;? measurements at the Morgantown power plant
(Weil, 1979).  The latter observations were screened to  remove buoyancy-
dominated cases.  The results of these comparisons show  a good deal  of
consistency, in convective scaling.   All  experiments support the tank-
numerical cjy/z-j curves, except for two BAD periods with  very large wind
direction shear; these had about double the ay.   The near-surface z-jIT/xdy/Q
for one of the surface chaff releases dropped to a value about 30% lower
than the tank minimum, but all  the remaining data were in better agreement.
The most significant result was for  the maximum  surface  value of /xdy
from elevated sources; there was a definite trend with increasing zs/z-j ,
with Cy1 = zs"u/xdy/0 = 0.48(1 + 2zs/Zi).   Conventional Gaussian models
give Cy1 = 0.48, regardless of source height. These maxima  were all
found in the neighborhood of x = 2 u" zs/w*.

     Another recent confirmation of  convective scaling validity comes
from photographic thickness versus distance of plumes from three power plants
in Australia (Carras and Williams, 1983).  Good  agreement with az/z-j  in
Lamb's numerical experiments is found for the less buoyant plumes
(F^ < 300 nr/s^).  Further testing of Willis and Deardorff's convective
scaling results was recently accomplished in a wind tunnel,  at Colorado
State University (Poreh and Cermak,  1984).  This was described in Section
3, under "Laboratory Experiments."  Rather good  agreement with the tank
results was obtained, in spite of lack of complete development of convective
turbulence intensity and other restrictions that accompany seal ing-down
to wind tunnel size.

                                   128

-------
    At present most theoreticians seem convinced that convective scaling
is the best way to go for modeling daytime, non-neutral  diffusion ("the best
thing since sliced bread" was heard in this context at a recent conference).
Many modelers have been fitting the experimental results with analytical
formulas for 
-------
LARGE EDDY SIMULATION MODELS

     Another numerical  modeling  approach  to boundary layer turbulence
that is really quite different  from  second order  closure models  has been
called the "subgrid closure"  or  large eddy simulation  (LES) model.  This
is the approach first applied long ago to prediction of long-range global
weather patterns (Smagorinski,  1963), then further developed for the CBL
by Deardorff (1970).  The equations  for the motions of fluids are believed
to be complete and can be solved exactly  for many non-turbulent  flows.
The "closure problem" arises  with turbulent flows because the equations
can be solved exactly only if initial  and boundary conditions are known
exactly and we attempt to calculate  every detail  of the motion,  down to
the smallest eddy.  This, of  course, exceeds the  computational capacity
of any computer so far envisioned (unless we regard wind tunnels and
water channels as analog computers!).  For tractability, we have to
introduce some kind of averaging, and this always introduces some new
unknown, such as entrainment  velocity or  shear  stress. Some assumption
must be made about how the new  unknown(s) relate  to quantities already
computed to "close" the problem  and  make  a solution possible.   In the LES
modeling approach the domain  of  computation is  divided up into grid cells
and the exact, or "primitive",  equations  are used to solve for the mean
velocity, temperature, water  vapor content, etc., of each grid cell for
each time step.  Thus, the details of the large eddy turbulent motions
are actually computed in time and space;  this gives a  "picture"  of the
turbulence, if desired.  Patterns of mean quantities are obtained by
applying time averaging, just as would be done  in a field or laboratory

                                   130

-------
experiment.  This contrasts to  higher  order closure models, which predict
mean patterns but none of the details  of motion.   What is lacking, however,
is exact computation of the way neighboring grid cells interact with each
other through turbulent exchange of momentum, heat, water vapor, etc., by
subgrid scale turbulent eddies.  The closure problem is still there, but
in this approach it  has been pushed down to the scale of the grid cells
used; this follows because bulk averaging  is done  on this scale with
nothing known about  the details of motion  within the grid cells.

     One advantage of this approach is that the structure of fine scale
turbulence is simpler than that of the large scales where turbulent energy
is produced.  The small  eddies  are more isotropic  and are driven only by
the "inertial" forces induced by the larger eddies.  This is called the
"inertia!  subrange"  of turbulence, and turbulent velocity spectra follow
a simple -5/3 power  law with frequency or  wave number in this regime.
This ought to make the closure  problem simpler on  this scale, with greater
universality of successful  solutions.  The closure that has been used is
a rather primitive eddy diffusivity assumption; the rate of subgrid
turbulent  exchange between cells is assumed proportional  to the difference
between the cells of the quantity under consideration.  This seems a
"safe" use of K theory because  if the grid cells are small enough, the
gradients  of mean quantities do not vary much on that scale.  K theory is
used only  to account for exchange done by  eddies of the grid cell  size
and smaller.
                                  131

-------
     The eddy diffusivity for subgrid  exchange  used  by  Smagorinski  (1963),
Lilly (1967), and Deardorff (1972)  is  based  on  the cell  size  and  velocity
differences between adjacent cells.   In  effect, it is assumed that  the
subgrid turbulence reacts instantly to changes  in  local  velocity  gradi-
ents.  More specifically, the subgrid  K  is taken to  be  proportional  to
the square of the grid size times the  local  deformation rate, which is
always positive.  The constant of proportionality  is chosen to produce
flux-gradient relationships in agreement with those  derived from  careful
field experiments, like the 1968 Kansas  program.  Deardorff (197?)  uses
an eddy diffusivity for heat, Kh, three  times as large  as  the one for momen-
tum, Km (except for the layer of grid  cells  adjacent to the surface); K^
and Km are different because of the pressure-velocity correlation.   Pres-
sure forces affect momentum exchange but not heat  exchange.   Presumedly,
this applies to material exchange also,  so on a subgrid scale the turbu-
lent exchange of concentrations between  cells should be described by Kn.

     Deardorff (1972, 1974) first applied this  type  of  model  to ARLs for
neutral and convective cases.  Many of these results were  mentioned in
the previous subsection on convective  scaling models, since they  led to
his convective scaling suggestion.  One  very instructive product  of this
model, not available with other models,  is the  mapping  of  fields  of turbulent
quantities computed for any given grid plane and time step.   You  can see
the eddies, so to speak.  It must be remembered that these are only model
predictions, however, and it must be asked how  well  the model simulates
what is known about the ARLs.  Heardorff's model results have compared
very well with aircraft and tethered balloon measurements  for convective

                                   132

-------
cases.  The greatest shortfall  has been in the magnitude of computed
lateral turbulent velocities.  This may be partly due to the limited
width of the domain of computation, 2 Zj  (the edges  of the  domain  are  not
treated like solid walls; cyclic boundary conditions are used,  so  whatever
advects in or out of one wall is balanced by advection through  the opposite
wall).  The model is not capable of computing subgrid details like entrain-
ment at the top of the mixing layer;  instead, a "solid"  lid  at  z = z-j  was
used, with w1 = 0 there.

     Deardorff's model  output for velocity fields was used  directly by
Lamb (1978, 1979) for numerical  diffusion experiments.   In  effect, Lamb
released thousands of numerical  "particles"  into the already-computed
turbulent velocity fields, computed their trajectories,  and  counted them
over grid volumes to get average concentration fields.   He  also added  to
the particle velocities a random subgrid  scale velocity  component, scaled
by subgrid-scale turbulent energy.  The results for  release  heights near
the surface, at z « z-j/4, and at z =  z^/2 in convective  cases are  in
satisfactory agreement with Willis and Deardorff's modeling  tank results,
as was discussed in the previous convective  scaling  subsection.  They  agree
especially on the "unorthodox"  qualitative behaviors of  the  surface and
elevated plumes.

    Large eddy models are at their best in large eddy turbulence,  which
is why the convective case has  been the most explored.   Deardorff  also
modeled the NBL.  In principal,  this  type of model can also  be  applied to
SBLs.  A much larger computational  capacity  will  be  needed,  however,
                                   133

-------
because the largest eddies (for vertical  turbulence,  at  least)  are much
smaller.  The nocturnal  mixing depth typically  goes to  10  L,  and  eddy
size is limited by L, rather than  z-,-.  Instead  of  a computational domain
20 cells high, as Deardorff used,  at least  100  are needed.  Much  smaller
time steps would be needed also because the rotation  period for the most
energetic eddies scales  to L/u* instead of  z-j/w*,  about  an order  of
magnitude shorter time period.  If Brost and  Wyngaard (1978)  are  correct,
the model would have to  be run a much longer  time  because  even  on slight
terrain slopes a steady  state is not achieved during  a  night.   In contrast,
the CRL approaches steady state in its turbulent properties after a few
z-j/w*, typically 0.5 to  1 hr.

STATISTICAL MODELS

     It was realized long ago that if enough  were  known  about the
statistics of turbulence and the mean flow  field,  it  should be  possible
to calculate the dispersion of passive substances  resulting from  the
turbulence.  The formal  foundation for this calculation  was laid  by G.  I.
Taylor in 1921 and is summarized by Pasquill  (1974, see  section 3.4).   Its
main limitation is that  it applies only to  homogeneous  turbulence,  i.e.,
the dispersant is assumed not to spread into  a  region where the statistics
of turbulence are different from the origin.  Thus, it  applies  best to
horizontal diffusion over terrain  flat and  homogeneous  enough to  not
cause large variations in the lateral turbulent velocity statistics.   It
can be applied to vertical diffusion close  to an  elevated  source, but  it
becomes mathematically unsound as the material  spreads  and approaches  the
                                   134

-------
ground, since vertical  turbulence statistics vary strongly with  height
near the ground.  This  is not to say that this  application should  not be
tried; with appropriate modifications,  useful  results  might be obtained,
as is proven by data comparisons with the "bouncing  ball"  models mentioned
in the subsection on convective scaling models  (these  models are based  on
turbulence statistics at just the source height).  The statistical
approach of Taylor has  been extended and adapted  for the case of vertical
diffusion from a ground source (Hunt and Weber, 1979)  by separating
the motion into mean rise velocity,  dT/dt, and  fluctuations about  this
mean and using the fact that aw  does not vary much  with height.

     The statistical  theory of Taylor and the simplified approach  to it
developed by Hay and Pasquill (1959)  need a capsule  review here because
they are so fundamental  to all  subsequent work with  statistical models.
Taylor's theory, as applied to turbulent diffusion,  begins by describing
the displacement of a single particle in terms of  its  time history of
velocity, assuming that it follows the  motion of  a single  parcel of fluid.
This is the Lagrangian  velocity, and  all  statistics  of such velocities
are called Lagrangian.   A second particle can be  added and the relative
displacement can be described in terms  of the Lagrangian statistics of
the two particles to compute the relative diffusion.   The  most frequent
application of this theory is to total  diffusion,  i.e., the variance of
displacement of a large number of particles released steadily from the
same point.  One fundamental  result,  written here  for  only the lateral
component of diffusion,  is
                                   135

-------
          da2/dt = 2 /  v'(t)v"(t  +  g) d£
            y         o                                            (5.5)
                     -  2  a/  /
                             o
where t is the travel  time, av  is the  total velocity variance, and R(£)
is the Lagrangian correlation coefficient  of velocity separated by a time
lag C«  Since R(£) * 1  at small  £,  i.e., there  is no significant
change in parcel  velocity over  a small time, it follows from the above that
ay = avt.  For very large travel times, if R(£) approaches zero faster
than 5"1 at large 5, ay = (Za^t^1/2, where tL is the limit of
the integral  of R(£) over a large time; t|_ is the "Lagrangian time
scale".  This is  the same result given by  constant eddy diffusivity, with
K = av2t|_.  Correlation coefficients have  a direct mathematical relation-
ship with spectra, so the basic  theory of  Taylor was later developed in
terms of integrals of Lagrangian velocity  spectra by Batchelor (1949) and
others.

     These results are exact  and relatively simple, but true Lagrangian
velocity statistics are rarely  obtainable. In  an approximate way, large
scale horizontal  velocities can  be  deduced from the trajectories of
tracked balloons, but these never follow the vertical air motions perfectly,
What is needed are cheap, small, neutrally buoyant air motion "trackers".
Because it is much easier to  obtain Eulerian velocity statistics, which
are measured at a fixed point such  as  a meteorological tower, Hay and
Pasquill (1959) tried a different approach.  They assumed that the
Lagrangian velocity correlation, R|_(£), is similar in shape to the

                                   136

-------
Eulerian velocity correlation, Rj:(t), except  for an expansion  of the  time
scale by a factor p; that is, R|_(£)  = RgU/p) .   Pasquill  (1974)  has
argued that the exact shape of the velocity correlation  is  not critical
anyway.  The value chosen for p is more important.   One  simple estimate  for
p is based on a scaling approximation for the two time scales.  The
Eulerian time scale, t£, is proportional  to the time it  takes  an eddy to
pass a fixed point, which is A/If, where a is the size of the dominant
eddies.  The Lagrangian time scale is proportional  to the time it takes  to
traverse an eddy with a turbulent motion, JL/0V, if  lateral  diffusion
is being considered.  Then p « t^/t^ « "u/crv = i~ ,  where i  is  called
the turbulence intensity.  Pasquill  (1974)  gives  estimates  of  (pi)
ranging from 0.35 to 0.8.

     The most useful result so far from the Hay-Pasquill  hypothesis has
been the following relationship,  written here  for only  the  lateral component
of spread,
Here, we use the approximations aa - crv/u" for  the  azimuth  angle  variance
and t = x/U for the time of travel;  T is  the sampling  duration;  and
t/p is the averaging time,  which acts like a filter.   By smoothing
over an averaging time t/p  before calculating  variances, the  contribution
to c?v of high frequency eddies that  only  move  the  particles back and
forth a number of times by  the time they  reach x is  suppressed.   Using
the above, all that is needed  to calculate dispersion  in homogeneous
turbulence is a record of the  velocity or azimuth  angle fluctuations at

                                   137

-------
the release point plus a good estimate of p;  this  method  can  be applied
to different sampling durations in a straightforward  manner.   This  approach
is, in fact, the main basis for PasquilTs curves  for ay, rather than
direct diffusion measurements.   Pasquill  (1974)  has  accumulated much
experimental support for this approach, especially for ay.

     Statistical theory goes back over 60 years, yet, since 1976, there
has been a marked increase of interest in it  as  a  practical  tool  for
diffusion modeling.  The first  of these recent attempts to make practical
simplifications in statistical  theory was by  Draxler  (1976),  who used  the
forms suggested by Pasquill (1971)
               ay = avt fi(t/tj_)  and                               (5.7)
               az = awt f2(t/tL)  ,                                 (5.8)

where av or aw are measured over the same sampling duration as ay or  az,
with an averaging time approaching zero (i.e., unfiltered fast-response
measurements).  Draxler attempted empirical  evaluations of f^, fp,  and t\_
separately  for elevated and surface sources,  horizontal and vertical
diffusion,  and stable and unstable conditions.  Altogether, 11 different
sets of field experiment observations were used.  Analytical  best fits
for ?i and  f% were suggested as were best values of Ty?  for each category,
where T^/2  is the time that f^ or f? is observed to drop  to 1/2 (the  derived
value of t|_ was 1.64 T^/2 ^or most cases).  ^1/2 ranged from 50 s for az  of
ground sources in stable conditions to 1000 s for  ay of elevated sources
for all conditions.  However, the scatter in  individual run best fits to
was very large, over several orders of magnitude.   The scatter in the
                                   138

-------
logarithms of the ratio of computed to observed ay and az, using the
suggested Ty2 values, ranged from a standard deviation of 1.29 to 2.05.
These were somewhat better than the comparison with the Pasquill-Gifford
predictions, but still represent errors of factors of 4 to a -  not a very
comfortable place to rest the "state-of-the-art".   Nevertheless, in 1977
this approach was recommended by an AMS workshop as a step toward improved
modeling (Hanna et al., 1977).  For ay, the simple form suggested by
Pasquill (1976a) was also recommended, with 0y/(aax)  a function of
only x.  It should be recognized that much of this scatter is due to
poorer instrumentation for measuring aa and ae in  the older experiments.
Draxler's method performed much better when compared  to recent  experiments
(Gryning and Lyck, 1984).

     Subsequently, there have been attempts to reduce the  amount of pure
empiricism in this approach,  to better model  the variability of quantities
like 0y/(aax).  Doran et al .  (1978) showed how this quantity,
plotted against x, systematically depends  on the sampling  duration and
the averaging time used in different experiments.   They qualitatively
related this dependence to velocity spectra considerations and  suggested
that wind speed and stability will  also have important effects.   The same
authors studied additional  data bases and  concluded that aa measured at
elevated release points is not a good indicator of surface level  ay;
near-surface aa orders the diffusion data  far better  (Horst et  al.,
1979).  Examining data for rough terrain experiments, they found
indications of problems with  cra measurements;  in the  older experiments,
wind vanes may have been either sluggish or insufficiently damped.

                                   139

-------
     Irwin (1979) suggested some forms for crz/(aex)  in which ae is the
variance of the elevation angle.  For neutral  and stable conditions he
presented them as empirical functions of travel  time, with one curve for
zs > 100 m and another for zs < 50 m, and a suggested interpolation;
these were based on averages from five experiments.   For unstable conditions,
he made the physically reasonable assumption that t[_ is proportional to
z-j/w* and presented the Willis and neardorff tank data, Lamh's numerical
results, and some limited field data as functions of tw*/Zi, the convective
scaling dimension!ess time scale.  Separate curves were fit to ground
source and elevated source data.  This approach  does appear to give
considerably less scatter than the purely empirical  analyses.

     A somewhat similar scientific approach was  recently used for elevated
source values measured in a SBL (Venkatram et al., 1983).  The data were
lidar measurements of 
-------
curve az = awt(l + t/2t|_)~1//2.  However, In the upper half-decade of
their t/t|_ values, some tendency towards a -1 slope in the plot for
cjz/(awt) can be seen.  This suggests az « aw t|_ as a possible limit,
a limit that has been suggested by Hunt (1982) for the case tL « N-1.

     The effect of sampling duration on plume width and average concentra-
tion can be readily determined by using statistical theory, a capability
that escapes most other modeling approaches.  All  that is needed is a set of
velocity spectra or a time series of fast-response wind speed or directions
from which quantities like (
-------
     "Cluster growth", or relative diffusion,  is  another  problem  area
that, in principle, is amenable to solutions  using  statistical theory.
In practice, this is a much more difficult  problem  to  solve because one
must know or assume properties of relative  Lagrangian  velocity statistics,
which depend on the spacial separation of particles as well as time
lags.  Taylor first addressed the problem (Batchelor,  195?),  and  Smith and
Hay (1961) attempted to simplify it by replacing  Lagrangian velocity
correlations with Eulerian ones, in the manner of Hay  and Pasquill  (1959).
This theoretical  matter then seemed to get  little attention until  Sawford
(1982a) discussed and compared both approaches.   Data  for testing predictions
for relative diffusion were limited and contained much scatter, but
Sawford believed that Taylor's approximation  gives  more realistic predic-
tions.

-------
RANDOM PERTURBATION MODELS

     In the last few years there has been  a resurgence of interest  among
theoreticians in mathematical  simulations  of turbulence diffusion that
use randomness as one element  of the simulation.   These are  variously
called random walk, Markovian  random walk, Monte  Carlo, random  force, and
Langevin equation methods.  Much of the mathematics  of these methods has
been borrowed from treatments  of Brownian  motion, which were made much
earlier in the century by a number of prominent scientists,  e.g., Einstein
(1905).

     Most of these methods also lean heavily on statistical  approaches to
diffusion modeling, since they depend on velocity correlation functions
for the non-random part of the modeling.  An exception is  Lamb's (1978)
modeling of the CBL, which uses Oeardorff's (1973) large-eddy,  subgrid-
closure model  to determine motions caused  by large eddies  and randomized
velocities for diffusion caused by eddies  on the  subgn'd  scale.  Undoubtedly,
computational  advances account for some of the renewed interest in  these
methods, especially the Monte  Carlo techniques that  "compute" the trajec-
tories of thousands of particles using random number generators.

     The most common starting  point is the assumption  that

          v(t + At) = R(At) v(t)  + v'(t)                         (5.9)

(Smith, 1968), where v is the  velocity of  a small  volume  of  air, At is
a small increment of time, and v'  is a random change in the  velocity,
with V1" = 0 and v1 v = 0.  It follows that  for stationary,  homogeneous flow,

                                   143

-------
R(At) must be the Lagrangian  autocorrelation  function  (Gifford,  1982).
Another form of the above uses the series  expansion  of v(t + At)  in terms
of At, making At small  enough that higher  order  terms  can be neglected,
so that v(t + At) = v(t)  + Atdv/dt.  Substitution  into the above yields

          dv/dt + [1 -  R(At)]At-!v(t)  =  v'/At.                 (5.10)

This is a form of Langevin's  equation, which  was widely utilized in
theories of Brownian motions; some of these theories are referenced by
Gifford (1982), who also mentions several  applications of the  equation to
turbulent diffusion made in the early 1960s.

     The above equations, while simple,  have  many  interesting  properties.
If the process is stationary, v7 must  be constant  in time; this  imposes
the condition that V'z   = v7!!! - (R(At))2], so that  the random velocity
changes do not cause an increase or decrease  in  the  total averaged kinetic
energy.  Quite a few investigators have  assumed  that R(At) has an
exponential form, as a  matter of convenience; Gifford  (1982, 1983) claims
that this form of R can be derived from  the fundamental  equation. On the
other hand, Pasquill and Smith (1983)  claim consistency of this  equation
with any form of R(At)  that decreases linearly  in  At for small At; they
also fit nonlinear forms of R, R = 1 - 9Atn for  small  At with  n  * 1, to
more complex velocity change  assumptions involving v at time steps prior to
t (double regression and triple regression models).  However,  the exponential
form is the simplest, and it  has been claimed that diffusion predictions
are not sensitive to the exact form of R (e.g.,  Pasquill,  1974).  Smith
(1968) notes that the exponential form for R  implies impulsive forces acting

                                   144

-------
on the parcel, as the analogy to Brownian motion would suggest,  but  this
may be of little consequence to the practical  outcome, the diffusion
prediction.

     Some controversy exists over the valid  applications  of this
assumption to diffusion prediction.  Gifford (1982)  applied the  Langevin
equation, with R = exp-(pt), to derive analytical  predictions  of the mean
displacement of particles with an initial  velocity v0  and the  mean-square
displacement of the same particles, ay2.   Applied  over all  initial
values of v0, with v0^ = "v2" = av2, he derived  the  same equation  for
total variance that Taylor (1921) derived from statistical  theory (Taylor
also assumed the exponential form of R).   Applied  to small  initial values
of v0, Gifford showed that the results approximate ay2 =  v02t2 at small
t, (?/3)av2pt3 at intermediate t, and 2 av2p~l at  large t (p'1
is the Lagrangian time scale of the turbulence and 
-------
that the ay equation applies to the averaged  diffusion  of  all  puffs
leaving the source with the velocity v0,  not  to  the  average  relative
of an isolated puff diffusion.

     Although we must be careful  about  the interpretation  of various
results, the Langevin equation with an  exponential  form of R is  very
mathematically malleable.  It was applied to  relative and  total  diffusion
from a finite-sized, finite-duration source by  Lee  and  Stone (1983a).
They also computed peak to mean concentration for a  point  source.   A
comparison paper explores relationships between  Eulerian and Lagrangian
time scales, so important in statistical  diffusion theory, using the
Langevin equation with Monte Carlo simulations  (Lee  and Stone, 1983a).
Sawford (1982) used a similar method, except  that he applied it  to  the
relative motion between a pair of particles,  a more  fundamental  approach
to the problem of relative diffusion.

     Most of the above results apply to horizontal diffusion because of
the assumption of homogeneous turbulence; this  assumption  is seldom
justifiable in the vertical dimension.   Ways  to  extend  the application of
the method to the case of a gradient in turbulent kinetic  energy were
developed by Wilson et al. (1982) and Ley and Thomson (1983);  this  requires
compensation for a bias velocity that develops  in the direction  of  the
gradient.  Some conclusions from these studies  are  given by  Pasquill
and Smith (1983); one is that about 5000 random-walk trajectories must be
calculated to get smooth profiles.  Davis (1983) used these  simulation
methods to obtain a number of practical results  for  elevated-source
                                   146

-------
diffusion in neutral  conditions.  He explored  the effects  of varying  the
release height, the surface roughness, and the friction velocity by making
Monte Carlo experiments for a number of parameter combinations.

     Very recent work on random walk modeling  in  inhomogeneous  flows
(Thomson, 1984) and inhomogeneous  and unsteady flows  (van  Dop,  Nieuwstadt,
and Hunt, 1985) has shown that the Langevin equation  can be  used under
these conditions, but a careful  analysis is necessary near boundaries.

     A more elaborate Monte Carlo  scheme using Eq.  5.9 was developed  by
Baerentsen and Rerkowicz (19R4)  for vertical diffusion in  convective
conditions.  Recognizing the highly skewed nature of  the vertical  velocity
distribution owing to the larger fraction of horizontal  area occupied by
downdrafts, they assumed separate  statistics for  updrafts  and downdrafts,
in effect, a double Gaussian distribution of w.   Particles in an updraft
or downdraft may stay in it until  reaching z = zi  or  z = 0;  then,
perfect reflection is assumed, with a switch in the updraft/downdraft
category that selects the statistical  parameters.   A  particle traveling
in an updraft, for instance, rises with the mean  velocity  of updrafts at
that height plus a deviation velocity based on the  Monto Carlo method and
the mean variance of  updraft velocities at the height  being  traversed.
All velocity statistics were estimated, in terms  of convective scaling,
on the basis of field data, water  tank data (Willis and  Deardorff,  1974),
and Lamb's (198?) numerical modeling.   The Monte  Carlo scheme provided a
very good duplication of the vertical  diffusion results  from the tank
experiments (Willis and Deardorff, 1976,  1978, 1Q81).
                                   147

-------
     The computation of thousands  of  particle trajectories of the Monte
Carlo method can be avoided  by means  of  the  "integral equation" approach,
which integrates all probabilities mathematically  by considering the
probability of a "collision" (change  in  v) in each grid square  (Smith and
Thomson, 1984).  This requires much less computation than the Monte Carlo
method.  With either approach, it  is  relatively  easy for today's computers
to use the random perturbation methods to calculate diffusion;  once
confidence in the assumed turbulence  parameterizations is established,
these methods can he used to explore  the effects of release height,
roughness, stability, etc, much more  easily  than by doing field experiments
for every combination of circumstances.
                                   148

-------
                                SECTION 6





                           WHAT ARE THE NEEDS?



                 Perceived data and modeling  deficiencies





EXPERIMENTAL VERSUS MODELING NEEDS





     If there is a overwhelming scientific  consensus  on  anything,  it  is



that we need well-conceived new experiments much more than we  need more



models.  This is not to say that there are  no needs for  new models or model



development.  There have been frequent calls  to replace  the P/R/T  curves,



especially for elevated-source modeling,  and  to replace  the Gaussian  plume



assumption in situations in which it seems  dubious, especially for vertical



diffusion in convective conditions.  However, alternative models are



available, as Section 5 of this paper amply shows.  The  real need  is  to



prove that, at least, some of them work better than models in  use, and



this requires a broader and more reliable experimental base than we have



at present.  Comparative studies like Irwin's (1983)  are handicapped  by



the doubtful  quality of wind variance measurements in pre-1970s experiments,



the likelihood of significant deposition  losses in the same, the sparsity



of direct vertical  diffusion measurements,  and the rapid decline in data



for distances greater than 1 or 2 km downwind.  In addition, scientists



have often expressed alarm and amazement  at how far accepted modeling



schemes extrapolate beyond the data bases supporting  them; Smith (1981)



compares this practice with extrapolating the parameters for a two-



story house to construct a building the size  of the World Trade Center.



This is hardly an  exaggeration;  P/G/T az  values based on Prairie Grass





                                   149

-------
measurements to 800 m have been  extrapolated  to  100,000 m.   More extended
experiments are needed both to support  better models  and  to  evaluate the
adequacy of models that have been  used  for over  20 years.

     A glance at the reference list  for Table 1  supports  the claim  of
Gifford (p. 4 of Smith, 1981) that experimental  efforts dried  up after
1968, at least for field experiments (the  Hanford-67  series  is  an exception,
spanning 1967 to 1973).  He and  Smith (1981)  attributed this to a misplaced
faith in numbers produced by computers, a  commodity that  became ever more
abundant and cheap.  In the period following  1968, considerably better
tracers, sampling equipment, remote  sensing techniques, and  turbulence
instrumentation were developed.   So  we  are now in a position (if we had
the money) to do much higher quality diffusion experiments.   The recent
elevated-source experiments in Europe,  in  Copenhagen  and  Cabauw, give us
some encouragement, as does the CONDORS experiment in Colorado  (Agterberg
et a!., 1983; firyning, 1981; Moninger et al., 1983).   These  were all
limited objective, but high quality, experiments.  There  has been moderate,
continuing support for laboratory  diffusion experiments through the 1970s
up to the present.  This support,  however, has shifted toward  studies of
complex terrain, acidic deposition and  buoyancy  effects;  there  is little
work at present on basic diffusion phenomena.

     On the model development side,  the opinion  expressed  by Smith  (1981)
that it needs no new support seems extreme; nor  do we share  his
confidence that "new sources of field data will  be devoured  by  the  modelers
as quickly as they become available."  Although  computer  power  becomes
                                   150

-------
steadily cheaper, it takes considerable human effort to assemble all  the
new and old pieces of information, process it effectively using computers,
and intelligently evaluate the results, including needs for model
revisions.  Comparison of one model  with one experiment is a relatively
modest task, but much more is learned, and more effort is required, when
many models are compared with that experiment (e.g., Gryning and Lyck,
1984).  Best yet are efforts to compare many models  with many experiments
(Turner and Irwin, 1982; Irwin, 1984).  These efforts require at least
the current level of support for model evaluation, as most scientists
engaged in this activity feel that it  will  take a long time to meet
current needs at the current rate of commitment.   Furthermore, it  takes
much well  thought out effort to translate research models, once validated,
into appliable models, given the fact  that we do  not normally have high-
grade meteorological measurements available.   This need will  be
addressed  at length in Section 7.  The need for better applied models is
indisputable.  The scientists participating in the Rural  Model  Review
were critical  of the eight models reviewed as being  too similar and
showing too little skill (Fox et al.,  1983).   The AMS committee con-
ducting this review strongly urged that "the  scientific community  submit
models that it considers technically better than  those available today."

     Because modeling improvements require suitable  data  bases for
validation, specific modeling needs  and experimental  needs tend to parallel
each other.  Specific areas needing  attention will be addressed in the
following  subsections.
                                   151

-------
WHAT KIND OF EXPERIMENTS  ARE  NEEDED?

     Suggestions for the  kind  of  experiments that are needed for improvements
in diffusion modeling are numerous.   This  section will discuss generalities,
leaving specific phenomena needing  investigation to the sections on the
SRL, the CBL, and on "calms and other inconvenient events".

     One thing is clear:  scientists want diffusion experiments with a
complete complement  of measurements.   It is frustrating to be unable to
use diffusion measurements to  test  a  model because some critical meteoro-
logical variable was not  measured.  Sometimes a climatological or other
estimate of the missing variable  can  be substituted, but this muddies
the analysis, making the  results  less certain.  There are many past
experiments that give only aa  and/or  cre; there is no hint of even the
P/fi/T stability.  In either case, no  matter how high the quality of the
diffusion measurements, the data  are  useful only for testing a certain
subset of diffusion  models.  Considering the trivial added cost and
bother of adding "cloud cover" to the list of measured variables, this is
ridiculous.  A number of  other meteorological parameters are more difficult
to obtain, but nevertheless are quite cost-effective in maximizing the
usefulness of the experiment.  The  goal should be to produce a set of
measurements that can be  used  to  test any  physically reasonable diffusion
model, present or future.  The Prairie Grass experiment is an outstanding,
solitary example of  what  can  be done; every possibly relevant meteorological
variable that could  be measured in  1956 was measured.  Who envisioned
then that the aircraft soundings  to 3000 m would be useful for determining
                                   152

-------
z-j and w*, when convective scaling theory was  developed  in  1970?   This
vintage experiment has been more useful  than  any other,  even  in model
evaluations of the most recent decade.

     There is also a strong consensus that diffusion  measurements  are
needed at much larger x than in past experiments, e.g.,  see Ching  et al.
(1983); no one is comfortable with extrapolations to  100 km based  on data
to about 3 km.  Many scientists mention  100 km as a goal  distance,  as
recommended by the AMS Review Panel  on Sigma  Computations (Randerson,
1979).  Especially needed are direct measurements of  az  or  °^  vertical
concentration profiles; in the CBLs, this need extends only to distances
where the material  becomes well-mixed from z = 0 to z\ ,  say to
x = 4 "u z-j/w*.  In neutral  and stable conditions, there  is  little  idea of
the distances at which vertical  growth terminates. There is also  uncer-
tainty about dy at large x; does it  approach  an asymptote <* x^,  <= x*,
or something in between? Extrapolating from 3  to 100  km,  the choice of
asymptote can make quite a difference.   There  are many who  believe that
wind direction shear is an important factor in ay growth  beyond x  = 10
km; thus, 9a(z) should be measured through the depth  of  the boundary
layer containing the plume.
     There have also been a number of calls for experiments over some
variety of terrain (e.g., Randerson, 1979).  The Prairie  firass experiment,
while very valuable, represented an  extremely  flat, smooth  terrain.  The
terrains in which most people live and most pollution problems occur
lie broadly between the prairie field of wheat stubble and  "complex"
                                   153

-------
terrain.  There is a  need  for  experiments  in  areas with typical surface
roughness, with trees,  buildings,  and  scattered clearings  (the Copenhagen
and Cabauw experiments, Table  1, are good  examples).   There is also a need
for experiments in rolling or  moderately hilly areas,  especially  in stable
conditions, when drainage  flow effects develop even on the slightest of
slopes.

     The need for plume measurements at larger distances parallels a
need for more complete  plume characterization.  A sparse sampler  network
merely results in questions about  representativeness, whether the peak
concentration was measured, and where  the  centerline of the plume was
located.  Where it is permissible  to create a highly visible plume, even
old-fashioned smoke photography can produce more useful results,  giving
both instantaneous and  time-averaged plume growth (Nappo,  1984).  This is
not feasible for large x,  however, except  in  stable conditions at a site
remote from development.  Remote sensors,  especially lidar, are seen by
many as the most satisfactory  way  to get three-dimensional plume  measure-
ments.  Lidar can either be surface-based  or  mounted on a  plane traversing
a plume from overhead (Johnson, 1982). Lidar signal processing techniques
have now developed to the  point that X/Q isopleths can be  mapped  (Eberhard
et al., 1985).  The overhead "burden"  (f^-dz)  of SC^ and other gases can
be measured by van mounted correlation spectrometers,  and  the ground
level concentrations can be measured with  fast-response gas analyzers by
using four to six traverses per hour across a plume on a highway; this
technique has been used to define  power plant plumes at distances up to
25 km (Weil, 1977).  For much  larger distances, several excellent tracers
                                   154

-------
have been developed that do not significantly deposit  and  that  can  be
detected at great distances, up to x = 1000 km (Johnson,  1982).   The
logistics of operating and maintaining a sampler  network  on  such  scales
is difficult and considerably increases the cost  of an experiment.

     A few other experimental  needs have been mentioned.   One is  for more
elevated-release experiments, with larger zs, typical  of  the taller sources
we are modeling.  The recent experiments discussed  in  Section 3 had source
heights ranging from 115 to 285 m, so there has been some  progress  for
more elevated release experiments.  The majority  of releases from taller
sources are also quite buoyant, which involves a  new range of phenomena
that is best studied with buoyant  plumes as the objects of measurements,
rather than passive releases.  One very good idea,  at  least  if  kept on a
scale that makes it not too expensive, is a multiple day  around-the-clock
experiment such as done in the Tennessee Plume Study (Schiermeier et al.,
1979).  Past experimenters have sought short periods of steady-state
meteorology in order to get data in the ideal  conditions assumed by
modelers.  The atmosphere is never in a steady state for long,  and  many
modelers and users wonder if transition periods,  turbulent "bursts" in
the SBL, and other transitory phenomena are being mismodeled.

     The above discussion has concerned needs for field studies.  While
some phenomena, like wind direction shear effects and  large-x diffusion,
may not be practical  objects of laboratory modeling studies, wind tunnels,
water channels, and tanks are very useful  for studying short-range  phenomena,
Since they are cheaper to operate  than field studies,  they ought to be
                                   155

-------
more fully exploited.  A few suggestions  regarding  these  studies were



made at the end of Section 3.





METEOROLOGICAL MEASUREMENTS





     As mentioned in the previous  subsection, diffusion field experiments



cannot be fully utilized for testing models  unless  they are accompanied



by complete meteorological  measurements.   There  are also  practical needs



concerning meteorological  measurements  to be used in applied modeling.



This subsection addresses  both  of  these concerns.





     The meteorological  parameters needed for testing current research



models are aa and ae at source  height,  u*, L, and z0 and, for daytime,



z.j and w* (there is one redundancy among  these,  because w*^/z^- = u*^/kL).



For large-x, lateral diffusion  experiments it is important to measure



mean wind direction, D, through the layer containing the  plume, so that



the rate of change of wind direction with height can be correlated with



 ay growth.  In stable conditions, potential  temperature  profiles, 9(z),



should be measured through the  layer containing  the plume because an



important time scale affecting  turbulence and vertical diffusion is N~l,



where N
-------
at some height representative of the plume;  better,  u(z)  profiles  should
be measured through the layer containing  the plume.   It  is  also  desirable
to have more than a minimal  number of U,  "ea, and  T measurement levels
in case an instrument becomes impaired  during the course of the  experiment,
which is a strong possibility.

     These measurements can  be made roughly  or with  high quality instrumen-
tation, depending on the overall  level  of effort  in  the  experimental
program.  The surface roughness,  z0, can  be  crudely  estimated by eye,  or it
can be more properly determined from wind profiles taken  near the  ground
in near-neutral  conditions.   An even better  method is  given by Wieringa
(1976, 1980).  This estimate of roughness uses the longitudinal  turbulence
intensity and gives an effective roughness.   The  surface flux quantities
u* and L require fast response measurements  of u,  w, and  T  to be measured
directly, and such measurements require some experience  in  quality assurance.
The next step down is to estimate u* and  L from established flux-profile
relationships and profiles of U and 9;  this  approach has  often been
used in analyses of the Prairie Grass experiments, as  excellent  profile
measurements were made on a  16-m  tower, but  fast-response turbulence
instrumentation  did not exist in  1956.  The  minimum measurements for
estimating L and u* are the  bulk  Richardson  number, IT^/A9,  and ZQ.
Particular care  should be taken in making these measurements; U  should be
measured near z  = 10 m, and  measurement between about  2 m and 10 m has
been recommended for A9 (Hoffnagle et al., 1981).  Potential temperature
is then calculated as a direct function of temperature and  ambient pressure.
The lateral  and  vertical  wind direction variances  can  be  replaced  by the
                                   157

-------
velocity variances av and aw,  respectively,  by  using  aa  »  av/u  and
ae = aw/TJ.  If u*, L, and Zi  or h  are measured, it  is now  possible to
estimate ov and ow from these  parameters  on  the basis of boundary layer
research, with the exception of av in stable conditions  (large  horizontal
eddies may depend more on the  terrain than on surface fluxes  in stable
conditions).  However, the atmosphere does vary from  its own  norm, so it
is much better to measure av  or aa and 
-------
     In making the sophisticated measurements,  one should  not  overlook


the simple ones.  To determine the P/G/T category, all  that  is needed


besides IT is a measurement of insolation, net radiation, or  a  cloudiness


estimate.  For slightly more refined systems, some notation  of surface


moisture (dry, lush plant cover, dew,  soil  set  from recent rain,  etc.)


and a relative humidity measurement are needed.



     This relates to another experimental  need.  Little work has  been


done on how to get the best estimates  of difficult to measure  parameters


that are from relatively simple measurements.  This must be  done  if  research


diffusion models are to be usable in improving  operational models.   Some


work along this line will  be discussed in Section 7; more work is needed.


It is especially valuable to try out the simpler  measurement schemes


in the context of a diffusion experiment, so  that we can discover how


much degradation of diffusion model  performance occurs  when  we use simple


meteorological measurements as proxies for the  difficult ones.  Can  a


research model applied in such a way beat out the P/G/T or Brookhaven


systems without a large increase in measurement effort? One step in this


direction was made by Weil  and Brower  (1984)  in the context  of power


plant plume modeling.  They demonstrated that a Gaussian model  using


improved dispersion parameters outperforms a  current operational  model.


Briggs (1982) showed that the Prairie  Grass values of u*/*dy/0 correlate


about as well with simple substitutes  for L as  with L itself,  the best


measure in previous comparisons (Briggs and McDonald, 1978).   The substi-

           	o         	o
tutes were u /Rs* and u /Rn*, where Rs* and Rn* are proportional  to  solar


insolation rate and net radiation, respectively,  rather easily measured



                                   159

-------
quantities.  Even IP alone provided  useful  correlation,  but  with  increased
scatter.

     Many scientists have been thinking about  the need to  get  the best
estimates of quantities like u* and  L for operational modeling purposes.
Another concern is to get representativeness,  because L  can  vary  by a
factor of 3 or more from field to forest (in this regard,  mean quantities
likeU3 have more spatial averaging  in them than  quantities  like  the
turbulent heat flux.  The AMS Workshop on Stability Classification Schemes
and Sigma Curves made recommendations on instrumentation for diffusion
modeling (Hanna et al., 1977).  More detailed  recommendations  resulted
from the U.S. EPA Workshop on On-Site Meteorological  Instrumentation
Requirements to Characterize Diffusion from Point Sources  (Strimaitis et
al., 1981); this report included recommendations  on instrument response
specifications.  Recent comparisons  of turbulence velocity responsiveness
and accuracy of six commercial systems were made  by Kaimal  et  al . (1984a);
comparisons were also made of four Floppier acoustic sounding systems
capable of remote measurements of wind speed,  direction, vertical velocity,
and vertical velocity variance, aw (Kaimal  et  al., 1984b).

EVALUATION AND CONCENTRATION FLUCTUATION ISSUES

     A surprising number of scientists are calling for more research on
fluctuations of concentrations about their mean value  in steady conditions,
x"  (defined x' = X - X~ )•  It is the  nature of  turbulence to be very
variable even when the conditions producing it are steady (for example,
note the gustiness a few feet in front of a steadily rotating  electric

                                   160

-------
fan).  Given this natural  variability,  the question  exists  of how to
know when x" , or If, or v'w',  etc,  has been determined.   The answer  is
that you cannot measure a  "true"  average quantity, but the  longer you
average in steady conditions  the  better the precision of the measurement.
The problem in the atmosphere is  that "steady"  driving forces do not
remain so for very long, so there is a  limit  to how  precisely x" and
and mean quantities can be measured no  matter how good the  instrumentation
(in a wind tunnel, you can maintain steady flow conditions  as long  as
your budget allows, so you can get much closer  to true averages).   However,
if one can measure the statistics  of the fluctuating quantities, it is
possible to state the probability  of the resulting finite-time average
being within a percentage  of  the  true average.   This should  be a major
consideration in the evaluation of the  worth  of experiments. (For instance,
each Prairie Grass run consisted  of only a 10-min release of tracer;
because large convective eddies take several  minutes to  advect past a
source, individual daytime trials  showed large  scatter with  regard  to
one another, regardless of the parameterization scheme being tested.
Scatter can be reduced by  combining runs with similar parameter values.)
At the other end of the modeling  process,  the consequence of turbulent
variability is that even with the  perfect  model  and  perfect  input para-
meters, a perfect prediction  of the results that will occur  for a particular
hour cannot be made.  Just as the  weather  is  rarely  "average" for the date,
time, and place, atmospheric  diffusion  routinely deviates about its own
mean, even given the same  set of  driving forces.
                                   161

-------
     Thus, knowledge of concentration variability is  essential  to  rational
model evaluation.  Given that even the perfect  model  cannot  be  in  perfect
agreement with data, except occasionally on chance, what  degree of agreement
should be our goal, and what degree of disagreement really indicates
modeling deficiencies?  This is the main motivation for  recent  interest
in X' measurements and models, as there has been  much interest  in  how
to properly evaluate diffusion models and how to  specify  error  bounds in
models (Fox, 1981; Fox et al., 1983).  Other practical  needs for X1
include prediction of flamability of accidental gas releases, the
response of insect pests to smells locating their host  plants,  and
prediction of lethal doses of fast-acting toxic gases.   The  U.S. EPA-AMS
Workshop on Updating Applied Diffusion Models (Weil,  1985) focused on
this issue.  This workshop recommended pursuit  of several  X1 modeling
approaches that appear promising, including second order  closure models,
large eddy models, and random perturbation (Monte Carlo)  models (these
are described in Section 5).  The greatest needs  are  for  X1  data.
Laboratory simulations were recommended as valuable,  as  X1 measurements
are easier to make in a wind tunnel or water tank than  in the field,
and X" can be determined more accurately by extending  the  "steady"  conditions,
Recent work by Willis and Deardorff (1984) in their modeling tank  began
exploration on X'/X for buoyant plumes in a CBL.

     A related need is to find the most effective ways  to evaluate
diffusion models with field data, which are always flawed and have unknown
deviations from the true x"  .  The AMS committee overseeing the  Rural
Model Reviews was obviously quite disappointed  in the results of extensive

                                   162

-------
statistical  comparisons between eight models  and  sampler data  from one



power plant; scatter of predicted-to-observed X ratios  was  so  large that



none of the sophisticated statistical  tests detected  significant  differences



in model  performance (Fox et al.,  1983).   Using a much  simpler statistic,



the fractional  error, and stratifying the same data by  distance and



stability, Irwin and Smith (1984)  were able to identify differences in



model  performance*and when and where they occurred.   A  number  of  scientists



have stated a need for more studies like  this, i.e.,  evaluations  of



evaluation methods for diffusion models.





     There is sometimes a need to  adjust  diffusion models  for  different



sampling times.  A number of authors have suggested power  laws in sampling



time to correct ay (Gifford, 1975).  These are mostly derived  from appli-



cation of statistical theory to wind direction fluctuation  measurements,



but such corrections should be functions  of stability and distance, i.e.,



existing methods tend to be overly generalized.  Measurements  and models



appropriate to  the X1 question will also  probably serve to  improve our



handling of the sampling time question.





     On a final note, a need frequently voiced that has much to do with



research is to  convincingly demonstrate how uncertainty in  predictions of



X can be usefully incorporated in  routine regulatory  decision  making.





STABLE BOUNDARY LAYER ISSUES





     Next to X1 and model  evaluation needs, the most  frequently voiced



research needs  have to do with SRLs.  Perhaps this is in part  because
                                   163

-------
there has been so much progress  in  the  last decade on understanding

diffusion in CRLs.  The SBL has  been  relatively  neglected partly because

it is more difficult to make turbulence measurements in  stable conditions,

when turbulence intensities are  quite low, and partly because SBLs are

more complex than CRLs.  They may have  very large wind direction shears

and are much more terrain sensitive.  This is because downslope gravita-

tional forces on cooled air near the  surface compete with large scale

pressure forces and momentum transferred  from above in forcing the flow,

even on very slight slopes.  At  the AMS workshop in 1977 it was flatly

admitted that "in very stable conditions  our present knowledge is in a

very poor state" (Hanna et al.,  1977).   Since the participants summarized

the problem so well, here is the rest of  the quote:

     "Our observations refer almost totally to "ideal" (flat, smooth)
     sites, and we have good evidence that roughness and terrain
     slopes greatly affect diffusion  in the stable boundary layer.
     It is recognized that turbulent  mixing normally occurs only in
     a shallow layer near the ground  in these conditions, but the
     thickness of this layer varies over  a very  wide range, from
     meters to hundreds of meters,  depending strongly on the upper wind
     and on site topography.  Whether this layer encompasses or falls
     short of a given source height obviously will have  a critical
     effect on diffusion from the source. We also need  to know much
     more about the frequency of occurrence of "turbulence episodes"
     that have been observed at  night.  These greatly increase the
     mixing height, evidently for significant periods of time such as
     several hours.  Except for  the urban nocturnal boundary layer,
     which is convective and somewhat more akin  to the daytime
     boundary layer, we have only begun to formulate possible ways
     to predict the nocturnal mixing  height.  Measurements of this
     height, for instance using  a remote  sensor  such as  an acoustic
     sounder, at a variety of sites and under a  variety  of wind and
     cloudiness conditions would greatly  improve our knowledge."


     There has been some progress in  understanding the turbulent structure

of momentum-driven SBLs since 1977  (Nieuwstadt,  1984a, b); this was greatly
                                   164

-------
helped by high quality meteorological  measurements  made on  a  200-m
mast at Cabauw, the Netherlands,  and  an  acoustic  sounder to measure
the height of turbulent mixing, h (Driedonks  et al.,  1978).  However,
this was a very flat site, with no significant driving  of the boundary
layer by downslope forces (Nieuwstadt  claims  that,  in contrast to "the"
CRL, there are many different  SBLs, depending on  the  balance  of driving
forces).  Even with this ideal  site, the investigators  decided to exclude
data from some time periods because of evidence of  gravity  waves; these
are long-period waves (several  minutes or more) that  cause  h  to fluctuate
up and down and also cause quantities  like  aw to  increase with height.
Normally ow decreases with height and  near  zero values  as z approaches h.
The waves are not "turbulence", cause  no mixing,  and  could  perhaps be
filtered out, but because they muddy the analysis of  an ideal  SBL, these
periods were discarded.  The effects of gravity waves and ways to predict
their occurrence in SBLs are other areas completely void of information.
The need to study SRLs in rolling or hilly  terrains is  also very great,
as no work has yet been done in this area.

     The most urgent need is for  observations of  h  in a variety of terrains
under different meteorological  conditions,  so that  ways to  model it  using
near-surface measurements may  he  tested;  h  has a  rough  correlation with
u* and L for flat sites, but we have no  appropriate measurements for
sloped sites and no theoretical  reason to expect  the  same correlations
to apply.  This is a most important parameter because turbulence drops
to zero at z = h; h acts as a  "lid" on diffusion  from sources below  h,
and there is no vertical diffusion for sources above  h.

                                   165

-------
     Lateral  diffusion  over  averaging  periods like an hour remains an
enigma for stable conditions.   Not  only  is  it very sensitive to terrain,
it is sensitive to the  history  of the  flow  for hours upwind.  This is
because when  turbulence is suppressed  due to static stability there  is
almost no dissipation of energy in  large horizontal eddies, which are not
suppressed by static stability.   It is like someone let the brake off.
This is especially true above h, where there is  no turbulence.  The
                                                                    •?•
horizontal eddies that  cause plumes to meander may originate from flow
around a mountain, or a thunderstorm a few  100 km upwind many hours
earlier or from the remnants of convective  eddies from the previous
afternoon.  Horizontal  eddies may be present in  a wide range of degrees
on different  nights at  the same site,  depending  on the history of the
flow.  For a  high source above  h at night,  there is no consequence at
night because there is  no downward  diffusion.  However, when the mixing
depth builds  up to the  plume level  in  the morning, fumigation occurs;
whether this  causes high concentrations  or  not depends on the lateral
diffusion that occurred during  the  night.   Relow h, large horizontal
eddies are very sensitive to terrain.  For  this  reason, Pasquill
recommended no cry curves except for 3-min averages, and the 1977 AMS
committee strongly urged aa  measurements at the  site in question (Hanna
et al., 1977).  Recently it  was found  that  nighttime av fluctuations on
a tower correlated with minute  pressure  fluctuations at the ground (Zhou
and Panofsky, 1983). With more investigation, this might turn out to be
an effective  basis for  categorizing ay at night  by using a surface
measurement.
                                   166

-------
     The need for testing the effects of wind  direction  shear  on  ay  was
already mentioned; these are especially large  effects  in SBLs.  Furthermore,
we have almost no grasp on the prediction of the rate  of change of wind
direction with height at night, even  in idealized terrain.   This  is  an
area that might be helped by second-order closure boundary  layer  models,
which have some demonstrated success  at simulating SBL flow on a  slope
(Brost and Wyngaard, 1978).

     Turbulence episodes or bursts have been mentioned as a phenomenon
requiring attention for nocturnal  diffusion.  These are  sudden jumps  in
h, accompanied by increased turbulence and downward mixing  of warmer
temperature, larger "u, and any pollutants that were "captured" by the
increase in h.  These jumps occur  when the stability in  the layer above h
becomes marginal, with a Richardson number Ri  = (g/e)(d9/az)/(alf/az)2
near a value of 0.5 or so.  This can  happen as the wind  profile
above h evolves slowly due to the  turning of the earth (before the
turbulent burst the layer is practically frictionless, so U responds  only
to inertia and the horizontal  pressure gradient);  it can also be  caused by
weakening of 59/az by more rapid radiative cooling from  above.
When Ri falls below the critical value 0.25 somewhere  in the layer,
turbulence develops spontaneously  and can spread at least through all
layers where Ri < 0.5.  These events  are extremely difficult to predict.
For the source below the original  height of h, more vertical dilution
will occur when h jumps, so such events would  serve to reduce surface
concentrations.  However, for the  source above the original  h, accumulated
pollutant may suddenly mix to the  surface in such an episode, somewhat

                                   167

-------
like morning fumigation,  (Evidence of such  events  have  been  seen  in
nighttime SO;? monitoring records  from near a power  plant,  but  high concen-
trations occurring during such an event have yet  to be observed.)

CONVECTIVE BOUNDARY LAYER ISSUES

     Although knowledge of diffusion characteristics of  CRLs  has advanced
more than that of SBLs, important research needs  remain.   These needs  are
primarily for more extended, three-dimensional  diffusion measurements,
studies of late afternoon transitions, and studies  of cloud effects.

     Except for surface measurements near a  surface source, measurements
of diffusion in the CBL to date are overwhelmingly  from  laboratory studies.
There were surface arc tracer measurements at one distance at  Cabauw and
at three distances in Copenhagen  for elevated SF^ releases in  rather weakly
convective conditions.  These satisfy only a small  part  of the need for
field data for comparison with the laboratory measurements, which  showed
such unexpected vertical dispersion behaviors.   Some CRL experimental
needs put forward by Ching et al. (1983) were (1) surface  X and x(z)
profiles to x = 6 km for both elevated and surface-source  releases, (2)
complete concurrent meteorological  measurements including  z-j,  (3)  long
averaging or release times of at  least several  hours, (4)  early morning
and late afternoon releases to study transition,  and (5)  studies of cloud
effects.  Participants at the U.S. EPA-AMS Workshop on  Updating Applied
Diffusion Models (Weil, 1985) also expressed the desirability of including
a dense surface sampling network  and using remote sensing  for three-dimen-
sional  plume measurements.  Some of these experimental  needs  are addressed

                                   168

-------
by the CONDORS experiment in Section 3.   But,  the CONDORS experiment  did
not address the issues of diurnal  transitions  and cloud  effects.   This
experiment was limited by (1) only a single surface sampling  arc  at
x a 1.2 km, (2) settling and deposition  of some of the chaff  plume, and
(3) oil fog detection and mapping  by lidar to  only x = 2 km or dimensionless
X = 1.5.  The critical area of maximum surface impact for elevated oil  fog
releases was well  covered by lidar, but  it could not distinguish  oil  fog
from background haze at much larger distances  due to highly diffusive condi-
tions and limits on the amount of  oil  fog that could be  released  (this
technique might be doubly effective at a remote, very low turbidity site).

     The midday CBL usually approximates a steady state, with slowly
increasing z-j and  nearly constant  w*,  so it is rather well  understood.
Many modelers feel  uneasy about the application of steady-state models to
transition periods early and late  in the daytime.   The growth of  z-f in
the early morning, with entrainment of initially stable  air aloft, has
been measured and  modeled with some success; inversion breakup fumigation
caused by z-j growth has been studied in  the laboratory (Deardorff and
Willis, 1982) and  in a field experiment  (TVA,  1970); this study predates
convective scaling, but only crude estimates of w* and z-j could be made
from the tabulated measurements).   What  needs  study the  most  is the
dying-out of convective turbulence in  the late afternoon and  the  concurrent
collapse of the mixing depth, which tends to leave material "frozen"  at
heights that it mixed up to early  in the afternoon (Ching et  al.,  1983).
Kaimal et al. (1982) suggest that  z-j may collapse rather suddenly in  the
late day, but evidence is sketchy.

                                   169

-------
     Cloud effects on turbulence  and  diffusion  have  received very little
attention in spite of the fact  that clouds  occur  rather commonly.  There
are many degrees and  types of cloudiness  and many questions.  Scattered
cumulus clouds often  form at  the  top  of vigorous  thermals; these have
ground shadows that create cool spots of  low or reversed surface heat
flux.  Does this phenomenon alter the balance of  updrafts and downdrafts
and affect vertical diffusion?  When  these  clouds develop significantly
into stable air above Zj, they  may "vent" pollutants that were drawn into
the supporting thermals;  this phenomenon  of cloud venting needs more
study (Ching et al.,  1983).

     When the sky becomes very  overcast,  solar  heating of the ground is
cut off and H* becomes very small, zero,  or slightly negative, cutting
off the buoyant forces driving  thermals from the  surface.  However, cooling
can occur at the top  of the boundary  layer  due  to radiative cooling of
cloud tops and cool  outflows; cooling from  above  can cause an "upside-down"
CBL driven from above.  This  may  sustain  convective  turbulence late into
the afternoon on days with afternoon  cumulus development.  There is a
basic need to study CBL turbulence in cloudy and  partly cloudy conditions,
which is probably done best with  turbulence-instrumented aircraft.  Heat
and moisture fluxes need to be  measured just below the cloud bases, as
well as close to the  ground,  so that  we can find  out if current CBL models
will work for top-driven CBLs merely  by turning them upside down and
using -H* at z « Zi in place  of H* at the surface.   Very likely, further
model modifications will  be needed.
                                   170

-------
     A good start in this direction has been made by the University of



Wisconsin Boundary Layer Research Team (Stull  and Eloranta,  1984;  Wilde



et al., 1985; and Stull, 1985)  who studied the CBL with cumulus  over



Oklahoma.  Albrecht et al. (1985) have studied the marine ARL topped with



stratocumulus over the eastern  Pacific.  This is  an example  of the ABL being



strongly cooled from the top.





CALMS AND OTHER INCONVENIENT EVENTS





     A perennially dodged question is, "how do you model  calms?" Most



models have X « Tj-1, so  their predictions for U = 0 are alarming.



One psuedo-answer to the question is that the wind never really  stops



blowing; it is just too  slow for instruments to respond.   When this



happens, use the Gaussian model  with u" = 1 m/s.  A second such answer  is



just like the first, except that "U = 0.5 m/s is the magic minimum  wind



speed that makes the model  work (this answer doubles X, of course).



These responses seem less than  scientific, at the least.   Plumes have  been



observed to rise and fan out in all  directions in stable, nearly calm



situations and have been observed to loop in all  directions  in nearly



calm, convective conditions; in fact, this condition sometimes gives the



highest observed mean surface concentrations at power plants (Bowne,



1984: personal communication).   A few physically-based models for  diffusion



in If = 0 conditions do exist, but very little has been published.   Recently,



a model for zero and very low u" diffusion in the  CBL was  proposed  (Deardorff,



1984).  Adequate testing of such models requires  a dense  sampler network



close to the source occupying all  compass directions or remote sensing of
                                   171

-------
a plume near the source and  some sensitive  wind  speed measurements  (e.g.,
sonic anemometers).

     Another "inconvenient event"  for which no  special modeling  techniques
exist is rain.  No diffusion experiments  have been  carried  out  in rain,
and some instruments perform well  only in dry weather.   Continuous  S&2
monitoring has been made at  many power plants,  but  accompanied  by only
standard, rather crude meteorological  measurements.  Furthermore, S02
is scavenged by rain and snow,  which adds a large complicative  factor to
the analysis.  Examination of such monitoring records should  help answer
the question of whether diffusion during  precipitation events is a  worst
case.  Scavenging and wet deposition are  very important  questions here,
but are beyond the scope of the present paper.   When  precipitation  comes
from stratus clouds, as is usual for mid-latitude winter storms, there  is
almost no positive or negative heat flux  at the surface, so we  usually
assume that "neutral" diffusion applies in  these conditions.  To my
knowledge, no one has tested this assumption.  When precipitation is
convective, as in thunderstorms and on-again-off-again showers,  large,
cool downdrafts are generated by evaporative cooling  in  the rain cells.
This condition undoubtedly affects diffusion in a profound  way,  but how?

RESOURCE ISSUES

     Efficient use of research funding is an issue  that  is  raised frequently
in the context of research needs.  As this  section  has shown, it is easy
to assemble a rather long list of areas in  which our  knowledge  is deficient.
These areas are not "ivory tower" concerns, but affect many practical

                                   172

-------
decisions involving investments that,  collectively,  total  billions  of
dollars.  Too much of the diffusion modeling  supporting  such  decisions
has been ad hoc or highly extrapolative,  lacking  appropriate  validation
for the application.   Usually this  reflects lack  of  appropriate  data, but
too often insufficient use is made  of  existing  data  that could provide
partial validation, at least.  More encouragement of evaluation  efforts
is needed, either through allocation of funds for contracted  work or
through allocation of scientific staff time.  These  efforts must also
include a careful  analysis of the range of applicability of any  new models
as well as outlining  the details of model operation.
     The long list of research needs necessitates prioritization, as
there are not enough  resources (including scientists)  to address all the
needs.  However, prioritization is  not enough;  there is  also  the strategic
problem of how to make the most of  resources.   Given a certain level
of funding, one could choose to direct it all towards  the top priority
need, or divide it among the top 10 or 20 priorities.  The consensus
among diffusion research scientists seems to be that more is  accomplished
by supporting a multiplicity of small  research  projects  with  diverse
goals, carried out by individuals or small teams, than by supporting a
large project with narrow goals.

     There are needs  for occasional  "big  experiments", but these need
not be budget busters.  The Prairie Grass experiment of  1956, the most
classic example, was  a multi-agency effort; four  universities and 60
scientists and technicians participated.  The recent CONDORS  convective
                                   173

-------
diffusion experiment was more modest,  involving  about  20  people during
data collection and costing about  1/4  million  dollars  (a  large fraction
of the cost was for computer processing of the remote  sensing  signals);
yet, it was a well-instrumented  experiment that  accomplished its basic
goal: to define diffusion in three dimensions  for elevated  and surface
sources for highly convective conditions.   Similar experiments on about
this scale or a little larger (with more surface sampling)  run in neutral
and stable conditions as well  as convective conditions at several  diverse
sites would give a far better experimental base  for testing models and
for checking against laboratory  modeling.   Laboratory  modeling is a very
cost-effective tool for basic diffusion research as well  as for complex
flow situations.  It cannot provide all  the answers, but  it can be used
to explore many fundamental phenomena  in a very  controlled  and systematic
way.  Numerical modeling, such as  second-order closure, LES, and Monte
Carlo models also can be very cost-effective tools for basic explorations,
if the models are reasonably efficient.  Furthermore,  it  would also be
cost-effective to support analyses of  many under-used  past  meteorological
and diffusion measurements in the  light of contemporary boundary layer
concepts.
                                   174

-------
                                SECTION 7







              IMPLEMENTATION OF IMPROVED MODELING PRACTICES



                      The technical  transfer problem





     There is a very large gap between diffusion modeling as  practiced



and the state of the science at the research level.   The Pasquill  curves



are still the basis of most diffusion modeling in the United  States;  they



were developed around 1958 for short-term averages from low sources,



contained many ad hoc assumptions, and were supported in a rather  spotty



way by diffusion and wind direction variance measurements. More informed



modelers prefer the Brookhaven curves for elevated sources, but  these too



were developed in the same era with limited diffusion measurements at one



field site and with many questionable assumptions (like Gaussian vertical



distribution and non-evaporation of oil  fog droplets).  Obviously, more



field and laboratory measurements on diffusion have  been made since then.



Tremendous progress has also been made in understanding turbulence in the



ABL and in successful mathematical modeling of many  aspects of atmospheric



diffusion.  At any conference or workshop concerning diffusion issues,



however, one can sense much frustration over this situation.





     It seems obvious that not enough effort is being made to translate



research into better practice.  What are needed are  major efforts  to



simplify research models sufficiently to make them viable as  practical



tools and to find adequate substitutes for the parameters that are difficult



to measure.
                                   175

-------
     The frequent lack of effective measurements  is  a  real  handicap  in
this endeavor.  If we must restrict our  input  variables  to  the type  of
measurements routinely made at  airports  up  to  now, we  can do  little  more
than define Pasquill  categories and make crude estimates of L on the
basis of wind speed sampled for 1  min  every 3  hr  at  an unusually smooth
site.  Of course, there is little  possibility  of  significantly improved
diffusion modeling using such restricted measurements.  For this reason,
diffusion scientists  have repeatedly recommended  effective  on-site
instrumentation and upgraded measurements at NWS  stations  (Hanna et  al.
1977; Strimaitis et al., 1981).

ESTIMATING NEEDED PARAMETERS FROM  MORE OBTAINABLE MEASUREMENTS

     Except for modeling diffusion at  short distances, < 1  km, from  sources
low in the ABL, the mixing height, zj  or h, must  be  measured  or estimated
to do any kind of diffusion modeling.  One cannot model  at  a  distance
pretending that there is no lid on diffusion.   For elevated sources  at
night, there is a good chance that there is no vertical  diffusion at all
if zs is above the nocturnal mixing depth,  h;  however, if  zs  < h, vertical
mixing down to the ground occurs and the whereabouts of  h  becomes a
critical question.  Other quantities required  depend on  the modeling
approach; these include D, ae,  u",  u*,  L, and w*,  discussed  in Section  6.
All of these should be measured in any diffusion  research  experiment for
maximum utility of the results, but direct measurement of many of these
parameters is too difficult to  be recommended  in  applied modeling.
Scientists are aware of the problem and  have proposed  a  number of schemes
                                   176

-------
for estimating the difficult parameters from those more easily measured.

     There are many ways to estimate the surface-layer parameters u* and L
(or H* = u*3/(0.4L), which also appears in the definition of w*).  The
profile method, using IT(z) and 9(z) profiles from a small tower, is
most often suggested (Rriggs and McDonald, 1978; Nieuwstadt, 1978;
Irwin and Binkowski, 1981; Berkowicz and Prahm, 1982).  However, site
uniformity and height of vegetation are major factors to be considered
when placing instruments.  Without profiles, a crude estimate of the
surface roughness is needed with any of the methods; this estimate can be
made from site inspection (Wieringa, 1980).  Pasquill and Smith (1983)
give nomograms for estimating u*, L, and H* from wind at a single height,
"uj, and potential temperature difference (A9) between 0.2 z\ and zj.
In convective conditions, H* is highly correlated with A9  alone (Dyer,
1965).  In general, u* and L can be determined from IT at a single height
if an estimate of H* is available.  Methods for doing this by using
cloudiness and some site-dependent parameters for surface moisture have
been developed (van Ulden and Holtslag, 1983; Holtslag and van Ulden,
1983).  An alternative method developed by Berkowicz and Prahm (1982)
requires net radiation and humidity deficit near the surface to estimate
H*; both are relatively easy measurements to obtain (humidity deficit
relates directly to temperature and relative humidity).  A similar method
for estimating L by using IF and net radiation or insolation was given by
Briggs (1982a); it is very simple, but lacks consideration of the surface
moisture.
                                   177

-------
     The variances of lateral  and vertical  wind directions are not as
difficult to measure as turbulent fluxes, and a good choice of
instrumentation is available (Kaimal  et al., 1984a).  Lacking measure-
ment, they can be estimated from aa = av/u~ and 
-------
observations from 10 days at Cabauw in the Netherlands (Driedonks, 1982);
excellent results were obtained from models that included a u* term in
addition to the above inputs.

     For the nocturnal h, Nieuwstadt and Tennekes (1981)  developed a
prognostic equation that tested well against observations at Cabauw, but
it requires many input parameters (including the geostropic wind).
Nieuwstadt (1984b) also compared some simpler schemes for h against
Cabauw observations; he found fairly good correlations with Zilitinkevich's
(1972) steady-state prediction, h = 0.4 (u*L/f)1/2, and with a semi-empirical
equation, h = 28 i^n, '  (mks units).  The constant in the latter equation is
dimensional, and should be regarded as site-dependent. It is very similar
to the semi-empirical equation of Venkatram (1980), h = 2400 u*3/2 (mks
units), which gave a good fit to Minnesota measurements (Caughey et al.,
1979).

     Except for a» in stable conditions, for which there is no substitute
for measurements of aa, it appears that all meteorological  parameters
for modeling diffusion can be estimated from a morning rawinsonde sounding,
U at a height of 10 m or so, and some type of sensible heating/cooling
estimate; the latter requires solar radiation, net radiation, or cloudiness
and solar elevation and possibly a humidity measurement.

     What is greatly lacking at present is independent testing of these
parameter substitutions at different sites.  For almost every technique
outlined above, testing was conducted at only a single site.  There is no
way of knowing whether the methods transfer to sites of different character.

                                   179

-------
Furthermore, in cases in which  there are many  approaches,  as  for  the
estimation of u* or L, we would like to  know which  one  is  best  and  which
is the simplest technique that  is  "good  enough."

     All  of these questions could  be settled nicely with comparative
meteorological  measurement experiments carried out  during  some  daytime
and some nighttime periods at three or four contrasting sites,  for  instance,
flat and rather smooth; suburban or forested;  gently rolling; and hilly.
Full meteorological measurements should  be made concurrently, so  that any
simplified scheme can be tested against  the direct  measurements.  Such  a
program would accelerate improved  modeling efforts  tremendously and save
years of guesswork, non-optimum on-site  measurements, and  modeling  attempts
with the wrong input parameters.

BUILDING BETTER OPERATIONAL MODELS

     Implementing improved models  requires research and commitment.
Adding an occasional experiment with partially complete meteorological
measurements to the potpourri of past experiments and the  continued
development of scientific, albeit  esoteric, models  is not  enough  for
progress in a practical sense if the last 20 years'  experience  is an
indicator.  Making the technical transfer happen requires  a  large effort
to sort through the numerous scientific  models (Section 5),  determine
the major features having the most practical significance, search for
effective substitutes for the difficult-to-measure  meteorological inputs,
sift through the available diffusion data for  experiments  with  the  required
measurements and no debilitating uncertainties, perform extensive

                                   1RO

-------
comparisons between the data and several  promising models plus  the old
models (both with and without various substitute meteorological  parameters),
evaluate the results for the most effective models given several  different
quality levels of input parameters, and then cast the favored model(s)
into operational  form with computer programs and/or nomograms.   Most  of
this hard work would be classified as applied research and a  prime research
priority (the only research deserving higher priority is conducting
experiments that are designed to facilitate meteorological  and  diffusion
model comparisons and that make complete and trustworthy measurements).

     The task so briefly stated above is so large, especially when the
variety of operational  modeling needs is considered,  that no  individual
or small group could carry it out in just a few years, even if  they were
relieved of all  other responsibilities.  A few scientists have  made
efforts in this direction, but their time is divided  among many competing
tasks, so that they cannot tackle the whole of the technical transfer
problem.  Another factor for research scientists to consider is that
there is more recognition for original  research than  for midwifing technical
transfer.  However, many scientists achieve a great sense of satisfaction
when their research results are translated into something practical,  and
they are willing to assist in the process.

     A few of the efforts towards better operational  models should be
mentioned.  Pasquill's  (1961) original  fiaussian scheme was such an effort,
on the part of a research scientist, and was a great  step forward at
the time.  Smith (1973) refined this system for vertical  diffusion from a
                                   181

-------
source near the ground by introducing continuous, rather than discrete,
stability categories and providing convenient nomograms, including one
for a correction for surface roughness (his curves were based on the
results of an eddy-diffusivity model).  Irwin (1979)  first advanced a
generalized scheme that used convective scaling for daytime diffusion and
on-site a a and ae data to characterize ay and az anc' provided meteorological
substitutes for nonavailable ora and ae measurements.   This has evolved
into a more complete approach that makes use of the latest advances in
knowledge of ABL structure, especially for stable conditions (Sivertsen
et al., 1985; Irwin et al., 1985).  Weil and Rrower (1982) developed a model
for power plant applications that uses u* and w* to characterize av and  aw;
these determine short-range dispersion in the usual way, based on statis-
tical theory (e.g., ay = crvt).  These forms for cjy and az are then modified
by simple, empirical functions of x.

     The willingness of scientists to seek model simplifications in order
to encourage practical applications was amply demonstrated at the 1984
Workshop on Updating Applied diffusion Models, cosponsored by the U.S. EPA
and the AMS (Weil, 1985; Briggs, 1985b).  This willingness should be
further encouraged by a program of sustained and systematic efforts to
translate research results into improved modeling practice.  This task
requires a thorough familiarity with models and the available data sets
for validation.
                                   182

-------
                                SECTION 8





                                ASSESSMENT





     The purpose of this paper is to review the current  state of science



concerning basic diffusion in the ABL and identify major research needs



in this area.  Basic diffusion in this paper means diffusion  of conserva-



tive, non-buoyant substances over reasonably homogeneous,  ordinary terrain,



Thus, the scope here excludes consideration of buoyancy, deposition,



chemical transformation, complex terrain, and shoreline  flow  issues.



Even so, a vast number of research papers concern  basic  diffusion, and



there are many complexities to consider as various forces  and surface



conditions come into play in the ABL.  This assessment section will



summarize the state of knowledge concerning the mechanisms driving turbu-



lent diffusion in the ABL, the scope of experimental  information that has



been obtained, the theoretical and experimental  basis for  diffusion



modeling as currently practiced, and the strengths and weaknesses of  a



half-dozen approaches to improved diffusion modeling.  It  will  then set



forth what most needs to be done to expand the experimental base for



model validation and further develop the most useful  modeling approaches.



Finally, it is suggested that more effort is needed to translate research



advances into modeling practice.  Model  simplifications  should be guided



by comprehensive evaluations using all reliable experimental  evidence.
                                   183

-------
ATMOSPHERIC BOUNDARY LAYERS

     Progress toward the understanding  of turbulent  and  mean  structures of
ABLs accelerated after the 1968 Kansas  surface-layer field  experiment
(Izumi, 1971), in which accurate profiles of potential temperature  and
wind speed components, including fast-response turbulence measurements,
were made on a 32-m tower.  This experiment, among others,  provided
verification of the Monin-Obukhov surface-layer similarity  theory,  (which
is discussed in Section 2).  This theory has been very successful in
predicting profiles of mean quantities  and vertical  turbulence in the
surface layer; it also works fairly well  for the horizontal turbulence
velocities in neutral conditions and the high-frequency  portion of  the
horizontal turbulence velocity in stable conditions.  It fails to correlate
with low-frequency horizontal velocity  changes in stable conditions,
which have periods ranging from about ~ 1/4 to 4 hr, if  they  are present
at all; it also does not correlate well  with horizontal  turbulent velocities
in unstable conditions.
     The situation with respect to CBLs vastly improved  after the convec-
tive similarity scaling was introduced  by Deardorff  (1970a,b) and was
tested in the 1973 Minnesota experiment (Izumi and  Caughey, 1976),  which
included tower measurements to 3?. m and extended turbulence and mean
measurements to 1200 m by mounting instruments on  a  cable tethering a
large balloon-.  Convective scaling is discussed in  Section  2.  Following
the analysis of the Minnesota experiment in terms  of convective scaling,
for the first time, the behavior of the major parameters of turbulent
                                   184

-------
structure was understood for the whole of the ABL.   Further results  and
confirmation of earlier results for CRLs have been  obtained at  many  sites
now, mostly by research aircraft.  One very significant  finding that
affects diffusion is that the mode of vertical  velocity  is  negative,
about -0.5w*, throughout the mid-mixing layer;  this explains  why the
center!ines of elevated plumes have been observed to descend  in laboratory,
numerical, and some field experiments.  Little  work has  been  done on the
effects of partial cloudiness on CRL structure, and no work has been done
on "upside-down" CBLs driven by radiative cooling at the top  of a fog or
heavy overcast of clouds.

     Progress has been made in understanding NRLs,  but mostly by means of
numerical models rather than by experiments.  Perhaps this  is because
truly neutral (H*/u*3 * 0) conditions are rather infrequent,  and overcast,
windy conditions are not the best flying weather.   There is ample infor-
mation for the lower part of the NBL, for which it  appears  that surface-
layer similarity works very well for the vertical component of  turbulence
and fairly well for the horizontal  components;  these may be influenced by
surface roughness or slope inhomogeneities.  Higher up,  Coriolis accelera-
tion affects NBL structure, including turbulence, and the Coriolis parameter,
f, is important.  Further details of NBLs are contained  in  Section 2. There
is need for more experimental measurements in the upper  part  of the  NBL.

     Much progress has been made in understanding SBLs since  the Cabauw
experiments begun in 1977, but understanding is far from complete because
of the varied nature of SBLs.  They are strongly influenced by  the slightest
                                   1R5

-------
terrain slopes and are also affected  by radiative cooling,  which  depends
on humidity and CO;? profiles.   Also,  they are subject  to  disruptive
events like gravity waves  and  turbulent bursts,  which  are hard  to forecast.
Much has been learned about undisrupted, pressure gradient-driven SRLs  in
flat terrain.  This work is discussed in Section ?..

     The diurnal cycle of ARLs is basically well understood,  except  that
there is not much experimental information on the transitional  periods
near sunrise and sunset.  The  least understood phenomenon is  the  collapse
or running down of CBL turbulence in  the late afternoon,  which  may even
overlap the beginning of S8L development at the surface;  this collapse
begins as soon as H* turns negative,  about 1 hr before sunset.  The  rate
of CBL collapse may depend on  humidity, which increases radiation flux
divergence and hastens the development of stable stratification aloft.

DIFFUSION EXPERIMENTS

     Several dozen field experiments  have been conducted  using  tracers  or
oil fog for basic diffusion studies,  but only a few of them included a
full range of desirable meteorological measurements, and  these  experiments
are quite limited in terms of  distance and plume sampling.   Until recent
experiments, the tracers released were not very conservative, which
confounds interpretation of their results for vertical diffusion  and
surface concentrations.  Laboratory modeling using tanks, water channels,
or wind tunnels has been used  with success in modeling diffusion  in  all
stabilities, but its potential has not been exploited very  much.   In
total, we have a patchwork quilt of experimental information that contains

                                   18fi

-------
many blank squares and many uncertain ones,  especially for vertical
concentration distributions and diffusion at distances greater than  3  km.

     Almost all  field experiment estimates of oz,  the vertical  diffusion
parameter, were made using only surface concentration measurements by  assuming
a Gaussian vertical  distribution of concentration, X, and  a conservative
tracer, i.e., //(Xu")dydz = 0, the source strength.  It has gradually
come to light that all the tracers used, except  for recent gaseous tracers
like SFfi, were not conservative; S02 gas and particulate tracers  deposit
significantly, and oil fog evaporates substantially within a travel  distance
of 2 km or less.  Unless reasonably good estimates can be  made of the
deposition or tracer loss rates and the data can be appropriately reanalyzed
(e.g., Gryning et a!., 1983), the az results from  these experiments
must be held questionable.  The same is true of  the crosswind-integrated
concentration, /Xdy.  We have reliable surface X measurements  for SFfi,
but only for neutral and slightly unstable conditions at x = 2 to fi  km
for elevated releases and only for very stable conditions  at x <  0.4 km
for surface releases.  In a few of the earlier experiments, tower or
balloon tethering cable measurements were obtained for vertical distribu-
tions from elevated  releases before significant  deposition could  occur,
but these were complete only for 26-m releases in  neutral  to stable
conditions and ~ 200-m releases in near-neutral  conditions.

     For lateral diffusion, ay, deposition is less of a problem and
there is a more complete collection of reliable  tracer measurements;
these include some to 26 km for surface releases and some  to 10 km for

                                   187

-------
elevated releases.   The meteorological  measurements  accompanying the
longer range experiments was sufficient for testing  only  certain classes
of models.  As we look at available plume width  measurements  at greater
distances, up to 100 km or so,  we find  rather scant  meteorological
information.  Many of these measurements are quasi-instantaneous plume
widths obtained by a single aircraft traverse at a given  distance.
Under some circumstances, observations  of buoyant plume parameters  can be
used to approximate passive plume dispersion, especially  for  ay after
vertical mixing is complete, as in the  CBL; this gives useable data in
the 15- 30-km range for ay in unstable  conditions.

     Most field experiments suffered from meteorological  instrumentation
deficiencies, so they cannot be used to test all  types of models.
The earlier turbulence instrumentation  was not accurate,  and  many
experiments lacked sufficient wind and  temperature profile measurements
and even simple measures like solar insolation and cloudiness.  Only one
of the earliest experiments included soundings adequate for obtaining z-j,
a key CRL parameter.  The most recent field programs have included  measure-
ments of Zi and a complete array of good meteorological measurements.
Hopefully, this is indicative of better experiments  to come.   Not only do
we now have good turbulence instrumentation available, but we have  much
better tracers and remote plume sensors like lidar,  which has been  used
in two very recent basic plume diffusion experiments.

     Fluid modeling has proven a reliable tool for studying many aspects
of diffusion in ARLs.  It is cheaper than field  experiments and  is  far
                                   188

-------
more controllable, although the drastic reduction in scale does  impose
some limits on what can be simulated, particularly when buoyancy is
involved.  For instance, it is not possible in present facilities to
simulate the effects of large horizontal  eddies or wind direction shear
on ay, but otherwise fluid modeling has proven quite versatile in
simulating vertical turbulent diffusion and the minimum cry in  the
absence of the above effects.

     A recent series of wind tunnel  experiments at Colorado State University
simulated basic diffusion for an elevated release and a surface  release
over a smooth surface and over a rough surface for CBLs, NBLs, and SRLs
(Cermak et al., 1983).  For the unstable runs, fairly good agreement was
obtained with other CRL diffusion experiments, in spite of short dimen-
sionless fetch and weak capping.  Moderately stable ARLs (h/L  =  4 to fi)
were obtained by using a strong temperature differential,  78°C,  between
cooled floor and heated inflow, which is  necessitated by the scale-down.
Other wind tunnels have given excellent comparisons with field measurements
in neutral conditions for az as a function of x and z0, the roughness
length, but with ay about 20% smaller than in the field under  similar
conditions (this is perhaps due to the constraining effect of  the tunnel
walls).  A similar decrease in ay, compared to field experiments inter-
preted in terms of convective scaling, has been noted for  the  convective
tank experiments of Willis and Deardorff (197fi, 1978, 1981).  Their unique
facility, basically just a water tank with a uniformly heated  bottom,
provided the pioneering results for diffusion in very convective conditions.
The somewhat controversial  results for vertical  diffusion  from this

-------
series of experiments have been  essentially duplicated,  by  using  convec-
tive scaling, in numerical experiments  in  the CS1I wind tunnel  and in the
few field observations for which adequate  measurements exist.

     Both wind tunnels and convective water tanks could  be  put to further
use to make systematic studies of source height  effects, surface  roughness
effects, surface roughness and surface  heat flux inhomogeneity effects,
and concentration fluctuation statistics,  a matter of great current
interest.  Also, as wind tunnels and  water channels have been  successfully
used to simulate diffusion in complex terrains,  it seems that  they could
also be used to study the effects of  typical  rolling or  hilly  terrains on
diffusion.

DIFFUSION MODELING AS PRACTICED

     The basic diffusion models  used  in standard practice today were
developed 25 to 30 years ago. The stability classification schemes that
they use, although crude, are not "unscientific".  They  have some basis
in the physics relevant to ABL diffusion,  but this basis is quite incomplete
in the light of what has been learned about ABLs since 1968.  The 
-------
This assumption, while far from rigorous, has proven amazingly resiliant.
For sufficient averaging times, 1 hr or so, it works as well  as any shape
for lateral diffusion.  Because vertical  turbulence is not homogeneous,
systematic deviations from the Gaussian shape have been observed for K.(z)
profiles.  In most cases, the consequences of using Gaussian  X(z)  for
ground-level  concentration predictions are not great and the  prevailing
mood seems to be in favor of not changing.  However, for vertical  diffusion
in unstable conditions in CBLs, the consequences appear to be considerable.

     The Brookhaven stability classification scheme (Singer and Smith,
1966), based  on wind direction fluctuations, was an adaptation of  a scheme
proposed at least as early as 1932 (Hiblett, 1932) and founded on  statistical
theory (Taylor, 1921).  The basic result  for short-range diffusion,
cjy = crax, where oa is the lateral wind direction variance, can be
stated succinctly as "the plume goes as the wind blows" and is so  direct
as to defy any contrary argument (as long as the wind vane faithfully
responds to the wind).  The correlation between ay and aa becomes
less direct in a few tens of meters of travel, and it was an  intuitive
assumption that some empirical correlation would remain at distances of
tens of kilometers.  This assumption has  held up, although it can  be
improved upon.  A more risky assumption was that vertical diffusion would
correlate with cra also; this has proven true at short ranges  in unstable
and neutral conditions when turbulence eddies are more or less isotropic
(high as the  they are wide); however, it  is definitely risky  for stable
conditions when site- and time-specific large horizontal  eddies, or
meanders, can make lateral  wind fluctuations and dispersion much larger

                                   191

-------
than vertical, at least  over 1-hr averages.   The  Rrookhaven curves were
based mostly on surface  measurements  of oil  fog droplets  from  a  plume
released at zs = 108 m in rolling, forested  terrain.   The usual  assumptions
of Gaussian *(z) and a conservative tracer were applied to infer az
values from surface concentrations.  As even heavy-oil fogs can  lose 50%
of their lidar cross-section in  about 2 km of travel,  there is reason to
doubt the conservative-tracer assumption and the  resulting az  values.
The surface ay values should not have been greatly  affected by oil fog
evaporation.

     Pasquill's (1961) stability classification scheme, which  was adopted
by fiifford (19fil) and adapted by Turner (1961) for  use with Pasquill's
curves, is based upon measurements usually available  at airports, wind
speed at z = 10 m and insolation or cloudiness.   It is an intuitive scheme,
but as was shown later,  it correlates significantly with  the physically
very important nbukhov length, L (Bolder, 1972; Rriggs, 198?).   It only
lacks adjustment for the effect  of local surface  roughness.  Like L, it
gives good correlation with vertical  turbulence and diffusion  near the
surface.  It correlates  with lateral  turbulence and ay only when the
turbulence is nearly isotropic,  as for 3-min averages  in  neutral  and
stable conditions.  In unstable  conditions,  turbulence is somewhat
isotropic at an elevation of 0.1 zj or more, but  not  at the ground; z-j
has as much influence on ay as does L, and Pasquill categories have
little to do with zj.  Pasquill's ay  and az  curves  were intended for
use with near-surface sources, as they are based  on near-surface data
(also, the categorization scheme, like L, has decreasing  relevance higher

-------
up in the ARL).  The  400 m, although the rapid drop in surface concentrations was quite


consistent with Willis and Deardorff's results (Nieuwstadt, 1980).



     Some brief comments are due on two other  stability classification

                                                   f\
schemes.  Hogstrom's (1964) suggestion of s =  A9/Uf as  an index in which


Uf is the "free wind" velocity at about 500 m  was based on testing of five


different indices for correlation with observed ay  and az values from


his oil fog puff experiments.  For his site, uf worked better than IT at


source height z = 87 m.  The potential temperature  difference was  measured


between 30 and 122 m, bracketing the source height. At  least in the surface


layer, A9/u? is a bulk Richardson number that  does  relate directly to L.


Using uf and A9 from higher up, there would not be  a direct correspondence,


but s would broadly indicate dynamic stability in the  elevated layer


encompassed by A9.  For Hogstrom's experiments, which  were made only in


neutral to very stable conditions, s correlated very well  with changes in


relative diffusion, 
-------
center!ines in the vertical,  0ZC.   There was  no  discernible  trend with
stability in ayC, the meandering part  of lateral  diffusion.   The suc-
cesses,  and failure,  of this  index  for Hogstrom's elevated releases
roughly  parallel  those of Pasquill's  scheme,  which is  likewise  loosely
related  to L.  However, if Hogstrom had been  able to use  his puff technique
in unstable conditions, he probably would not have found  much correlation
with s,  as A9 is  very small  in  CBLs above 30  m (for good  correlation with
L, A9/u"2 should be measured  below 20  m; the bottom level  should be as
low as 1 m, if practical).  For the same reason,  the use  of  AT  at consider-
able elevation as a stability index could not distinguish at all between
unstable and neutral  conditions.  Measured near  the surface, AT or A6
correlates well with  H* in unstable conditions,  hut not with IT  or u*,
which have much more  effect  on  diffusion. Thus,  AT alone is inferior
as a stability index; even u" alone  broken into a  "day" and a "night"
category works much better (Briggs  and McDonald,  1978).

ALTERNATIVE MODELING  APPROACHES

     Many papers have been published  concerning  more scientific approaches
to basic diffusion modeling.   Thankfully, they sometimes  include testing
of a model against some kind  of diffusion observations.   However, compre-
hensive testing of a  variety of alternative approaches has not  been done;
meteorological measurements  in  data sets make such comparisons  very diffi-
cult, if not impossible.  This  paper  has divided  the approaches into  six
generic  categories for discussion and attempts to identify the  strengths,
weaknesses, and potentials of each  method.
                                   194

-------
     Surface-layer similarity theory expresses gradients of mean quantities
in terms of (u*z)~l times their surface fluxes times functions of z/L
and expresses turbulence variances as u* times functions of z/L.  This
theory has been quite successful  at predicting profiles of IT,  9, and other
mean quantities in the lower ABL.  Near the surface, aw/u* - 1«3; 1 15 or so,
there is little doubt that convective scaling is a better approach.  This
replaces u* and L with w* and Zj  as principal  scales.  Convective scaling
has been demonstrated in numerous measurements programs to correlate well
with turbulent velocities throughout  CBLs (surface-layer similarity is
superior only for aw below z = |L|).   Experience has shown that approaches
that order turbulence measurements well  also order diffusion well.  The
fact that similar results are obtained for  diffusion in such diverse
media as water tanks, wind tunnels, numerical  experiments, and field

                                   195

-------
experiments when the results  are nondimensionalized  with  convective
scaling does much to instill  confidence  in  the  method.  The  correct
scaling, in itself,  does  not  constitute  a diffusion  model; it merely
provides a way to summarize data efficiently  and  in  a more universal
format to allow comparisons of data  from diverse  experiments.   It  also
suggests a simplified format  for diffusion  models.   A number of equations
for ay and az and non-Gaussian vertical  diffusion models  using  w*  and
z-j scaling have been proposed (Briggs,  1985b).   Some of these also include
neutral asymptotes that use u* as a  second  velocity  scale.

     Gradient-transfer, or "eddy diffusivity",  models assume that  the  flux
of a quantity is proportional to its gradient,  and that the  quantity moves
toward the region of lower concentrations.  The constant  of  proportionality,
K, is called the eddy diffusivity in turbulent  flow  and has  units  of
Iength2/time.  This  theory is built  on  the  same analogy to molecular
diffusion as the Gaussian plume assumption  and  shares some of its  virtues
and weaknesses.  It  is rarely valid  for  lateral  diffusion, as ay must
be larger than the largest horizontal  eddies  for the assumption to hold
true.  In the CBL, Ky = O.lw*z-j does provide  a  supportable asymptotic
prediction for the minimim a    =* 0.5 (w*z1-x/tl)1/2, but usually  horizontal
motions on a larger scale than z^ are present and an x*'2 asymptote  is
not observed (Briggs, 1985b).  Gradient  transfer theory has  the most validity
for vertical diffusion from a surface source  because the  largest eddies
affecting this diffusion are  not much larger  than az (except in CRLs at
x > 0.5 "u" z.j/w* when the mode of *(z)  lifts off the  ground against
the predictions of gradient-transfer models).  It has been a useful  tool

                                   196

-------
for producing practical  az predictions  when  good  estimates  of  K(z)  are
made by using surface-layer similarity  scaling  (Smith,  1973).   Some more
advanced models, like spectral  diffusivity theory,  attempt  to  account for
non-local effects on fluxes of  concentration; these have not been developed
enough to encourage practical application.

     A more advanced technique  is  second-order  closure  modeling, which
assumes that third-moment quantities  like u'w'x'  are  functions  of
local second-order or mean quantities or their  gradients.   Any  local
closure technique works best for stable conditions  when  turbulent  eddies
are small, and second-order closure models have been  highly successful at
predicting the behaviors of SRLs, including the effects  of  terrain slope
and radiative cooling.  They have hardly begun  to be  applied to turbulent
diffusion in SBLs, but it seems that  they would provide  an  excellent
exploratory tool, especially for the  effects of zs/h  and terrain slopes.

     Large eddy simulation (LES) models were first  applied  to prediction of
global  weather patterns and have provided very  good results for turbulent
structure in CBLs and NRLs; both of these contain large  eddies, of the order
of their height.  Unlike the second-order closure models, which predict
time-averaged properties, the big-eddy models compute the instantaneous
structure of eddies larger than the computational grid size.  This provides
actual  "pictures" of the turbulence structure.   To  get averaged quantities,
one computes the time averages by "sampling" over many time steps, much
as in a field experiment.  To account for the interactions-between adjacent
cells,  an eddy-diffusivity assumption is applied  only for small scales,
                                   197

-------
the "subgrid" scales.  This permits counter-gradient  diffusion  to  occur
on the large scales, which does occur in  simulated  CBLs.   These models
have given very good results for diffusion  in  CBLs.  They  have  further
potential for this case and probably would  also be  a  good  tool  to  explore
the effects of zs/z-j and z-jf/u* on diffusion  in NRLs.

     Statistical  models in various forms  have  long  been workhorses of
diffusion modeling.  They are based on the  idea that  if enough  is  known
about the statistics of turbulent motions,  both mean  instantaneous and
time-averaged diffusion of passive substances  following those motions can
be predicted.  The most immediate results,  like 
-------
sources in CBLs have used observed vertical  velocity statistics at source
heights to predict *(z) versus distance essentially by assuming that tL
is large.

     Since 197fi, there has been a move to improve ay and 0Z predictions
by going back to the basic statistical  theory prediction that cjy/(avt)
and crz/(awt) are functions of t/t|_.  The motivations are that
(1) this is surely correct at small t/t|_, (?) this is simple, (3)  the
functions of t/tj_ are weak, not strong, so their form and the estimates
of t|_ are not critical , and (4) we now know enough about ARL structures
to be able to estimate av and 0W at source height in many ways, if
they are not measured directly.  A number of functional  forms and  ways  to
approximate t|_ have been suggested, and much testing against data  sets
has been performed for this method.  It seems a  desirable approach for
generalized applied diffusion modeling and,  once optimized, should give
adequate predictions except when the fiaussian plume assumption itself is
inadequate, i.e., for vertical diffusion in  the  CRL.

     A further development along the statistical  line of approach, but
with "artificial" statistics, has been the random perturbation/random
force/Markovian random walk/Monte Carlo/Langevin equation methods  that
have been revived and extended in recent years.   The basic idea is to
compute the velocity of particles in a turbulent flow assuming for each
particle the velocity of the previous time step  reduced  by whatever is  the
drop in the autocorrelation function for the time step At plus a random
velocity perturbation.  This is provided by  a random number generator
                                   199

-------
in a computer, of course,  and  is  distributed  in  a  way  that maintains the
proper average velocity statistics  for the flow.   Many results  can  be
produced by mathematical manipulation  of the  basic equations.   Monte Carlo
diffusion simulations, using thousands of computed particle  trajectories,
have produced further results, such as peak-to-mean concentrations, relative
diffusion, and relationships between t(_ and tp.  Methods  have been  developed
to allow application of these  techniques to vertical diffusion  i.e., with
inhomogeneous turbulence.   For instance, the  technique has been applied
to vertical diffusion in the CRL  with  good simulations of experimental
results.  However, such applications require  quite a few  assumptions
about the turbulence statistics and particle  behaviors at boundaries.
It would seem that these methodologies need considerably  more testing
against research-grade observations, perhaps  wind  tunnel  observations,
before they can be relied  upon as general  purpose  research tools.

NEEDS FOR DIFFUSION EXPERIMENTS

     Without question, there is a broad consensus  that we must  broaden
the experimental base if we are to make progress towards  better diffusion
modeling.  The many flaws  and  omitted  measurements in  most past experiments
make them impossible to use for even-handed comparisons with both  old  and
new modeling approaches, although they are useful  as auxiliary  tests of
certain models.  The field experiments with adequate,  or  nearly adequate,
meteorological measurements were too limited  in  distance  range  or  stability
range to provide for most  of our model validation  needs and  usually had
only surface concentration measurements.  Loss of  tracer  by  deposition was
                                   200

-------
a problem in the great majority of experiments.   Laboratory modeling  of
diffusion appears to be capable of fairly accurate simulations  of  most
ABL diffusion aspects and could be used to a  much greater  degree to
systematically explore the effects of source  height and  changed surface
conditions.

     The usefulness of any diffusion  field experiment  is determined by
the completeness of the accompanying  meteorological  measurements as
much as by the scope of concentration measurements.  Most  past  experiments
included only a selection of meteorological parameters,  so they cannot be
tested against most modeling approaches.  A complete set of meteorological
parameters includes (1) variances of  vertical  and lateral  velocities  or
wind directions, 
-------
For experiments in stable conditions, the  slope of the site is important,
and in uneven areas the topography  surrounding the site may be important.
The site description should  include inhomogeneities  such as tree lines
and varied surface cover.  For  the  majority  of past  experiments, the
Pasquill category cannot be  determined  because solar insolation was not
measured and no notation was made of cloudiness.  Such simple measurements
should not be neglected.  For alternative  schemes, these measurements of
net radiation and some notation or  measure of surface humidity conditions
should also be included.

     In all conditions, plume concentration  measurements are needed to
larger distances.  Some vertical  profiles  of concentration, even if very
sparse horizontally, are greatly needed to at least  the distance where
the plume fills the boundary layer  (from about 10 km in very convective
conditions to perhaps 100 km in neutral  conditions). Lateral plume
width, or ay, measurements are  needed to distances of the  order of 100 km;
even if measured at only a few  distances,  it seems that interpolation
through large distance factors  would be preferable to the  present practice
of extrapolating from 3 to 100  km.   Three-dimensional plume concentrations
can now be obtained by lidar within a few  kilometers of the source, and
excellent tracers with negligible deposition loss now exist that can be
distinguished from background hundreds  of  kilometers away.

     For stable conditions,  in  SBLs, the greatest need at  present is for
both boundary layer measurements and a  few diffusion experiments at a
variety of site types.  Almost  all  that we know now  pertains to flat
                                   20?

-------
sites or to sites with very slight slopes,  ~0.1°;  even  this  much  slope
has been observed to cause quite significant effects  on h  and  on  the
turning of wind with height and time.   Because of  the great  significance
of h to diffusion modeling, it would be very helpful  to have a "traveling
boundary-layer experiment" (even without diffusion measurements)  to
measure h at sites of differing character - broad  slopes,  short slopes,
hilly, rolling, etc.  To correlate measured h with parameters  measureable
at the surface, we would also want u*,  L, IT, z0, net  radiation, cloudiness,
and humidity.  A portable instrumented  tower, a tethersonde, and  an
acoustic sounder would be adequate basic tools for such experiments.  We
also need to develop some sort of climatology of nighttime 
-------
extend to about 25 km.  There is  especially  a  need  for turbulence
measurements aloft in the late afternoon,  as we  have  almost no  informa-
tion on CRL collapse; in addition to  the usual CRL  parameters,  these
measurements aloft should be accompanied by  humidity  profiles,  as radiation
flux divergence may greatly affect the process.   Finally, as clouds are
rather common in afternoon CRLs,  we need to  stop avoiding them  and make
turbulence measurements when clouds are present  in  varying degrees to see
how correlations with CRL parameters  are affected.   In overcast, daytime
conditions, there may be more cooling at the top of cloud layers than
heating at the surface, and heat  fluxes at the surface and at the cloud
base may be in delicate balance.   These conditions  need  to be measured
at the same time as other CRL turbulence measurements.   It is likely that
some of these ARLs would turn out to  be "neutral",  which is good because
there is so little information on the upper  part of NRLs and the effect
of Zif/u* on their turbulent structure. These latter tasks could be
performed by research aircraft.

     There have been many calls for studies  of concentration fluctuations,
X1, and their statistics.  These  are  needed  for  rational evaluation
of modeling uncertainties, for determining sampling time effects, and for
flamability and toxicity predictions  for some  gases.  "Raseline" experiments
on X'/X" could most easily be performed in  a  wind tunnel, e.g.,  for an
elevated and a surface source, over a rough  and  a smooth floor, and in
unstable, neutral, and moderately stable conditions.  For CRLs, such
measurements could also be made in convective  tanks,  as  some have already.
In addition, because of uncertainties introduced by laboratory  scale-down

                                   204

-------
and the finite sample and probe sizes,  some X1  measurements  should  be
made in conjunction with full-scale diffusion  experiments, even  if  only
at a few locations per experiment.

     Large voids exist in our understanding of  and  appropriate  modeling
of diffusion during stability transition,  calms,  and rain.   It would be
helpful, in the course of diffusion experiments in progress, to  include a
few runs that span the transition  from  unstable to stable conditions
near sunset.  Inversion breakup fumigation has received  some laboratory
tank study, but no full-scale experiment has been attempted with  full CBL
turbulence and profile development  measurements.  Calms  are difficult
to find in the atmosphere, especially during short periods of readiness
to launch a field experiment, but they  do  occur.  It seems likely that
some case-study calm periods  could  be found among existing power  plant
sn£ and meteorological monitoring archives;  this, at least, is a  place to
begin.  There is a real  need  to study calms, as near-zero wind periods
during unstable conditions have produced "worst case"  surface concentrations
at some power plants.  A study of this  phenomenon could  be made  in  a
convective laboratory tank, but it  probably would have to be two  or three
times wider than the one used up to now.   As for  rain, we have usually
assumed that passive diffusion proceeds as it  does in  dry, neutral  condi-
tions, but there is no evidence that this  is so; often,  showery conditions
are accompanied by convective downdrafts produced by evaporative  cooling.
                                   205

-------
RESEARCH MODELING NEEDS

     Because research diffusion  models  are numerous  and  seem  somewhat
ahead of the game at present in  that there are generally inadequate
diffusion observations for proper model  validations, we  have  chosen  to
somewhat emphasize the experimental  needs.  We believe that this  reflects
the prevailing opinion in the scientific community at present.  However,
we do this with expectations that research model  development  will  continue
and will improve as better experimental  data become  available.  Perhaps
some approaches will stand out as more  clearly rewarding when this happens.
In the meantime, there are some areas in which we see more  immediate
benefits possible from some of the models.

     We are beginning to understand the rudiments of SRL turbulence
structure, but have not yet performed diffusion experiments concurrently
with needed basic measurements like h.   It seems  that second-order closure
modeling of SRLs so far has produced very good simulations  of SRL evolution
in time and of slope effects, at least  on broad slopes.   Surely these models
could also be used to advantage to explore vertical  diffusion in  such
boundary layers (these models presently only compute the vertical  dimension)
Slope effects, radiative cooling effects (probably small),  and  zs/h
effects could be readily explored, given adequate computer  time.   These
models will also give the wind direction shear, but  they cannot be used
to predict lateral diffusion.  For ay,  further use could be made  of
statistical theories, including random  perturbation  modeling, especially
with good wind direction fluctuation measurements.
                                   20fi

-------
     For CBLs, it seems possible that radiative cooling could  be added  to



a LES model.  In conjunction with diminishing H* at the lower  boundary,



such a model could simulate CBL collapse, at least  until  the big eddies



are no longer present.   The overcast boundary layer, with various balances



between surface H* and  cloud base H*, could  also be simulated  with LES



models.  One advantage  of these models is that they can be readily used



to predict vertical  and lateral  diffusion from any  height within the



resolution of the grid  cells; all  it takes is sufficient  computer time  to



compute numerous particle trajectories.  A more speculative possibility



for these models is  simulation of turbulence and diffusion in  CBLs topped



by scattered clouds, which form at the top of more  vigorous thermals;



this would probably  require some form of cloud updraft  parameterization



and some accounting  of  subsidence outside the clouds.





TECHNICAL TRANSFER NEEDS





     Even without further research advances, we have gained so much new



knowledge about boundary layer turbulence structure and diffusion since



1968 that it is greatly perplexing to many scientists that so  little of



this has filtered down  to applied modeling.   Of course, this is not an



automatic process.  Research models contain  some parameters that are



difficult to measure.  The models are often  complex entities that do not



lend themselves well  to routine usage, at least not in  their original



format.  Furthermore, many operational modelers do  not  feel  inclined to



switch from the old  schemes until  they are proven inferior to  an alterna-



tive scheme. This requires, first of all, support of a scientific effort
                                   207

-------
to validate, simplify, revalidate,  and  then  package  preferable models  in
a way that simplifies application  and  requires  only  minimal meteorological
and source inputs.

     In Section 7, we reviewed  efforts  that  have  been made to find adequate
ways to estimate the required,  difficult meteorological  parameters by  using
simpler meteorological measurements.   A rather  large array of methods  has
been suggested, but, in most  cases  each method  has only  been tested with
data from one site.  This is  inadequate, especially  in stable conditions.
One reason that full-meteorology diffusion  experiments have been  recommended
here is that they allow any number  of  schemes for input  simplification to
be tested both against the fundamental  ABL  parameters, such as L, and
diffusion data in a modeling  context.   Furthermore,  any  applicable model
or model simplification can be  tested  against the same diffusion  data.
This is rarely achievable with  past experiments,  except  for very  recent
ones.  Only testing of old and  new models  side  by side against "universal"
data sets like the above will  settle the question of how much modeling
improvement, if any, can be obtained for various  levels  of effort.
     Apart from diffusion experiments,  traveling  boundary-layer  experiments
like the one suggested here for SBLs would  settle the question of which
input simplification methods work  best  at  sites of various characteristics.
Any modeling approach, old or new,  will work vastly  better if we  can make
a better estimate of the turbulent  mixing  depth,  h or z-j.
                                   208

-------
     Such efforts will  bear fruit,  to the satisfaction  of research
scientists and model-users alike,  only after some period  of sustained
support for researchers who can concentrate on  the job  of transforming
validated research diffusion models into  efficient and  more accurate
operational models than those in common use.
                                   209

-------
                                REFERENCES


Agterberg, R., F. T. M. Nieuwstadt,  H.  van  Duuren,  A.  J.  Hasselton,
     and G. D. Krijt, 1983:   Dispersion Experiments with  Sulphur  Hexa-
     fluoride from the 213m  HighMeteorological  Mast at Cabauw in the
     Netherlands.Royal  Netherlands Meteorological  Institute, DeBilt,
     The Netherlands, 129 pp.

Albrecht, B. A., R. S. Penc, and W.  H.  Shubert,  1985:  An Observational
     Study of Cloud Topped Mixed Layers. J. Atmos.  Sci.,  42:  800-822.

Arya, S. P. S., 1981:  Parameterizing the Height  of the Nocturnal  Boundary
     Layer.  In: Fifth Symposium on  Turbulence,  Diffusion,  and Air Pollution,
     Amer. Meteor. Soc. Boston,  MA,  pp. 111-112.

Baerentsen, J. H., and R. Berkowicz, 1984:   Monte Carlo Simulation of  Plume
     nispersion in the Convective Boundary  Layer.  Atmos. Environ.,18:
     701-712.

Barad, M. L., 1958:  Project Prairie Grass  - A  Field Program  in Diffusion,
     Volumes I and IT.  Geophysical  Research Paper  No. 59.  Air Force
     Cambridge Research Center,  Bedford, MA. [NTID  PB151424 and PB1514251.

Batchelor, G. K., 1949:  Diffusion in a Field of Homogeneous  Turbulence,
     I. Eulerian Analysis.  Aust. J. of Sci. Res.,2: 437.

Batchelor, G. K., 1952:  Diffusion in a Field of Homogeneous  Turbulence,
     II. The Relative Motion of  Particles.   Proc. Cambridge.  Phil. Soc.,
     48:345.

Berkowicz, R., and L. P.  Prahm,  1980:  On the Spectral Turbulent  Diffusivity
     Theory for Homogeneous  Turbulence.  J. of  Fluid Mech.,100: 433-448.

Berkowicz, R., and L. P.  Prahm,  1982:  Evaluation of the  Profile  Method  for
     Estimation of Surface Fluxes of Momentum and Heat.   Atmos. Environ.  16:
     2809-2819.

Binkowski, F. S., 1983:  A Simple Model for the Diurnal Variation of the
     Mixing Depth and Transport  Flow.  Boundary-Layer  Meteor., 27-236.

Bowne, N., 1960:  Measurements of Atmospheric Diffusion From  an Elevated  Source.
     Proc. 6th A.E.C. Air Cleaning Conference,  Atomic  Energy  Commission,
     76-88. [NTIS TID-75931.

Briggs, G. A., 1981:  Canopy Effects on Predicted Drainage Flow Characteristics
     and Comparisons with Observations.  In: Fifth  Symposium  on Turbulence,
     Diffusion and Air Pollution, Amer. Meteor. Soc.,  Boston, MA, pp.  113-115.

Briggs, G. A., 198?a:  Simple Substitutes for the Obukhov Length. Third
     Joint Conference on  Application of Air Pollution  Meteorology,
     Amer. Meteor. Soc.,  Boston, MA, pp. 68-71.

                                   210

-------
Briggs, G. A., 1982b:   Similarity Forms for Ground-Source Surface-Layer
     Diffusion.  Boundary-Layer Meteor.,23: 489-502.

Briggs, G. A., 1985a:   Diffusion Modeling with Convective Scaling and  Effects
     of Surface Inhomogeneities.  Specialty Conference on Air Quality
     Modeling of the Urban Boundary Layer, Amer.  Meteor.  Soc.,  Boston,  MA,
     (in press).

Briggs, G. A., 1985b:   Analytical  Parameterization  of Diffusion:  the Convective
     Boundary Layer.  J. of Climate and Appl.  Meteor, (in press).

Briggs, G. A., and K.  R. McDonald, 1978:  Prairie Grass Revisited:  Optimum
     Indicators of Vertical  Spread.  Proceedings  of the Ninth International
     Technical Meeting on Air Pollution Modeling  and  Its  Application,  No. 103,
     NATO/CCMS, pp. 209-220.

Rrost, R., and J. C. Wyngaard, 1978:   A Model  Study of the Stably-Stratified
     Planetary Boundary Layer.  .1. Atmos. Sci .,35:  1427-1440.

Businger, J. A., .). C. Wyngaard, Y. Izumi, and E. F.  Bradley, 1971:  Flux-
     profile Relationships in the Atmospheric  Surface Layer.  J. of  Atmos.
     Sci., 28: 181-189.

Busse, F. H., 1970:  Bounds for Turbulence Shear  Flow. J. Fluid  Mech.
     41: 219-240.

Calder, K. L., 1949:  Eddy Diffusion  and Evaporation  in Flow  over Aerodyna-
     mical ly Smooth and Rough Surfaces: A Treatment Based on  Laboratory
     Laws of Turbulent Flow with Special Reference  to Conditions  in  the
     Lower Atmosphere.  Quart. J. Mech. and Appl. Math.,II; 153.

Carpenter, S. B., T. L. Montgomery, J. M. Leavitt,  W. C.  Colbaugh,  and
     F. W. Thomas, 1971: Principal °lume Dispersion Models: TVA Power
     Plants. J. Air Pollut. Control Assoc., 21: 491-495.

Carras, J. N., and D.  J. Williams, 1981:  The  Long  Range  Dispersion  of  a
     Plume from an Isolated Point Source.  Atmos. Environ., 15: 2205-2217.

Carras, J. N., and D.  J. Williams, 1983:  Observations of Vertical  Plume
     Dispersion in the Convective Boundary Layer.  Sixth  Symposium  on
     Turbulence and Diffusion, Amer.  Meteor. Soc.,  Boston, MA,  pp.  249-252.

Caughey, S. J., 1982:   Observed Characteristics of  the Atmospheric  Boundary
     Layer. In: Atmospheric Turbulence and Air Pollution  Modeling,
     F. T. M. Nieuwstadt and H. Van Dop, eds., n. Reidel  Publishing  Company,
     Dordrecht, Holland, pp. 107-158.

Caughey, S. J., and S. G. Palmer, 1979:  Some  Aspects of  Turbulence  Structure
     Through the Depth of the Convective Boundary Layer.   Quart.  J.  Roy.
     Meteor. Soc.,105: 811-827.


                                   211

-------
Caughey, S. J., J. C. Wyngaard, and J. C. Kaimal , 1979:   Turbulence in the
     Evolving Stable Boundary Layer., Atmos. Sci .,6:  1041-1052.

Cermak, J. E., 1963:  Lagrangian Similarity Hypothesis Applied to Diffusion
     in Turbulent Shear Flow.  J. Fluid Mech.,15: 49-64.

Cermak, J. E., P. K. Shirvastava, and M. Poreh, 1983:   Wind Tunnel  Research
     on the Mechanics of Plumes in the Atmospheric Surface Layer: Part I.
     (ER 83-84JEC-PKS-MP12), Colorado State University,  Fort Collins, CO,
     140 pp.

Chaudhry, F. H., and R. N. Meroney, 1973:  Similarity  Theory of Diffusion and
     the Observed Vertical Spread in the Diabatic Surface Layer.  Boundary-
     Layer Meteor.,3: 405-415.

Ching, J. K. S., ,J. F. Clarke, J. S. Irwin and J. M.  Godowitch, 1983:
     Relevance of Mixed Layer Scaling for Daytime Dispersion Rased on
     RAPS and other Field Programs.  Atmos. Environ.,17:  859-871.

Christensen, n., and L. P. Phahm, 1976:  A Pseudospectral  Model for
     Dispersion of Atmospheric Pollutants.  J. Appl.  Meteor.,15: 1284-1294.

Counihan, J. 1975: Adiabatic Atmospheric Roundary Layers:  A Review and
     Analysis of Data from the Period 1880-1972.   Atmos.  Environ.,9:  871-905.

Cramer, H. E., 1957:  A Practical Method of Estimating the Dispersal  of
     Atmospheric Contaminants.  Proceedings of the Conference on Applied
     Meteorology.  Amer. Meteor. Soc., Roston, MA, pp. 33-55.

Cramer, H. E., 1976:  Improved Techniques for Modeling the Dispersion of
     Tall Stack Plumes.  Proceedings of the 7th International Technical
     Meeting on Air Pollution Modeling and Its Application, No. 51,
     NATO/CCMS, pp. 731-780.

Cramer, H. E., F. A. Record, and H. C. Vaughan, 1958:   The Study of the
     Diffusion of Rases of Aerosols in the Lower  Atmosphere. AFCRL-TR-58-239,
     MIT, 133 pp. [NTIS AD 152 5821.

Davis, P. A., 1983:  Markov Chain Simulations of  Vertical  Dispersion from
     Elevated Sources into the Neutral Planetary  Roundary Layer.  Roundary-
     Layer Meteor., 2fi: 355-376.

Deardorff, J. W., 1970a:  Preliminary Results from Numerical Integrations
     of the Unstable Roundary Layer.  .1. Atmos. Sci.,  27;  1209-1211.

Deardorff, >1. W., 1970b:  Convective Velocity and Temperature Scales for
     the Unstable Roundary Layer for Rayleigh Convection.   .1. Atmos. Sci.,
     27: 1211-1213.

Deardorff, J. W., 1972:  Numerical Investigation  of Neutral and Unstable
     Planetary Roundary Layers.  J. Atmos. Sci.,  29:  91-115.

                                   212

-------
Oeardorff, J. W., 1973:  Three Dimensional  Numerical  Study of Turbulence
     in an Entraining Mixed Layer.  Boundary-Layer Meteor. 1: 169-196.

Deardorff, J. W., 1974:  Three Dimensional  Numerical  Study of the Height
     and Mean Structure of a Heated Planetary Boundary Layer.  Boundary-
     Layer Meteor.,7: 81-106.

Deardorff, J. W., 1984:  Upstream Diffusion in the Convective Boundary
     Layer with Weak or Zero Mean Wind.   Fourth Joint Conference on
     Applications of Air Pollution Meteorology, Amer. Meteor. Soc.,
     Boston, MA, pp. 4-7.

Deardorff, J. W., and G. E. Willis, 1974:   Computer and Laboratory Modeling
     of the Vertical Diffusion of Nonhuoyant Particles in the Mixed Layer.
     In: Advances in Geophysics, 18B, H.  E. Landsberg and J.  van Mieghem, eds.,
     Academic Press, New York, pp. 187-200.

neardorff, J. W., and G. E. Willis, 1975:   A Parameterization of Diffusion
     into the Mixed Layer.  J. Appl. Meteor., 14:  1451-1458.

Deardorff, J. W., and G. E. Willis, 1982:   Ground  Level  Concentration Due to
     Fumigation into an Entraining Mixed  Layer.  Atmos. Environ., 16: 1159-
     1170.

Deardorff, J. W., and G. E. Willis, 1984:   Ground-level  Concentration
     Fluctuations from a Buoyant and a Non-Buoyant Source within a Laboratory
     Convectively Mixed Layer.  Atmos. Environ., 18:  1297-1309.

Deardorff, J. W., and G. E. Willis, 1985:   Further Results from  a Laboratory
     Model of the Convective Planetary Boundary Layer.  Boundary-Layer
     Meteor.,  32: 205-236.

DeBaas, A. F., and A. G. M. Driedonks, 1985:  Internal  Gravity Waves in  a
     Stably Stratified Boundary Layer. Boundary-Layer Meteor.,  31: 303-323.

Doran, J. C., T. W. Horst, and P. W. Nickola, 1978a:   Variations in Measured
     Values of Lateral  Diffusion Parameters.  J. Appl. Meteor.,  17: 825-831.

Draxler, R. R., 1976:  Determination of Atmospheric Diffusion Parameters.
     Atmos. Environ.,10: 99-105.

Draxler, R. R., 1984:  Diffusion and Transport Experiments.  In: Atmospheric
     Science and Power Production, Darryl  Randerson,  ed.   U.S. Dept. of
     Energy, NTIS No. DE84005177, pp. 367-422.

Driedonks, A. B. M., 1982:  Models and Observations of the Growth of the
     Atmospheric Boundary Layer.  Boundary-Layer Meteor.,23:  283-306.
                                   213

-------
Dn'edonks, A. G. M.,  H. Van Dop,  and W.  Kohsiek,  1978:   Meteorological
     Observations on  the 213m Mast at Cabauw in The Netherlands.
     Fourth Symposium on Meteorological  Observation and  Instrumentation.
     Amer. Meteor. Soc., Boston,  MA, pp. 41-46.

Druilhet, A., J. P. Frangi, D. Guedalia, and J. Fontan,  1983:   Experimental
     Studies of the Turbulence Structure Parameters of the Convective
     Boundary Layer.   J. Climate  Appl. Meteor.,22:  594-608.

Dyer, A. J., 1965:  The Flux-Gradient Relation  for  Turbulent  Heat  Transfer
     in the Lower Atmosphere.  Quart. J. Roy. Meteor. Soc., 91:  151-157.

Eberhard, W. L., W. R. Moninger,  T. llttal,  S. W.  Troxel, and  J.  E. Gaynor,
     and G. A. Briggs, 1985:   Field Measurements  in Three Dimensions  of
     Plume Dispersion in the Highly Convective  Boundary  Layer.   Seventh
     Symposium on Turbulence and  Diffusion.  Amer.  Meteor. Soc.,  Boston,
     MA, (in press).

Einstein, A., 1905:  Uber die van Molekular Kinetischen  Theorie  der Warme
     Geforddr Bewegung von in Ruhenden Flussigkeiten Suspendierten Teilchen.
     Ann. Phys. 17: 549.

Elderkin, C. E., W. T. Hinds, and N. E.  Nutley, 1963:   Dispersion  from
     Elevated Sources.  Hanford Radiological  Sciences Research  and Develop-
     ment Annual Report for 1963, 1.29-1.36. [NTIS  HW-81746].

Ellison, T. H., and J. S. Turner, 1959:   Turbulent  Entrainment  in  Stratified
     Flows.  J. Fluid Mech.,6; 423-448.

Ferber, G. J., and R. J. List, 1973:  Long-Range  Transport and  Diffusion
     Experiments, NOAA Air Resources Laboratories,  Silver Spring,  MD, unpublished,

Fiedler, B. H., 1984:  An Integral  Closure  Model  for the Vertical  Turbulent
     Flux of a Scalar in a Mixed  Layer.  J.  Atmos. Sci.,41: 674-680.

Fox, D. G., 1981:  Judging Air Ouality Model  Performance.  Summary of the
     Amer. Meteor. Soc. Workshop  on Dispersion  Model Performance.   Bull.
     Amer. Meteor. Soc., 62:  599-608.

Fox, D. G., n. Randerson, M.  E. Smith, F. D. White, and  J. C. Wyngaard, 1983:
     Synthesis of the Rural Model Reviews.   Environmental Sciences Research
     Laboratory, U.S. Environmental Protection  Agency,  Research  Triangle
     Park, NC.  4? pp.

Fuquay, J. J., C. L.  Simpson, and W. T.  Hinds,  1964:  Prediction of Environ-
     mental Exposures from Sources near  the Ground  Based on Hanford Experi-
     mental Data. J.  Appl. Meteor., 3, 761-770.

Garratt, .]. R., 1982:  Observations in the  Nocturnal Boundary Layer.
     Boundary-Layer Meteor.,  22:  21-48.


                                   214

-------
Garratt, J. R., and R. A. Brost, 1981:   Radiative Cooling Effects within and
     above the Nocturnal  Boundary Layer.  .1. Atmos. Sci.,38:  2730-2746.

Giblett, M. A., et al., 1932:   The Structure of Wind over Level  Country.
     Meteorological Office Geophysical  Memoirs No. 54.

Gifford, F. A., 1961:   Use of  Routine Meteorological Observations for
     Estimating Atmospheric Dispersion.  Mud. Safety.2:  47-51.

Gifford, F. A., 1962:   Diffusion in the Diabatic Surface Layer.  J. Geophys.
     Res.,67:3207.

Gifford, F. A., 1975:   Atmospheric Dispersion  Models for Environment Pollution
     Applications.  In: Lectures on Air Pollution and Environment Impact
     Analysis.  Amer.  Meteor.  Soc., Boston, MA, pp. 35-58.

Gifford, F. A., 1976:   Turbulent Diffusion Typing Schemes: a  Review. Nucl.
     Safety.,17: 68-86.

Gifford, F. A., 1982:   Horizontal  Diffusion in the Atmosphere: A Lagrangian-
     Dynamical Theory.  Atmos. Environ.,16: 505-512.

Gifford, F. A., 1983:   Atmospheric Diffusion in the Mesoscale Range: The
     Evidence of Recent Plume  Width Observations.  Sixth  Symposium on
     Turbulence and Diffusion.  Amer. Meteor.  Soc.  Boston, MA,pp. 300-304.

Golden, D., 1972:   Relations among Stability Parameters  in the Surface Layer.
     Boundary-Layer Meteor.,3: 47-58.

Gryning, S. E., 1981:   Elevated Source SF^-Tracer Dispersion  Experiments in
     the Copenhagen Area.  RISO-R-446.   Riso National  Laboratory, Denmark.

Gryning, S. E., and E. Lyck, 1984:  Atmospheric Dispersion from  Elevated
     Sources in an Urban  Area:  Comparison between Tracer Experiments and
     Model Calculations.   J. Climate Appl. Meteor.,23: 651-660.

Gryning, S. E., P. Van Mlden,  and S. E. Larson, 1983:   Dispersion from a
     Continuous Ground-Level Source Investigated by a K  Model .   Quart. J.
     Roy. Meteor.  Soc.,109: 355-364.

Manna, S. R., 1981:  Turbulent Energy and  Lagrangian Time Scales in the
     Planetary Boundary Layer.  Fifth Symposium on Turbulence, Diffusion,
     and Air Pollution.  Amer. Meteor.  Soc., Boston, MA,  pp.  61-62.

Hanna, S. R., 1983:  Lateral Turbulence Intensity and Plume Meandering
     During Stable Conditions.  J. Climate Appl. Meteor.,22:  1424-1430.

Hanna, S. R., G. A. Briggs, J. Deardorff,  B. A. Egan, F.  A. Gifford, and
     F. Pasquill,  1977:  Meeting Review.  AMS  Workshop on Stablility
     Classification Schemes and Sigma Curves - Summary of Recommendations.
     Bull. Amer. Meteor.  Soc.,58:  1305-1309.


                                   215

-------
Hanna, S. R., G. A. Briggs, and R. P. Hosker, Jr., 198?:   Handbook on
     Atmospheric Diffusion.  U.S. Dept. of Energy, NTIS No. DE82002045,
     102 pp.

Haugen, D. A., Ed., 1959:  Project Prairie Grass, a Field  Program in Diffu-
     sion.  Geophys. Res.  Pap., No. 59, Vol. Ill, Rep. AFCRC-TR-5S-235,
     Air Force Cambridge Research Center, 673 pp. [NTIS PB 161 101].

Haugen, n. A., and J. J. Fuquay, Eds., 1963:  The Ocean Breeze and Dry Gulch
     Diffusion Programs, Vol. I.  Air Force Cambridge Research Laboratories
     and Hanford Atomic Products Operations, HW-78435, 240 pp.  (Available
     from Technical Information Center, P.O. Box 62, Oak  Ridge, TN.)

Hay, J. S., and F. Pasquill, 1957:  Diffusion from a Fixed Source from a
     Height of a Few Hundred Feet in the Atmosphere.  J.  Fluid Mech.,?: 299.

Hay, J. S., and F. Pasquill, 1959:  Diffusion from a Continuous Source in
     Relation to the Spectrum and Scale of Turbulence.  Atmospheric Diffusion
     and Air Pollution, F. N. Frenkiel and P. A. Sheppard, eds.,Advances in
     Geophysics, 6, Academic Press, NY, 345-365.

Hicks, B. B., 1985:  Behavior of Turbulent Statistics in  the Convective
     Boundary Layer.  J. Climate and Appl. Meteor., 24: 607-614.

Hilst, G. R., and C. L. Simpson, 1958:  Observations of Vertical Diffusion
     Rates in Stable Atmospheres.  J. Meteor., 15: 125-126.

Hino, M., 1968:  Maximum Ground-Level Concentration and Sampling Time.
     Atmos. Environ.,2: 149-165.

Hoffnagle, G. F., M. E. Smith, T. V. Crawford, and T. J.  Lockhart, 1981:
     On-site Meteorological Instrumentation Requirements  to Characterize
     Diffusion from Point  Sources - A Workshop.  Fifth Symposium on
     Turbulence, Diffusion, and Air Pollution.  Amer.  Meteor. Soc.,
     Boston, MA, pp. 10-11.

Hbgstrb'm, U., 1964:  An Experimental Study on Atmospheric Diffusion.
     Tellus.16: 205-251.

Hogstrbm, II., 1985:  Von Karman's Constant in Atmospheric Boundary Layer
     Flow: Reevaluated.  J. Atmos. Sci.,42: 263-270.

Holtslag, A. A. M., and A. P. Van illden, 1983:  A Simple  Scheme for Daytime
     Estimates of the Surface Fluxes from Routine Weather Data.  J. Climate
     Appl. Meteor.,2?: 517-529.

Holtslag, A. A. M., 1984b:  Estimates of Diabatic Wind Speed Profiles from
     Near Surface Weather Observations.  Boundary-Layer Meteor.,29; 225-250.

Horst, T. W., 1979:  Lagrangian Similarity Modeling of Vertical Diffusion
     from a Ground-Level Source.  J. Appl. Meteor.,18; 733-740.
                                   216

-------
Horst, T. W., J. C. Doran, and P.  W. Nickola, 1979:   Evaluation of Empirical
     Atmospheric Diffusion Data.   U.S. Nuclear Regulatory Commission.
     NUREG/CR-0798.  137 pp.

Hosker, R. P., Jr., 1984:  Flow and Diffusion Near Obstacles.  In: Atmospheric
     Science and Power Production.  Parry!  Randerson,  ed.  U.S. Dept. of
     Energy, NTIS No. DE84005177,  pp. 241-326.

Hunt, J. C. R., 1982:  Diffusion  in the Stable Boundary Layer.   In:  Atmospheric
     Turbulence and Air Pollution  Modeling, F. T.  M.  Nieuwstadt and  H. Van
     Hop, eds., n. Reidel Publishing Company, Dordrecht,  Holland,  pp. 231-274.

Hunt, J. C. R., J. C. Kaimal,  and  J. E. Gaynor, 1985:   Some Observations of
     Turbulence Structure in Stable Layers.  Quart. J. Roy. Meteor.  Soc.,
     Ill: 793-815.

Hunt, J. C. R., and A. H. Weber,  1979:  A Lagrangian  Statistical Analysis
     of Diffusion from a Ground-Level Source in a  Turbulent Boundary  Layer.
     Quart. J. Roy. Meteor. Soc.,105: 423-443.

Irwin, J. S., 1979:  Estimating Plume Dispersion - A  Recommended Generalized
     Scheme.  Fourth Symposium on  Turbulence, Diffusion and Air Pollution.
     Amer. Meteor. Soc., Boston, MA, pp. 62-69.

Irwin, J. S., 1983:  Estimating Plume Dispersion - A  Comparison of Several
     Sigma Schemes.  J. Climate and Appl. Meteor.,22:  92-114.

Irwin, J. S., and F. S. Binkowski, 1981:  Estimating  of the Monin-Obukhov
     Scaling Length Using On-Site  Instrumentation. Atmos.  Environ.,
     15: 1091-1094.

Irwin, J. S., S. E. Gryning, A. A. M. Holtslag, and B. Sivertsen,  1985:
     Atmospheric Diffusion Modeling Based on Boundary  Layer Parameterization.
     EPA 600/3-85/056, Atmospheric Sciences Research  Laboratory, II.  S.
     Environmental Protection  Agency, Research Triangle Park, NC,  43  pp.

Irwin, J. S., and M. Smith, 1984:   Potentially Useful  Additions to the Rural
     Model Performance Evaluation.  Bull. Amer. Meteor. Soc.,65; 559-568.

Irwin, J. S., 1984;  site to Site  Variation in Performance  of Dispersion
     Parameter Estimation Schemes.  In: Air Pollution  Modeling  and Its
     Application III, C. De Wispelaere, ed., Plenum,  New York,  NY, pp. 605-617.

Islitzer, N. F., 1981:  Short-range Atmospheric Dispersion  Measurements  from
     an Elevated Source. J. Meteor., 18, 443-450.

Islitzer, N. F., and R. K. Dumbauld, 1963:   Atmospheric Diffusion-Deposition
     Studies over Flat Terrain.  Int. J. Air Water Pollut.,7: 999-1022.
                                   217

-------
Iztimi , Y., and S. J. Caughey, 1976:   Minnesota 1973 Atmospheric Boundary Layer
     Experiment Data Report.   Air Force Cambridge Research  Laboratory,
     Bedford, MA. 28 pp.

.Johnson, W. B., 1982:  Meteorological  Tracer Techniques for Parameterizing
     Atmospheric nispersion.   Workshop on the Parameterization of Mixed
     Layer Diffusion.  20-23  October 1981, Las Cruces, NM,  pp. 123-132.

Kaimal , J. C., N. L. Abshire, R. B.  Chadwick, M.  T. Decker, W. H. Hooke,
     R. A. Kropfli, W. D. Neff, and  F. Pasqualicci, 1982:   Estimating the Depth
     of the Daytime Convective Boundary Layer. .1 . Appl . Meteor., 21:  1123-1129.

Kaimal, J. C., J. C. Wyngaard, D. A. Haugen, 0. R. Cote, Y. Izumi , S. J. Caughey
     and C. J. Readings, 1976:  Turbulence Structure in the Convective  Boundary
     Layer.  J. Atmos. Sci.,33: 2152-2169.

Kaimal, J. C., J. C. Wyngaard, Y. Izumi, and 0. R. Cote, 1972:  Spectral
     Characteristics of Surface Layer Turbulence.  Quart. J. Roy. Meteor.
         ,98: 563-586.
Kaimal, J. C., J. E. Gaynor, P. L. Finkelstein, M.  E.  Graves,  and T.  J.  Lockhart,
     1984a:  A Field Comparison of In-Situ Meteorological  Sensors.  Boulder
     Atmospheric Observatory Report Number Six. NOAA/ERL,  Wave Propagation
     Laboratory, Boulder, CO, 96 pp.

Kaimal, J. C., J. E. Gaynor, P. L. Finkelstein, M.  E.  Graves,  and T.  J.  Lockhart,
     1984b:  An Evaluation of Wind Measurements by  Four Doppler Sodars.   Boulder
     Atmospheric Observatory Report Number Five.  NOAA/ERL, Wave Propagation
     Laboratory, Boulder, CO, 110 pp.

Kantha, L. H., and R. R. Long, 1980:  Turbulent Mixing with Stabilizing  Surface
     Buoyancy Flux.  Phys. Fluids, 23: 2142-2143.

Krishna, J. and S. P. S. Arya, 1981:  Wind Structure in Neutral, Entraining
     PBL Capped by a Low-Level Inversion.  Fifth Symposium on  Turbulence,
     Diffusion, and Air Pollution.  Amer. Meteor. Soc., Boston, MA, pp.
     65-fifi.

Lamb, R. G., 1978:  A Numerical Simulation of Dispersion from  an Elevated
     Point Source in the Convective Planetary Boundary Layer.   Atmos. Environ.,
     12: 1297-1304.

Lamb, R. G., 1979:  The Effects of Release Height on Material  Dispersion in
     the Convective Planetary Boundary Layer.  Fourth Symposium on Turbulence,
     Diffusion and Air Pollution^  Amer. Meteor"!  Soc., Boston, MA,  ppT
     27-33.

Lamb, R. G., 1982:  Diffusion in the Convective Boundary Layer.  In:  Atmospheric
     Turbulence and Air Pollution Modeling.  F. T.  M.   Nieuwstadt and H. van
     Dop, eds., D. Reidel Publishing Company, Dordrecht, Holland, pp. 159-230.


                                   218

-------
Lavery, T. F., A. Bass, n. G. Strimaitis, A. Venkatram, B. R. Greene,
     P. J. Drivas, and B. Egan, 1982:   EPA Complex Terrain Model  Development
     Program:  First Milestone Report  - 1981".  EPA-600/3-82-036,  Environmental
     Protection Agency, Atmospheric Sciences Research Laboratory, Research
     Triangle Park, NC, 304 pp.

Lee, J. T., and G. L. Stone, 1983a:  The Use of Eulerian Initial  Conditions
     in a Lagrangian Model of Turbulent Diffusion.  Atmos. Environ.,
     17: 2477-2481.

Lee, J. T., and G. L. Stone, 19835:  Eulerian-Lagrangian Relationships in
     Monte Carlo Simulations of Turbulent Diffusion.  Atmos. Environ.,
     17: 2483-2487.

Lewellen, W. S., and M. Teske, 1976:   Second-Order Closure Modeling of
     Diffusion in the Atmospheric Boundary Layer.  Boundary-Layer Meteor.,
     10: 69-90.

Ley, A. J., and D. J. Thomson, 1983:   A Random Walk Model  of Dispersion in
     the Diabatic Surface Layer.  Quart. J. Roy. Meteor. Soc., 109: 847-880.

Li, W. W., and R. N. Meroney, 1984:  Estimation of Lagrangian Time Scales
     from Laboratory Measurements of  Lateral Dispersion.  Atmos.  Environ.,
     18: 1601-1611.

Lilly, D. K., 1967:  The Representation of Small-Seale Turbulence in Numerical
     Simulation Experiments.  Proceedings, IBM Sci. Comput. Symp. Environ-
     mental Sci., IBM Form No. 320-1951, pp. 195-210.

Lu, N., W. D. Neff, and .1. C. Kaimal,  1983:  Wave and Turbulence  Structure
     in a Disturbed Nocturnal Inversion.  Boundary-Layer Meteor., 26: 141-155.

Lumley, J. L., and H. A. Panofsky, 1964:  The Structure of Atmospheric
     Turbulence.  John Wiley and Sons, Inc., New York, NY, 239 pp.

Mahrt, L., and R. C. Heald, 1979:  Comments on "Determining Height of the
     Nocturnal Boundary Layer. J. Appl. Meteor.,18: 383.

Mahrt, L., J. C. Andre, and R. C. Heald, 1982:  On the Depth of the Nocturnal
     Boundary Layer.  J. Appl. Meteor..21: 90-92.

McElroy, •). L., 1969:  A Comparative  Study of Urban and Rural Dispersion.
     J. Appl. Meteor., 8: 19-31.

McElroy, J. L., and F. Pooler, 1968:   St. Louis Dispersion Study, Volume
     II, Analysis.  AP-53, U.S. Public Health Service, National Air
     Pollution Control Administration, Cincinati, Ohio, 51 pp.

Monin, A. S., 1959:  Smoke Propagation in the Surface Layer of the Atmosphere.
     In: Advances in Geophysics, 6, F. N. Frenkiel and P.  A. Sheppard, eds.,
     Academic Press, New York, pp. 331-343.

                                   219

-------
Monin, A. S., and A. M. Yaglom, 1971:   Statistical  Fluid Mechanics:
     Mechanics of Turbulence.  Vol.  1,  MIT Press,  Cambridge,  MA,  769  pp.

Monin, A. S., and A. M. Obukhov, 1954:   Basic Laws  of Turbulent  Mixing  in
     the Atmosphere Near the Ground.  Tr. AK ad.  Nauk. SSSR  Reofiz.  Inst.
     No. 24 (151): 163-187.

Moninger, W. R., W. L. Eberhard, G. A.  Briggs, R. A.  Kropfl i,  and  J.  C.
     Kaimal, 1983:  Simultaneous Radar and Lidar  Observations  of Plumes
     from Continuous Point Sources.  21st Radar Meteorology  Conference,
     Amer.  Meteor.  Soc., Boston,  MA,  pp. 246-250.

Nappo, C. J., 1984:  Turbulence and Dispersion Parameters Derived  from
     Smoke-Plume Photoanalysis.  Atmos. Environ.,18:  299-306.

Nickola, P. W., 1977: The Hanford 67-Series: A Volume of Atmospheric  Field
     Diffusion Measurements.  Battelle Pacific Northwest Laboratories,  454 pp.
     [NTIS PNL-2433].

Nieuwstadt, F. T. M., 1978:  The Computation of the Friction Velocity u*
     and the Temperature Scale T* From Temperature and Wind  Velocity  Profiles
     by Least-Square Methods.   Boundary-Layer Meteor.,14: 235-246.

Nieuwstadt, F. T. M., 1980:  Application of Mixed-Layer Similarity to the
     Observed Dispersion from  a Ground-Level Source.   J. Appl. Meteor.,
     19: 157-162.

Nieuwstadt, F. T. M., 1984a:  Some  Aspects of the Turbulent  Stable Boundary
     Layer.  Boundary-Layer Meteor.,30: 31-55.

Nieuwstadt, F. T. M., 19S4b:  The Turbulent Structure of the Stable  Nocturnal
     Boundary Layer.  J. Atmos. Sci .,41; 2202-2216.

Nieuwstadt, F. T. M., and H. Van Duuren, 1979:  Dispersion Experiments  with
     SFs from the 213m High Meteorological Mast at Cabauw in The Netherlands.
     Fourth Symposium on Turbulence, Diffusion and Air Pollution,  Amer.
     Meteor. Soc., Boston, MA, pp.  34-40.

Nieuwstadt, F. T. M., and H. Tennekes,  1981:  A Rate  Equation  for  the Nocturnal
     Boundary-Layer Height.  J. Atmos. Sci .,38: 1418-1428.

Nieuwstadt, F. T. M., and H. van Dop,  1982:  Atmospheric Turbulence  and  Air
     Pollution Modeling. F.T.M. Nieuwstadt and H. van Dop, eds., D.  Reidel
     Pub!ishing Company, Dordrecht, Holland.

Ogawa, Y., P. G. Diosey, J. llehara, and H. Ueda,  1985:  Wind Tunnel  Observa-
     tions of Flow and Diffusion under Stable Stratification.   Atmos. Environ.,
     19: 65-74.

Panofsky, H. A., and B. Prasad, 1965:   Similarity Theories and Diffusion.
     Int. J. Air and Water Pollution,  9: 419-430.


                                   220

-------
Panofsky, H. H., H. Tennekes, D. H., Lenschow, and J. C. Wyngaard, 1977:
     The Characteristics of Turbulent Velocity Components in the Surface
     Layer under Convective Conditions.  Boundary-Layer Meteor.,11:  355-361.

Pasquill, F., 1961:  The Estimation of the Dispersion of Windborne Material,
     Meteor. Mag.,90: 33.

Pasquill, F., 1971:  Atmospheric Dispersion of Pollution.  Quart. J. Roy.
     Meteor. Soc.,97: 369-395.

Pasquill, F., 1974:  Atmospheric Diffusion, 2nd Ed.,  Ellis Norwood Ltd.,
     Chichester, England, 429 pp.

Pasquill, F., 1976a:  Atmospheric Dispersion Parameters in Gaussian  Plume
     Modeling - Part II; Possible Requirements for Change in the Turner
     Workbook Values.EPA-600/4-76-030b, U.S. Environmental  Protection
     Agency, Research Triangle Park, NC, 44 pp.

Pasquill, 1976b:  The "Gaussian-Plume" Model with Limited Vertical Mixing.
     EPA-600/4-76-042, U.S. Environmental Protection  Agency, Research Triangle
     Park, NC, 13 pp.

Pasquill, F., 1978:  Atmospheric Dispersion Parameters in Plume Modeling.
     EPA-600/4-76-042, U.S. Environmental Protection  Agency, Research
     Triangle Park, NC, 13 pp.

Pasquill, F., and F. B. Smith, 1983:  Atmospheric Diffusion, Third Edition.
     Ellis Horwood Ltd, Chichester, England, 437 pp.

Pooler, F., Jr., 1966:  A Tracer Study of Dispersion  over a City.  J. Air
     Pollut. Control Assoc., 16: 677-681.

Poreh, M., and J. E. Cermak, 1984:  Wind Tunnel  Simulation of Diffusion in
     a Convective Boundary Layer. Boundary-Layer Meteor.,30:  431-455.

Prahm, L. P., R. Berkowicz, and 0. Christensen, 1979:  Generalization of
     K-Theory for Turbulent Diffusion.  Part II. Spectral Diffusivity Model
     for Plume Dispersion.  J. Appl. Meteor.,18: 273-282.

Randerson, D., 1979:  Review Panel on Sigma Computations.  Bull . Amer.
     Meteor. Soc.,60: 682-683.

Randerson D., editor, 1984:  Atmospheric Science and  Power Production.
     U.S. Dept. of Energy, NTIS No. DE84005177, 850 pp.

Rao, K. S., and H. F. Snodgrass, 1978:  The Structure of the Nocturnal Boundary
     Layer.  NOAA, Air Resources Laboratory, Oak Ridge, TN, 44 pp.

Roberts, 0. F. T., 1923:  The Theoretical Scattering  of Smoke in a Turbulent
     Atmosphere.  Proc. Roy. Soc. A, 104; 640.


                                   221

-------
Robins, A. G., 1978:   Plume Dispersion from Ground-Level  Sources  in  Simulated
     Atmospheric Boundary Layers.   Atmos.  Environ.,1?:  1033-1044.

Sagendorf, J. F., and C.  R. Dickson,  1974:   Diffusion  under  Low Windspeed,
     Inversion Conditions.  NOAA TM ERL ARL-52,  NOAA,  Air Resources  Laboratory,
     Idaho Falls, ID, 89  pp.

Sawford, B. L., 1982a:  Comparison  of Some Different Approximations  in  the
     Statistical Theory of Relative Dispersion.   Quart. J. Roy. Meteor.
     _§££.,108: 191-206.

Sawford, B. L., 1982b:  Lagrangian  Monte Carlo  Simulation of the  Turbulent
     Motion of a Pair of  Particles. Quart. J. Roy.  Meteor. Soc.,108:  207-213.

Schiermeier, F. A.,  W. E. Wilson,  F.  Pooler, J.  K.  S.  Ching, and  J.  F.  Clarke,
     1979:  Sulfur Transport and Transformation  in  the Environment  (STATE):
     A Major EPA Research Program.   Bull.  Amer.  Meteor. Soc., 60:  1303-1312.

Schiermeier, F. A.,  1984:  Scientific Assessment Document of Status  of
     Complex Terrain  Dispersion Models for EPA  Regulatory Applications.
     EPA-600/3-84-103, U.S. Environmental  Protection Agency, Atmospheric
     Sciences Research Laboratory,  Research Triangle Park, NC,  132  pp.

Simpson, C. L., 1961:  Some Measurements of the Deposition of Matter and  its
     Relation to Diffusion from a Continuous Point  Source in a  Stable
     Atmosphere.  Report  NW-69292.   Rev.,  Hanford Atomic Products  Operation,
     Hanford, Wash.

Singer, I. A., and M. E.  Smith, 19fi6:  Atmospheric  Dispersion at  Brookhaven
     National Laboratory.  Int. J.  Air and Water Pollut.  10: 125-135.

Sivertsen, B., S. E.  Gryning, A. A. Holtslag, and J. S. Irwin,  1985:
     Atmospheric Dispersion Modeling Based Upon  Boundary Layer  Parameteri-
     zation.  15th Int. Tech. Meeting on Air Pollution and Its  Application.
     St. Louis, Missouri.  16 pp.

Sklarew, R. C., and  A. Joncich, 1979:  Survey of Existing Data  Bases For  Plume
     Model Validation Studies.  Final Report, EPRI  EA  1159 TPS  78-806,
     Electric Power  Research Institute, Palo Alto,  CA, pp  .

Slade, D. H.,ed., 1968: Meteorology and Atomic  Energy, U.S.  Atomic  Energy
     Commission, Germantown, MD, 446 pp.

Smagorinsky, J., 1963:  General Circulation Experiments with the  Primitive
     Equations:! The Basic Experiment.  Mon. Wea. Rev.,91: 99-164.

Smith, F. B., 1968:   Conditioned Particle  Motion in a  Homogeneous  Turbulent
     Field.  Atmos.  Environ.,2: 491-508.
                                   222

-------
Smith, F. B., 1972:   A Scheme for Estimating the Vertical  Dispersion of a
     Plume From a Source Near Ground Level.  NATO/CCMS No. 14.  In:
     Proceedings of the Third Meeting of the Expert Panel  on Air Pollution
     Modeling, NATO/CCMS, Brussels, Belgium, pp. XVII-1 -  XVII-14.

Smith, F. B., 1981:   The Significance of Wet and Dry Synoptic Regions on
     Long-Range Transport of Pollution and its Deposition.  Atmos.  Environ.,
     15: 863.

Smith, F. B., 1983:   Discussion of 'Horizontal  Diffusion on the Atmosphere:
     A Lagrangian-Dynamical  Theory1 by F. A, Gifford, Atmos. Environ.,
     17: 194-197.

Smith, F. B., and J. S. Hay, 1961:  The Expansion of Clusters of Particles
     in the Atmosphere.  Quart. J. Roy. Meteor. Soc.,87: 82.

Smith, F. B., and D. Thomson, 1984:  Solutions of the Integral  Equation of
     Diffusion and the Random Walk Model  for Continuous Plumes  and
     Instantaneous Puffs in  the Atmospheric Boundary Layer.  Boundary-Layer
     Meteor.,30: 143-157.

Smith, M. E., 1956:   The Variation of Effluent Concentrations from  an
     Elevated Point  Source.   A. M. A. Arch. Indust. Health, 14:  56-68.

Smith, M. E., 1981:   Transport and Diffusion Modeling and  its Application.
     In: Air Quality Modeling and the Clean Air Act, Amer. Meteor.  Soc.,
     Boston, MA, pp. 228-270.

Snyder, W. H., 1972:  Similarity Criteria for the Application of Fluid
     Models to the Study of  Air Pollution Meteorology, Boundary-Layer
     Meteor..3: 113-134.

Snyder, W. H., 1981:  Guideline for Fluid Modeling of Atmospheric Diffusion.
     EPA 600/8-81-009, U.S.  Environmental  Protection Agency, Atmospheric
     Sciences Research Laboratory, Research Triangle Park, NC,  185  pp.

Strimaitis, D., G. Hoffnagle, and A. Bass, 1981:  On-Site  Meteorological
     Instrumentation Requirements to Characterize Diffusion from Point
     Sources:Workshop Report.EPA-600/9-81-020, U.S. Environmental
     Protection Agency, Research Triangle Park, NC, 124 pp.

Strimaitis, D., A. Venkatram, B. R. Greene, S.  Hanna, S. Heisler, T. F. Lowery,
     A. Bass, and B. A. Egan, 1983:  EPA Complex Terrain Model  Development:
     Second Milestone Report - 1982, EPA-600/3-83-015, U.S. Environmental
     Protection Agency, Atmospheric Sciences Research Laboratory, Research
     Triangle Park,  NC, 402  pp.

Stull, R. B., 1984:   Transilient Turbulence Theory.  Part  I: The Concept of
     Eddy Mixing Across Finite Distances.  J. Atmos. Sci., 41:  3351-3367.
                                   223

-------
Stull, R. B., 1985:   A Fair Weather Cumulus  Cloud  Classification  Scheme
     for Mixed-Layer Studies.   J.  Climate and  Appl.  Meteor.,  24:  49-56.

Stull, R. B., and E. W. Eloranta,  1984:   Boundary  Layer  Experiment  -  1983.
     Bull. Amer. Meteor. Soc.,  65:  450-456.

Stull, R. B., and T. Hasegawa,  1984:   Transilient  Turbulence  Theory.
     Part II: Turbulent Adjustment.  J.  Atmos. Sci.,  41:  3368-3379.

Sutton, 0. G., 1953:  Micrometeorology,  McGraw-Hill,  New York.

Taylor, G. I., 1921:  Diffusion by Continuous  Movements.   Proc. London.
     Math. Soc.. Ser. 2, 20: 196.

Tennekes, H., 1984:   Discussion in Boundary-Layer  Meteor., 30:  104.

Tennessee Valley Authority, 1970:   Tennessee Valley  Authority Environmental
     Research and Development Report on  Full-Seale Study of Inversion
     Breakup at Large Power Plants, Muscle Shoals, AL.

Thomson, D. J., 1984:  Random Walk Modeling  of Diffusion in Inhomogeneous
     Turbulence.  Quart. J. Roy. Meteor. Soc., 110:  1107-1120.

Thomas, P. W., and K. Nester,  1976:  Experimental  Determination of  the
     Atmospheric Dispersion Parameters over Rough  Terrain, Part  II,
     Evaluation of Measurements.Rep. KFK2286, Central  Division  of Nuclear
     Research, M. R. H., Federal Republic of Germany, 116 pp.

Thomas, P. W., W. Hubschmann,  L. A. Konig, H.  Schuttelkopf, S.  Vogt,  and
     M. Winter, 1976:  Experimental Determination  of the Atmospheric  Dispersion
     Parameters over Rough Terrain, Part I,  Measurements at the Karlshruhe
     "Nuclear Research Center,  Report KFK2285,  Central Division  of Nuclear
     Research, M.B.H., Federal  Republic  of Germany,  132  pp.

Turner, D. B., 1961:  Relationships Between  24-Hour  Mean Air  Ouality
     Measurements and Meteorological  Factors in Nashville, TN.  J.  Air
     Pollut. Control Assoc.,11: 483-489.

Turner, D. B., 1970:  Workbook  of Atmospheric  Dispersion Estimates.
     U. S. Environmental Protection Agency,  Atmospheric  Sciences  Research
     Laboratory, Research Triangle Park, NC, 84 pp.

Turner, D. B., and J. S. Irwin, 1982:  Extreme Value Statistics  Related to
     Performance of a Standard  Air Quality Simulation Model Using Data at
     Seven Power Plant Sites.   Atmos. Environ., 16:  1907-1914.

van Dop, H., F. T. M. Nieuwstadt, and J. C.  R. Hunt, 1985: Random Walk
     Models  for Particle Displacements in Inhomogeneous  Unsteady  Turbulent
     Flows.  Phys. Fluids, 28:  1639-1653.
                                   224

-------
Van Duuran, H., and F. T. M. Nieuwstadt, 1980:   Dispersion from the 213m High
     Meteorological Mast at Cabauw in the Netherlands. In: Studies in
     Environmental  Science, 8.  Elsevier, Amsterdam, The Netherlands, 77-90.

van Ulden, A. P., 1978:   Simple Estimates for Vertical Diffusion from Sources
     Near the Ground.  Atmos. Environ.,12: 2125-2129.

van Ulden, A. P., and A. A. M. Holts!ag, 1983:   The Stability of the Atmospheric
     Surface Layer  During Nighttime.  Sixth Symposium on Turbulence and
     Diffusion.  Amer. Meteor. Soc. Boston, MA.,pp. 257-260.

van Ulden, A. P., and A. A. M. Holtslag, 1985:   Estimation of Atmospheric
     Boundary Layer Parameters for Diffusion Applications.  J. Climate and
     Appl. Meteor.,(in press).

Venkatram, A., 1980:   Estimating the Monin-Obukhov Length in the Stable
     Boundary Layer for Dispersion Calculations.  Boundary-Layer Meteor.,
     19: 481-485.

Venkatram, A., 1982:   A Semi-Empirical  Method to Compute Concentrations
     Associated with  Surface Releases in the .Stable Boundary Layer.
     Atmos. Environ.,16: 245-248.

Venkatram, A., D. Strimaitis, and D. Dicristofaro, 1984:  A Semi-Empirical
     Model to Estimate Vertical Dispersion of Elevated Releases in the
     Stable Boundary  Layer.  Atmos. Environ.,18: 923-928.

Venkatram, A., D. Strimaitis, and W. Eberhard,  1983:  Dispersion of Elevated
     Releases in the  Stable Boundary Layer.  Sixth Symposium on Turbulence
     and Diffusion.  Amer. Meteor. Soc. Boston, MA,  pp. 297-299.

Vogt, K. J., and H. Geis, 1974:  Tracer Experiments on the Dispersion of Plumes
     over Terrain of  Major Surface Roughness, Report Jul-113, Central Division
     for Radiation  Protection, Kernforshungsanlage Julich GmbH, Julich,
     Federal Republic of Germany.

Voloshchuk, V. M.,  1976:  Nonlocal Parameterization of Vertical Turbulent
     Exchange in the  Surface Layer.  Meteorologiva i Gidrologivia, 6: 35-42.

Walker, E. R., 1965:   A Particulate Diffusion Experiment. J. Appl. Meteor.,
     4, 614-621.

Ueber, A. H., 1976:  Atmospheric Dispersion Parameters in Gaussian Plume
     Modeling: Part 1 - Review of Current Systems and Possible Future
     Developments.   EPA-600/4-76-030a.   U.S. Environmental Protection Agency,
     Research Triangle Park, NC, 59 pp.

Weber, A. H., K. R. McDonald, and G. A. Briggs, 1977:  Turbulence Classifi-
    cation Schemes  for Stable anH Unstable Conditions.  Joint Conference
     on Applications  on Air Pollution Meteorology.  Amer. Meteor. Soc.
     Boston, MA,  pp. 96-102.


                                   225

-------
Weil, J. C., 1977:  Evaluation of the Gaussian Plume Model  at Maryland Power
     Plants.  PPSP-MP-16.   Maryland Power Plant  Siting Program,  Department
     of Natural  Resources, Annapolis, MD.

Weil, J. C., 1979:   Assessment of Plume Rise and Dispersion Models Using Lidar
     Data. PPSP-MP-27^Maryland Power Plant Siting Program.Department of
     Natural Resources, Annapolis, MD.

Weil, J. C., 1983:   Application of Advances in Planetary Boundary Layer
     Understanding  to Diffusion Modeling. Sixth  Symposium on Turbulence
     and Diffusion.  Arner. Meteor. Soc., Boston, MA,pp. 42-46.

Weil, J. C., 1985:   Updating Applied Diffusion Models.  J. Climate. Appl.
     Meteor, (in press).

Weil, J. C., and R. P. Brower, 1982:  The Maryland PPSP Dispersion Model for
     Tall Stacks.  PPSP-MP-3fi.  Maryland Power Plant Siting Program.  Depart-
     ment of Natural  Resources, Annapolis, MD.

Weil, J. C., and R. P. Brower, 1984:  An Updated Gaussian Plume  Model for
     Tall Stacks.  J. Air Pollut. Control Assoc. 34: 818-827.

Weiringa, .]., 1976:  An Objective Exposure Correction Method for Average Wind
     Speeds Measured at a Sheltered Location.  Quart. J. Roy. Meteor. Soc.,
     102: 241-253.

Wieringa, J., 1980:  Representativeness of Wind  Observations at  Airports.
     Bull. Amer. Meteor. Soc., 51: 962-971

Wetzel, P. J., 1982:   Toward Parameterization of the Stable Boundary Layer.
     J. Appl. Meteor.,21:  7-13.

Wilde, N. P., R. B. stull, and E. W. Eloranta, 1985:  The LCL and Cumulus
     Onset.  J. Climate and Appl. Meteor.,24: 640-657.

Willis, G. E., and  J. W. Deardorff, 1974:  A Laboratory Model of the Unstable
     Planetary Boundary Layer.  J. Atmos. Sci .,31: 1297-1307.

Willis, G. E., and  J. W. Deardorff, 1976a:  A Laboratory Model  of Diffusion
     into the Convective Planetary Boundary Layer.  Quart. .). Roy. Meteor.
     _Soc.,102: 427-445.

Willis, G. E., and  J. W. Deardorff, 19876b:  Visual Observations of Horizontal
     PIaniforms of Penetrative Convection.  Third Symposium on Atmospheric
     Turbulence, Diffusion, and Air Quality.Amer. Meteor. Soc., Boston,
     MA, pp. 9-1?.

Willis, G. E., and  J. W. Deardorff, 1978:  A Laboratory Study of Dispersion
     from an Elevated Source in a Convective Mixed Layer.  Atmos. Environ.,
     12: 1305-1313.
                                   226

-------
Willis, G. E., and J. W. Deardorff, 1981:   A Laboratory Study of Dispersion
     from a Source in the Middle of the Convectively Mixed Layer.  Atmos.
     Environ.,15; 109-117.

Willis, G. E., and J. W. Deardorff, 1983:   On Plume Rise Within a Convective
     Boundary Layer.  Atmos. Environ., 17:  2435-2447.

Willis, 6. E., and N. Hukari, 1984:  Laboratory Modeling of Buoyant Stack
     Emissions in the Convective Boundary  Layer.  Fourth Joint Conference
     on Applications of Air Pollution Meteorology.Amer.Meteor. Soc.,
     Boston, MA,  pp. 24-26.

Wilson, 0. J., A. G. Robbins, and J. E. Fackrell, 1982a:  Predicting the
     Spatial Distribution of Concentration  Fluctuations from a Ground-
     Level Source.  Atmos. Environ.,16; 497-504.

Wyngaard, J. C., 1975:  Modeling the Planetary Boundary Layer - Extension
     to the Stable Case.  Boundary-Layer Meteor.,9:  441-460.

Wyngaard, J. C., 0. R. Kote, and K. S. Rao, 1974:  Modeling the Atmospheric
     Boundary Layer.  In: Advances in Geophysics, Vol. ISA: 193-211.

Wyngaard, J. C., and R.  A. Brost, 1984: Top-Down and Rottom-llp Diffusion
     of a Scalar in the  Convective Boundary Layer.  J. Atmos. Sci., 41:  102-112.

Yamada, T., and G. Mellor, 1975:  A Simulation of the Wangara Atmospheric
     Boundary Layer.  J. Atmos. Sci., 32:   2309-2329.

Yokoyama, 0., M. Gamo, and S. Yamamoto, 1977:   On the Turbulence Quantities
     in the Neutral  Atmospheric Boundary Layer.  J.  Meteor. Soc. Japan,
     55: 312-318.

Zhou, L., and H. A. Panofsky, 1983:  Wind  Fluctuations in Stable Air at  the
     Boulder Tower.  Boundary-Layer Meteor.,  25: 353-362.

Zilitinkevich, S. S., 1972:   On the Determination of the Height of the Ekman
     Boundary Layer.  Boundary-Layer Meteor.,  3: 141-145.
                                   227

-------