co
co
      UJ
      CD
Effective
Stack Height

Plume Rise
 Sl:406

-------
                SI406
               Effective
               Stack Height
               Plume Rise
                1974
UNITED STATES ENVIRONMENTAL PROTECTION AGENCY
Office of Air and Waste Management
Office of Air Quality Planning and Standards
Control Programs Development Division
Air PollutionTraining Institute

-------
This manual has been reviewed by the Control Programs
Development Division, EPA,  and approved for publication.
Approval does not signify that the contents necessarily
reflect the views and policies of the Environmental
Protection Agency,  nor does mention of trade names or
commercial products constitute endorcement or recommen-
dation for use.

-------
                            Objectives


    EFFECTIVE STACK HEIGHT/PLUME RISE
                             OBJECTIVES
To introduce the student visually  to  elevated pollutant emissions
and to present the physical principles most important in determin-
ing the rise of a plume.

To present the background assumptions of  several commonly used
equations for computing plume rise together with major advantages
and disadvantages or limitations of each.

To show visually several meteorological and topographical conditions
which limit the use of any plume rise equation.

To require the calculation of effective stack height by three
common equations so that the student  realizes what input data is
needed;  he can then compare the different answers obtained by
using the same data.  The second exercise treats Briggs' equations
in more  depth, while the third presents current EPA calculation
procedures.

To present an in-depth discourse on the development of the Briggs'
equations together with recent developments and modifications
suggested by Briggs and the Meteorology Laboratory, EPA.

-------
                             Introduction
        EFFECTIVE  STACK HEIGHT/PLUME RISE
   Effective Stack Height/Plume Rise is  a self-instructional package
designed by the Air Pollution Training Institute, Environmental Protection
Agency.  An Air Pollution Training Institute  Certificate of Completion will
be awarded if the learner achieves a satisfactory level on the problem
sets included in the package.  The suggested  involvement time is eight hours.

   The package contains:

         Plume Rise/Effective Stack Height;  a work  manual
         Flume Rise; a text by Gary A. Briggs, Ph.D.,  Research Meteorologist
              Atmospheric Turbulence and Diffusion  Laboratory, NOAA

         Plume Rise; an audio tape presentation by  Briggs
         Effective  Stack Height; an audio-slide presentation by
              James L. Dicke, Meteorologist,  Meteorology Laboratory,
              Air Pollution Training Institute, EPA

A complete listing of the components is  on page 7.

   The package consists of three exercises.   Exercise  one  is made up
of a narrated slide series and an APTI article, both entitled Effective
Stack Height and both writen by Mr. Dicke.  Exercise two is made up of
the text Plume Rise and an audio tape presentation  by  Dr..  Briggs with
accompanying lecture notes in the work manual.  Exercise three contains
a summary of Dr. Briggs'  lastest analyses and the current  EPA calculation
procedures  as stated by D.  Bruce Turner,  Environmental Applications Branch,
Meteorology Laboratory, EPA.   Problem sets conclude  each exercise.

   Those who desire a certificate should complete  each component in the
order of presentation in the exercise.  Problem sets may be submitted
individually or the three sets may be submitted on  completion of the
package.  All problem sheets must be returned to receive credit.  Adequate
space is provided for calculations.  Extra paper may be used if additional
space is required, but these calculations must also be submitted to receive
credit toward a certificate.

-------
   A critique form is provided in the appendix.   It will take but  a moment
to complete the basic questions.   The Air Pollution Training Institute
welcomes evaluation.  For extended comments,  please use the back of the
critique form.  As an effort towards  maintaining a viable and current pack-
age, it will be necessary for the participant to complete the critique
before a certificate will be awarded.

   Pre-addressed envelopes are provided.   If  these are not enclosed in the
package, please mail your problem sets and critique to:

                                  Air Pollution Training Institute
                                  National Environmental Research  Center
                                  Research Triangle Park, North Carolina  27711
                                  Attention:  Plume Rise Instructor SI - 406

-------
                           COMPONENTS
•THE PACKAGE                                                           PAGE

 A.  EFFECTIVE STACK HEIGHT	11
       An audio-slide presentation by James L. Dieke

 B,  EFFECTIVE STACK HEIGHT	13
       James L. Moke

 C.  PROBLEM SET OXE       	_,	27
 D.  PLUME RISE       -		,	37
       Gary A. Briggs,  Ph.D.

 E.  PLWiE RISE	,	39
       An audio  presentation with supplementary
       lecture materials bv G.A.
     PROBLEM  SET  TWO
 G.  SOME RECENT ANALYSES OF PLUME RISE OBSERVATIONS      	 55
       Gary A. Briggs

 H.  ESIIMAT10K OF PLUME RISE      -	77
       Bruce  Turner

 I.  PROBLEM  SET IHRE-E	SI

     APPESMX	—	37
       Critique fora
       Effective Stack Height, a cued script


 JLEARXSR RESPONSE TO THE PACKAGE

 1.  Submission of Problem Set One to AFII
 2.  Submission of Problem Set Two to APTI

 3.  Submission of Problem Set Three to APTI
     Critique of the package (submit with Problem Set Three)
>AIR POLLUTICX 1RAIXIXG INSTITUTE
   RESPONSIBILITY TO THE LEARXER

 1.  Instructor grading of the Problem Sets;  return of
       evaluation to the learner with set of  correct answers

 2,  Instructor certifying satisfactory completion.
 3.  Registrar issuing Certificate of Completion to learner

-------
Exercise  One
      A.   EFFECTIVE STACK HEIGHT
            An audio-slide presentation by  James L. Dicke

      B.   EFFECTIVE STACK HEIGHT
            James L.  Dicke

      C.  ' PROBLEM SET ONE
    Please complete the exercise in the given order.   Upon completion
of components A and B, please work the problems in set one and  forward
to the Air Pollution Training Institute.  A pre^addressed envelope  is
provided.  A critique of your calculations will be returned with a  copy
of the correct.calculations.  Please be certain that  your name  and  address
is on each sheet.  ALL CALCULATIONS MUST BE RETURNED  TO RECEIVE CREDIT.
                              Air Pollution Training Institute
                              National Environmental Research Center
                              Research Triangle Park, North Carolina 27711
                              Attention: Plume Rise Instructor SI - 406

-------
                                                                            A
                 EFFECTIVE  STACK HEIGHT

                              James  L.  Dicke *

Component A, "Effective Stack Height", is an audio-slide presentation.
It is made up of 54 35mm slides with an accompanying 50 minute narration
on cassette tape.  The slides and cassett2 tape are standard  format  and
should be suitable for any 35mm slide projector and audio cassette  recorder-
playback unit.  For your convenience, a cued script of the audio  portion of
the presentation is included in the appendix.

When you have completed the audio-slide series, please turn your
attention to Mr. Dicke's article on effective stack height (component  B).
Problem Set One, which may be forwarded toward credit for an  Air  Pollution
Training Institute certificate, completes the  exercise.
*James L. Dicke,  Meteorologist,  Meteorology  Laboratory and
 Air Pollution Training Institute,  Environmental  Protection Agency
                                                                        11

-------
                  EFFECTIVE  STACK HEIGHT

                               James L. Dicke *

    In  any  consideration  of  concentrations  downwind  from a source, it
    is  desirable  to  estimate the  effective  stack height, the height at
    which  the  plume  becomes  level.   Rarely  will this height correspond
    to  the  physical  height of the stack.

    A high  velocity  of  emission of the  effluents and a  temperature
    higher  than that of the  atmosphere  at the  top  of the stack will
    act to  increase  the effective stack height above the height of
    the actual stack.   The effect of aerodynamic downwash, eddies
    caused  by  the flow  around buildings or  the stack, and also the
    evaporative cooling of moisture  droplets in the  effluent may
    cause  lowering of the plume to the  extent  that it may be lower
    than the physical stack  height.

    FFFECT  OF  EXIT VELOCITY  AND STACK GAS TEMPERATURE

    A number of investigators have proposed formulas for the esti-
    mation  of  effective stack height under  given conditions: Davidson
    (1949), Sutton (1950), P.osanquet et al. (1950), Holland (1953),
    Priestley  (1956) .

    A recent comparison of actual plume heights and  calculations using
    six of  the available  formulas was made  by Moses and Strom (1961) .
    The formulas  used were Davidson-Bryant, Holland, Scorer, Sutton,
    Bosanquet-Carev-Halton,  and Bosanquet (1957).  They found that
    "There  is  no  one formula which is outstanding  in all respects."
    The formulas  of  Davidson-Brvant,  Holland,  Bosanquet-Carey-Halton,
    and Bosanquet (1957)  appear to give satisfactory results for many
    purposes.   It must  be pointed out that  the experimental tests made
    by  Moses and  Strom  used  stack gas exit  velocities less than 15 m/sec
    and that temperatures of the  effluent were not more than that of the
    ambient air.

    Stewart, Hale, arid Crooks  (1958) compared effective stack heights
    for the Harwell reactor  emitting radioactive Argon with computations
    using the  formula of Bosanquet et al.  (1950).   The temperature of
    the gases  was 50°C above  that of the ambient air and stack gas
    velocity was  10 m/sec.  At low wind speeds, agreement between formula
    and plume height were quite good.  At wind speeds greater than 6  m/sec
    and distances greater than 600 meters from the stack the formula  under-
    estimates  effective stack height.

    Two of the formulas for estimation of effective stack height are  given
   below.   Both the Davidson-Brvant formula and Holland's  formula fre-
    quently underestimate the effective stack height.  Therefore,  a slight
    safety factor is frequently made by using the following formulas.
*James L. Dicke, Meteorologist, Meteorology Laboratory, and
 Air Pollution Training Institute, Environmental Protection Agency
                                                                         13

-------
 Effective  Stack Height/Plume Rise
  The  Davidson-Bryant  formula  is:
                       I/A
                / v    ^
          AK  =


  Where:

    AR =   the  rise of  the plume above the
           stack

    d =   the  inside  stack diameter

    v   =   stack  gas velocity

    u =   wind speed

    AT =   the  stack gas temperature minus the
           ambient air  temperature  (°K)

    T   =   the  stack gas temperature ( K)

  Any  consistent  system of units for AH,  d, v  ,  and  u  may  be used.   It
  is recommended  that vs and u be in meters/sec  and  d  in meters  which
  will give AH  in meters.

  The  Holland stack rise equation is:

                 v d    /                        AT
         AH  = 	  I 1.5   +  2.68 X 10   p 	
                                                T
                                                s
  Where:
    AP  =  the rise of the plume above the stack
           (meters)

    v   =  stack gas velocity (m/sec)

     d  =  the inside stack diameter (meters)

     u  =  wind speed (m/sec)

     p  =  atmospheric pressure (mb)

    Tg  =  stack gas temperature (OK)

    AT  =  as in equation (1) and 2.68 X 10~3 is a constant having
                  units of (m-1 mb~l).
14

-------
                                          Effective Stack Height
                                                           (B)
It is recommended that the result from the above equation be used
for neutral conditions.   For unstable conditions a value between
1.1 and 1.2 times that from the equation should be used for AH.
por stable conditions a value between 0.8 and 0.9 times that from
the equation should be used for AH.

Since the plume rise from a stack occurs through some finite distance
downwind, these formulas should not be applied when considering effects
near the stack.  Note that these formulas do not consider the stability
of the atmosphere but only the ambient air temperature.  Actually,
stability should have some effect upon the plume rise.

Lucas et al., 1963, have tested Priestley's theory on stack rise at
two power stations.  These investigators write the formula for stack
rise:
                                         (3)
  Where:
   max
effective rise above the top of
the stack (feet)
     a   =   a variable  affected  by  lapse  rate
           and  topography  (ft2  MW~lM  sec~l)

     Q   =   the  rate  of heat  emission  from the
           stack  in  megawatts  (MW)

     u   =   the  wind  speed  (ft/sec)

  They  determined a  to be  4900  and  6200  for the  2  power  stations  under
  neutral  conditions.   Clarke  (1968)  states that if  the  exit velocity
  of the stack  gases exceeds 2000  fpm  (23mph), no  rain will enter the
  stack.   As  an example, he  found  from a ten year  summary  of U.S.
  Weather  Bureau  wind  data for  Chicago that 98%  of the days had a
  maximum  wind  velocity <_ 20 mph.


  EFFECT OF  EVAPORATIVE COOLING

  In  the washing  of  effluent gases  to  absorb certain gases before  re-
  lease  to the  atmosphere, the  gases are  cooled and  become saturated
 with water vapor.  Upon  release from the absorption tower further
  cooling  is likely  due to contact with  cold surfaces of ductwork  or
  stack.   This  causes  condensation of  water droplets in  the gas stream.
  Upon release  from  the stack,  the water  droplets  evaporate withdrawing
  the latent heat  of vaporization from the air and consequently cooling
                                                                      15

-------
 Effective  Stack Height/Plume Rise
  the plume, causing it to have negative buoyancy, thereby reducing
  the stack height.  (Scorer 1959).
  The practice of washing power plant flue gases to remove sulfur
  dioxide is practiced at Battersea  and Bankside power stations near
  London, where frequent lowering of the  plumes to ground level is
  observed.

  EFFECT OF AERODYNAMIC DOWNWASH

  The influence of the mechanical turbulence around a building
  or stack may significantly alter the effective stack height.  This
  is especially true under high wind conditions when the beneficial
  effect of high stack gas velocity  is at a minimum and the plume
  is emitted nearly horizontally.  The region of disturbed flow
  surrounds a building generally to  twice its height and 5 to 10
  times  its height downwind. Most of the knowledge about the
  turbulent wakes around stacks and buildings have been gained through
  wind tunnel studies.  Sherlock and Stalker (1940), (1941), Rouse
  (1951)j Sherlock (1951), Sherlock  and Leshner (1954), (1955), Strom
  (1955-1956),  Strom et al. (1957),  and Halitsky (1961), 0-962), (1963).
  By using  models of building shapes and stacks the wind speeds re-
  quired to cause downwash for various wind directions may be determined.

  In the use of a wind tunnel the meteorological variables that may
  most easily be taken into account  are the wind speed and the wind
  direction (by rotation of the model within the tunnel).   The plant
  factors that may be taken into consideration are the size and
  shape  of the plant building, the shape, height, and diameter of
  the stack, the amount of emission, the stack gas velocity, and
  perhaps the density of the emitted effluent.  The study of the re-
  leased plume from the model stack has been done by photography
  (Sherlock and Lesher, 1954), decrease in light beam intensity (Strom,
  1955) ,  and measurement of concentrations of a tracer gas (Strom
  et al., 1957, Halitsky, 1963).

  By determining the critical wind speeds that will cause  downwash
  from various directions for a given set of plant factors,  the
  average number of hours of downwash annually can then be calculated
  by determining the frequency of wind speeds greater than the critical
  speeds for each direction (Sherlock and Lesher, 1954).  It is assumed
  that climatological data, representative of the site considered,
  are available.

  It is  of interest to note that the maximum downwash about a rec-
  tangular structure occurs  when the direction of the wind is at
  an angle of 45 degrees from the major axis  of the structureand
  that minimum downwash occurs with wind flow parallel  to the major
  axis of the structure (Sherlock  and  Lesher, 1954;.


16

-------
                                          Effective Stack Height       (B)
It has been shown by Halitsky (1961), (1963) that the effluent
from flush openings on flat roofs frequently flows in a direction
opposite to that of the free atmosphere wind due to counter-flow
along the roof in the turbulent wake above the building.  In
addition to the effect of aerodynamic downwash upon the release
of air pollutants from stacks and buildings, it is  also necessary
to consider aerodynamic downwash when exposing meteorological
instruments near or upon buildings so that representative measure-
ments are assured.

In cases where the pollution is emitted from a vent or opening on
a building and is immediately influenced by the turbulent wake of
the building, the pollution is quite rapidly distributed within this
turbulent wake.  An initial distribution may be assumed at the
source with horizonal and vertical variances of 6y2 and 62 in the
                                                         z
form of a binormal distribution of concentrations.  These variances
are related to the building width and height.

The resulting equation for concentrations from this source has
(oy2 + 6y2)1/2  in place of a  and (crz2 + <5Z2)'/2  in place of
az in the point source equations.

EFFECT OF VERY LARGE POWER PLANTS

A power generating plant in the range of 1000-5000 megawatt capacity
emits heat to such an extent that its own circulation pattern will
be set up in the air surrounding the plant.  It is doubtful that
extrapolation of dispersion estimates from existing smaller sources
can be applied to these large plants.  Fortunately, the effluent
plume will rise far above the ground and surface concentrations of
pollutants downwind will increase by only a rather small amount
most of the time.  Such large plants will usually be engineered to
minimize the effects of the two preceding topics.  The "2 1/2" rule
will tend to eliminate downwash and the "4/3's" rule for stack gas
velocity, i.e. v  should exceed u by at least  a  third, will tend  to
eliminate entrainment of the effluent into  the wake of the stack.(Pooler,  1965)

Three weather conditions, however, can still bring ground level
fumigations:  high winds, inversion breakup, and a limited mixing
layer with light winds.   The climatolopy of these conditions will
determine the magnitude and frequency of the pollution problem.
Pooler (1965) has presented nomograms for estimating groundlevel
S02 concentrations for these three situations together with effective
stack height formulas.  Bripgs (1965) has also presented a plume
rise model which has been compared with data from various TVA
power plants.  Pooler (1967) introduced a slight modification to
one of Briggs's equation and suggests this latter equation be sub-
stituted for the Pooler (1965) inversion breakup fumigation equation.
Thus the basic equation for this important weather condition is:

                                                                      17

-------
Effective Stack Height/Plume Rise
                                       S    dz

   Where:   u  =  wind speed (m/sec)

           r  =  inside stack radium (m)

           T  =  ambient air temperature (°K)
           d6 =  potential temperature lapse rate
           dz
   Other symbols are defined as  in equation (2)  above.
  A correction term for the additive effect of multiple-stack sources
  is also presented in Pooler (1967) .

  Other industries are also utilizing tall stacks for pollutant dis-
  persion and anticipate results similar to the low measured ground
  level concentrations found near tall British power plant stacks.
  A case in point is the 1250 foot stack constructed in 1971 at a
  cost of $5.5 million to serve the Inco Copper Cliff smelter in the
  Sudbury district of Ontario,  Canada.  This gigantic stack is 116
  feet in diameter at the base, tapering to just under 52 feet in
  diameter at the top.  The interior diameter is 45 feet.
  STATE OF THE ART

  Considerable research is being conducted to further quantify the
  dilution effects of tall stacks and to develop better models for
  predicting the dispersion of power plant effluent in complex terrain
  and meteorological regimes.   An extensive series of field experiments
  is being conducted, called the Large Power Plant Effluent Study
  (LAPPES),  near Indiana,  Pennsylvania and a report has been published
  by Schiermeier and Niemeyer  (1968).  Field measurements include de-
  termining plume geometry by  laser-radar, in-plume and ground level
  S02 concentrations, vegetation damage and meteorological conditions
  during the experiments.

  A summary of recent European studies dealing with plume rise and
  stack effluent dispersion is contained in the July 1967 issue of
  the journal Atmospheric Environment.

  In addition  another recent publication which contains  a specific
  chapter  on calculating effective stack height is  the ASME  Recommended
  Guide  for the  Prediction  of  the dispersion  of Airborne Effluents(1973;.
18

-------
                                          Effective Stack Height      (B)
                                     mean velocity profile
                                                            *"   potential
                                                                flow
Typical flow pattern around a cube with one face normal to the wind.
                                                                     19

-------
 Effective Stack Height/Plume Rise
  A comprehensive literature survey in this field was conducted by
  NAPCA (EPA) and incorporated into an annotated bibliography of
  over 200 references.  The publication,  "Tall Stacks, Various
  Atmospheric Phenomena and Related Aspectsvl (1969),  includes
  articles published through mid-1968.

  A very recent series of EPA publications deal with air pollution
  aspects of emission sources.  Of particular interest is GAP Pub.
  No. AP-96, "Electric Power Production - A Bibliography with Abstracts".
  Section D on air quality measurements and Section E on atmospheric
  interactions contain many specific references to plume rise deter-
  mination, plume behavior, and pollutant concentrations associated with
  this class of sources.

  Briggs in his publication, Plume Rise (1969), has presented
  both a critical review of the subject and a series of equations
  applicable to a wide range of atmospheric and emission conditions.
  These equations are being employed by an increasing number of
  meteorologists and are used almost exclusively within EPA.  An
  important result of this study is that  the rise of buoyant plumes
  from fossil-fuel plants with a heat emission of 20 megawatts (MW) -
  4.7 x 10° cal/sec - or more can be calculated from the following
  equations under neutral and unstable conditions.


           AH  =  1.6 F1/3 u-1 x 2/3         (5)
           AH  =  1.6 F1/3 u V (10 h )2/3    (6)
  where:
           AH  =  plume rise
            F  =  buoyancy flux
            u  =  average wind at stack level
            x  =  horizonal distance downwind
                  of the stack
           h   =  physical stack height

  Equation  (5)  should be applied out to a distance of 10 hs from the
  stack;  equation  (6)  at further distances.

  The buoyancy  flux  term,  F, may be  calculated  from:
                          =  3.7 x  10-5m4/sec3[Q         (7)
                                              J
we pT               Leal/sec
20

-------
                                          Effective Stack Height      (B)
where:

     g= gravitational acceleration
    Q = heat emission from the stack, cal/sec
    c = specific heat of air at constant
        pressure
     p= average density of ambient air
     T= average temperature of ambient air

Alternatively, if the stack gases have nearly the same specific heat
and molecular weight as air, the buoyancy flux may be determined from:


        J^   g v  r2                        (8)
          T       S
           s
where the notation has been previously defined.

In stable stratification with wind equation (5) holds approximately
to a distance x = 2.4 u s~l/- where:

          g   3e   ,  a  stability parameter     (9)
          T   3z

         =  lapse rate of potential  temperature
Beyond this point the plume levels off at about

   AH  = 2.4    F   1/3                       (10)
However, if the wind is so light that the plume rises vertically,
the final rise can be calculated from:


   H = 5.0 F1/A s-3/8                         (ID

For other buoyant sources, emitting less than 20 MW of heat, a con-
servative estimate will be given by equation (5) up to a distance of:
       x = 3x                                 (12)

    where:

       x*=0.52  fsec6/5 ]  F2/5h  3/5     (13)
                   ft.  6/b
                                                                       21

-------
  Effective Stack Height/Plume Rise
  This  is  the  distance at which atmospheric turbulence begins to dominate
  entrainment.

  Anyone who is responsible  for making  plume  rise  estimates  should
  familiarize  himself thoroughly with Briggs' work.
  REFERENCES

    1.  Bosanquet, C. H., Carey, W. F. and Halton, E. M.
       Dust Deposition from Chimney Stacks.  Proc. Inst.
       Mech. Eng. 162:355-367.  1950.

    2.  Bosanquet, C. H.  The Rise of a Hot Waste Gas Plume.
       J.  Inst. Fuel.  30:197. 322-328.  1957.

    3.  Davidson, W. F.  The Dispersion and Spreading of Gases
       and Dust from Chimneys.  Trans. Conf. on Ind. Wastes.
       14 Annual Meeting, Industrial Hygiene Found. Amer. 38-55.
       November 18, 1949.

    4.  Gifford, F. A., Jr. Atmospheric Dispersion Calculations
       Using the Generalized Gaussian Plume Model.  Nuclear
       Safety. 2:56-59.  December, 1960.

    5.  Halitsky, James.  Diffusion of Vented Gas Around Buildings.
       J. of APCA. 12:2.  74-80.  February, 1962.

    6.  Halitsky, James.  Wind Tunnel Model Test of Exhaust Gas
       Recirculation at the NIH Clinical Center.  New York Univ.
       Tech. Report No. 785.1. 1961.

    7.  Halitsky, James.  Some Aspects of Atmospheric Diffusion
       in Urban Areas. Air Over Cities.  R.obert A. Taft Sanitary
       Engineering Center Technical Report A 62-5.  1962.

   8.  Halitsky, James, Gordon, Jack, Halpern, Paul, and Wu, Paul.
       Wind Tunnel Tests of Gas Diffusion From a Leak in the Shell
       of a Nuclear Power Reactor and From a Nearby Stack.  New
       York Univ.  Geophysical Sciences Laboratory Report No. 63-2.
       1963.

   9.  Halitsky,  James.   Gas  Diffusion Near Buildings,  Theoretical
       Concepts and Wind Tunnel Model Experiments  with  Prismatic
       Building Shapes.   New  York Univ.   Geophysical  Sciences
       Laboratory  Report  No.  63-3.   1963.
22

-------
                                           Effective  Stack Height       (B)
10.  Holland, J. Z.  A Meteorological Survey of the Oak Ridge
     Area.  AEC. Washington, Report ORO-99.  554-559,  1963.

11.  Lucas, D.  H., Moore, 0. J., and Spurr, G.  The Pvise of Hot
     Plumes from Chimneys.  Inst. J. Air Wat. Poll. 7:473-500.
     1963.

12.  Moses, Harry, and Strom, Gordon H.  A Comparison of Observed
     Plume Rises with Values Obtained from Well-Known Formulas.
     J.  APCA. 11:10.  455-466.  October, 1961.

13.  Priestley, C. H. B.  A Working Theory of the Bent Over Plume
     of Hot Gas.  Quart. J. Roy. Met. Soc. 82-352.  165-176.
     1956.

14.  Rouse, Hunter.  Air-Tunnel Studies  of Diffusion in Urban
     Areas.  On Atmospheric Pollution.   Meteorol.  Monogr.
     1:4.  39-41.  November, 1951.

15.  Scorer, R.. S. Natural Aerodynamics.  Pergamon.  London.
     186-217.  1958.

16.  Scorer,~ R. S.  The Behavior of Chimney Plumes.  Int.  J.
     of Air Poll. 1:3. 198-220.  January, 1959.

17.  Sherlock,  R. H., and Stalker, E. A.  The Control of Gases
     in the Wake of Smokestacks.  Mech.  Eng.  62:455. 1940.

18.  Sherlock,  R. H., and Stalker, E. A.  A Study of Flow Phenomena
     in the Wake of Smokestacks.  Univ.  of Mich.   Eng. Res. Bulletin.
     29:   1941.  49 pp.

19.  Sherlock,  R. H.  Analyzing Winds for Frequency and Duration.
     On Atmospheric Pollution.  Meteorol.  Monogr.  1:4. 42. 1951.

20.  Sherlock,  R. H., and Leshner, E. J. Role of  Chimney Design
     in Dispersion of Waste Gases.  Air  Repair.  4:2. 1-10.
     August 1954.

21.  Sherlock,  R. H., and Lesher, E. J.  Design of Chimneys  to
     Control Downwash of Gases.  Trans.  Amer. Soc.  Mech. Enp.rs.
     77:1.  1955.

22.  Stewart, N. G., Gale, H. J., and Crooks, R.  N.  The
     Atmospheric Diffusion of Gases Discharged from the
     Chimney of the Harwell Reactor BEPO.  Int. J.  Air Poll.
     1:   1/2.  87-102.  1958.

-------
 Effective  Stack Height/Plume Rise
   23.   Strom,  G.  H.  Wind  Tunnel  Scale Model  Studies  of Air  Pollution
        from Industrial  Plants. Ind. Wastes.  September -  October
        1955, November - December 1955,  January  - February 1956.

   24.   Strom,  G.  H., Hackman, M., and Kaplin. E. J.  Atmospheric
        Dispersal  of  Industrial Stack Gases Determined by Con-
        centration Measurements in Scale Model Wind Tunnel Ex-
        periments.  J. APCA.  7:3  November, 1957.

   25.   Sutton,  0. G.  The Dispersion of Hot  Gases in the Atmospheric.
        J.  Meteorol.  7:307-312. 1950.

   26.   Briggs,  G. A.  A Plume Rise Model Compared with Observations.
        J.  APCA  15:9. 433-438.  1965.

   27.   Pooler,  F., Jr.  Potential Dispersion of Plumes from Large
        Power Plants, PHS Publication No. 99-AP-16.   1965.

   28.   Pooler,  F., Jr.  Derivation of Inversion Breakup Ground Level
        Concentration Frequencies  from Large  Elevated Sources.
        Air Resources Field Research Office,  ESSA, NCAPC, Cincinnati,
        Ohio, 1967.   (Unpublished Manuscript)

   29.   Symposium  on Plume Behavior.  Air and Water Pollut.  Int. J.
        Vol. 10  Nos. 6/7, 393-409.  1966.

   30.   Moore, D.  J.  Physical Aspects of Plume Models.   Air and Water
        Pollut.  Int. J.  Vol.  10 Nos. 6/7, 411-417.  1966.

   31.   Nonhebel,  G. British  Charts for  Heights of Industrial Chimneys
        Air and  Water Pollut. Int. J. Vol. 10 No. 3,  183-189.

   32.   Gartrell,  F. E., Thomas, F. W.,  and Carpenter, S. B.  Full-
        Scale Study of Dispersion of Stack Gases - A  Summary Report.
        Chattanooga, Tennessee, 1964.  (Reprinted by  the  U.  S. Dept.
        HEW, Public Health Service.)

   33.   Carson,   J.  E.  and Moses, Harry.  Calculation  of Effective
        Stack Height.   Presented at 47th Annual Meeting. Amer.
       Meteor.   Soc. New York, Janaury 23, 1967.

   34.   Culkowski,  W.  M.  Estimating the Effect of Buildings on
        Plumes from Short Stacks.   Nuclear Safety.   8:  257-259.
        Spring 1967.

   35.   Symposium on Chimney Plume Rise and Dispersion.  Atmos.
        Environ. 1:351-440.  July  1967.
24

-------
                                          Effective Stack Height       (B)
36.   Recommended Guide for the Prediction of the Dispersion of
     Airborne Effluents.  2nd Ed. M.E. Smith, Ed. ASME,  345 E.
     47th St. New York, N. Y.  10017, 1973.

37.   Clarke, John H.  Effective Stack Design in Air Pollution
     Control, Heating, Piping and Air Condition.  125-133,
     March 1968.

38.   Tall Stacks, Various Atmospheric Phenomena and Related
     Aspects.  National Air Pollution Control Administration,
     Pub. No. APTD 69-12, May 1969.

39.   Briggs, G.  A. Plume Rise.  AEC Critical Review Series.
     1969.  Avail, as TID-25075 from CFSTI, NBS, U.S. Department
     Commerce, Springfield, Virginia 22151. $6.00.

40.   Fay, J. A., Escudier, M., Hoult, D. P.  A Correlation of
     Field Observations of Plume Rise, JAPCA Vol. 20, No.  5,
     pp. 391-397, June 1970.

41.   TVA.  Report on Full-Scale Study of Inversion Breakup at
     Large Power Plants.  Div. of Env. R&D Muscle Shoales,
     Alabama.  March 1970.

42.   Schiermeier, F. A., Niemeyer, L. E. Large Power Plant
     Effluent Study (LAPPES) Vol. 1 (1968). NAPCA Pub. No.
     APTD 70-2,  June 1970.

43.   Air Pollution Aspects of Emission Sources:  Electric Power
     Production A Bibliography with Abstracts.  EPA, Office
     of Air Programs, Pub. No. AP-96, May 1971.

44.   Moses, Harry and Kraimer, M. R. Plume Rise Determination -
     A New Technique Without Equations.  JAPCA Vol. 22,  No. 8,
     pp. 621-630, August 1972.

45.   Bowman, W.  A. and Biggs, W. G. Meteorological Aspects of
     Large Cooling Towers.  Paper 72-128 pres. APCA, June 1972.

46.   Briggs,  G.  A.  Some Recent Analyses of Plume Rise Observa-
     tions,   pp.  1029-1032 in Proceedings of the Second  Interna-
     tional Clean Air Congress.  Ed. by H. M. Epland and W. T.
     Berry.  Academic Press, New York,  1971.

47.   Briggs,  G.  A.  Discussion on Chimney Plumes in Neutral and
     Stable Surroundings.  Atmos. Environ. 6:507-510, July 1972.
                                                                      25

-------
Exercise Two
      D.  PLUME RISE
            Gary A. Briggs
      E.  PLUME RISE
            An audio presentation with supplementary
            lecture materials by G.A. Briggs

      F.  PROBLEM SET TWO
    Please complete the exercise in the given order.   Upon completion
of components D and E, please work the problems in set two and  forward
to the Air Pollution Training Institute.  A pre-addressed  envelope  is
provided.  A critique of your calculations will be returned with a  copy
of the correct calculations.  Please be certain that  your  name  and  address
is on each sheet.  ALL CALCULATIONS MUST BE RETURNED  TO RECEIVE CREDIT.
                              Air Pollution Training Institute
                              National Environmental Research Center
                              Research Triangle Park, North Carolina  27711
                              Attention: Plume Rise Instructor SI - 406
                                                                        35

-------
                                                                              D
                             PLUME RISE

                             Dr. Gary A. Briggs *

 Component D, Plume Rise, is a separate 80 page text.
 Briggs     1
 5.7, p. 58:
Briggs     has recommended the following correction be made  in  Eq.
      The constant 2.9 should be 2.4
  1.  Briggs, G. A. 1972.  Discussion on Chimney Plumes in Neutral and
     Stable Surroundings.  Atmos. Environ. 6:507-510, July 1972.
 When you have completed the text, please turn to Component E, a taped
 lecture by Dr. Briggs.  The audio cassette is standard format and should
 be suitable for any audio cassette recorder-playback unit. Please do not
 attempt to use the lecture cassette without referring to the supplementary
 written material.

 Problem Set Two,  which may be forwarded toward credit for an Air  Pollution
 Training Institute certificate, completes the exercise.
*Dr. Gary A. Briggs, Research Meteorologist
 Atmospheric Turbulence and Diffusion Laboratory,
 National Oceanic and Atmospheric Administration
                                                                         37

-------
                               PLUME  RISE

                                Gary  A.  Briggs  *

   Component  E  is  an audio  tape cassette of  a lecture  by  Dr.  Briggs.  The
   cassette is  standard  format  and  should be suitable  for any audio recorder-
   playback unit.

   Please do  not  attempt to use the lecture  cassette without  referring  to
   the  following  supplementary  material.
              momentum Flux/ IT p


              (po/p)  w02 r02


              buoyancy Flux/ TT p                     -t-   good

              g (1-pQ/p) w0 r02                     -*-   better

              g (1-mo/m) w0 r02  (T/TO)              -s-   best*

             +g QH/(  TI  Cp p T)



              Avg.  wind speed  at plume  height

              Avg.  wind speed  at stack  height
              (g/T)   39/3 z

              (g/T)  (AT/Az   +   l°C/100m)

                               From  top  of  stack  to  top  of  plume
              QH              m4      1013 mb
         TT  Cp p  T           secJ        P       MW
*Dr. Gary A. Briggs, Research Meteorologist
 Atmospheric Turbulence and Diffusion Laboratory,
 National Oceanic and Atmospheric Administration.
                                                                          39

-------
Effective Stack Height/Plume Rise
    Should heat of condensation be included in QH?
               Usually not, but —
         1.  Calculate rise Ah on basis of no condensation.

         2.  Calculate max. volume flux Vmax  =  T u (0.5 Ah)2
             (u ^0.2 Fl/4 sl/8  =  0.08  (F/sec)l/4 « 0.5 m/sec)

         3.  Calculate moisture capacity  of entrained air:
               Q    =  q ,  - q~  V
                cap   V sh    a/  max

               where q ,  is the saturation specific humidity at

               the height of the plume  (h = h  + Ah) and "q" is the
                                             S            3.
               average specific humidity through the plume rise
               layer,  q ,  is a function of the ambient temperature
                        s ri
               at height h.

         4.  Subtract Qcap from efflux  rate of water vapor, Qwv

             If remainder is positive,  assume condensation adds

             (Qwv ~ Qcap) L to heat emission QH, where L is the

             heat -of fusion.

         5.  Recalculate buoyancy parameter F and new Ah.

         6.  Start at Step 2 again, using new value of Ah.  Repeat

             calculation until desired degree of convergence is

             obtained.
40

-------
                                      Plume Rise, Lecture Notes       (E)
Should evaporative cooling be included in QJJ?







            Yes!







To be conservative, assume complete evaporation of water droplets




in plume.  Subtract Qwi L from the heat emissions, where Qwl




is the efflux rate of liquid water.






If the total buoyancy turns out to be negative, so is the plume




rise  —  the plume falls to the ground, and:
        
-------
 Effective  Stack Height/Plume Rise
    'Rule  of  lowest  plume  rise":   consequences  for buoyant  plumes







                                                      F  \ 1/3    „
    1.   Ah   =   5.0  FlM s -3/8   or   Ah  =  2.4
                                                     us
        "Calm"  formula gives  lowest  rise  if u  < ( T-Q" )     F1/^  sl/8









       At most, upper limit  is about 0.5 m/sec  (F  «  Ifl3 	3





           s  « 10-3 sec-2).






       Therefore, for practical purposes, "calm" formula only




          applies to zero wind speed.
2.  Ah  =  1.6 Fl/3 u-l (3x*)2/3  or  Ah  =   2.4









    "Neutral" formula gives lowest rise if s






        or if IT  us "1/2  >  7x*.
                                                        us
42

-------
                                       Plume Rise,  Lecture  Notes       (E)
Buoyant plume rise in stable conditions






                         F  I/3
     Since Ah  =  2.4  	       applies very widely to buoyant



     plumes in stable air, it is seen that Ah is not very



     sensitive to variations in u and s, particularly since



     they tend to be negatively correlated:  on windy nights,



     the stability is small; wind speed increases with height,



     but stability decreases.  Hence, "ballpark" estimates of



     Ah in such cases need not include u and s.  On the



     basis of TVA and Bringfelt data, I recommend:









            Ah  =  17 pl/3  (very stable or high wind)





            Ah  =  35 F1/3  (slightly stable or low wind)
            ...  where Ah is in m, and F is in
                                                                     43

-------
Effective Stack Height/Plume Rise
  Buoyant plume rise in neutral conditions
       "Plume Rise" recommends the "2/3 law" cut off at a distance
       xmax as a simple approximation to buoyant plume rise in
       neutral conditions, where xmax  =  10 hs (only_ for fossil
       fuel plants with QH > 20 MW) or xmax  =  3x*. with x* being
       a function of F and hs given by Eq. 4.35.
       Recent examination shows that, within the range of present
       data, a simple cut-off distance proportional to /]? works
       just as well.  This also avoids a problem in the case of
       very small source heights.  Specifically,
                     max
                          =  63 m
                          =  1250 ft /QH/(106 cal/sec)
       This results in a "final" rise
                    Ah  =  25  (sec-m -2/3)  F2/3/u
   Note:
the ~K   =10
     max
h  formula
 s
   was  NOT intended  to  be  applied
   to gas turbine  plumes.
 44

-------
                                        Plume Rise,  Lecture Notes
(E)
Comparison of 10 hs,  3x* (Eq. 4.35) and 1250 >'QH for Table 5.1 of
"Plume Rise".
     SOURCE


  Harwell

  Bosanquet

  Darmstadt

  Duisburg

  Tallawarra

  Lakeview

  Barley



  Castle Donington



  Northfleet



  Shawnee

  Colbert

  Johnsonville

  Widows Creek

  Gallatin



  Paradise
10hs (ft)
N.A.
N.A.
N.A.
N.A.
N.A.
4930
N.A.
2500
4250
4250
4900
4900
2500
3000
4000
5000
5000
5000
6000
3x* (ft)
1110
1450
1140
2100
2040
4890
1450
2300
4500
5100
4200
5000
2400
2920
4200
5700
5SOO
4400
7100
1250 ,'QH (ft)
1310
1550
1150
1700
2140
4250
1550
2700
4300
5000
3500
4300
2900
3250
4100
5100
5100
3650
5800
                                                                       .45

-------
Exercise Three
      G.  SOME RECENT ANALYSES OF PLUME RISE OBSERVATIONS
            Gary A. Briggs

      H.  ESTIMATION OF PLUME RISE
            D.  Bruce  Turner
      I.  PROBLEM SET THREE
     Please complete the exercise in the given order.   Upon  completion
 of components G and H,  please work the problems  in set three  and  forward
 to the Air Pollution Training Institute.   A pre-addressed envelope is
 provided.   A critique of your calculations will  be returned with  a copy
 of the correct calculations.   Please be certain  that  your name and address
 is on each sheet.   ALL CALCULATIONS MUST BE RETURNED  TO RECEIVE CREDIT.
                              Air Pollution Training Institute
                              National Environmental Research Center
                              Research Triangle Park, North Carolina  27711
                              Attention: Plume Rise Instructor SI - 406
                                                                        53

-------
                                                                             G
                   SOME  RECENT ANALYSES
               OF PLUME  RISE OBSERVATIONS
                             Gary A. Briggs *
 Introduction

 Good estimates of plume rise are required to predict  the  dispersion  of
 continuous gaseous emissions having large buoyancy or a high  efflux
 velocity.  The rise of such emissions above their source  height often
 account for a considerable reduction of the concentration experienced
 at the ground .

 According to a recent critical survey on the subject,  several dozen
 programs of plume rise observations have been carried out and the results
 published.  This alone does not solve the problem,  however.   The quality
 of these observations varies considerably,  and in some cases  important
 parameters were not measured.   The picture  is also  blurred by the presence
 of turbulence in the atmosphere, which causes the plume rise  to fluctuate
 rapidly in many situations.  The great number of  empirical plume rise
 formulas in the literature reflect these uncertainties.   Each formula is
 based on an analysis of one or more sets of observations, but each time a
 different style of analysis or a different  collection of  observations is
 used, a different empirical formula results.  When applied to new situa-
 tions, the predictions of these formulas sometimes  differ by  a factor of
 ten.  Obviously, great care in the analysis of available  observations is
 required.

 The present paper is a summary of some of the analyses of observations
 made in my recent critical review on plume  rise and in my doctoral disser-
 tation.  Both of these works include extensive comparisons of observations
 with formulas; care was taken  to categorize the data  according to the type
 of source and the meteorological conditions, and  to weight the data  accord-
 ing to the quality and quantity of observations they  represent.
:Dr.  Gary  A.  Briggs, Research Meteorologist
 Atmospheric  Turbulence  and Diffusion Laboratory,
 National  Oceanic  and Atmospheric Administration.
 Atmospheric Turbulence and Diffusion Laboratory Contribution Number  38.
 This paper was presented at the 1970 International Air Pollution Conference
 of the International Union of Air Pollution Prevention Associations  and
 portions have been extracted for publication by the Journal of the Air
 Pollution Control Association.

                                                                        55

-------
Effective Stack Height/Plume Rise
Momentum Conservation and the "1/3 Law" for Jets	

One of the major findings of researchers in the field of plume rise  is  that
the radius of a plume bent over in a wind is approximately^proportional to
the rise of the plume centerline above its source height: - '   This is true
for a considerable distance downwind of the source, at least several stack
heights.  Mathematically, we can express this by
where r is a characteristic plume radius, y is a constant  (dimensionless),
and z is the rise of the plume centerline.  Surprisingly,  this simple re-
lationship accounts very well for the great bulk of observed  plume  rises,
when it is used with appropriate conservation assumptions.

For instance, no outside forces act on a non-buoyant plume (jet)  rising
through unstratified surroundings, so we might expect  that the total flux
of vertical momentum in the plume is conserved.  If the  plume is  only
slightly inclined above the horizontal, is nearly  the  same density  as the
ambient air, p-, and has a  horizontal component of  motion nearly  equal to
the mean wind at that  height, u, then  the  flux of  mass is approximately
pirr2u.  The flux of vertical momentum  is  then w(pirr2u) ,  where w  = udz/dx,
the vertical velocity  of the centerline of a plume segment moving down-
wind  at a  speed u;  x  is  the  distance downwind  of the  source.  We then
have
                       2    222
                 w u r  =  u y z dz/dx = F  = constant,                 (2)
                                         m

where F  is the initial vertical momentum flux divided by  irp.  For  a jet
having the same density as the ambient air, which  must be  true if it is non-
buoyant, F  is given by
                        22     22
                 F  = w  r   = w  D /4,                                (3)
                  m    oo     o     '                                ^ '

where WQ is the mean efflux velocity, r  is the radius of  the stack, and D
is its diameter, assuming a circular, vertical source.   Integration of Equa-
tion 2 yields the prediction that

                 z3 =  (3Fm / Y2u2)x


                 Ah =  z =  (ewV/4Y2u2) 1/3 x 1/3                      W


                 Ah/D = (3 / 4Y2) 1/3 R 1/3 (x/D)  1/3  ,

where R = WQ / u, the ratio of the efflux velocity to  the  wind speed   In
this paper, the  plume rise", Ah, is identified with the height of  the plume
centerline above the source (Ah = z).

56

-------
                    Some Recent Analyses of Plume Rise Observations     (G)
The above prediction that the rise of a jet is proportional to the one-third
power of distance downwind, the "1/3 law", is very well confirmed by the
available observations on jets.  These are plotted in Figure 1 for the data
in which R = 2, 4, 8, 16, and 40.  (The code identifying the six experiments
is given in Reference 2.)  Surprisingly, the dotted lines representing the
"1/3 law" give fair agreement with observations even in the upper-left part
of the figure, where the plumes are more nearly vertical than horizontal
(the derivation of Equation 4 utilizes the assumption that the plumes are
only slightly inclined).  However, the data indicate a stronger dependence
on R than the two-thirds power.  Specifically, the dotted lines represent
Equation 4 with


                 Y = 1/3 + R"
                                                                       (5)


Thus, it appears that the "entrainment constant",  y, varies with the ratio
of efflux velocity to wind speed for a jet.  This turns out not to be true for
a buoyant plume.  Substituting this expression for v into Equation 4, we
have finally
                              1+3R
Buoyancy Conservation and the "2/3 Law" for Buoyant Plumes

If a buoyant plume is rising through unstratified surroundings and it neither
gains nor loses buoyancy through radiation, ordinarily the total flux of buoy-
ancy is conserved (for an exemption, see Reference 5).  Applying Newton's
Second Law to a segment of a  bent-over plume moving downwind with the mean
speed of the wind, we find that the rate of vertical momentum flux increase
equals the buoyancy flux:
                        2               22
                 d(w u r )/dt = ud(w u r )/dx = b u r   ,               (7)
                                                                     2
where b is a characteristic buoyant acceleration of the plume and bur   is the
buoyancy flux divided by irp.  The initial value of bur  is given by

                 F = g(l-p0 / P) worQ2  ,                               (8)

where g is gravitational acceleration and p  is the density of the effluent
at the stack.  A better determination of F, that accounts for alteration of
buoyancy due to dilution with ambient air, is given by

                 F = g(l-mo / m) (T/To)worQ2 + gQH /  (TTC  pT)  ,        (9)


                                                                        57

-------
                                                                                     ORNL-DWG 69-13875
<1J
en
QJ
B

rH
PH
fl
W>
•H
0)
X

A:
a
n)
                                  0.5     1       2         5      10     20       50
                                         DISTANCE  DOWNWIND/JET DIAMETER
100    200
           Figure  1.   Rise of jet centerlines versus distance downwind for R = 2,  4,  8,  16, and 40.
                       (dotted lines are Equation 6)(see Reference 2 for code identifying experiments)

-------
                                                                             ORNL-DWG 68-12899
    1000

-------
Effective Stack Height/Plume Rise
where m is the mean ambient molecular weight (28.9), T is the ambient absolute
temperature, c  is the specific heat capacity of air (0.24 cal/gm - 1C), and
Qu is the heatpemission; subscript "o" denotes values for the efflux gas,
instead of the ambient air.  The quantity g/ (ircppT) is just a constant,
3.7  10~5 (m4/cal-sec2), times the ratio of standard sea level pressure to
the ambient pressure.  If the effluent has considerable latent heat due to
water vapor and condensation of the plume is likely to occur near the source,
as would be expected in cold or wet weather, this latent heat may be inclined
in the determination of QR; otherwise, only dry heat should be considered.
If the process producing the effluent is uniform, F is proportional to the
rate of production.  For instance, for modern fossil fuel power plants, F
is about 1.5 m4/sec3 times the megawatts per stack generated.
                         2
                 When bur  = F = constant, Equation (7) integrates to

                 wur2 = F  + F x/u                                     (10)
                         m

This relation implies that buoyancy becomes more important than the initial
momentum flux when x •» uFm / F.  For a hot effluent with about the same heat
capacity and mean molecular weight as air, this occurs at x = uw / (g (T0/T-1)),
a distance of the order of 5 seconds times the wind speed for most hot plumes.
Then the effect of Fm quickly becomes negligible, and for the region in which
Equation (1) applies, we have


                 z3 =  (3F/2Y2u3)x2


                 Ah = z =  (3F/2Y2)1/3  iT1 x273                          (11)


                 Ah/L = (3/2Y2)1/3 (x/L)2/3



Where L = F/u ,  a characteristic length for the rise of buoyant plumes.

Most of the observations available on buoyant plume rise approximate this
"2/3" of rise with distance downwind.  This is illustrated in Figure 2,
which is a superposition of curves  hand drawn through scatter diagrams of
Ah/L versus x/L for 16 individual sources1 (the sources are identified in
Reference 1).  Only data for stable atmospheric conditions are omitted.
There is considerably greater scatter about the "2/3 law" in this figure
than there is about the "1/3 law" for jets in Figure 1.  The difference is
probably due to the fact that all the experiments represented in Figure 1
were made under controlled conditions in wind tunnels or modeling channels,
while all the observations shown in Figure 2 were made on plumes from real
stacks in the atmosphere.   This introduces the possibility of aerodynamic
effects caused by buildings near the stack and uneven terrain, and also

60

-------
LJ -
uo 10
CC
_l
<
z
z
UJ
5
o
o
10 " 2
^
X
PO
^E
_l
< 0.





• R= 8
A R= 4
R= EFFLU>

R- 16
• *
R = 8
A R=4

o







' V


A





EL

(






OC
n
^






IT








Y








/








WIND SPE
DO
	 *<>~<% 	 '
	 —£*
A A





ED


'"3 —
A







°J"!2








» —








•or
4








nr
8<
2/








_
3






Q_
-c
A
L;






f
A
av






-«XJ 	
*.«LjP^-
^ _ 0^—0^
A ^-
V" FOR BU







*A*'
OYANT






-stf.
^
PLl






^

JME






r5

s






5^








9>







^
-f


01 0.02 0.05 0.1 0.2 0.5 1 2 5 10
                      Lx/L,
MOMENTUM  FLUX  ENHANCEMENT  DUE  TO BUOYANCY
                                                                                                     en
                                                                                                     o
                                                                                                     S
                                                                                                     rt>
                                                                                                     rt

                                                                                                     >
                                                                                                     3
                                                                                                     CD
                                                                                                     0>
                                                                                                     en
                                                                                                     H-
                                                                                                     0)
Figure 3.   Nondimensionalized  rise  of model  plume  centerlines  versus ratio of buoyancy-

            induced momentum  to initial  momentum.      (solid lines are Equation 13)
                                                                                                     O
                                                                                                     a'
                                                                                                     01
                                                                                                     (D
                                                                                                     H

-------
Effective Stack Height/Plume Rise
assures greater fluctuations about the mean plume rise due to large turbulent
eddies in the atmosphere.  Atmospheric turbulence should also lead to more
rapid mixing of plumes with ambient air,  and therefore a downward departure
from the "2/3 law" should occur at some point downwind; however, Figure 2
offers no particular support for this expectation.  This means either that
L does not correlate well with the distance of downward departure (leveling
off), or that "leveling off" occurs at greater values of x/L than those
measured up until now.

There is no evidence that y is dependent on R for buoyant plumes6, at least
when R > 1.2.  Below this value, downwash of the plume into the low pressure
region  in the wake of the stack is likely to occur.  This reciprocal wind
speed relationship predicted by Equation 11 with y constant is well estab-
lished for buoyant plumes in neutral conditions.  An analysis'of photo-
graphed plume diameters and concurrent plume rises of TVA plumes from single
stacks showed that y = 0.5.  Bringfelt3obtained similar results, finding
an average value of y of 0.53 for eleven plumes in slightly stable or windy
conditions and 0.46 for ten plumes in strongly stable or weak wind conditions.
The behavior of buoyant plumes in stable conditions is well predicted by
y = 0.5, as will be shown, but the optimum fit' to the "2/3 law" at large
distances downwind in neutral and unstable conditions corresponds closer to
y = 0.6, or

                 Ah = 1.6 F1/3 u-1 x2/3                                 (12)
Transition to Buoyancy-Dominated Rise

Equation 10 implies that a transition from the "1/3 law" for momentum-
dominated rise to the "2/3 law" for buoyancy-dominated rise occurs  as Fx/uFm
grows from small to large values.  This prediction is confirmed  by  Figure 3,
which plots observations of plume rise modeled in a channel by Fan.7 The rises
are divided by      2/3  1/3    1/3  _2/3   f
                 L     x    = F    u     x
                  m
so there should be no variation with Fx/uF  = Lx/I2 in the momentum-dominated
                                          m
region.  This is seen to be approximately true in the left-hand side of  the
figure, where Lx/Lm  < 0.3.  However, there is a separation of the points
for different values^ of R in this region; this is easily accounted for by
letting y - 1/3 + R  , as was done for jets.  There is a clear upswing and
some convergence of  the points in the right-hand side of the figure  and
these appear to asymptotically approach the line representing the "2/3 law"
Equation 11, with y  = 0.5.

The simplest way to  describe this transition mathematically is to integrate
Equation 10, after substituting r = yz and w = udz/dx, and then substitute
the empirical value  of y for jets in the momentum term (y = 1/3 + R~') and

62

-------
                   Some Recent Analyses of Plume Rise Observations       (G)
the empirical value for y for buoyant plumes in the buoyancy term (y = 0.5).
The result is
                                  (1/3 + R l)2 u2       2(0.5)2  u3


This equation is represented as  solid lines  in Figure  3, for R = 4, 8, and
16.  It is seen to describe the  transition region  fairly well.


Stability-Limited Rise

When a plume rises in a stable environment,  it entrains air and carries it
upward into regions of relatively warm ambient air.  Eventually, the plume's
buoyancy becomes negative and its rise is terminated.  If heat is conserved,
that is, the motion is adiabatic, the rate at which each plume element loses
temperature relative to the ambient  temperature is just its rate of rise
times the ambient potential temperature gradient  (potential temperature, 0,
is the temperature that air would acquire if it were compressed adiabatically
to a standard pressure, usually  the  mean pressure  at sea level; the potential
temperature gradient is just the real temperature  gradient plus the adia-
batic lapse rate, i.e., 30/3z =  3T/3z + 1°C/100m  in our lower atmosphere).
The resulting decay of the buoyancy  flux is  expressed  by
                      222
                 d(bur )/dt = ud(bur )/dx =  -w sur                      (14)

where s =(g/T) 30/3z.

If we differentiate Equation 7 with  respect  to t  and substitute  in  equation  14,
we find that
                  2222222          2
                 d (wur )/dt  =  u d  (wur )/dx  =  -s(wur )   .            (15)

                                         -1/2
This equation establishes the fact that s     is  a characteristic time  for
the decay of the momentum flux.   If  s is positive  and  approximately constant
with height, the momentum flux is a  harmonic function  of s/2  t:
                    2   „  _,  1/2^  ,  -1/2 „ _,_  ,1/2^
                 wur
                      = F  cos(st) +  s"     F  sin  (St).            (16)
                         m
Since r always increases, the plume centerline behaves like a damped harmonic
oscillator.  If the wind speed is constant with height, a jet  (F =  0) reaches
its maximum rise at x = ut =  (rr/2) us  /2  , and a buoyant plume  (F   = 0)
                                   - I/,                          m
reaches its maximum rise at x = ITUS   •< .

The above conclusions are based on conservation assumptions alone,  and do not

                                                                        63

-------
                                                                             ORNL-DWG  68-12903A
0)
en
•H
PS
 BO
•H
 CD
 O
 id
 4-1
 cn
                          (TVA, CARPENTER 61/1 a/., 1967)

                          	 PLUME  TOP

                                  PLUME  CENTER  LINE

                             •    THEORY
1           23456

       x/x', NONDIMENSIONAL DISTANCE DOWNWIND
            Figure  4.   Nondimensionalized rise versus  nondimensionalized distance  downwind for single-
                       stack TVA plumes in stable conditions.          (dotted  line  is Equation 18)

-------
                                                                    ORNL-DWG 69-13872
 LJ
 CO
 o
 CO

 UJ


 Q

 O
 2

JT>~
\j
 x
     0
                                                           BRINGFELT,1968

                                                           PLUME CENTER  LINE

                                                           THEORY
       0
                      234567

                         x/x' ,NONDIMENSIONAL DISTANCE DOWNWIND
10
                                                                                                  to
                                                                                                  0>
                                                                                                  O
                                                                                                  cr
Figure 5.  Nondimensionalized rise  versus nondimensionalized distance downwind  for plumes

          measured by Bringfelt in stable conditions.     (dotted line is Equation 18)

-------
Effective Stack  Height/Plume Rise
      ,.                           «"---
point, as it would imply a shrinking  plume.  With u constant  w = udz/dx, and
r = yz, Equation 16 can  easily  be  integrated.  For a  jet we fmd that
                                           /  •   ,  /
                 Ah = z =  |	      (sin (x/x'
                              2  1/2
                             Y us
                                                                       (17)
                                          /  Fm   \l/3
                    = 3 (1 + 3R"1)"       (	 )      (maximum rise)   ,
where x' = us 1   and y = 1/3 + R  .   For a buoyant plume we find that
                              3F  \ 1/3                 ,,,
                 Ah = z =   ( —	j     (1- cos (x/x'))  '
                               us
                                                                       (18)


                    = 2.9  /  F  \ 1/3 = 2.9 L1/3 x'2/3  (maximum rise) .
 There are sufficiently detailed data to verify Equation 18, which is shown
 as dotted lines  in Figure 4 and 5.  Both these figures show plume rise,
 divided by L     x'    , versus x/x'.  To determine s, the measured potential
 temperature gradients  were  averaged throughout the layer of plume rise
 (from the top  of the  stack to the  top of the plume).  The first figure shows
 centerlines of TVA  plumes  from single  stacks  that  were  observed  to  level
 off  in stable air.  It also shows the observed rises of  plume tops.   The
 second  figure shows the longest plume centerlines observed in very stable
 air  by Bringfelt.3  Both of  these figures  give excellent support  to  the
 prediction that the maximum rise is obtained at  x  = TTX' .   There  is  only a
 little evidence of oscillation beyond this point; evidently,  most plume
 centerlines experience considerable damping through continued mixing beyond
 this point.  The leveled-off  plume rises,  which range  from 140 to 290m for
 the  TVA data and from 60 to  160m for  the Bringfelt  data,  seem to  be  well
 approximated by Equation 18  on  the average.  For  the TVA data,  the scatter
 about  the predicted rise seems  to be  greatest  in  the rising stage, which

 66

-------
                   Some Recent Analyses ot Plume Rise Observations
                                                                        (G)
                                 Table 1.
   Ratios of Calculated to Observed Plume Rises in Stable Conditions
Formula Bring
Holland 0.
Priestley ! 0.
Bosanquet
Schmidt, m = 0
Schmidt, m = 1/2
Equation 18
1.
0.
0.
0.
33
74
09
29
94
89
;felt :
± 73% : 0
± 221 0
±24% 1
± 072
± 27%
± 07%
0
0
0
TVA
.81 r
. 4-4 ±
. 22 ±
.28 r
.85 t
.96 ±
Bring felt
and TV A
07%
05%
12%
24";
25%
08%
0.72 =
0.47
1.20 ±
0.28
0.90 =
0.93 ±
39
35
18
16
27
-
?/
»
c,<
0 !
08%
approximates the "2/3 law" when x -- 2x'.  There is less scatter in this
stage in the Bringfelt data, which utilized more representative wind speed
measurements and were taken during much more stable conditions.

Reference 2 also compares these observations with other formulas for buoy-
ant plume rise in stable conditions, namely; the Holland formula,9 minus
20% as suggested for stable conditions; Priestley's equation,10 reduced to
the case of a buoyant, point source; Bosanquet's formula,11 similarly re-
duced; and Schmidt's formula,1"1 with his parameter "m" set equal to 0 and 1/2.
The centerline plume rises at a standard distance x = 5x' were interpolated
and averaged for periods in which there were at least five photographs of
the plume at this distance.  This yielded five periods each from the TVA and
the Bringfelt observations.  The ratios of calculated to observed plume rises
were then calculated for each formula and each period.  The median ratio and
the average percentage deviation from the median ratio for each formula are
shown in Table 1.

The TVA heat emissions were substantially higher than those observed by
Bringfelt, so the high percentage deviation exhibited by the Holland and
Priestley formulas may be because of too much and too little predicted de-
pendence of rise on heat emission, respectively.  Clearly, Equation 18 for
maximum rise gives the most consistently good predictions for buoyant,
stability-limited plume rise.
Turbulence-Limited Rise

The simple plume rise model outlined in the preceding section, based on r = yz
and conservation assumptions, succeeds in predicting the  approximate rise be-
havior of all available observations of plumes bent over in a wind.  It is
very similar to the successful model for nearly vertical plumes  suggested by
G.I. Taylor in 1945 and later developed by Morton, Taylor, and Turner.  How-
ever  as it stands, it predicts unlimited rise in neutral  (s = 0) and unstable
environments (s < 0).  This is contrary to the expectations of many plume rise

                                                                        67

-------
Effective Stack Height/Plume Rise
observers, some of whom have assumed that the plume rise is the rise of the
plume at the point that it becomes hard to follow.  Yet, no observations made
so far show any leveling-off tendencies, except in stable conditions.

Nevertheless, it is quite reasonable to expect more rapid growth of  the plume
radius in neutral and unstable conditions, due to the presence of considerable
environmental turbulence.  This, in turn, leads to a reduced rise velocity
and perhaps to a limited plume rise, at least in neutral conditions.  The ques-
tion is how to account for the enhanced growth of the plume radius.  One way
is to assume that only eddies of the same order of size as the plume radius
are effective at mixing ambient air into the plume, and that these eddies are
predominantly in the inertial subrange of the atmospheric turbulence spectrum.
This part of the spectrum is adequately characterized by the eddy energy
dissipation rate, e, and eddies of the order of r in size have velocities of
              1/3  1/3
the order of e    r    .  This suggests the relationship

                 dr/dt = Be1/3r1/3  ,                                   (19)


where 6 is a constant (dimensionless).  This should not apply at small
distances, where r is small and w is large; Equation 1 gives a faster growth
rate at first (r = yz implies that dr/dt = yw).  In References 1 and 2, I
developed a model identical to the one outlined so far, except for the
assumption that Equation 19 applies instead of Equation 1 beyond the dis-

tance at which Be    r    becomes equal to yw-  Since this model is based
on an inertia! range atmospheric turbulence entrainment assumption, I call
this the "IRATE" plume rise model.2
                                            2
For rise in neutral conditions, in which bur  = F = constant, the "IRATE"
model predicts a very gradual leveling of the plume centerline beyond the
distance that 3e    r    becomes equal  to yw> designated by x*.  For a

buoyant plume, the "2/3 law", Equation 11, applies to the first stage of
rise.  The distance of transition to the second stage is then given by
                 x* - (2/3)7/5 (YF)2/5 u3/5 (BE^) -y/>                (20)

If E is approximately constant above the height at which this transition
occurs, the second stage rise (x > x*) is given by
                                                                        (21)

                 Ah = (3F/2Y2) 1/3 u-1 x* 2/3   \H _ ^   (x/x* + 5/8)
                                                  6    32
While this is not a very simple formula, note that it is just the "2/3 law"
times a function of x/x*.  A final rise, equal to 55/16 times the rise at

-------
                    Some Recent Analyses of Plume Rise Observations       (G)
the transition point, is approached, but only at a great distance; 90% of
the asymptotic rise is achieved at x = 20x*, and x* can be as large as a
kilometer for a very buoyant or very high source.  It is unlikely that the
maximum ground concentration occurs well before this point, especially since
Equation 19 predicts an extremely large radius at x = 20x*.  If y = 0.5 and
the bottom of the plume is taken to be a distance (Ah - r) above the source
height, we find that the plume bottom begins to descend at about x = 2x*
and spreads down to the source height (r = Ah) at x = 5x*.  Since the growth
of the plume radius is quite rapid at this point, the highest ground concen-
tration should ordinarily occur in this neighborhood.  It therefore seems
prudent to use the rise at x = 5x*, 2.3 times the rise at the  transition
point, as the "final" plume rise in neutral conditions.  This rise is the
same as that given by the "2/3 law" with x = 3.5x*,  and Equation 21 deviates
from the "2/3 law" by only 11% at x = 3.5x*.  This suggests a much more
practical prediction procedure for buoyant plume rise in neutral conditions:
                 Ah =  (3F/2 Y2) 1/3 u'1 x273    when x < 3.5x*

                             2  1 /T  -1         ? /I
                 Ah =  (3F/2 Y )  '  u   (3.5x*)  '  when x > 3 . 5x* .    (22)


With this approach, in a simple way we recognize the observed fact that plume
rise is substantially a function of distance, yet we have a usable "final"
rise formula.

In order to use the formulas based on the "IRATE" model, an estimate of x*
is needed; this requires values of 6 and E for substitution into Equation
20.  In Reference 1, B was conservatively estimated to be about unity, on
the basis of observations on the growth rates of puffs and particle clusters
(in order to infer the value of 3 from Equation 19,  it was necessary to esti-
mate values of E, as this quantity was not measured in  the puff and particle
                                              O _
experiments).  It is well known that e = 2.5u* /z  in the neutral  surface  layer,13
but  this  layer extends only  to  a  height of  the order of  10   u*/f  in  neutral
conditions  (u* is  the  friction velocity,  z  is the height  above  the  ground,
and  f  is  the  Coriolis parameter); in mid-latitudes,  this height  is of  the
order  of  10  seconds times the wind speed.  However, most  plumes rise  to  heights
above  this layer, where E is less dependent  on wind speed and height.

In a convective mining layer,  such as  exists in  the  lowest  few hundred  to  few
 thousand  meters on any sunny,  non-windy day, the  average  value of e  is  about 2
 (1/2)  gH/(cppT), where H  is  the  heat  flux transported  upward from the  ground;
a fairly  strong heat  flux, 1 cal/cm2 -min,  corresponds  to e  = 30  cm2  /  sec3 ;
the  eddy  dissipation  rate is relatively constant  with height, except  near
the  ground, where  the  neutral  surface layer  expression  dominates.  Less  is
known  about  the variation of  E  above the surface  layer  in neutral  conditions.

                                                                         69

-------
Effective Stack Height/Plume Rise
If  e ceased to decrease with height at the  top of  this  layer,  it;  would be
proportional to fu*2, which is approximately proportional  to  fu2.  In Refer-
ence 1, measurements of E at heights from 15 to 1200m were shown  to fit e<*u
slightly better than e<*u2 or e = constant (measurements made  in stable or
convective conditions were  excluded from this analysis).   This result was
especially convenient for application to the IRATE plume rise model,  as e«u
cancels out the wind speed in Equation 20;  above the neutral  surface layer,
x*  is approximately independent of u.  However, the great  variation in E at
these heights,  of the order of ±50%, can be expected to account  for consider-
able variations in the plume rise.

In  the above analysis, definite decrease of E values with  height  above the
ground was also noted.  Above a height somewhere between 100  and  300m the
variation  becomes much less.  The best fit in the 15 to 300m range was given
by

                 e = 0.068 (m/sec)2 u/z  .                              (23)

Substituted in Equation 20, this gives

                 x* = 2.16 m(F/m4/sec3) 2/5 (z/m)  3/5   .                (24)


In References 1 and 2, it was suggested that conservative  values  of x* and
Ah would result by_ evaluating E at the source height, hs .   Thus,  z = hs  ,
but no more than z = 300m, was substituted  in Equation  24;  for the present,
let us designate this estimation of x* by x*  .  A simpler  estimate of x* is
suggested by a plot of total plume height, h1 + Ah, versus  buoyancy flux.
Using TVA8, _CERL1,4'15 Bringfelt'6 and other observations, a  quite  conservative
value for z is given by
                 —
                 z = 22 m  (F/m /sec )  '  < 100 m   .                    (25)

The 100m maximum value of  z may be overly conservative when  the  stack height
itself is greater than 100m, but consider also the  fact  that e does  not  diminish
very much with height above this  elevation.  The predicted  final  value  for Ah
only depends on  z °'4 , and the scatter in the few  plume rise data at large
distances renders tentative any conclusion about the best  evaluation of  x* for
these purposes.  Equations 23 and 25 give a particularly simple  way  to evaluate
x*, and can even be applied to ground sources without difficulty.  Let us
designate this estimate of x* by x *:



                 X2* = 14m (F/m4/sec3)  5/8    when F < 55 m4/sec3
                                                                        (26)
                 x2* = 34m (F/m4/sec3)  2/5    when F > 55 m4/sec3
70

-------
                    Some Recent Analyses of Plume Rise Observations     (G)
A number of plume rise formulas for buoyant plumes in neutral conditions were
compared with all available observations in Reference 1; TVA8 and CERL14'15 data
each comprised about one-third of the data analyzed.  The more recent obser-
vations by Bringfelt3-16 were added to the comparisons in "Reference 2.  A simi-
lar analysis is summarized in Table 2.  Since the relationship Ah<*u~  seemed
well verified1 for neutral conditions, the average value of uAh for each source
was calculated at the greatest distance downwind that  was represented by at
least three 30-120 min periods of observations with at least five Ah deter-
minations each.  The Bringfelt data had to be handled differently, because
there was only one period  of observation for many of the sources; accord-
ingly, these are weighted only one-third as much as the "Reference 1" obser-
vations in the last column of Table 2.  When appropriate measurements were
available, it was required that x < 2x', in order to exclude cases of sta-
bility-limited plume rise.  Cases of probable downwash, terrain effects,
etc. were eliminated in "select'1 set of observations in Reference 1, and
agreement with almost every formula improved.  Of 25 periods chosen by
Bringfelt for analysis,3  only 12 are selected here (periods 7, 8, lib, 15,
18, 27a, 27b, 31, 29, 41, A7a, and 27b); the periods rejected greatly in-
crease the scatter in the ratios of calculated to observed values for every
formula tested, tending to obscure the comparison.  Table 2 shows the median
value of the ratio of calculated to observed plume rises, and the average
percentage deviation of these ratios from the median, for eight  different
formulas of the Ah= u~'  type.  The last column combines the two  sets of
select data, with appropriate weighting.

Of  the first three formulas, which are empirical, it is seen  that the Moses
and Carson18formula in which Ah ^Q,,''2 gives the most consistent fit  to  the
select data; the fit would be optimized by multiplying by a correction  factor
of  2.  The next three formulas are based on the Priestley10 model, the first

being the asymptotic prediction of the first stage19 that Ah ^QR
u   x
 This formula and the "2/3 law", Equation 11, are very  similar,  and  neither
 predict any "final" rise; yet, both of these formulas  give better agreement
 than the empirical formulas.  The scatter in the Bringfelt data makes  it
 difficult to conclude that any one of the six formulas  is superior,  as  about
 ± 15% seems to be the lowest possible scatter.  The differences between the
 last five formulas are also slight in the Reference 1  data,  except  in  the
 select set.  In this set, as well as in the weighted select  data, it is seen
 that Equation 22 gives the best fit; this equation is  simply the  "2/3  law",
 Equation 11, terminated at a distance x = 3.5x*.  The  second estimate  for
 x*  x *, as given by Equation 26, seems to have a slight edge over  x*  = x *;
 the amount of scatter and the scarcity of data at large values  of x/x*
 makes this comparison of x* = x * and x* = x* inclusive.  Which  value  of x*
 to be preferred is mostly just a matter of convenience.

 When it is not clear whether plume rise is turbulence-limited or  stability-
 limited  an analysis2 of  the  IRATE model with both factors  included shows
 that Equation 18 or Equation  22, whichever gives  the lowest  rise  in a  given
                                                                        71

-------
                                                     Table 2.

                         Ratios  of  Calculated to Observed Plume Rises  in  Neutral Conditions
Formula
Holland (9)
Stumke (17)
Moses and Carson (18)
Priestley (10, 19)
Lucas, et al. (14)
Lucas (20)
Equation 11 *
Equation 22 * (x* = x *)
Equation 22 * (x* = x *)
Reference 1
(all data)
0.44 ± 37%
0.79 ± 27%
0.54 ± 34%
1.44 ± 26%
1.36 ± 21%
1.18 ± 20%
1.17 ± 23%
1.12 ± 21%
1.12 ± 17%
Reference 1
(select)
0.47 ± 26%
0.72 ± 24%
0.48 ± 19%
1.41 ± 18%
1.24 ± 22%
1.16 ± 14%
1.17 ± 12%
1.13 ± 08%
1.13 ± 06%
Bringfelt
(select)
0.26 ± 32%
0.82 ± 35%
0.53 ± 28%
1.42 ± 15%
1.36 ± 13%
1.12 ± 17%
1.08 ± 13%
0.99 ± 15%
1.00 ± 13%
Weighted,
Select Data
0.40 ± 35%
0.74 ± 29%
0.48 ± 23%
1.41 ± 17%
1.35 ± 19%
1.16 ± 15%
1.17 ± 13%
1.11 ± 10%
1.11 ± 09%
60
•H
Q)
o
OJ
M-l
            * with y = 0.5

-------
                   Some Recent Analyses  of Plume Rise Observations
situation, offers a good approximation of the very complicated prediction
that results when both e and s are greater than zero.  In unstable conditions,
there is no strong evidence that the average plume rise differs much from its
value at the same wind speed in neutral conditions, but the rise is much more
variable.1'2

Summary and Conclusions

Direct analysis of plume rise observations and several comparisons of obser-
vations with a number of empirical and theoretical formulas have shown that
very satisfactory predictions of  plume rise are given by a rather spare phy-
sical-mathematical model.  This model was briefly outlined here and is more
rigorously developed in Reference 2; it basically consists of the assumptions
that momentum, buoyancy, and potential temperature are conserved, that the
horizontal component of motion of plume elements is essentially equal to the
mean wind speed, u, and that r = yz in a first stage of rise and dr/dt =

Be   r    in a second stage of rise (r is the characteristic plume radius,
z is the rise of the plume centerline above the source height, e is the eddy
dissipation rate of ambient atmospheric turbulence, and y and 0 are dimen-
sionless constants).  Empirical guidance is used in evaluating y, 6, and e.

The assumptions that r = yz and that  momentum is conserved in a non-buoyant
plume (jet) in unstratified surroundings lead to a simple "1/3 law" of rise
that fits a large variety of observed jet center lines:



                             /   R  \2/3       1/3
                 Ah/D = 1.89(	 j     (x/D) /J   ,                   (6)
                             ^ 1+3R


where x is the distance downwind, D is the stack diameter, and R is the ratio
of efflux velocity to wind speed.  To derive Equation 6, one must assume that
y = 1/3 + R  .  On the other hand, there is no evidence that y is a function
of R for buoyant plumes.  The assumptions that buoyancy is conserved and that
the initial plume momentum is negligible for a very buoyant plume in unstratified
surroundings lead to the often-cited "2/3 law" of rise:

                 Ah = 1.6F1/3 u-1 x 2/3 ,                               (11)


where F is the initial buoyancy flux divided by irp; complete expressions for
F are given by Equations 8 and 9.  The constant in Equation 11 is based on  the
best fit  to data shown in Table 2, and corresponds to y = 0.6.  Only equations

that include the second stage entrainment assumption that dr/dt = Be    r
give a better fit to observations of the rise of hot plumes in near-neutral
conditions.  For plumes in which both momentum and buoyancy are significant,
Equation 13 gives a semi-empirical transition between Equations 6 and 11.

                                                                        73

-------
Effective Stack Height/Plume Rise
Buoyancy becomes the dominant factor for most hot plumes at a distance down-
wind of the order of five seconds times the wind speed.

The assumption that the potential temperature of entrained air is conserved
leads to the prediction that a buoyant plume attains a maximum rise at a
                —1/2
distance x = ITUS     in stable air (s = (g/T)39/3z, g is gravitational
acceleration, T is the absolute ambient temperature, 6 is  the ambient
potential temperature, and 38/3z = 3T/3z + 1° C/100m) .  This prediction is
very well  confirmed by plots of plume rise versus distance in stable con-
ditions (38/3z and u are averaged from the top of the stack to the top of
the plume).  These plots also indicate that the "2/3 law" is approximated
            -1/2
when x < 2us     and that the plume centerlines level off at a height
                 Ah = 2.9 (-)                                          (18)
This corresponds to the maximum rise given by r = jz with y = 0.5.  Equation
18 gives substantially better agreement with observations than other formulas
tested for buoyant, stability-limited rise.  It should be noted that in very
light winds the well-proven'-2 formula of Morton, Taylor, and Turner  best
applies if it gives a lower plume rise than Equation 18:

                 Ah = 5.0 F1M a'378                                   <27>


In neutral conditions, a limited rise results only after the second stage en-
entrainment assumption is utilized.  A good approximation to the complete
prediction for buoyant plumes in neutral conditions is given by

                           1/3  -1  2/3
                 Ah = 1.6 F '  u   x '     when x < 3.5x*

                                                                       (22)
                           1/3-1        9/1
                 Ah = 1.6 F /J u   (3.5x*)Z/J    when x > 3.5x* ,


where x* is the distance of transition from the first stage to the second
stage of rise.  This equation gives a somewhat better fit to observations
than any other formula tested when x* is estimated by:



                 x* = 14m (F/mA/sec3) 5/8    when F < 55 m4/sec3
                                                                       (26)
                 x* = 34m (F/m /sec )        when F > 55 m
74

-------
                   Some Recent Analyses of Plume Rise Observations       (G)
This equation for x* should be considered tentative, since it is based on
limited empirical determinations of 6 and e, and there is too much scatter
in the few observed plume rises at large values of x/x* to make any strong
conclusions about x*.  Equations 22 and 26 apply satisfactorily to the mean
rise in unstable conditions as well,  and also in slightly stable conditions
if they give a lower rise than Equation 18.
REFERENCES

 1.    Briggs, G.A., Plume  Rise, AEC  Critical  Review  Series,  TID-25075  (1969)

 2.    Briggs, G.A., "A  Simple Model  for  Bent-Over  Plume  Rise", Doctoral
      Dissertation, The Pennsylvania State  University  (1970).

 3.    Bringfelt, B.,  "A Study of  Buoyant Chimney Plumes  in Neutral and
      Stable Atmospheres", Atmos.  Environ.  3:  609-623  (1969).

 4.    Scorer, R.S., Natural Aerodynamics, pp.  143-217, Pergamon Press, Inc.,
      New York  (1958).

 5.    Gifford,  F.A.,  "The  Rise of  Strongly  Radioactive Plumes", J. Appl.
      Meteorol. 6:644-649  (1967).

 6.    Fay, J.A., Escudier, M. and  Hoult, D.P., "A  Correlation of Field
      Observations of Plume Rise", Fluid Mechanics Lab. Pub. No. 69-4,
      Massachusetts Institute of Technology (1969).

 7.    Fan, L.,  "Turbulent  Buoyant  Jets into Stratified or Flowing Ambient
      Fluids",  Califoria Institute of Technology,  Report KH-R-15 (1967).

 8.    Carpenter, S.B. et al.,"Report on  a Full Scale Study of Plume Rise
      at Large  Electric Generating Stations", Paper 67-82, 60th Annual
      Meeting of the Air Pollution Control  Association, Cleveland, Ohio
      (1967).

 9.    U.S. Weather Bureau, "A Meteorological  Survey of the Oak Ridge Area:
      Final Report Covering the Period 1948-1952", USAEC Report ORO-99,
      pp. 554-559 (1953).

10.   Priestley, C.H.B., "A Working  Theory  of the  Bent-Over Plume of Hot
      Gas", Quart. J. Roy. Meteorol. Soc.,  82:165-176  (1956).

11.   Bosanquet, C.H.,  "The Rise of  a Hot Waste Gas Plume'1, J. Inst. Fuel,
      30:322-328 (1957).

12.   Schmidt,   F.H., "On the Rise  of Hot Plumes in the Atmosphere", Int.
     J. Air Water Pollut., 9:175-198 (1965).

                                                                        75

-------
Effective Stack Height/Plume Rise
13.  Lumley, J.L. and Panofsky, H.A., The Structure of Atmospheric
     Turbulence, pp. 3-5, John Wiley & Sons, Inc., New York  (1964).

14.  Lucas, D.H., Moore, D.J. and Spurr, G., "The Rise of Hot Plumes
     from Chimneys" Int. J. Air Water Pollu., 7:473-500 (1963).

15.  Hamilton, P.M., "Plume Height Measurements at Two Power Stations",
     Atmos. Environ., 1:379-387 (1967).

16.  Bringfelt, B., "Plume Rise Measurements at Industrial Chimneys",
     Atmos. Environ., 2:575-598 (1968).

17.  Stumke, H., "Suggestions for an Empirical Formula for Chimney
     Evaluation", Staub, 23:549-556 (1963); translated in USAEC Report
     ORNL-tr-997, Oak Ridge National Laboratory.

18.  Moses, H. and Carson, J.E.,  "Stack Design Parameters Influencing
     Plume Rise", Paper 67-84,  60th Annual Meeting of the Air Pollution
     Control Association, Cleveland, Ohio (1967).

19.  Csanady,  G.T.,  "Some Observations on Smoke Plumes", Int. J. Air
     Water Pollut.,  4:47-51 (1961).

20.  Lucas, D.H.  "Application and  Evaluation of Results of the Tilbury
     Plume Rise and Dispersion  Experiment", Atmos. Environ.,  1:421-424
     (1967).

21.  Morton, B.R.,  Taylor, G.I.,  and Turner,  J.S., "Turbulent Gravita-
     tional Convection from Maintained and Instantaneous Sources, Proc.
     Roy.  Soc. (London), Ser. A.,  234:1-23 (1956).
76

-------
                                                                             H
                  ESTIMATION OF PLUME RISE
                             D. Bruce Turner *
   The Environmental Application Branch has used the equations  of ,
   Dr. Gary Briggs for a number of years to estimate plume  rise.
   Dr. Briggs has revised the equation several  times.(1,2,3)
   The following procedures are consistent  with the  way  in  which the Meteorology
   Laboratory calculates Briggs1 plume rise:
        The following symbols are  used:
        IT   A Constant - 3.14
        g   Graviational Acceleration = 9.80  m  sec
        T   Ambient Air Temperature,  ^K
        u   Average wind speed  at  stack level,  m sec
        v   Stack gas exit velocity,  m sec
        d   Top inside stack diameter,  m
        T   Stack gas exit temperature,   K
         S                          3-1
        V   Stack gas volume flow,  m  sec
                                      4-3
        F   Buoyancy flux parameter,  m sec
        x   Distance at which atmospheric turbulence begins to  dominate
              entrainment, m
        AH  Plume rise above stack top,  m.
        x   Downwind distance from the source,  m.
        x,-  Distance downwind to final rise,  m.
    38/3z   Vertical potential  temperature  gradient  of atmosphere,  K m
        s   Restoring acceleration per unit vertical displacement for
              adiabatic motion  in  the atmosphere - a stability  parameter,
                  O
              sec-
   If T is not given, we have been using:

        T = 293°K (20°C) for design calculations
       V  = —^  vsd~   =  0.785  v,d
        f
                         -T \             / T -T  \         ,„.
                                             s    \         (2)
 *D. Bruce Turner, Chief, Environmental Applications Branch
  Meteorology Laboratory, Environmental Protection Agency
May, 1973

-------
Effective  Stack Height/Plume  Rise
      unstable or neutral conditions:
       x" = 14 F5//8    For F less than 55                   (3)
        A       ? / S
       x  = 34 F       For F greater than or  equal  to  55    (4)
  The distance of the final rise is:  x  = 3.5 x*           (5)
  The final plume rise:




            1.6 F       (3.5 x--)                            ,,,

       AH = 	                        (6)
                       u



  For x less than the distance of final rise:



            n  , ,-,1/3   2/3
            1.6 F    x                                      .  .

       AH = 	                                 (  '
                   u







  For stable conditions,  36/3z is needed





       If 86/3z is not given use:
            0.02 °K m l for stability E



            0.035 °K m'1 for stability F
                       =  9.806
Calculate





     AH = 2.4   /jM  1/3


                \us
                                                               (9)
78

-------
                                          Estimation of Plume Rise      (H)
    and
                 1/4
       AU _  5 F      (plume rise for calm conditions)   (10)
               s 3/8
    Use the smaller of these two AH's

       This is the final rise.

  The distance to final rise is:


          =  3.14   u
       Xf        1/2                                     (11)
  If you want to calculate rise for a downwind distance x less  than
  x ,  this  is given by
       AH -  1-6 F    x 2/3                               (12)
       which is  the same equation used for unstable and neutral
       conditions.
   Although (under stable conditions) the plume begins to rise
   according to  the 2/3 power with distance,  it does  not  continue  the
   same  rate of  rise to the distance of final rise, xc, given by
   equation (11).   Therefore equation (12)  will give  a iH higher
   than  the final  rise  at distances beyond  about 2/3  xf.   It  is
   therefore recommended that when using equation (12), the result
   be  compared with the final rise and the  smaller value  used.
   In  effect then,  for  determining the plume  rise at  a distance,
   x,  (during stable conditions)  the minimum  value of the three
   values of AH  determined by equations (9),  (10) and (12) should
   be  used.
Problem set three (component I) follows this article.  An Air Pollution
Training Institute certificate will be awarded upon satisfactory completion
of the three problem sets and the return of a completed critique.  All
calculations must be returned to the Air Pollution Training Institute.  A
set of answer sheets will be returned to the learner.
                                                                        79

-------
 Effective Stack Height/Plume Rise
                                REFERENCES

    1.   Briggs,  Gary A.,  1969:   Plume Rise.USAEC Critical Review
        Series  TID-25075,  National Technical Information Service,
        Springfield,  Va.  22151

    2.   Briggs,  Gary A.,  1971:   Some Recent  Analyses of Plume Rise
        Observation pp.  1029-1032,  in Proceedings of the Second
        International Clean  Air  Congress,  edited by H.  M. England and
        W.  T. Berry.   Academic Press,  New  York.

    3.   Briggs,  Gary A.,  1972:   Discussion on Chimney Plumes in Neutral
        and Stable  Surroundings.   Atmos. Environ.  6,  507-510 (Jul 72)
80

-------
Appendix
     CRITIQUE FORM




     EFFECTIVE STACK HEIGHT SCRIPT
                                                                  87

-------
               EFFECTIVE  STACK  HEIGHT
       A cued script of James  L. Dicke's audio tape in Component A.
     The topic I am going to discuss  is effective stack height and its
calculation.  Before proceeding any further, we should define exactly
what is meant by effective stack height.  2Effective stack height is the
height at which the plume centerline  from a smoke stack becomes essentially
level.  To make sure that everyone  understands, let's go through that again.
The effective stack height is the height above ground at which the plume
centerline becomes essentially level.  3^^ j_s true whether we are talk-
ing about the plume from a large stack or  ^the emission from a small pipe
or vent near rooftop,  with an effective stack height not much above the
top of the building.
     ^Effective stack height, which we will denote by capital H,  is equal
to the physical height of the stack,  small h, plus the rise of the plume
above the stack, delta h.  &In this diagram, the effective stack  height is
reached near the right end of the slide, where the centerline of  the plume
is leveling out.  When an effective stack height is specified, a  virtual
origin for the plume has been assumed.  It is assumed that the plume is
not emitted from the stack, but from  some point above and possibly upwind
of the stack.  This virtual origin  is such that a plume with dimensions
similar to those of the real plume  results at the distance where  the
effective stack height is reached.  7Here you can see the virtual origin
is slightly behind the stack.  It is  at height H, which is equal  to the
physical stack height, small h,  plus  the plume rise, delta h.
     ®Next, we should answer the question, "Why do we need to know the
effective stack height?"  Primarily, the effective stack height is required
because of its effect on the ground-level concentration of contaminants.
Its practical use in the diffusion  equation can be demonstrated to you in
other sessions.  As the effective stack height increases, the concentration
at ground level becomes less due to the fact that there is more atmosphere
in which to dilute the pollutants.  Also, as effective stack height increases,
the distance to the maximum concentration is moved further downstream from
                                                                      91

-------
the source and the maximum concentration becomes smaller.
     9As you can see from the example, if we assume that emissions from
both stacks are equivalent, the concentration at some point downwind from
the lower stack will be greater than that from the taller stack, since the
emissions from the taller stack have a greater volume of the atmosphere in
which to dilute before they significantly affect the ground.  Also, the
maximum concentration, due to the lower stack, will be greater and closer
to the source than that for the taller stack, which will have a smaller
maximum concentration with the point of maximum concentration off the slide
some place to the right. ^A practical example is considered in this slide
taken near Brilliant, Ohio.  Suppose that all three sets of stacks have
equivalent emissions.  It is apparent that the largest concentration will
be caused by the middle plant.  The plant on the right will cause the
smallest concentrations since it has the greatest effective stack height.
    " Again, effective stack height is equal to the physical stack height
plus plume rise.  Usually, the physical stack height can be measured or
is already known.  Therefore, we will be mainly concerned with calculating
the plume rise.
           stacks have some plume rise, though there are cases where plume
rise is zero or negative.  The momentum and buoyancy of gases in a plume
cause the plume to rise, sometimes to several times the physical stack
height.  As you can see in this slide, the effective stack height, which
is near the top of the slide, is at least three tines the physical stack
height. '3 With no wind the plume might rise straight up until it
reaches a level where there is some wind, and then level out quite rapidly,
or expending all its buoyancy and momentum, it just remains stationary,
forming a cloud.  When there is some wind, the plume will gradually slope
upward and over until it levels out. 14In any case, we are interested
priinarily in the height at which the plume becomes level at some point
downstream from the source.
92

-------
     15If  there  is no buoyancy or momentum,  the plume beccrres  level almost
 iimediately, especially if wind speeds are  moderate.  As you  can see in
 this case,  the  plume does not rise at all after  leaving the stack, but
 becomes level immediately. 16Here  plume rise is  zero, and effective stack
 height is equal to  the physical stack height.  ^Haaever, in cases of very
 strong winds or inproper stack design, aerodynamic downwash on the lee side
 of the stack can cause the plume to lower below  the top of the stack. ^In
 fact,  where there are short  stacks,  or the  effluent is  being  emitted right
 at rooftop, aerodynamic downwash on the lee side of the building is ccmmon.
^Here plume rise can be thought of as being negative and, as  a result, the
 effective stack height is less than the physical stack  height.  Another
 factor which can contribute  to the lowering of a plums  below  the physical
 stack height is evaporative  cooling of the  moisture droplets  in the plume.
 We will consider mainly cases where there is some  plume rise, but we will
 discuss aerodynamic downwash and evaporative cooling briefly  near the end
 of this session.
    20we  are interested in cases where there is  rise of the plume. 21 Most
 of the plume-rise equations  are obtained either  empirically or through a
 theoretical, derivation.   The empirical equations are obtained by arranging
 the stack and atmospheric parameters so that visual observations or physical
 measurements of plume rise will be duplicated by the equation.  Theoretical
 equations are derived from physical principles including dimensional analysis.
 However,  both types of equations are dependent on  data  collected in a labora-
 tory,  a wind tunnel or in the field, since  the theoretical equations usually
 include at  least one constant which is dependent on sampled data.
     All  the equations which have  been derived to  date  are subject to certain
 criticisms.  These  criticisms boil down to  the fact that there is no one
 equation  which  is universally accurate and  reliable.  No one  equation applies
 to all sizes of sources and  under  all atmospheric  conditions. Some equations
 give a generally good estimate of  plume rise, but  only  when the equation  is
 used with the specific stack and under the  atmospheric  conditions for which
                                                                        93

-------
the plurte-rise equation was derived.  Also, there is a general lack of plume-
rise data on a large variety of sources and atmospheric conditions; this
makes it difficult to cone up with a widely applicable equation.  Also, even
when sufficient data are available for obtaining an equation, there is still
the problem of little independent data with which to check the equation.
    22in the past 20 years there have been many attempts at expressions for
plume-rise - twenty equations and more.  The first two references which you
see give a very good evaluation of plume-rise equations and demonstrate whid:
equations give the most reasonable results.  Some additional factors which
affect plume rise are also discussed.  The third reference, by Stem, gives
a good review of plume-rise equations to 1968.  These references are among
those listed in Part B.
            rise is a result of the buoyancy and momentum of the stack
effluent and the manner in which they are affected by the atmosphere.
4ihe individual parameters which affect buoyancy and momentum with regard
to the stack are:  the stack diameter at the top of the stack in meters;
stack gas velocity, meters per second; stack gas temperature, degrees
Kelvin; and the heat content of the effluent, frequently expressed as
calories per second.  We will consider units only in the MGS system.
5The individual atmospheric parameters which affect the buoyancy and
momentum are:  wind speed, meters per second; potential temperature lapse
rate - degrees Kelvin per 100 meters; atmospheric stability - whether
unstable, neutral' or stable; atmospheric temperature - degrees Kelvin; and
atmospheric pressure in millibars.  In this session we will assume the mean
molecular weight of the atmosphere and the stack gas are essentially the
same.
    26Various combinations of these elements contribute individually to the
momentum and buoyancy.  The parameters which contribute to momentum are the
stack gas exit velocity, the stack diameter at the top, and the atmospheric
wind speed.  The parameters which contribute to buoyancy are the stack gas
exit temperature, the atmospheric temperature, potential temperature lapse

94

-------
rate, and  the heat  content of  the  effluent.  Usually,  the buoyancy  term is
expressed  either as a  temperature  difference between the gas and  the atmos-
phere or as the heat content of  the effluent.
     2'Two  of the plume-rise equations which frequently appear are the
Bryant-Davidson and the Holland  equations.  These are  of interest because
they are widely used;  they provide simple computation  schemes; and  they are
conservative, thus  providing a safety factor.  By conservative, I nean  that,
if anything, they underestimate  the amount of plume rise; this allows an
over-calculation of the pollutant  concentration  at the ground.  Thus, by
using these equations, a  ground  level concentration will never be under-
estimated  if the source conditions are  similar to those for which the
equation was derived.  Fran the  health  and safety standpoint this is the
desirable  situation.
     The Bryant-Davidson  equation  has,  as you see here, the momentum term
on the left and the buoyancy term  on the right.  The momentum term is based
on wind tunnel experiments  and the buoyancy term was added later  to take
buoyancy into consideration.   The  authors state  that this equation should
be applied only to  stacks of moderate or great height; however, the stacks
referred to are actually  rather  small by today's standards.  Other problems
with the equation are, there is  no allowance for different atmospheric
stabilities, and the equation  is based  on wind tunnel  data rather than  on
data collected in the  field.
    28The  Holland equation, which  we will go through rather thoroughly,  is
presented  here.  The first  portion of the equation on  the right is the  momen-
tum  term,  the other, with the  temperature difference,  is the buoyancy term.
As you can see, there is a  correction allowable  for unstable and  for stable
conditions.  The equation,  as we see it here, is for neutral stability  con-
ditions.   To apply  it to unstable  atmospheric conditions, it is appropriate
to multiply the answer by  1.1  or 1.2; for stable atmospheric conditions,
the effective stack height  should  be multiplied  by 0.9 or 0.8.  The data
on which the Holland equation  are  based are:  physical stack heights from

                                                                        95

-------
approximately 50 to 60 meters, diameters in the range of 2 to 4 meters,
stack gas exit velocities from 2 to 20 raps, exit temperatures from about
350 to 450°K, atmospheric wind speeds from approximately 0.5 to over 9 raps,
and all atmospheric stabilities.  The criticisms of this equation, which, I
might mention, apply to many of the equations we have seen, are:  it is
empirical; the coefficient 2.68 x 10~  has units which are per meter per
millibar; the data used in obtaining the equation are limited; and the
equation is conservative.  The last item, as I stated earlier, is actually
desirable.
          Bosanquet equation is the complex expression which you see here.
The terms A, X, and X  require separate calculations and the functions f,
and £„ are evaluated from tables.  As you might imagine, this equation is
rather difficult and time consuming to use.  Although it gives a generally
good approximation of plume rise, the approximations are consistently on
the high side.
    30i"he Lucas equation was developed by the Central Electricity Research
Laboratory in Great Britain.  It is the first of several equations we will
consider that does not have a momentum term,  only a buoyancy term.  In these
equations it is assumed that the upward momentum of a plume is negligible
compared to its buoyancy.  The Lucas equation, as you can see, is expressed
as the heat content of the effluent divided by wind speed multiplied by an
empirical coefficient, where the coefficient depends on the source being
considered.  This coefficient is variable for different source types and also
can vary for sources of the same type.  In other words, two power plants may
have different coefficients.  The approximations given by this equation are
consistently on the high side; also this equation is not very sensitive to
changes in stack diameter.
    3 1 New, having covered the background, we are prepared to discuss some
of the more commonly used plume-rise equations and to look at some of their
desirable and undesirable qualities.  The Stumke equation is a supposed
improvement of the Holland equation.  The first term on the right hand side

96

-------
of the equation can be considered the momentum term, stack gas exit velocity
times diameter divided by wind speed; and the second term, which expresses
the temperature difference between the gas and the atmosphere, can be thought
of as the buoyancy term.  This equation gives a good approximation of plume
rise, but it predicts a bit high for some of the observed data.  However, for
a widely applicable equation, it probably gives as good a result as any other
equation.
          G3NCAWE equation also has a European background.  It was developed
by a group concerned with emissions from oil refineries.  The equation gives
reasonably good estimates of plume rise, but it is not recommended for use
with large power plants.
    "Briggs has derived an equation from dimensional considerations.  The
general equation states that the plume rise is equal to the buoyancy flux,
F, to the 1/3 power, times the downward distance, to the 2/3 power, divided
by the wind speed measured at stack height.  The first equation should be
applied out to distances up to ten times the physical stack height.  At
greater distances the plume is assumed to have leveled off and if there is
no change in wind speed or stability, the plume centerline will remain at
an essentially constant height.  In addition, these two equations should be
applied only to sources emitting at least 20 megawatts of heat or at least
5 million calories per second.
     Under stable conditions a stability parameter, s, is introduced, pro-
portional to the potential temperature lapse rate/
     If there is no wind, the above equations become meaningless and thus
the last equation should be used.
    3^In this slide we can see how to calculate the buoyancy flux and sta-
bility parameter terms.  The buoyancy flux is equal to the temperature
difference between the stack gases and the atmosphere divided by the stack
gas temperature times the acceleration of gravity, the stack gas exit velocity
and the square of the stack radius , or it may be approximated by a constant
                                                                        97

-------
times the heat emission expressed in calories per second.  The stability
pararreter 's' is equal to the acceleration of gravity divided by the atmos-
pheric temperature at stack height multiplied by the lapse rate of potential
temperature through the layer in which the plume is dispersing.  Anyone who
is responsible for making plume rise estimates should be thoroughly familiar
with Briggs' work.
     The last three equations we have considered, the Lucas equation, the
CCNCAWE equation, and the Briggs equation, have been getting quite a bit of
discussion recently.  The series of Briggs equations are considered to be
the most up-to-date equations and those most applicable to large power plants,
which are the source of many air pollution problems.  However, as I indicated,
they are not all-encompassing, and do have certain deficiencies.
    35Qne other thing we should consider with regard to plume-rise equations
is some of the work done by Moses and Carson.  They have taken the basic form
of many plume-rise equations with a momentum term to some power times a co-
efficient plus a buoyancy term to some power times a coefficient plus a
constant.  They have taken this equation, used much of the aval .Table plume-
rise data, and by using regression techniques, have determined values for
the coefficient.  They have done this in three ways for the equation as you
see here, with C,- = 0, and with just the buoyancy term.  In most cases the
plume-rise equations which resulted are accurate to within approximately 30
meters.  One other item that should be noted about this technique is that
for some of their evaluations they got a negative coefficient for C-, this,
in effect, says that the momentum of a plume detracted from the plume rise,
which may be unrealistic.  This indicates the problems involved in deriving
an equation where it is fitted only to describe the data, without proper
consideration of the physical realities.
    36Now, let us take a look at the results that several of these equations
give in predicting plume rise.  We will consider the Lucas, the Briggs, and
the Holland equations, as applied to heat emission data from power plants.
As you can see, the Lucas equation overpredicts for most of the data.  The
98

-------
Briggs equation goes through approximately the middle of the data.  The
Holland equation underpredicts for all the data except for sources with very
large heat emissions.  Remember, however, the original data used to develop
the equation did not include heat emissions of this magnitude.  Therefore,
if the Lucas equation .is used, you would be predicting more plume rise than
actually occurred, the Briggs equation would give a good average, and with
the Holland equation, you would be obtaining an underestimate except for
very large sources.  Also, another problem with plume rise data is demon-
strated here.  These plume-rise observations are supposedly for the same
stack under similar atmospheric conditions, and you can see the wide range
of plume rises which are obtained.
    3'Now, let us consider an example in which we can use the Holland equation
to calculate plume rise. 3oTnese data are from the TVA's Shawnee power plant
near Paducah, Kentucky.  The stack gas exit velocity is 14.7 mps, the diameter
of the stack at the top is 4.27 meters, the height of the stack is 77 meters,
its exit temperature is 416 degrees Kelvin.  The atmospheric parameters are
wind speed five meters per second, atmospheric temperature 288 degrees Kelvin,
and pressure 1,000 millibars.  In evaluating this equation you can see the
computed plume rise is 63 meters.  The effective stack height is equal to the
physical stack height plus plume rise.  The physical stack height was 77 meters
and the calculated plume rise was 63 meters, giving us an effective stack
height for this plant and these atmospheric conditions of 140 meters. ^^A
similar calculation using the Briggs equation results in a plume rise of 158
meters.  This is 2 and 1/2 times the rise we calculated using the Holland
equation and, based on our previous discussion, is about what you might expect.
         have discussed various ways to estimate effective stack height
through plume-rise computations.  It should be pointed out that there are
adverse effects of meteorology and terrain which can make these plume-rise
calculations unrealistic.  The conditions which can adversely affect plume
rise are elevated inversions, irregular terrain, and changing thermal
regimes .

                                                                        99

-------
    41If -there is an inversion based at sane level above the ground where
the temperature of the air starts increasing with height, the inversion
base will act as a lid on the rise of the effluent, not allowing it to
penetrate through the inversion. 42Some actual scenes in the New York City
area when there were data to indicate an elevated inversion based at 1,500
feet, shows the effect of inversions on plume rise.  Hare we see a plume
which appears to be rising and gradually leveling out.  Then, all of a
sudden, its top becomes level, as if sheared off, or as if somebody put
their hand down on top of it. ^Another good example is this slide where,
in the upper left hand corner, the plume, which is relatively small, can
be seen to be rising gradually, then suddenly flattens out.
    44In cases of complex terrain, under reasonably stable flow on the wind-
ward side of a hill, the plume actually rises over the hill instead of im-
pacting into the side of the hill.  This is where a plume-rise calculation
would indicate that the plume should impact on the side of the hill where,
in fact, it rises over the hill.  On the downwind side, the turbulence
induced by the terrain causes the plume to be downwashed, so an effective
stack height calculation here would be of no use since the plume is actually
lowered by induced turbulence and is not allowed to rise naturally.45jn
cases of complex terrain and highly unstable meteorological conditions, as
you can see here, a calculation of an effective stack height would be of no
use in calculating the impact on the hill to the right, since the plume is
looping and has a much stronger impact on the hill than a calculation using
an effective stack height would indicate.  The plume impinges right on the
hill.
   46In this case, we have a plume emitted over grassy terrain behind the
hangers.  An effective stack height is established then, as the plume moves
out over the concrete runways where the convective turbulence is much
greater, the plume seems to be lifted to a new effective stack height.
    47Now we will consider what can be called negative plume rise, consisting
of evaporative cooling of the effluent and aerodynamic downwash.  A good

100

-------
example of the negative plume rise caused by evaporative cooling is what
happened to a power plant in Great Britain several years ago.  In this
case the effluent was put through a spray tower to absorb gases.  The
gases were thus cooled and saturated with water vapor.  Contact with the
cold surfaces of the duct work caused further cooling and condensation of
the water vapor.  When the effluent was released into the atmosphere, the
condensed water droplets evaporated, withdrawing the latent heat of vapori-
zation from the surrounding plume and air, thus causing the plume to cool
below the atmospheric temperature.  The plume thus had negative buoyancy and
the effective stack height was reduced to below the physical stack height.
In this case, the plume actually came right down to the ground.  For this
plant the adverse effect caused by evaporative cooling resulted in ground-
level concentrations which were greater than before anything had been done
to the stack effluent.
    48 With regard to aerodynamic dcwnwash, eddies which are the result of
mechanical turbulence around a building or low stack can affect the effective
stack height.  They cause the plume to be downwashed.  This is especially so
when the wind speed is high, momentum is small, and the plume is emitted
horizontally.  Most knowledge about this situation has come from wind tunnel
studies such as those conducted by Halitsky at New York University.  The
results of these investigations show that maximum downwash around a rectan-
gular building occurs when the wind direction is at a 45° angle to the major
axis of the structure and is a minimum when it is parallel to this axis.
Maximum downwash would occur when the wind blows from one corner of the
building to the opposite corner, and the least downwash would be when the
wind was blowing parallel to the building. ^Also, it has been shown that
effluents from flush openings on rooftops frequently flow in a. direction
opposite to the wind, due to counterflow induced by turbulence along the
roof in the turbulent wake above the building. ->OThe region of disturbed
flow extends up to twice the building height and 5 to 10 times its height
downwind as indicated by the streamlines. 51 TWO rules of thumb which are
                                                                      101

-------
  commonly used to prevent downwash are  (1) the stack gas exit velocity should
  be 1.5 times the average wind speed to prevent dcwnwash in • the wake of the
  stack, and  (2)  a stack should be 2.5 times the height of the building
  adjacent to the-stack, to overcome building turbulence.  These items
  should be kept in mind whenever one deals with a situation where down-
 wash could be taking place, although there is no really good quantitative
 way to handle all situations. 52jjere is an example of a power plant which
 has short stacks certainly less than two times the height of the adjacent
 building, and you can see the inferior stack which is subject to aerodynamic
 dcwnwash. ^IXMnwash is not readily apparent here but I have seen the situ-
  ation when the effluent came down directly on top of the river, flowed across
 the river right above its surface and up the hill on the opposite side.
->T?Vgain, in this instance, to assume an evalated emission from this  source
 would, be a mistake.
 102

-------

-------
Hum©
G
   ,
Air Resources Atmospheric Turbulence and Diffusion Laboratory
Environmental Science Services Administration
Oak Ridge, Tennessee
Prepared for
Nuclear Safety Information Center
Oak Ridge National Laboratory
U.S. ATOMIC ENERGY COMMISSION
Office of Information Services
1969

-------
Available as TID-25075 from
  National Technical Information Service
  U. S. Department of Commerce
  Springfield, Virginia 22151
Library of Congress Catalog Card Number:  72-603261
Printed in the United States of America
USAEC Technical Information Center, Oak Ridge, Tennessee
November 1969; latest printing, December 1972

-------
                                                                FOREWORD
Scientists  and  technologists  have  been  concerned  in  recent  years  about the
"explosion" of original literature engendered by the staggering volume of research and
development being undertaken throughout  the  world. It has proved all but impossible
for scientific workers to keep up with  current progress even in quite narrow fields of
interest. Automated retrieval systems for identifying original literature pertinent to the
interests  of individuals are being developed. These systems are only a partial solution,
however, because the original literature is too large, too diverse, too uneven in quality,
to fully satisfy by itself the information needs of scientists.
    In  this situation of vastly expanding knowledge, there is increasing recognition of
the valuable  role that can be  played by critical  reviews of the literature and  of the
results of research in specialized fields of  scientific interest. Mr. Briggs's study, the
third  published in  the AEC Critical Review Series, is an excellent  example of this
genre.
    This review is also significant as a further step in the unceasing effort of the AEC
to assure that nuclear plants operate safely. Plume Rise is a much needed addition in a
field in which a meteorologist must choose  from over 30 different plume-rise formulas
to predict  how  effluents from  nuclear plants are dispersed into the atmosphere. Mr.
Briggs  presents and  compares all alternatives, simplifies and combines results whenever
possible,  and makes clear and practical recommendations.
    The Atomic Energy  Commission  welcomes any comments about  this  volume,
about  the  AEC Critical Review  Series in general, and about other subject areas that
might beneficially be covered in this Series.
                                       in

-------
                                             CONTENTS
FOREWORD                                     iii
SYNOPSIS                                        1


1. INTRODUCTION                                 2
2.BEHAV/OR OF SMOKE PLUMES                   5
      Downwash and Aerodynamic Effects                  5
      Plume Rise                                    8
      Diffusion                                     11
3. OBSERVATIONS OF PLUME RISE                 16
      Modeling Studies                               16
      Atmospheric Studies                            18

-------
4. FORMULAS FOR CALCULATING
   PLUME RISE                                    22
      Empirical Formulas                              22
      Theoretical Formulas                             25
5. COMPARISONS OF CALCULATED
  AND OBSERVED  PLUME BEHAVIOR              38
      Neutral Conditions                              38
      Stable Conditions                               50
6. CONCLUSIONS AND RECOMMENDATIONS        57


APPENDIX A                                     61
      Effect of Atmospheric Turbulence on Plume Rise
APPENDIX B                                     65
      Nomenclature
APPENDIX C                                     67
      Glossary of Terms


REFERENCES                                     69


AUTHOR INDEX                                  77
SUBJECT INDEX                                   80
                        VI

-------
                                                                  SYNOPSIS
The  mechanism  of plume rise and dispersion is described in qualitative  terms with
emphasis  on  possible  aerodynamic,  meteorological,  and  topographical  effects.
Plume-rise observations and  formulas in the literature are  reviewed, and  a relatively
simple theoretical model is developed and compared with other models. All available
data are used to test the formulas for a number of idealized cases.
   The inverse wind-speed relation, Ah <* u ', is shown to be generally valid for  the
rise  of a hot plume at a fixed distance downwind in near-neutral  conditions. Nine
formulas of this  type are  compared with data from sixteen different sources, and  the
best  agreement is obtained  from the "2/3 law," Ah = 1.6F'6 u~! x^,  modified by  the
assumption  that  a ceiling  height  is  reached at  a distance  of ten stack heights
downwind. The term F is proportional to the heat emission. In uniform stratification
buoyant plumes  are seen to follow  the 2/3 law until a ceiling height of 2.9 (F/us)^ is
approached, where  s is proportional  to the potential temperature gradient.  In calm
conditions the formula Ah = 5.OF A s' " is in excellent agreement with a wide range of
data.
   Formulas of  a similar  type are recommended for nonbuoyant plumes on the basis
of much more limited data.

-------
1
INTRODUCTION
      The calculation of plume rise is often a vital consideration in predicting dispersion of
      harmful  effluents into the atmosphere, yet such a calculation is not straightforward.
      The engineer or meteorologist must choose from more than thirty different plume-rise
      formulas, and a casual search through the literature for help in choosing is likely to be
      confusing. The purpose of this survey is to present an overall view of the pertinent
      literature and to simplify and combine results whenever possible, with the objective of
      setting down clear, practical recommendations.
          The importance of stack height and buoyancy in reducing ground concentrations
      of effluents has been recognized for at least 50 years.1 In a 1936 paper Bosanquet and
      Pearson2  showed that under certain  conditions the maximum ground concentration
      depends  on the  inverse  square of stack  height, and experience soon confirmed this
      relationship.3  Later the  stack height in this formula came  to  be  replaced by the
      "effective stack height," which was defined as the sum of the actual stack height and
      the rise of the plume  above the stack. Since  smoke plumes from large sources of heat
      often rise several stack heights  above the top  of the stack even in  moderately high
      winds, plume rise  can  reduce  the  highest  ground  concentration  by  an order of
      magnitude or more.
          In spite of the importance of plume rise in predicting dispersion, there is much
      controversy about  how  it should be calculated. A  recent  symposium on  plume
      behavior,4 held in 1966, summarizes the current state  of affairs. Lucas expressed a
      desire for better agreement between empirical results and stated flatly, "There are too
      many theoretical formulae and  they  contradict one another!" Spurr lamented, "The
      argument for and against different plume rise formulae can be discussed clinically by

-------
INTRODUCTION

physicists and theoreticians, but the engineer who has to apply the formulae is obliged
to make  a choice." He then compared five recent formulas for a specific example and
concluded that the results varied by a factor of 4 in the calculated maximum ground
concentration. Even worse examples were given in the same symposium.
    There are reasons for the lack of agreement. Different  techniques for measuring
plume height and wind speed can account for some of the  disparity in the data, but
the differences  in  the  results  are  due primarily to  the  different  concepts of what
constitutes effective stack height. A recent paper by  Slawson  and Csanady states:
         With an ostrich-like philosophy, the effective stack height is often defined to be
      the point where the plume is just lost sight of. It is then not  very surprising to find
      that the observed thermal rise of the plume depends, for example, on a power of the
      heat flux  ranging from  % to 1.0, influenced by  a number of factors  including,
      presumedly, the observer's eyesight.j
It was natural for early plume-rise observers to assume that a smoke plume leveled off
in all conditions and that the plume was near the  height of leveling off when it was
inclined only slightly above  the horizontal; subsequent observations suggest otherwise.
The early formula of Holland,6  sometimes called the Oak Ridge formula, was based on
photographic data that  followed the plumes only 600 ft downwind,7 yet recent data
of the Tennessee Valley Authority (TVA) show plumes  still  rising at 1  and even  2
miles downwind.  Over  this  distance even a  small  inclination above  the horizontal
becomes important. The plume height normally of  greatest  concern is  that above the
point of maximum ground  concentration, and it seems logical to define this as the
effective stack height, as suggested by Lucas.4 A major difficulty with this definition  is
that none of the present observations  goes that far downwind. In practice we must
choose formulas for plume  rise on the  basis  of agreement with data on hand and, at
the same time, be aware of the limitations of the data.
    General  plume  behavior, which is discussed briefly in the next chapter, has been
described in greater  detail  in  other publications.  The  textbook by Sutton8  first
reviewed all aspects of diffusion, including plume rise. Pasquill9  surveyed the subject
in considerably more  detail and  on the  basis of more data than was previously
available. The first edition of Meteorology and Atomic Energy1 Q adequately covered
the qualitative  aspects  of  plume  rise and diffusion,  but  the new edition11   is
quantitatively more up-to-date.  An excellent survey by Strom12  reviewed all aspects
of plume behavior,  including the  potential  for modeling  dispersion. Smith briefly
reviewed the main qualitative considerations  in plume rise and  diffusion13 and more
recently discussed the practical aspects  of dispersion from tall stacks.14 The practical
experience of TVA has  been described by Thomas,15 by  Gartrell,1 6 and by Thomas,
Carpenter,  and  Gartrell.17  The British experience  with diffusion  from  large power
plants and their tall-stack policy has been analyzed by Stone and Clark.1 8
    Several  attempts  have  been  made at  setting  down  definite procedures  for
calculating diffusion, including the plume rise. The first, primarily concerned with dust
   fRef. 5, page 311.

-------
                                                                INTRODUCTION

deposition,  was  by  Bosanquet,  Carey, and  Halton.19  Hawkins and  Nonhebel
published a procedure based on a revised formulation for plume rise by Bosanquet.
More recently, Nonhebel2 '  gave detailed recommendations on stack heights, primarily
for small plants, based on the Bosanquet plume-rise formula and the Sutton diffusion
formula.8'9  Many of these recommendations were  adopted in the British Memoran-
dum on Chimnev Heights,22 which has been summarized by Nonhebel.23 Scorer and
Barrett24  outlined  a  simple   procedure   applicable  to  long-term  averages.  A
CONCAWEf publication25'26 presented a method for determining stack height for a
plant built on flat, open terrain with a limited  range of gas emissions; this method
included a formula for plume rise based on regression  analysis of data. The American
Society of Mechanical Engineers (ASME)27  has prepared a diffusion manual with
another formula for plume  rise. The implications of this formula and the CONCAWE
formula are discussed in  Ref. 28. Further discussions  of plume-rise questions can be
found in Refs. 29 to 33.
   fCONCAWli (Conservation of Clean Air and Water, Western Europe), a foundation established
by the Oil Companies' International Study Group for the Conservation of Clean Air and Water.

-------
2
BEHAVIOR
OF SMOKE  PLUMES
      Flume dispersion is  most easily  described by discussing separately three aspects of
      plume behavior: (1)  aerodynamic effects due to the presence of the stack, buildings,
      and topographical teatures; (2) rise  relative to the mean motion of the air due to the
      buoyancy and  initial vertical  momentum  of  the  plume;  and  (3) diffusion due to
      turbulence in the air. In reality all three effects can occur simultaneously, but in the
      present state of the  art they are  treated separately and are generally assumed not to
      interact. This is  probably not too unrealistic an assumption. We know that undesirable
      aerodynamic effects  can  be avoided with good chimney design. Clearly  the rise of a
      plume is impeded  b\ mixing  with  the an, but there is not much agreement on how
      important a role atmospheric tuibuleuce plays. It is known that a rising plume spreads
      outward from its center line fastei  than a passive  plume, but this increased  diffusion
      rate usually  icsults in an only negligible decrease of ground concentrations.
          The following sections discuss the three aspects ol plume dit fusion. Symbols and
      frequently used metcoiological  leims are delined in  Appendixes B and C.
                                                                           DOWIMWASH
                                                         AND AERODYNAMIC EFFECTS
         Oownwash of the plume into the low-pressure region  in the wake of a stack can
     occui if the efflux velocity is too low. If the stack is too low, the plume can be caught
     m the wake of associated buildings, where it will bring high concentrations of effluent

-------
                                                   BEHAVIOR OF SMOKE PLUMES
     (a] STACK DOWNWASH
                                             (t>) BUILDING DOWNWASH
                                                    n.
                             (c) TERRAIN DOWNWASH

                      Fig. 2.1 Undesirable aerodynamic effects.


to the ground and even inside the buildings. A similar effect can occur in the wake of a
terrain feature. These three effects are illustrated in Fig. 2.1.
   The wind-tunnel studies  of  Sherlock and Stalker34  indicate that down wash is
slight as long as w0 > l.Su,  where w0 is  the efflux velocity of gases discharging from
the stack  and u  is the average wind speed  at the top of the stack. These results are
consistent with elementary theoretical considerations: when w0 > l.Su, the upward
momentum  of  the stack gases  should  overcome the downward  pressure gradient
produced  by the  wind  blowing around the  stack on  the  basis of the pressure
distribution  around an infinite circular cylinder  in a crosswind  given by Goldstein;35
when w0 <  0.8u, the  smoke  can be sucked into the lower pressure region  across the
entire back  of the chimney.  If the plume is very buoyant, i.e., if the efflux Froude
number, Fr, is 1.0 or less, the buoyancy forces are sufficient to counteract some  of the
adverse pressure  forces, and  the  preceding criterion for  w0 could be relaxed. This
factor probably abates downwash at the Tallawarra plant, cited in Table 5.1, where
                              Fr =
                                  g(AT/T)D
                                            = 0.5
Experiments  are  still needed to determine  quantitatively the  effect of the efflux
Froude number on  the abatement  of downwash, unfortunately, the experiments of

-------
OO\\\\\ASH AND \1 RODYNA.MIC 111 TLCTS

Sherlock and  Stalker involved  only high values of Fr. and thus buoyancy was not a
significant factor.
    Nonhcbel''  recommends that \v0 be at least 20 to 25 ft/sec for small plants (heat
emission less than 10" cal/sec)  and that  \v0 be in the neighborhood of 50 to oO ft/sec
tor a large plant (e.g., with  a  heat  emission greater than  107  cal/sec). Larger efflux
velocities are  not necessary since such high winds  occur  very rarely, in fact, much
higher velocities may be  detrimental to the rise of  a buoyant plume because they are
accompanied  by more  rapid entrainment of ambient air into the  plume. Scorer3"
reports that, when efflux velocity must be low, placing a hori/ontal disk that is about
one  stack  diameter  in  breadth  about  the rim  of the  chimney  top  will prevent
down wash.
    One  of  the  most enduring rules of  thumb  for stack design was a recommenda-
tion3'7 made in  1^52 that stacks be built at least 2.5 times the height of surrounding
buildings, as illustrated in Fig. 2.2.  If such a stack is designed with sufficient  efflux
velocitx  to  avoid down wash,  the plume is  normally carried above the region of
downflow in the wake of the building.  If the stack height or efflux velocity is slightly
                       Fig. 2.2  I lo\\ past a u pical po\\cr plant.
lower, in  high  winds the plume  will get  caught  in the downflow  and be efficiently
mixed to  the ground b\ the increased turbulence in the wake of the building. If the
stack is less than twice the building height, at least part of the  plume is likely to be
caught in  the  cavity  of aii  circulating in the lee of the building; this can bring high
concentrations  of effluent to the ground near the building and even into the building.
The  streamlines in Fig.  2.2 also illustrate the advantage of constructing a chimney on
the side of the building facing the prevailing wind, \\here the air is still  rising.
   Still, this  is only a rough rule, because  the air-flow  pattern  around a  building
depends on the  particular shape of the building and on the wind direction. Details on
these effects are given by Halitsky.3^ Also, for sources with very small emissions, the
rule  for stacks  2.5 times higher than nearby buildings  may be impracticable.  Lucas39
suggested  a correction factor  for smaller stacks, and this has been  incorporated into
the  British  Mcm^rjnJmu  su Chimney  Heights.22  The  correction factor  is also
reported b\ Ireland'10  and  Nonhebel.23  The behavior of effluents from  very short
stacks has  been discussed by Barry."*' Culkowski.42 and Davies and Moore.43 For such
sources plume  rise is probably negligible.

-------
                                                         BEHAVIOR OF SMOKE PLUMES

         It is much more difficult to give any  rules about the effect of terrain features,
      partly because of the great variety of possibilities. Fortunately the general effect of
      terrain and buildings on a plume can be  fairly well modeled in a wind tunnel, such as
      the one at New York University  or at the D.S.I.R.  (National Physics  Laboratory,
      England).  Stumke44'45  gives a  method for correcting  effective stack height for a
      simple step in the terrain, but only streamline flow is considered.
         A curious aerodynamic effect sometimes  observed  is bifurcation, in which the
      plume splits into two plumes near the source. This is discussed by Scorer,   and a
      good  photograph of the phenomenon appears in Ref. 46. Bifurcation arises from the
      double-vortex  nature  of a plume in  a  crosswind, but it is not clear  under what
      conditions the two vortices can  separate. However, bifurcation is rare and appears to
      occur only in light winds.
         Scriven47  discusses  the  breakdown  of  plumes  into  puffs due  to  turbulent
      fluctuations in the  atmosphere. Scorer46 discusses  the breakdown into puffs of
      buoyant plumes with low exit velocity and  includes a photograph. The process appears
      to be associated with a low efflux Froude number, but  a similar phenomenon could be
      initiated through an organ-pipe effect, e.g., if the vortex-shedding frequency of the
      stack corresponds to a harmonic mode of the column of gas inside the stack.
PLUME RISE
         Although quantitative aspects of plume  rise  are the concern of the bulk of this
      report,  only the qualitative behavior is  discussed in  this section.  More  detailed
      discussions can be found in a paper by Batchelor48 and  a book by  Scorer.46 It is
      assumed that the plume is not affected by the adverse aerodynamic effects discussed in
      the previous section since these effects can be effectively prevented.
         The gases are turbulent as they leave the stack, and this turbulence causes mixing
      with the ambient air; further mechanical turbulence is then generated because of the
      velocity shear between the stack gases and the air. This mixing, called entrainment, has
      a critical effect on plume rise since both the upward momentum of the plume and its
      buoyancy  are greatly diluted by this process. The initial vertical velocity of the plume
      is soon greatly reduced, and in a crosswind the plume acquires horizontal momentum
      from the entrained air and soon bends over.
         Once the plume bends over, it moves horizontally at nearly the mean wind speed
      of the air it has entrained; however, the plume continues to rise relative to the  ambient
      air, and the resulting vertical velocity shear continues to produce turbulence and
      entrainment. Measurements  of the mean velocity distribution in a cross section of a
      bent-over plume show the plume to be a double vortex, as shown in Fig. 2.3. Naturally
      the greatest  vertical velocity and buoyancy occur near the center of the plume, where
      the least mixing takes  place. As the gases encounter ambient  air above the plume,
      vigorous mixing occurs all across the  top of the plume. This mixing causes the plume
      diameter to grow approximately linearly with height as it rises.

-------
PLUME RISE

    If the plume is hot or is of lower mean molecular weight than air, it is less dense
than air  and is therefore  buoyant. If the heat is not lost and the atmosphere is well
mixed, the  total buoyant force  in a given segment of the moving plume  remains
constant.  This causes the total vertical momentum  of  that segment  to increase at a
constant  rate, although  its vertical velocity may decrease owing to  dilution of the
momentum through entrainment.
          Fig. 2.3  Cross section of mean velocity distribution in a bent-over plume.
    At some point  downwind of the stack, the turbulence and vertical temperature
gradient of  the atmosphere begin to affect plume rise significantly. If the atmosphere
is well mixed because of vigorous turbulent mixing, it is said to be neutral or adiabatic.
In such an atmosphere the temperature decreases at the rate of 5.4°F per 1000 ft. This
rate of decrease, which is  called the adiabatic lapse rate (F), is the rate at which air
lifted adiabatically  cools  owing to  expansion as the  ambient atmospheric pressure
decreases. If the temperature lapse of the atmosphere  is less than  the adiabatic lapse
rate, the  air is said  to be stable or stably stratified. Air lifted  adiabatically in such an
environment becomes cooler than  the surrounding air and thus tends to sink back. If
the temperature actually increases with height, the air  is quite stable. Such a layer of
air  is called  an inversion. If the temperature lapse of the atmosphere is greater than the
adiabatic  lapse  rate,  the air is said to  be  unstable or unstably stratified.  Air lifted
adiabatically in  such  an environment becomes warmer than the surrounding air, and
thus all vertical motions tend to amplify.
    The potential temperature, 8, is  defined as the temperature that a  sample of air
would  acquire if it  were compressed adiabatically to some  standard pressure  (usually
1000 millibars). The  potential  temperature is a convenient measure  of atmospheric
stability since
where  F =  5.4°F/1000  ft =  9.8°C/km. Thus the  potential temperature gradient is
positive for stable air, zero for  neutral air, and negative for unstable air.

-------
10                                                       BEHAVIOR OF SMOKE PLUMES

          If the ambient air is stable, i.e., if 5d/8z> 0, the buoyancy of the plume decays as
      it rises since the plume entrains air from below and carries it upward into regions of
      warmer ambient air. If the air is stable throughout  the layer of plume rise, the plume
      eventually becomes negatively buoyant and settles back to a height where it has zero
      buoyancy relative to  the ambient air. The plume may  maintain this height for a
      distance of 20 miles or more from the source. In stable air atmospheric turbulence is
      suppressed and has little effect on plume rise.
          If the atmosphere is neutral, i.e.,  if 50/5z = 0, the buoyancy of the plume remains
      constant in a given segment of the plume provided  the buoyancy is a conservative
      property. This assumes no significant radiation or absorption of heat by the plume or
      loss  of heavy particles.  Since a  neutral atmosphere usually comes about  through
      vigorous mechanical mixing, a neutral atmosphere is normally turbulent. Atmospheric
      turbulence then increases the rate of entrainment; i.e., it helps dilute the  buoyancy
      and vertical momentum of the plume through mixing.
          If the atmosphere is unstable, i.e., if 50/5z < 0, the buoyancy of the plume grows
      as it rises. Increased entrainment due to convective  turbulence may counteract this
      somewhat, but the net effect on plume rise is not well known. The few usable data for
      unstable situations seem to indicate slightly  higher  plume rise than in comparable
      neutral  situations. On warm, unstable afternoons with light wind, plumes from  large
      sources rise thousands of feet and even initiate cumulus clouds.
          Measurements are made difficult  by fluctuations  in plume rise induced by
      unsteady  atmospheric conditions. On very  unstable days  there  are large vertical
      velocity  fluctuations  due to convective eddies that  may cause a  plume to loop, as
      shown in Fig. 2.5d. Figure 2.4 illustrates the large  variations in plume rise  at a fixed
      distance  downwind during unstable  conditions. On  neutral, windy days the plume
      trajectory at any one moment appears more  regular, but there still may be  large
      fluctuations in  plume rise due to  lulls and peaks in the horizontal wind speed. Since
      the  wind is responsible  for  the horizontal  stretching  of  plume  buoyancy and
      momentum, the wind strongly affects plume rise.  In stable  conditions there is very
      little turbulence,  and plume rise is also less sensitive to wind-speed fluctuations. This
      can be  seen in  Fig. 2.4. In this case the plume leveled off in stable air, and its rise
      increased in a smooth fashion as the air became less and less stable owing to insolation
      at the ground.
          One might ask whether plume rise is affected by  the addition  of latent heat that
      would occur if  any  water vapor in the stack gases were  to condense.  This is  an
      important question because there  may be as much  latent heat as there is sensible heat
      present in a plume from a conventional power plant.  It is true that some water vapor
      may condense as the  plume entrains cooler air, but  calculations  show that in most
      conditions the  plume  quickly entrains enough air to cause  the  water to  evaporate
      again. Exceptions occur on very cold days, when  the air has very little capacity for
      water vapor, and in layers of air nearly saturated with water vapor, as when the plume
      rises through fog. Observations by Serpolay49  indicate that on days when cumulus
      clouds  are  present condensation  of water  from  entrained  air  may  increase  the
      buoyancy of the plume  and enhance its ability to penetrate elevated stable layers

-------
DIFFUSION
                                                                                   11
  1000
   800
-  600
   400
   200
                                         TIME
   Fig. 2.4 Fluctuations of plume rise with time (Gallatin Plant, Tennessee Valley Authority).
Ordinarily only the sensible heat of the plume should be used in calculations.
    One  might also  ask  whether thermal radiation  can significantly alter the heat
content  of  a plume,  i.e., its  buoyancy. Not  much is known  about the radiative
properties of smoke  plumes, but crude calculations show that radiation is potentially
important only for very opaque plumes some thousands of feet downwind and should
have little  effect on clean  plumes from modern power plants or on plumes  from
air-cooled reactors.  Plumes from TVA plants have been observed to maintain a
constant height for 20 miles downwind in the early morning; thus there appears  to be
negligible heat loss due to radiation.
                                                                          DIFFUSION
    Detailed diffusion calculations are beyond the scope of this review, but the main
types of diffusion situations should be discussed with regard to plume rise. On a clear

-------
12
BEHAVIOR OF SMOKE PLUMES
               \/ACTUAL TEMPERATURE PROFILE

                 ,ADIABATIC LAPSE RATE
                                           (a) FANNING
                                          (b) FUMIGATION
                                           (c) CONING
                                           (d) LOOPING
          TEMPERATURE—"-                 (e)  LOFTING




                    Fig. 2.5  Effect of temperature profile on plume rise and diffusion.

-------
DIH'USION                                                                         13

night  the ground radiates heat, most of which passes out into space. In this process the
air near the ground is cooled,  and an inversion is formed. The  stable  layer may be
several thousand feet deep; so  most plumes rising through it lose all their  buoyancy
and level off. This behavior is called fanning and is pictured in Fig. 2.5a.  When the sun
comes up,  convcctive eddies develop and penetrate higher and higher as the ground
warms up.  As the eddies reach the  height  at which  the plume has  leveled off, they
rapidly  mix the smoke toward the ground while the  inversion  aloft  prevents upward
diffusion. This  phenomenon, called fumigation, can bring heavy concentrations of
effluent to the  ground (Fig. 2.5b). Just after an inversion has been broken down by
convectivc  eddies  or in cloudy, windy conditions, the atmosphere is well mixed and
nearly neutral. Then the plume rises and diffuses in a smooth fashion  known as coning
(Fig. 2.5c). As  the heating of the ground  intensifies,  large convective eddies may
develop and twist and fragment  the plume in a looping manner (Fig. 2.5d). Diffusion is
then more  rapid than in a  neutral atmosphere. The convection dies out as the sun gets
lower, and  an inversion again  starts to build from the ground up. This  ground inversion
is weak enough  at first that the plume can penetrate it, and the plume diffuses upward
but is prevented by the stability below from diffusing downward. This lofting period
(Fig. 2.5e)  is  the most ideal time to release  harmful effluents since they  are then least
likely to reach ground.
    The meteorological conditions that should be considered in stack design depend on
the size of the source,  the climatology   of  the  region, and the   topography. In
reasonably  flat terrain, high wind with neutral stratification usually causes the highest
ground concentrations since there is the least plume rise in these conditions. The mean
concentration  of the effluent in the  plume is reasonably well described by a Gaussian
distribution, for which the maximum ground concentration is given by

                                    2Q            Q
                                          = 0.164—                        (2.2)
                                ay Treuh          uh

where  Q is  the  rate at  which pollutant  is emitted, u is the mean wind speed at the
source height, and  h is the effective stack  height  (defined as  the sum of the actual
stack height, hs, and the plume rise, Ah); aijay is the ratio of the vertical dispersion to
the horizontal dispersion  and is  equal to about 0.7  in a neutral atmosphere for an
averaging period of 30 min.25 Variation with distance has been neglected in deriving
Hq. 2.2. This equation is valid only when the atmosphere is neutral from  the ground
up to at least twice the effective  stack height. Inversions may exist below this height
even in windy conditions.  A diffusion  model  for this case is given  by  Smith  and
Singer.50 If the  plume reaches the height of the inversion and penetrates it, as can be
predicted by Eq. 4.30,t  none of the effluent reaches the ground. If the plume does not
penetrate, the inversion acts as an  invisible ceiling and prevents upward diffusion.
    A good  measure of the efficiency of the diffusion process  on a given  occasion is
        "Basic Theory Simplified" in Chapter 4.

-------
14                                                        BEHAVIOR OF SMOKE PLUMES

      Q/X, the effective ventilation, which has the dimensions of volumetric flow rate (/  /t).
      For the case just described,


                                -=6.1uh2 =6.1u(hs +Ah)2                       (2.3)
                                A.

      Naturally the effective ventilation is large for extremely high wind speeds, but it is also
      large at low values of u because of very high plume rise. It is at  some intermediate
      wind speed  that Q/x attains a minimum, i.e., x attains a maximum; this wind speed is
      called the critical wind speed. If the dependence of Ah on u is known, Eq. 2.3 can be
      differentiated and set equal to zero .to find the critical wind speed. The result can be
      substituted into the plume-rise equation  and into Eq. 2.2 to find the highest expected
      ground concentration for the neutral, windy case, Xmax-
          There is evidence that fumigation during calm conditions may  lead to the highest
      ground-level concentrations at large power plants. This type of fumigation can occur
      near the  center of large slow-moving high-pressure  areas in so-called "stagnation"
      conditions.  Such  high-pressure  systems  usually  originate  as  outbreaks  of cold,
      relatively dense air, and, as these air masses slow  down, they spread out much in the
      manner of cake batter poured into a pan.  Since the air underneath the upper surface of
      these air masses is appreciably colder than the air above it, a subsidence inversion forms
      and presents a  formidable barrier to upward mixing; such an inversion normally occurs
      1500 to 4000 ft above the ground.51 In combination with a near-zero wind speed, a
      subsidence inversion severely limits atmospheric ventilation, and the Little ventilation
      that occurs  is due to convective mixing from the ground up  to the inversion.
          Fortunately such circumstances are rare except in certain geographical areas. The
      southeastern United States, one such region, averages 5 to 15 stagnation days a year
      with the higher figure occurring in the Carolinas and Georgia.5:  Nevertheless, there is
      only  one  outstanding  case  of  fumigation  during  stagnation  in all  the  years  of
      monitoring  S02 around TVA power plants. In this instance ground concentration near
      an isolated  plant was 50% higher than the  maximum  observed in  windy, neutral
      conditions,  and this condition continued for most of one afternoon. The wind speed
      was 0 to 1  mph, and the effective ventilation, as defined above, was 1.5  X  108  cu
      ft/sec  (4.3 x 106 m3/sec). This value is adequate  for  a small plant but too small for a
      large  plant.  There is not much hope of improving the effective ventilation in this rare
      condition, for  a stack would have to be thousands of feet high  to ensure that the
      plume  could  penetrate  a subsidence inversion.  The only  way  to  reduce ground
      concentrations  in  this  case  seems  to  be to reduce the  emission  of pollutants;
      accordingly, TVA stockpiles low-sulfur coal for use when the Weather Bureau predicts
       stagnation conditions.
          Similar  conditions occur  under marine inversions, such as  are  found  along the
      Pacific coast of the United States. The inversions there are sometimes less than 1000 ft
      above the ground,5l and plumes from high stacks can often penetrate them. Such
      penetration can be predicted by equations presented in later chapters.

-------
DIFIUSION                                                                          15

    Fumigation associated with inversion breakdown may be serious when topography
is prominent. If the  plume does not rise out of a  deep valley during the period of the
nighttime inversion,  the pollutant will mix fairly uniformly down to the ground during
fumigation; therefore concentration is given by
where  u is the average velocity of the along-valley drainage flow at  night, h is  the
effective stack height at night, and W is the average width of the valley up to height h.
    An  elevated plateau can also  be  subjected to intensified fumigation  if during an
inversion the plume rises slightly  higher than the plateau and drifts over it. This  has
occurred at a plant on the Tennessee River where the river cuts a 1000-ft-deep gorge
through the Cumberland Plateau.17  Careful consideration  should be given to this
possibility  at such  a site. Topographic effects are discussed  by Hewson. Bierly. and
Gill.52

-------
3
OBSERVATIONS
OF  PLUME RISE
      Dozens of plume-rise observations have been made, and each is unique in terms of type
      of source  and technique  of  measurement.  Observations  have  been made  in the
      atmosphere, in wind tunnels, in towing channels, and in tanks. Brief descriptions of
      these experiments are given in this chapter.
 MODELING STUDIES

         Plume rise  is a phenomenon of turbulent fluid mechanics and, as such, can be
      modeled; i.e., it can be simulated on some scale other than the prototype. There are
      obvious advantages to modeling plume rise. For example, the model plume can be
      measured much less expensively than the real plume since it is not necessary to probe
      high above the  ground, and the variables can be controlled at will. The main difficulty
      is in ensuring that the behavior of the model plume essentially duplicates that of a real
      smoke plume. The most obvious requirements are that all lengths be scaled down by
      the same factor and that the wind speed and efflux velocity  be scaled down by
      identical factors. For exact similarity the Reynolds number has  to be the same in
      model and in prototype. The Reynolds number is defined by

                                         Re =17                             (3.1)

      where v is a characteristic velocity, I is a characteristic length, and v is the kinematic
      viscosity of air or the fluid in which the model is measured. Exact similarity is seldom

                                          16

-------
MODELING STUDIES                                                                 17

possible in modeling since Re is of the order of 106 for a real plume. Fortunately fully
turbulent  flow is not very dependent on Reynolds number so long as it is sufficiently
high. In most experiments Re is at least 103 on the basis of efflux velocity and stack
diameter, but the adequacy of this value is not certain.
    For buoyant  plumes  the  Froude number must be the same in model  and in
prototype. Since we are unable to scale down gravity, which is a prerequisite  for the
existence of buoyancy, the basic requirement is that

                                                                           (3.2)
                                model  * ' 'prototype

provided the temperature or density ratios are kept unchanged.
    Numerous measurements  have been  made  on  the  simple circular jet.53'54
Schmidt5 s  first  investigated  the heated plume with zero wind. Yih5 6 studied the
transition from laminar to turbulent flow in a heated plume. Later, Rouse, Yih, and
Humphreys57  studied the detailed distribution of vertical velocity and temperature in
a  fully  turbulent  hot  plume  from   a  gas flame  near  the floor  of an  airtight,
high-ceilinged  room. They measured temperature with a thermocouple  and velocity
with a 1 '/4-in. vane on jeweled bearings. The important result of all these investigations
is  that  both jets and hot plumes are cone shaped in  calm,  unstratified air. The
half-cone angle is smaller for  the heated plume than for  a jet, and the decreases of
temperature and velocity with distance above the source are consistent with heat and
momentum conservation principles. Also, the cross-sectional  distributions  of vertical
velocity  and temperature  excess  are  approximately Gaussian except close to the
source. The characteristic radius describing  the  temperature distribution in a heated
plume is  16% greater than that for the velocity distribution.
    Several modeling studies have been made on heated plumes rising through a stable
environment. Morton, Taylor,  and Turner58  confirmed predictions by using measured
releases of dyed  methylated spirits in a  1-m-deep  tank  of stratified salt solution.
Crawford and Leonard5 9  ran a similar  experiment with a small electric heater to
generate a plume on the floor of an ice rink. The invisible plume was observed with the
Schlieren  technique, and  convection thermocouples were  used to measure the
temperature profile  of the air  above the  ice. Their results are, in fact,  in  good
agreement with  those of Ref. 58,  although they miscalculated the  constant in the
equation of Ref. 58 by  a factor of  21/,. Vadot50 conducted experiments  with an
inverted  plume of heavier  fluid  in a tank of salt solution. His inversions were  quite
sharp in contrast to the smooth density gradients used in the preceding studies.
    A number of wind-tunnel investigations of jets in a crosswind have  been made. The
early  study of Rupp and his associates6 '  has been used as the basis for a momentum
contribution to plume rise by several investigators. Callaghan and Rugged62 measured
the temperature  profile  of heated  jets in  experiments in  which the efflux velocities
were of the order of the  speed of sound. Keffer and Baines63  measured rise for only
four stack diameters downwind and obtained some  velocity and turbulence intensity
measurements within the jets. Halitsky64  and  Patrick65 summarized  the work of

-------
18                                                     OBSERVATIONS OF PLUME RISE

      previous  investigators. In addition,  Patrick presented  new measurements to  about
      20 stack diameters downstream, including detailed profiles of velocity and concentra-
      tion of a tracer (nitrous oxide).
          The effect of buoyancy on plume rise near the stack was studied by Bryant and
      Cowdrey66'67 in low-speed wind in a tunnel. Vadot60 made a study of buoyancy
      effects in a towing channel  with both stratified and unstratified fluids. This study was
      unusual in that the  ambient fluid  was at  rest  and the effect of crosswind was
      incorporated  by towing  the source at a constant  speed down the channel. This is a
      valid experimental technique  since motion is only relative. However, Vadot's source
      was a downward-directed stream  of  dense fluid. There is some question whether a
      bent-over plume from such a source  behaves as a mirror image  of a bent-over plume
      from  an  upward-directed stream  of light fluid.  Subtle changes in the entrainment
      mechanism could take place owing to centrifugal forces acting on the more dense fluid
      inside the plume. The recent treatment by Hoult, Fay, and Forney68 of past modeling
      experiments tends to confirm  this. The bent-over portion of a hot plume  behaves
      much like  a  line thermal, which  was modeled for both dense  and  light plumes by
      Richards,6 9 who found that the width of the thermals increased linearly with vertical
      displacement  from their  virtual origins, just as had been observed  for jets and plumes
      that were not bent over. The line thermal was also modeled numerically by  Lilly.70
      Lilly did not have enough  grid points to reach the shape-preserving stage found in
      laboratory  thermals,  but,   as larger  computers are  developed, numerical modeling
      should be quite feasible.  Extensive experiments made recently by Fan71  in a modeling
      channel included plume rise both into a uniform crossflow and into a calm stream with
      a constant  density gradient. In the latter case most of  the plumes were inclined; i.e.,
      the stacks were not vertical. Although the buoyancy of these plumes was varied, they
      were momentum dominated for the  most part. The behavior of plumes with  negative
      buoyancy in a crosswind  was modeled by Bodurtha.72
ATMOSPHERIC STUDIES

          The first full-scale plume-rise data were given in  an appendix to the Bosanquet,
       Carey, and Halton paper19 of 1950. The center lines of plumes from four chimneys
       were  traced  from  visual observation onto a Perspex screen. The observations were
       carried only as far as 800 ft downwind of the stacks, where apparently the visibility of
       the plumes was  lost. These observations  also appear in  a  paper  by Priestley.73
       Holland6  published some of the details of the observations that he used in deriving the
       Oak Ridge formula, but  the distance of observation was not mentioned. According to
       Hawkins and Nonhebel7 the plume heights were measured at only two or three stack
       heights  downwind and were obtained from photographs. Holland found only a small
       correlation between  plume rise and the temperature gradient, which was measured
       near the ground. However, the plume is affected only by the temperature gradient of
       the air  through which  it  is rising, and the gradient near the ground is not a good

-------
ATMOSPHERIC STUDIES                                                            19

measure of the gradient higher up. Stewart. Gale, and Crooks74'73 published a survey
ot plume  rise  and ditfusion parameters at the Harwell pile. Vertical surveys  of the
invisible plume were conducted by  mounting up to ten Geiger counter units  on the
cable of a mobile barrage balloon. The stack was a steady, known source  of radioactive
argon (4 J Ar),  and the Geiger units were arranged to measure the disintegration of beta
particles, which have  a maximum range of only 3 m in air. Again, the temperature
gradient was measured well below plume level except for a few runs that were made in
neutral conditions. Most of the wind-speed measurements were also made at a height
well below the plume height.  Since wind speed generally increases with height, the
reported wind speeds are probably too low for such runs.
    Ball76 made measurements on very small  plumes from lard-pail-type oil burners.
The heights were estimated at 30 and t>0 ft downwind by visual comparison with 10-ft
poles and were averaged  over  2 or 3 min. There was some tendency  for the burning
rate to increase with wind speed. Moses and Strom7 7 ran experiments on a source with
about  the same  heat emission, but  here  the effluent  was fed  into  an  111-ft
experimental  stack with a blower.  Plume-rise  data  at 30  and  t>0 m downwind were
obtained photographically and averaged over  4 min. Wind speed was interpolated at
plume  level from measurements from a nearly 150-ft meteorological tower.  The
temperature gradient was measured between the 144- and 5-ft levels of the tower.  This
provided only a fair measure of the actual gradient at  plume level since the gradient
above  111 ft  may be quite different  from that near the ground. In only 2  of the
3b runs, the plume appeared to level off owing to stable conditions. These data tend to
be dominated b\ momentum rise.
    Danovich and  Zeyger'8 published some  plume-rise data  obtained from  photo-
graphs. However,  the effective rise was assumed to occur when the plumes were still
inclined at 10 to  15"  above horizontal, and plumes  have been  observed to rise many
times the height  at  this  point.  Some  interesting data  were obtained  from exhaust
plumes  of rocket motors  by van Vleck and Boone,'9  including some  runs with
complete temperature profiles  furnished. The sources ranged up to 1000 Mw, which is
about ten times the  heat-emission rate of a large power plant stack. However, they
were not true continuous sources since burning times varied from 3 to 60 sec.
    Extensive plume  photography was  carried  out at two moderate-size power plants
in Germany by Rauch.50 Plume center  lines were determined for 385 runs at Duisburg
and  for 43 runs at Darmstadt. Each determination  consisted  of two or three time
exposures of about 1 min each, together with  five instantaneous pictures taken at set
time intervals. The horizontal speed of  the plume was calculated by following irregular
features of the plume from one  negative  to the next. This method should provide a
good  measure  of the wind  speed  experienced  by  the  plume. In  most  of  the
photographs, the plume center line could not be determined for a distance downwind
of more  than 1000 ft, although a  few could be  determined out to  3000 ft.  The
accuracy   of  the  temperature-gradient measurements  was  such  that  only general
stability classifications could be made. In practice no measurements in  very unstable
conditions were made because of looping, and no measurements in stable conditions
were made far enough downwind to show the  plume leveling off.  In fact, not one of

-------
20                                                     OBSERVATIONS OF PLUME RISE

      the  428 plume  center  lines  leveled  off. It would therefore  seem  that  Rauch's
      extrapolation of these center lines to a final rise is rather speculative.
          Much more extensive observations, consisting of about 70 experiments on more
      than 30 smoke stacks in Sweden, were recently made by Bringfelt,8 l and some of the
      preliminary data have been reported  by Hogstrom.82 Each experiment consisted in
      taking about one  photograph a minute  for 30  to 60 min.  The  center lines  were
      measured up  to  9000 ft downwind, and wind  speed and temperature  gradient  were
      measured at the plume level.
          Some observations of plume rise at a small  plant were reported by  Sakuraba and
      his associates.83 The best fit to the data was given  by Ah <* u~%, but downwash was
      likely at the  higher wind speeds since the wind speeds exceeded the efflux velocity.
      The temperature gradient and distance downwind were not given. More observations
      were  carried  out  by the Central Research Institute of Electric Power Industry,
      Japan,84 in which temperature and wind profiles were measured, as well as the vertical
      profile of SO2 concentration at 1 km downwind.
          Several groups  have  shown continuing interest  in plume-rise measurements. The
      Meteorology  Group at Brookhaven  National  Laboratory  has conducted  several
      programs by  burning rocket fuel  on  the  ground near their well-instrumented 420-ft
      meteorological  tower. Limited data85  were  published in 1964  from tests in which
      there was some difficulty in obtaining a constant rate of heat release. This problem has
      been overcome, and more detailed  data are available.86
          Csanady published plume-rise observations8 7 from the Tallawarra power station in
      New South Wales in 1961. Plume  rise was measured photographically, and wind speed
      was determined from displacement of plume  features in a succession of photographs.
      Csanady has been conducting a continuing program of plume-rise and dust-deposition
      research at  the University of Waterloo in  Ontario  since  1963. More-elaborate
      photographic measurements of plume rise made at the  Lakeview Generating Station
      were  published  by  Slawson  and Csanady.5'88 Tank,  wind-tunnel, and  small-scale
      outdoor studies are now  in progress.89'90
          The  Central Electricity Research Laboratory in England has been conducting
      plume-rise studies for some time. In 1963 they published results from the Earley and
      Castle Donington power stations.9  1 The measurements were unique in that the plumes
      were  traced a long  distance  downwind by  injecting balloons into  the base of the
      chimney.92 The balloons were observed to stay within  the plumes when the plumes
      were  purposely made visible, but  there may have been systematic errors due to the
      relative  inertia  and buoyancy of the balloons.  Although  some  of  the  balloons
      continued to  rise  beyond  2 miles downwind,  the  reported  rises were in the range
      3600 to 6000 ft  downwind. The   motion of  the  balloons  provided  a convenient
      measure of wind speed. More recently measurements were made by Hamilton9 3 at the
      Northfleet Power Station by using lidar to detect the plume.  Lidar is an optical  radar
      that  uses  a  pulsed ruby  laser. It  measures  the  range  and  concentration  of
      light-reflecting particles and can detect smoke plumes even when they become invisible
      to  the  eye.94'95  Some  searchlight determinations of the height  of the Tilbury  plant
      plume are also given in Refs. 93 and 96.

-------
ATMOSPHERIC STUDIES                                                             21

   The Tennessee Valley Authority has also conducted plume-rise measurements over
many years. The plume-rise and dispersion results97'98 published in  1964 were based
on  helicopter probes  of S02  concentrations  in  the plume.  The helicopter  also
measured  the  temperature gradient up  to the top of the plume. Plumes in inversions
were  observed to  become  level and maintain a nearly constant elevation as far  as
9 miles downwind. Much more detailed studies at six TVA plants have recently been
completed.99  Heat emissions ranged from  20 to 100 Mw per stack with one to nine
stacks operating.  Complete  temperature  profiles were obtained by helicopter, and
wind profiles were obtained from pibal releases about twice an hour. Such intermittent
sampling of wind  speed  does not provide  a good  average value, however, and may
account for some of the scatter in the results. After several different techniques were
tried, with good agreement among them,  infrared photography was used to detect the
plume center  line. Complete plume trajectories as  far  as  2 miles  downwind  were
obtained from the photographs.
    There have been a few atmospheric studies concerned particularly with plume rise
in stable air. Vehrencamp, Ambrosio, and  Romie100  conducted tests on the Mojave
Desert,  where very  steep surface inversions  occur in the early morning. The heat
sources were shallow depressions, 2.5, 5,  10, and 20 ft in  diameter, containing ignited
diesel oil.  Temperature profiles were measured  with  a thermocouple  attached to a
balloon, and the dense black plumes were easily  photographed. Davies101 reported a
10,000-ft-high plume rise from an oil fire at a refinery in  Long Beach, Calif. The heat
release was estimated to be of the order of 10,000 Mw;1 °2  i.e., about 100 times the
heat emission  from a large power plant stack. Observations of plume rise into multiple
inversions over New York City were presented recently  by Simon  and Proudfit.103
The plumes were  located with a fast-response S02  analyzer borne by helicopter, and
temperature profiles were also obtained by helicopter.

-------
4
FORMULAS
FOR  CALCULAT/NG
PLUME  RISE
      There are over 30 plume-rise formulas in the literature, and new ones appear at the
      rate of about 2 a year. All require empirical determination of one or more constants,
      and some formulas are totally empirical. Yet the rises predicted by various formulas
      may differ by a factor greater than 10. This comes about because the type of analysis
      and the selection and weighting of data differ greatly among various investigators.
         Emphasis is given  here on how the formulas were derived and on the main features
      of each. Complicated formulations are omitted since readers may check the original
      references. For convenience all symbols are defined in Appendix B.


 EMPIRICAL FORMULAS

      Formulas for Buoyant Plumes

         Of the purely  empirical plume-rise formulas, the first to be widely used was that
      suggested by Holland6 on the basis of photographs  taken at three steam plants in the
      vicinity of Oak Ridge, Tenn.  The observed scatter was large, but the rise appeared to
      be roughly proportional to the reciprocal of wind speed. Holland used the wind-tunnel
      results of Rupp and his associates6 ' for the momentum-induced part of the rise and,
      by assuming a linear combination of momentum and buoyancy rises, found the best fit
      to the data with
                      Ah = 1 .5       D + 4.4 X  10-            H             (4
                                u                [ cal/sec                  l
                                         22

-------
EMPIRICAL FORMULAS                                                            23

The dimensions of constants are given in brackets. Thomas1 5 found that a buoyancy
term twice as large as that in Eq. 4.1 gave a better  fit  to observations  at the TVA
Johnsonville plant, and Stumke104 recommended a rise nearly three times that given
by Eq. 4.1 on the basis of comparisons with many sets of observations.
    Another early empirical  formula  was suggested by Davidson105 in 1949 on the
basis of Bryant's66 wind-tunnel data:

Equation  4.2, although a dimensionally homogeneous formula, is  physically  over-
simplified in that the buoyancy term (AT/TS) docs not take into consideration the
total heat emission  or the effect of gravity, without which buoyancy does not exist.
The main weakness  of Eq. 4.2 is that it  is based on data obtained at  only seven  stack
diameters downwind and often greatly underestimates observed rises.
    Berlyand, Genikhovich, and Onikul1 °6 suggested

                                    /w0\         F
                           Ah = 1.9  ( — )  D+ 5.0—                       (4.3)
                                    V  u /         u

where  F is a quantity that is proportional to the rate of buoyancy emission from the
stack.  This formula is dimensionally consistent, but few details are  given about the
observations on which it is based. The constant in the buoyancy term, 5.0, is curiously
almost  two  orders  of  magnitude  smaller  than  the  constants  recommended by
Csanady,87 by Briggs,1 ' •' °7 and by the new ASME manual.27
    On the basis of  data from  four stacks, namely, the  Harwell stack,74'75 Moses and
Strom's experimental stack,77 and the two stacks reported by Rauch,80  Stumke108
derived the formula
              AH = 1.5    >    D+118        o     l+1          (4.4)
                       \ u /          [ sec

The  argument for omitting emission velocity from the buoyancy term is not clear.
The  constants and exponents for the various terms resulted from applying the method
of least squares to the observed and calculated rises.
   Lucas, Moore, and Spurr9 '  fitted observed plume rises at two of their plants with


                                                °"                      (4.5)
                                                  u                      v   '

The  heat emissions varied from 4 to 67 Mw, and the plumes were traced to about a
mile  downwind  by releasing  balloons  in the stacks  (see "Atmospheric Studies" in
Chap. 3). The  formula is based on a simplification of Priestley's theoretical plume-rise
model.73 The best values  for  the constant  in  Eq. 4.5 differed by 25'"- at  the two

-------
24                                          FORMULAS FOR CALCULATING PLUME RISE

      plants, and further  variations have been observed at other plants.93 Lucas     noted
      some correlation with stack height and suggested a modification of Eq. 4.5:


                         Ah = (134 + 0.3 [ft'1 ] h.)
                              v                s
          Recently a CONCAWE working group2 5 '2 6 developed a regression formula based
      on  the assumption  that plume rise depends mainly on some power of heat emission
      and some  power  of wind  speed.  The  least-squares fit  to  the logarithms  of the
      calculated-to-observed plume-rise ratios was

                                          [ft-(ft/sec)3'4l  Q$
                                Ah =1.40  —	H  -J-                    (4-7)
                                          L (cal/sec) *  J  u/4

      Data from eight stacks were used,  but over 75% of the runs came from Rauch's80
      observations at Duisburg, i.e.,  from just one stack.  Most of these data fall into a small
      range of QH and of u, and therefore it is difficult to establish any power-law relation
      with confidence.
          Even more recently Moses and Carson11 ° developed a formula of the  same  type as
      Eq. 4.7 with data  for  ten different stacks,  but again the Duisburg  observations were
      heavily weighted.  A momentum term of the type that appears in the  formulas of
      Holland,6  Berlyand and his associates,106  and Stumke108  was included, but the
      optimum value of  the  constant turned out to be very small. The least-squares fit was
      given by
                                       •rir-                        -
                                           |_ (cal/sec) <* J  u                       v

       Actually, changing the exponent of QH to V3 or %  increased the standard error very
       little. This insensitivity is due partly to the small range of QH into which the bulk of
       the data fell. Another  shortcoming of this  analysis, as well as of the analysis by
       Stumke, is  that absolute  values of the error in predicted rises were employed. This
       tended to weight the analysis in favor of situations with high plume rise; cases with
       high wind speed counted very little since both the predicted and the observed rises,
       and hence their differences, were small. Relative or  percentage error, such as used by
       CONCAWE by means of logarithms, results in more even weighting of the data.

       Formulas for Jets

           One of the first empirical relations for the rise of pure jets was  given by Rupp et
       al.6' This relation was determined from photographs of a plume in a wind tunnel. The
       investigators found the height of the jet center line at
                                     Ah =1.5  -^
                                                                                 (4.9)

-------
THEORETICAL TORMULAS                                                         25

the point at which the plume became substantially horizontal, i.e., when its inclination
was only 5  to S1"'
    Subsequent investigators have all given empirical relations for the jet center line as
a (unction  of downwind distance. The results are summarized in Table 4.1 for the case
in which the density of the jet is the same as that of air. A theoretical formula to be
given later in the chapter is included for comparison.


                                   Table 4.1
                 COMPARISON OF EMPIRICAL RESULTS FOR JET
             CENTER LINES AS A FUNCTION OF DOWNWIND DISTANCE
Range of
Investigator R = (.WQ/U)
Eq. 4.33
Rupp et al.61 2 to 31
Callanhan and
n '62
Ruggen
Gordier (b\
Patrick65)
Shanelorov (by
Abramovich11 ') : to 22
Patrick65
Concentration 6 to 45
Velocity S to 54
Maximum
x/D Ah/D
1.44 R°'6V,
47

SI 1.91 R0-61 1.\,

1.31R°'74ix

0.84 R°'7S(v

22 1.00R°-85(.x.
34 1.00R°-S5(.x
Ah/D at 5.7°
Inclination
.•Dl0-33


^Dl0'30

Dl0-37

,D)0.39

D10.34
D)°-3S
^.:RI
>1.5 R1

4.0 R°

3.3 R1

l.S R1

1.9 R1
2.3 R!
.00
.00

.8 7

.1 7

.2 S

.2 9
.37
    The early Callaghan and Rugged62 experiments involved heated, supersonic jets in
a very narrow  wind tunnel; so application of their results  to  free,  subsonic jets is
questionable. Since the penetration was determined as the highest point at which the
temperature was 1°F above the free-stream temperature, the rises given represent the
very top of the plume  and are  noticeably  higher than in  other experiments. The
Gordier formula was obtained from total-head traverses in a  water tunnel as reported
by Patrick."5 The formula attributed to Shandorov by Abramovich1 i'  was based on
experiments that included various angles of discharge and density ratios. The Patrick""
formulas were based  both on the height or maximum concentration  ot nitrous  oxide
tracer and on the height of maximum velocity as determined by a pilot tube.
                                                        THEORETICAL FORMULAS


   There are main  theoretical approaches to the problem of plume rise, and some of
them are quite  complex. To reproduce them  all here would be tedious and of little
help  to most readers.  Instead,  the  various theories are  compared with a relatively
simple  basic plume-rise  theon, based on assumptions common to most of the theories.

-------
26                                          FORMULAS FOR CALCULATING PLUME RISE

      It will be shown later that this basic theory in its simplest form gives good agreement
      with observations.
      Basic Theory

          In most plume-rise theories, buoyancy is assumed to be conserved; i.e., the motion
      is  considered to be  adiabatic.  This  means that the potential  temperature  of each
      element of gas remains constant. It is also assumed that pressure forces are small and
      have little net effect  on the  motion,  that they  merely redistribute  some  of the
      momentum within the plume. Molecular viscosity is also negligible because the plume
      Reynolds   number is very high, and  local density changes  are  neglected. These
      assumptions lead to three conservation equations:

                          ^  pp vp = 0        (continuity of mass)                (4.10)

                               d0p
                               	= 0        (buoyancy)                        (4.11)
                               dt

                               ^2 = ! 0' £    (momentum)                      (4.12)

      where   vp = the local velocity of the gas in the plume
              pp = the local gas density
              6p = the local potential temperature
       6'= 9p — 6 = the departure of the potential temperature from the temperature of the
                   environment at the same height
               fc = the unit vector in the vertical direction (buoyancy acts vertically)

          Equations 4.10, 4.11, and 4.12 can be transformed to describe the mean motion of
      a plume by integrating  them over some plane  that intersects the plume. It is most
      convenient to integrate over a horizontal plane because then the mean ambient values
      of potential temperature (0), density (p), and velocity (ve) can be considered constant
      over the  plane  of integration and  are assumed  to  be functions of height only.
      Furthermore, if ve is assumed to be horizontal, the  vertical component of vp,  denoted
      by w', is due entirely to the presence of the plume. Thus w' is a convenient variable
      with which to identify the plume.
          A further simplification results from assuming that the vertical velocity and the
      buoyancy  are everywhere proportional to each other in a  horizontal section of the
      plume since it is then unnecessary to assume any specific distribution of either. This
      assumption is approximately true for measured cross sections of vertical plumes.5 7
      Admittedly it does  not hold  near  the height of  final rise in  a  stable atmosphere,
      because buoyancy  decays more rapidly than vertical velocity  in such a situation.
          A steady state  is assumed. To obtain Eq. 4.13, we combine Eq. 4.10 times 6 'with
      Eq. 4.11  times pp and integrate  the resulting equation  over a horizontal plane,

-------
THEORETICAL FORMULAS                                                          27

assuming that the vertical velocity and the buoyance are everywhere proportional to
each other. Similarly, to obtain Eq. 4.14, we combine Eq. 4.10 times vp with Eq. 4.1 2
times pp  and integrate the resulting equation  over the  same  horizontal plane. The
plane of integration  must completely intersect the plume so that 9'=0 around the
perimeter of the  plane.  The  resulting  equations  for  the  net  buoyancy  flux and
momentum flux in a plume are
                                                                         (4.13)
where

                  v = /j_Pp_w dxdV        (vertical volume flux)           (4.15)
                           ?rp

                      g 30
                   s = = ;r—                (stability parameter)             (4.16)


                      //(g/T)<9'p w dxdy
                 Fz= - - -    - '(buoyancy flux)                (4.17)
                            •np
                                          (momentum flux)               (4.18)
    The vertical volume flux of the plume, as defined in Eq. 4.15, is the total vertical
mass  flux  divided by rrp, where  p  is the  environmental density. The stability
parameter,  s,  can  be  interpreted  as  the restoring  acceleration  per unit vertical
displacement for  adiabatic motion in  a  stratified  atmosphere  (either stable  or
unstable); in an unstable atmosphere, s is  negative; Fz is the vertical flux of the
buoyant  force  divided  by  -np; v is an average plume velocity  at a given height, as
defined by  the total velocity field at that height weighted by the normalized vertical
mass flux;  w is the vertical component of tf and is the velocity of plume rise at any
given height. The drag term in Eq. 4.14 is not written  out since it will be dropped
later, but it can be interpreted as the net horizontal  advection of momentum deficit
across the boundary of the plane of integration.
    The initial conditions are

                                   v=w0£                             (4.19a)


                                            =Fm£                      (4.19b)

-------
28                                          FORMULAS FOR CALCULATING PLUME RISE

      and


                                 FZ= (l-^)gWor^F                     (4.19c)


      For a hot source
      where cp is the specific heat of air at constant pressure.
          Equations 4.13 and 4.14 can be solved for the mean motion of a plume through
      any  atmosphere,  including one with  stability varying with height and wind shear.
      However, the equations cannot be solved until some specific assumption is made about
      the growth of volume  flux  with  the height  (dV/dz). This  assumption,  called an
      entrainment assumption, is  necessary to describe the bulk effect of turbulence in
      diffusing momentum and buoyancy in a plume.
      Basic Theory Simplified

          It is  desirable to reduce the basic theory to the simplest form that works. To be
      more specific, we would like to derive  from the basic theory simple formulas that
      agree with data. To do this, we must make the simplest workable entrainment and
      drag-force assumptions, assume simple approximations for the atmosphere, treat the
      stack as  a point source, and treat the plume as being either nearly vertical or nearly
      horizontal, i.e., ignore the complicated bending-over stage.
          When the  wind speed is sufficiently  low, a plume rises almost vertically, and the
      drag force and mechanically produced  atmospheric turbulence are  negligible. The
      turbulence that causes entrainment of ambient air is generated within the plume by the
      shear between the vertical plume motion and the almost stationary environment. The
      simplest  workable  entrainment  hypothesis  for  this  case  is that the  entrainment
      velocity, or the  average  rate at which outside air enters  the  plume surface,  is
      proportional to the  characteristic  vertical  velocity (w) at any  given  height. This
      assumption, based  on dimensional  analysis, will be called  the Taylor  entrainment
      assumption  after the author1 12  who suggested it in 1945. If (V/w)1'4 is defined as a
      characteristic  plume  radius, the  rate at which  the volume flux grows in a given
      increment of height is  then  2rr(V/w)^ aw, where a is called the entrainment constant

-------
THEORETICAL FORMULAS                                                           29

and is dimensionless. The complete set of equations governing the vertical plume are
then
                                    dz

which was given as Eq. 4.13,
                                                                          (4.21)

and

                                 — = 2a(wV)H                            (4.22)

This set of equations is equivalent to the relations given by Taylor112 in 1945 and
further developed in  1956 in a classic paper58 by Morton, Taylor, and Turner, who
found that a value of 0.093 for the entrainment constant gave the best  fit to observed
profiles of heated plumes. Briggs1' 3  found that a: = 0.075 gives the best predictions of
the height of the top of stratified plumes in stable air, based on the height at which the
buoyancy  flux decays  to  zero. The latter value is used here.  The direct empirical
determination of entrainment in jets by Ricou and Spalding114 yields a comparable
value of 0.080.
    The case of a bent-over plume, in which the vertical velocity  of the  plume is much
smaller than the horizontal velocity, is simpler. Both the total plume velocity and its
horizontal  component  are then very close to the ambient wind  speed, u, which is
assumed constant; wind shear  is neglected.  It is more  reasonable in this case  to
integrate Eqs. 4.10 to 4.12 over a vertical plane intersecting the plume  since a vertical
plane is more nearly perpendicular to the plume axis. When this is done, the resulting
equations are identical  to  Eqs. 4.13 and  4.21, provided that s  is constant over the
plane of integration,  that Fz, V,  and wV are defined  as fluxes of plume quantities
through a vertical plane, and that  the drag term is zero. Measurements  by Richards69
of the mean streamlines near horizontal  thermals  suggest  that the drag term is zero
provided the chosen plane of integration is large enough. This is also intuitively evident
since one would  not  expect  a vertically rising plume to leave a very extensive wake
underneath it.
    In the  initial stage of rise of a bent-over plume, the  self-induced  turbulence
dominates  the mixing process, and  the Taylor entrainment  hypothesis can  be used
again. The  main difference from a vertical plume is that in this case the velocity shear
is nearly perpendicular  to  the plume axis, rather than  parallel to  it. This apparently
results in  more  efficient  turbulent  mixing  since the entrainment constant for a
bent-over plume  is  about 5 times  as large as that  for  a vertical plume. With a
characteristic plume radius defined as (V/u)^ , the rate at which the volume flux grows

-------
30
                                            FORMULAS FOR CALCULATING PLUME RISE
      in  a given increment of axial distance is 27r(V/u)^  ?w, where 7 is the entrapment
      constant  for  a  bent-over plume.  If this is transformed to vertical  coordinates,  the
      plume rise is governed by Eqs. 4.13, 4.21, and
                                      dV  .  ,,n
                                      — = 27(uV)
(4.23)
      which  is comparable  to  Eq. 4.22. Since u  is a constant,  Eq. 4.23  can readily be
      integrated.  For a point source this  yields a characteristic radius  equal to yz. The
      relation is confirmed by modeling experiments of Richards69  and by photographs of
      full-scale plumes made by TVA99  (see Fig. 4.1). On the basis of these photographic
      plume diameters, 7 = 0.5.
             1400
             1200
             1000
           Q_
           UJ
           O
              800
              600
              400
              200
                  0      200     400     600      800     1000     1200    1400
                                         PLUME RISE (ft)

        Fig. 4.1  Photographic plume depth (top to bottom) vs. plume rise (center line) at TVA plants.

          Atmospheric turbulence is  small in  a stable environment and can be neglected, in
      which  case  Eq. 4.23 is valid  up to the  point  where a bent-over plume reaches its
      maximum rise. However, in a  neutral or unstable atmosphere, turbulence is vigorous
      enough to  eventually  dominate  the entrainment process. This occurs some  distance

-------
THLORET1CAL I-ORMULAS                                                          31

downwind  of the stack when the  vertical velocity  of the  plume  becomes small
compared with ambient  turbulent velocities. The simplest measure  of the effective
intensity of  atmospheric turbulence  is the eddy energy dissipation,  e, because  it
adequately  describes the part of the  turbulence spectrum that is most effective at
diffusing the  plume relative to its axis, i.e., the inertial subrange. The characteristic
radius of the  plume, ( V/u) * , determines the range of eddy sizes that most efficiently-
diffuse  the  plume.  If  these  two  terms  are adequate  enough  to characterize
entrainment,  the effective entrainment velocity must be given by /3e'^(V/u) '6 . where 0
is  a dimensionless constant; the exponents of the terms result  from dimensional
considerations. Since the  entrainment velocity in the initial stage of plume rise is 7W,
for the  simplest  model of a bent-over plume an abrupt transition  to an entrainment
velocity of (3e ^ (V/u) 1/6 is  assumed to occur when yw = (3e^(V/u)'f
    The solution for  the  bending-over stage of a plume in a crosswind is less certain
because both shear parallel  to the plume axis and shear perpendicular to the axis are
present. Both mechanisms operate at once to cause turbulent entrainment. Drag force
could contribute to the bending over of the plume since there could be an extensive
wake downwind  of the plume  in this case, but the drag force will have to be neglected
at present owing to insufficient knowledge. In  the early stage  of bending over, the
vertical-plume model is applicable except that there  is a perpendicular shear velocity
nearly  equal  to  u. If the two contributions to entrainment can be summed  in the
manner of vectors, the resultant entrainment velocity becomes (a2w2 + 72u2)1/4, and
the plume center line is given by Eqs. 4.13, 4.21, and


                            =2        («2w2+72u2)H                    (4.24)
    Before  applying  models of the vertical  plume and  bent-over plume  to specific
cases, some approximations about the source can be made. Usually it is reasonable to
assume that either the initial vertical momentum or the buoyancy dominates the rise.
in the former case the plume is called a jet, and we set F  equal to 0. Unheated plumes
composed mostly  of air are in this  category. Most hot plumes  are dominated  by
buoyancy, and we can neglect the initial vertical momentum flux, Fm. At a sufficient
distance from the stack, e.g., beyond 20 stack diameters downwind, we can  neglect the
finite size of the source  and treat the stack merely as a point source of momentum
flux or buoyancy flux.
    Some of the approximations that come out of the simplified theory are given in
Eqs. 4.25  to 4.34. Vertical  plumes are indicated by the term "calm"  and bent-over
plumes by "wind." For rise into stable air in which s is constant, we have

                    Ah = 5.0FV3(!           (buoyant, calm)               (4.25)

                    Ah =2A(L\*         (buoyant, wind)              (4.26)
                             \us/

-------
32                                           FORMULAS FOR CALCULATING PLUME RISE


                                                   (jet, calm)                    (4.27)


                          Ah=1.5          s       Get, wind)                    (4.28)
                            (— )
    In the calm case, Eq. 4.25 gives the height at which the buoyancy goes to zero. In
the windy cases for a bent-over plume, the equations are integrated to the point where
w = 0, and the plume is assumed to fall back to the level at which the buoyancy is zero
with no further mixing. More  details are given by Briggs.1 1 3 The plume will penetrate
a ground-based inversion or stable layer if the preceding formulas  predict a rise higher
than the top  of  the  stable air. If the  air is neutrally stratified above this level, a
buoyant  plume will continue  to rise  since it still has some buoyancy. A jet will fall
back and level off near the top of the stable air because it acquires negative buoyancy
as it rises.
    The model predicts penetration  of a sharp, elevated inversion of height Zj through
which the temperature increases by ATj if
                                      F   6
                     Zi<7.3F°'4 b"j°-6     (buoyant, calm)                 (4.29)


                                           (buoyant, wind)                 (4.30)
                     Zi<1.6t (FjnY    Get, calm)                      (4.31)
                              Ui /
where b; = g AT,/T.  The buoyant plume is assumed to penetrate if its characteristic
temperature excess, given by (T/g)Fz/V, exceeds AT; at the height of the inversion.
    For the first stage of rise, the bent-over model predicts plume center lines given by

                     Ah= l.SF^if'x-54    (buoy ant, wind)                (4.32)

                     Ah = 2.3Fmu~ V*    (jet, wind)                      (4.33)

    For the general case where s  is positive and constant, Eqs. 4.13 and 4.21 can be
combined with the transform dz = (w/u) dx to give
This is the equation of a simple  harmonic oscillator. Since V always increases, the
plume center line behaves like a damped harmonic oscillator (the author has observed
   tEmpirical; numerical value difficult to determine from present model.

-------
THEORETICAL FORMULAS                                                          33

such behavior at a plant west of Toronto in the early morning). Since V ~ U72z2 , the
preceding expression can be integrated and satisfies the initial conditions when

            [(72/3)us^j Ah3 = Fm sin (xs'^/u) + Fs~^ [1 — cos (xs'^/u)]

This equation is valid only up to the point of maximum rise because beyond this point
a negative entrainment velocity would be implied. According to this equation a jet
(F = 0) reaches  its maximum height at x = (n/2) us" ^ and a buoyant plume (Fm = 0)
reaches its maximum height at x = TTUS"^. At much smaller distances the plume center
line is approximated by

                                               Fx   \*
                                                    )"
From this equation it is seen that the ratio Fx/Fmu is a general criterion of whether a
bent-over  plume is dominated  by buoyancy or by momentum at a given distance
downwind. It, in fact, represents the ratio of buoyancy-induced vertical momentum to
initial vertical momentum.
    For the buoyant bent-over  plume in neutral conditions, the first stage of rise is
given by Eq. 4.32 up to the distance at which atmospheric turbulence dominates the
entrainment. The complete plume center line is given by Eq. 4.32 when x < x* and by
            Ah=1.8FV'x*«   |+^4+ i(-*)     1+1-^        (4.34)
                               LJ   ^J x    j  \x /  J \   j  x /

when x >x*, where x*  is the distance at which atmospheric turbulence begins to
dominate entrainment. This distance is given by

                            x* = 0.43F!503~3ue~1)3/5

Results from puff and cluster diffusion data and from measurements of eddy energy
dissipation  rates,  given in Appendix A, show  that j3 = 1 is acceptable as a somewhat
conservative approximation. In  the  surface  layer of the  atmosphere defined by
constant stress, e.g., the lowest 50 ft or so, it is well established11 s  that e = u*3/0.4z,
where  z  is  the  height  above the ground and  u* is  the  friction  velocity. If we
approximate z by z= Ah, the final plume rise  given by Eq. 4.34 is  Ah = 4.5 F/uu*2 ;
since u cc u* and  changes only gradually with  height in the neutral surface layer, this
result is  similar   to those  of  earlier  theories36'46'107  that predict  Ah  °c F/u3
Unfortunately, this clear  relation between e  and u* breaks down at  heights more
typical of smoke plumes.  In Appendix A, data from 50 to 4000 ft above the ground
give more support to the empirical relation
                                        sec

-------
34                                          FORMULAS FOR CALCULATING PLUME RISE

      up  to z  ^ 1000  ft, then  becoming  constant  with height.  If  we  conservatively
      approximate z with the stack height, the resulting estimate for x* becomes
                         x* = 0.52   p   F'*h*    (hs<1000ft)

                         x* = 33 f-^1 F'«         (hs > 1 000 ft)              (4.35)
      Other Theories

          There is such a variety of plume-rise theories in the literature that only the briefest
      discussion of each must suffice. One can only be amazed, and perhaps perplexed, at
      the number of different approaches to the solution of this fascinating fluid-dynamics
      problem. The theories will be discussed chronologically, first  for the calm case and
      then for the crosswind case.
          The first theoretical treatment was of a jet in calm surroundings and was given by
      Tollmien116 in  1926. Rather than making an entrainment assumption, he used the
      Prandtl mixing-length hypothesis to derive a specific velocity-profile law  that agrees
      quite well  with  data. A similar approach was used for heated  plumes in calm air by
      Schimdt55 in  1941.  Rouse,  Yih,  and Humphreys57 treated  the  same problem by
      assuming  eddy  viscosity diffusion of the buoyancy and momentum by a process
      analogous to molecular diffusion. They determined experimentally that the  mean
      temperature and velocity  profiles are approximately  Gaussian with the characteristic
      plume radius growing linearly with height. Yih5 6 also considered the case of a laminar
      plume, which does not apply to full-scale plumes.
          Batchelor48  considered the same problem in 1954  by dimensional analysis. He
      included the case of a stratified environment and found power-law expressions for the
      mean plume velocity and temperature as functions of height in an unstable atmosphere
      whose potential  temperature gradient is  also approximated by a power law. The first
      theoretical model  for a vertical plume rising through any type of stratification was
      given by Priestley and Ball117 in 1955. Their equations are similar to the preceding
      equations for the vertical plume except that the  entrainment assumption, Eq. 4.22, is
      replaced by an  energy equation involving  an assumption about the magnitude and
      distribution of the turbulent stress. Vehrencamp, Ambrosio, and Romie100 were the
      first to apply the results from an entrainment model to final rise in  stable air by using
      the Taylor entrainment assumption. A general  model involving this assumption and
      complete with   experimental  verification  was  put  forth by  Morton, Taylor, and
      Turner58 in 1956. This model is  called  the M,T,&T  model  in the discussion that
      follows. The M,T,&T model is virtually identical to the vertical-plume model presented
      in the section "Basic Theory Simplified" of this chapter and differs  from the Priestley
      and Ball117 model  mainly by predicting a wider half-cone angle for jets than for
      buoyant plumes. This is actually observed in the laboratory. Both the M,T,&T model
      and the Priestley and Ball model predict a linear increase of radius  with height  in the

-------
THEORETICAL FORMULAS                                                          35

unstratified  case  and  give similar  results for the final  plume height but disagree
somewhat on  the values of the numerical constants. Estoque118  further compares
these two theories.
    Morton11 9  extended the numerical integrations of the M,T,&T  model to the case
of  a buoyant  plume with  nonnegligible initial  momentum and concluded  that
increasing the  efflux velocity can actually lessen rise in stable  conditions because of
increased entrainment  near the  stack  level. In another paper,120  he extended the
theory to include  augmented buoyancy due to the condensation of moisture of the
entrained air.   Hino121'122  made  further calculations  with the  M,T,&T  model,
including the effects of a finite source radius. Turner123  coupled the M,T,&T model
with a vortex  ring model to predict the speed of rise for a starting plume in neutral
surroundings. Okubo124  expanded  the M,T,&T model to the  case  of a plume rising
through a salinity gradient in water.
    A generalized theory for steady-state convective flow incorporating several of these
solutions was given by Vasil'chenko.1 2S Recently Telford126  proposed another type
of entrainment assumption in which the entrainment velocity  is proportional to the
magnitude  of  turbulent fluctuations in the plume as  calculated  from  a turbulent
kinetic energy  equation. Telford's results are similar to those of the M,T,&T model for
a buoyant plume, except near the stack, but his model predicts  too-rapid growth for a
jet. This happens because the model is, in effect, based  on the assumption that the
scale of the energy-containing  turbulent eddies is proportional to the  plume radius, but
this is not true for a jet, because most  of the turbulent energy is generated while the
jet  radius is relatively  small.  Morton127  has  further  criticized Telford's model in a
recent note.
    Lee128 developed  a model  for  a turbulent swirling plume. He used the  Prandtl
mixing-length  hypothesis.  Still  another problem was  explored  by  Fan,7 ]  who
extended the M,T,&T theory  to the  case of nonvertical emissions and tested the result
in a modeling tank with linear  density stratification.
    One of the  earliest theories for a bent-over buoyant plume was given by Bryant66
in 1949. A drag-force assumption was included, and the entrainment assumption was
in the form of a fairly complicated hypothesis about how the plume  radius grows with
distance  from  the source  along  its  center line. Eventually  the radius in  this model
becomes proportional  to  x^ ,  which  is too  small  a  growth rate compared with
subsequent observations.
    In 1950 Bosanquct, Carey, and Halton1 9 published a  well-known theory that was
later revised by Bosanquet.20 The  entrainment assumptions  were similar to  those
made in  the simplified theory here  except that the  same entrainment constant was
applied to both the vertical and the bent-over stages of plume development, i.e., 7 = a.
In  addition,  a  contribution  to  the  entrainment  velocity due  to  environmental
turbulence  was assumed that  was proportional to the wind speed. This assumption
eventually led to a linear growth of plume radius with  distance downwind and resulted
in a final height for a bent-over jet and rise proportional to log x for  a buoyant plume.
The theory tends to underestimate rise at large distances downwind (see Fig. 5.3 in the
next chapter).

-------
36                                          FORMULAS FOR CALCULATING PLUME RISE

          About the same time, Sutton129  developed a simple theory for a buoyant plume
      in a. crosswind which was based on Schmidt's55 result for a vertical plume, i.e., w oc
      (F/z)^ . Sutton replaced z in this relation with the distance along the plume center line
      and  took the horizontal speed of the  plume  to  be  equal to u. The expression is
      dimensionally correct and, at  large distances, approaches the form given by Eq. 4.32.
          Priestley73 adapted his  and Ball's vertical-plume model to the bent-over case. The
      average radius of a horizontal section was assumed to  grow linearly with height, and
      the entrainment constant  was modified by a factor proportional to u^ . Thus the
      equations of rise were identical to  those  for  a vertical  plume except  for the
      entrainment  constant modification. Priestley coupled this  first-phase  theory with a
      second phase in which atmospheric turbulence dominates the mixing. This latter phase
      is  complicated and yields some unrealistic results, as was mentioned by Csanady.8 7
      The first phase leads to an asymptotic  formula identical to  Eq. 4.32 times  a factor
      proportional to (F/x)~'/lz ; namely,


                              Ah = 2.7 [(jL)M] FV'x3/<                    (4.36)


      Lucas, Moore, and Spurr91  were able to simplify Priestley's theory considerably. For
      the first stage of  rise,  they  obtained a plume rise 15% greater than that given  by
      Eq. 4.36, and, for the atmospheric-turbulence-dominated stage, they obtained
      where xt is the distance of transition to the second stage. It was estimated that x1 =
      660 ft, in contrast to the transition distance x* given by Eq. 4.35, which depends on
      both the source strength and the height in the atmosphere.
         Scorer36'1 30  introduced a simple plume-rise model for which he assumed that the
      plume  radius  grows linearly with  height (see  Fig. 4.1). The constant governing the
      growth rate depends on whether the plume is nearly vertical or bent over and also on
      whether it is  dominated by momentum or by buoyancy in a given stage. Scorer
      considered all  the separate possibilities and then matched them at the bend-over point
      to get  a complete set of formulas for rise in neutral conditions. The predictions for
      transitional rise, the plume  center line before final height is reached,  are similar to
      those given by Eqs. 4.32 and 4.33. In addition, he postulated that the active rise
      terminates  when the vertical velocity of the plume reduces to the level of atmospheric
      turbulence velocities, which he took to be some fraction of the wind speed. This led to
      the prediction that Ah  cc p/u3 for a very buoyant plume. This  type of formula has
      been given by many authors, but the leveling off of the plume in neutral conditions
      has not yet actually been observed. Furthermore, it now appears that atmospheric
      turbulence velocities are less  strongly  related  to wind speed at typical plume
      heights.131

-------
I'M! ORMK'AL 1 OKMUL AS                                                            37

    A groat variety  of work has been done in the last 0 \eais. Lilly70 constiueted a
lumiencal model ol  the two-dimensional vortex pair seen in a vertical cross section of a
bent-over  buoyant plume. KetTei and Barnes"'1 presented a model for the bonding-over
stage ot a jot with an entrammont assumption similar to the one in  the bonding-over
model  given  in this review except  that  only  the  hori/ontal shear was  included.
Oanovich  and 7e\'ger7h developed a theory along the lines of the Priestle\  theory tor
the first  stage  but  with the second-stage  dynamics determined by  the  diffusion of
buoyancy by atmosphenc  turbulence. The typo of diffusion assumed was  essentially
the same  as  that observed for  total diffusion of gases in a passive plume.  However,
total diffusion  includes the meandering of the  plume axis  caused by shifts in wind
direction, whereas  the  action  of  buoyancy  on  the  plume  is affected  only  by  the
diffusion  ot  buoyancy relative  to  the plume axis. Only  relative diffusion should be
used. The same criticism applies to a theory developed b\  Schmidt,1'': which is based
on  the assumption that  the spread of material equals that given by the total diffusion
ot a passive  plume.  There is also  the criticism  that  the  diffusion of a rising plume,
especially in  its early stages, is  not the same as for a passive plume, because the rising
plume  generates  its own tuibulonco in addition to the  ambient turbulence. These
problems were also pointed out  by  Moore.1
    Fquations 4.25  through 4.28 and Fqs. 4.32 and 4.33 were proposed by Briggs107
on the  basis of rather elemental}, dimensional analysis as an extension of Batcheloi  s4 ^
and  Scorer's4'1   approaches.   Briggs113  recentK  considered  in  some  detail  the
penetiation  of inversions by  plumes  of all types by  using  a model based  on  the
simplified theory given here. Clifford1'14 extended this type of model to  the case of a
bent-over plume whoso total buoyancy  flux increases linearly \\ith  time as it mo\es
away  trom  the source,  again  using  the l'a> lor entrainment  assumption.  Modeling
experiments  of  Turner1 ''^  with  thermals  of  increasing  buov anoy  support   this
assumption.
    A  model  by Csanady1 3p  for the bent-ovei buoyant plume included the effect ot
eddy-oneigy dissipation  and of  ineitial subrange  turbulence in  the relative diffusion of
plume  buoyancy. In a later paper by Slawson and Csanadv / a three-stage  model  was
proposed. In  the first stage, self-generated turbulence dominates, and the governing
equations are in fact the same as those given in  the bent-over plume model here. The
second stage  is dominated by  ineitial subrange  atmospheric turbulence, and, in the
third stage, the  plume is supposed to be large enough for the eddy  diffusivity to be
essentially constant, as is the case  for moleculai  diffusion. This model \ields a radius
pioportional  to  x1^  and a  constant  rate  of rise in  the  final  stage  rathei  than  an\
limiting height of rise.
    Very   recentK  a model  along  the lines of the basic  theor\  presented here  was
developed by Hoult, Fay, and Forney,1'1'  in which entrainment velocity depends on
the longitudinal  and transverse  shear velocities. This theor\ is more elaborate than the
simplified theorv  presented hoio. in that )  ma\  be a function of vv0 u and the  Fronde
numbei at the  stack but does  not  take  into  account  the effect  of  atmospheric
turbulence.

-------
5
COMPAR/SONS
OF  CALCULATED
AND  OBSERVED
PLUME  BEHAVIOR
NEUTRAL CONDITIONS

      Buoyant Plumes in Neutral Conditions

      Some previous comparisons of plume-rise formulas with data  for the  case of hot
      bent-over plumes in  near-neutral conditions were  reviewed by Moses,  Strom, and
      Carson13 8 and are only summarized here. Moses and Strom1 3 9 compared a number of
      formulas with data from their experimental stack. However,  there was much scatter in
      the data, and only the absolute differences between observed and calculated values
      were used in the analysis, rather than their ratios. The results of the comparisons were
      rather inconclusive. Rauch80  made a brief comparison of the Holland6 formula and
      that of Lucas, Moore, and Spurr9 1 with his own data and  found the latter formula,
      multiplied by a factor of 0.35, to  be a better fit.  Stu'mke104 made more extensive
      comparisons between 8  different  formulas and  the data of Bosanquet, Carey, and
      Halton,19 of Stewart, Gale, and Crooks,75 and of Rauch.80 By  computing the ratios
      of  calculated to observed  rises,   Stu'mke  concluded  that the  Holland formula,
      multiplied by a factor of 2.92, works best.
         Since these comparisons were made, a number of new formulas have appeared,
      including those by Stu'mke,108  Moses and Carson,110 CONCAWE,25'26 the modified
      Lucas formula,109 and Eq. 4.34, published for the first  time here. In addition, more
      data are now available, especially the data from three Central Electricity plants and six
      TVA plants;  therefore comparisons can now be mad-  over a much wider range of
      conditions.

                                        38

-------
NEUTRAL CONDITIONS                                                              39

    First,  a  simple wind-speed relation would be convenient since this would allow
some reduction  of a large amount of data that covers a wide range  of wind speeds,
source  strengths,  and  measuring  distances.  Many formulas, both empirical  and
theoretical,  suggest that plume rise is inversely proportional to wind speed, at least at a
fixed point downwind. In  Fig. 5.1, data from a large number of  sources tend to
confirm this. In each graph the plume rise at one or more fixed distances is plotted
against wind speed on logarithmic  coordinates so that Ah <* u~! is represented by a
straight line with a slope of — 1; such lines are indicated for reference.
    For most of the sources, Ah ^ u~: is the  best elementary relation. It would be
difficult to  make a case for Ah  °= u^, as appears in the CONCAWE25 '26  formula. A
better fit would result only for the Duisburg data, upon which the CONCAWE formula
is very largely based. A few of the sources, in  particular Shawnee and Widows Creek,
show a greater decrease of Ah with increasing u, which probably indicates some form
of downwash at higher wind speeds.  However, the Davidson—Bryant1 °5 prediction
that Ah is proportional to  u"1 A would not fit most of the data.
    With the inverse wind-speed law reasonably well established for neutral conditions,
we can now average the product of plume rise and wind speed for all wind speeds to
greatly  reduce  the volume of  data.  Such a  presentation was first employed by
Holland.6  In Fig. 5.2, u Ah is plotted as a function  of x for all available data sources.
The average heat  efflux  per stack, in  units of  106 cal/sec, is given in parentheses
following  each identification code, along with  the number of stacks if more than one.
The key  to the code is  given  in  Table 5.1.  In  system A at Harwell  (HA), wind
measurements were at a height of 27 m, whereas in  system B (HB) the measurements
were at 152m, which  is much  closer to the height of the plume. A considerable
amount of data is presented in Fig. 5.2. A general criterion was that each point plotted
should represent at least three periods of 30 to 120 min duration each and that each
period should be represented by at least five samples of plume rise or  some equivalent
amount of data.
    The outstanding feature of Fig. 5.2 is mat all  the plume center lines continued to
rise  as far as measurements were made; there is no evidence of leveling off. In general,
the  plume center  lines approximate a 2/3 slope, as predicted by  the "2/3  law" in
Eq. 4.32.  This means  that the final rise  has not definitely been measured in neutral
conditions,  and  therefore  we will have  to find some other way of defining effective
stack height.
    The same data as in Fig. 5.2, along with the data of Ball,76  are plotted in  Fig.  5.3.
Both the  rise and  the distance downwind are made  nondimensional by means of the
length L = F/u3  The result is a somewhat entangled family of curves  that lie  between
1.0 and 3.0 times F^u^x'4. Rise for  a buoyant plume  according to the Bosanquet
theory20  and  the asymptotic plume  rises according to Csanady87  in 1961 and to
Briggs1 °7  in 1965 are shown. They all underestimate rise at large values of x/L.
    The  Bosanquet20  formulation  underestimates plume rise  when  x/L>103   The
CONCAWE2 5 '2 6 relation  mat Ah is proportional to u~'4 and the Davidson-Bryant1 °5
relation that  Ah is proportional tou'1'" are not valid for most data sources. Formulas of
the type Ah oc L = F/u3 are difficult to test because they apply only to final rise in

-------
40
                                         CALCULATED AND OBSERVED PLUME BEHAVIOR
35
30
£
20
UJ
CO
a:
LJ
5
5 to
0-
8
6
QH = 1.0 X to4 cal/sec
• x = 30 ft A x = 60 ft

\
•





IN
»
•^





AA
1%

A
'S
N
— • —








I <
It



A
A
L
>4
»-
><



I
t
J-

>_ 5
BALL



i
J
A
H


AA
>t—
F
?
r~ ' "
10 1
1000
800
600
400
200
160
6 2
QH = 1.2 X 107 col/sec
• x = 1000 ft
A x = 3000 ft

^A
^


•
\
*




\
N














\

s


\
0 40 6
LAKEVIEW
200
100
80
60
40
0 1
QH = 1.2 X 106 to
2.9 X 106 cal/sec
• x = 330 ft
A x = 820 ft

_^ 	
* 1

•
(

_\A
^
\

i





2 20 3
DUISBURG
                                       WIND SPEED (ft/sec)
300
UJ
ir
^ 100
3 80
a.
60
50
1
QH = 1.1 X 106 cal/sec
• x = 1260 to 2t50 ft
s
• \
•






.
*\














1






2 20 30 40
HARWELL
200
100
80

30
1
OH = 1.6 X 107 cal/sec
• x = 200 ft
A x = 600 ft
N
^p
\
\»




A\
\

	 A^
•
\
\



k^


^
1200
1000
800
400
200
• WIDOWS CREEK,
QH = 1.7 X 107 cal/sec
A JOHNSONVILLE, QH = 1.1 X 107
cal/sec X 2 STACKS
N

.






•
^
s
-9







~



10
A


^S



0(




\ A
\w
\^
^\
3ft 4






i

\
N
(A
'• —
2 20 30 40 68 10 20
BOSANQUET WIDOWS CREEK AND

                                                                  JOHNSONVILLE
                                      WIND SPEED (ft/sec)
                     Fig. 5.1  Plume rise vs. wind speed in near-neutral conditions.

-------
NEUTRAL CONDITIONS
                                                                                        41
       900
       700

       500
    ^  300
    ce
       ZOO
            • SHAWNEE, OH = 5.5 X 106
              col/sec X 8 STACKS
            A COLBERT, QH = 6.7 X 10s
              col/sec X 3 STACKS
        100
        90

1 '
-X
f. -
•
>_
\
0

A
A ^
\
\
\
•
00 ft 5






•S
3ft 67






fl
           7 8 9 10        20    30
              SHAWNEE AND COLBERT
  0H = 1.7 X (O7 cal/sec TOTAL
  O x = 1000 ft, 1 STACK
  • i = 1000 ft, 2 STACKS
  A x = 2500 ft, 1 STACK
800
600
400
200
100
£












0




A


\ A

°\ ^
^\ o o
" (J




\
•)
\







^
o
0
]\
3 10 20 30 40
GALLATIN
                                  WIND SPEED (ft/sec)
                                                   • EARLEY, OH = 1.0 X 106 lo 5.1 X 106
                                                     col/sec X 2 STACKS, x = 3600 to
                                                     6000 ft
1200
1000
800
LU
in
cc 400
UJ
5
0_
200
i?n
0H = 2.1 X I07
O x = 1000 ft,
• x = 1000 ft,
Ax = 3000 ft,
Ax = 3000 ft,
V x = 5000 ft







A \ V
4^
Nj*
s
"\ • A .


i^


col/sec per stock
STACK
2 STACKS
1 STACK
2 STACKS
V
\
A V\(
A A
V A A
^A
,
^*d
^
0



\
^\

NA^
K

• o
0





\
- A-
\
N
O






A

O
2000
1000
800
600
400
POO
T CASTLE DONINGTON, 0H = 0.8 >
107 to 1.6 X 107 cal/sec X 2
STACKS, x =3600 to 6000 ft
A NORTHFLEET, 0H = 0.8 X 107 to
1.2 X 107 col/sec X 2 STACKS,
x = 4000 to 8000 ft
\
1









T \

1
\
V s
\
»



r\ _
V
\
\A
\
V






x

\
^\x








>s








A
             10
                       20   30  40  50 60
                       PARADISE
10         20    30  40  50 60
 EARLEY, CASTLE DONINGTON,
   AND  NORTHFLEET
                                  WIND SPEED (ft/sec)

-------
42
CALCULATED AND OBSERVED PLUME BEHAVIOR
        20,000 h
          500
             100
                       200
                                     500      1000       2000
                                      DISTANCE DOWNWIND (ft)
                              500O
                                        10,000
     Fig. 5.2  Plume rise times wind speed  vs.  downwind distance in  near-neutral conditions. The
     average heat efflux per stack, in units of 106 cal/sec, and the number of stacks, if more than one,
     are given in parentheses. See Table 5.1 for identification of sources and for additional data.
     neutral conditions, which has not yet been clearly observed. Therefore only relations
     of the type Ah oc if1 have been chosen for the comparison shown in Table 5.1. Data
     are given for the plume rises at the maximum distance downwind for which there was
     sufficient information to  meet  the  data  criterion set up for Fig. 5.2. The  ratio  of
     calculated to observed plume rises times wind speed was computed for each source and
     each  formula, and the results  were analyzed on a one-source  one-vote basis. The
     exceptions to this  rule were plants  that were run both with one stack and with two
     stacks emitting  (Paradise  and Gallatin) and  plants at which  there were substantial
     amounts of  data  for different rates of  heat emission  (Earley,  Castle Donington,
     Northfleet).  The median value  of the ratio was  also  computed  for each plume-rise
     formula, along with  the  average percentage  deviation from the  median. The same
     computation was repeated for a selected set of data that  excluded the following data
     sources: Ball, source very small; Harwell A, wind speed measured much below plume
     and obviously lower than  that measured with system B;  Bosanquet, no stack heights
     indicated and length of runs uncertain; Darmstadt, low efflux velocity and insufficient

-------
  1000
   500

-------
                                                         Table 5.1

                   COMPARISON OF CALCULATED VALUES WITH OBSERVATIONS FOR NEUTRAL CONDITIONS
Code
B
HA
HB
BO
DS
DB
T
L

E
E
CD
CD
N
N

S
c
J
we
G
G
P
P
Source
Ball*
Harwell A}
Harwell B
Bosanquet^:
Darmstadt:):
Duisburg
Tallawarra:):
LakeviewJ
CEGB plants
Earley
Earley
Castle Donington
Castle Donington
Northfleett
North fleet:f
TVA plants
ShawneeJ
Colbert^
Johnsonville
Widows Creek:f
Gallatin
Gallatin
Paradise
Paradise
Reference
76
75
75
73
80
80
87
88

91
91
91
91
93
93
99








Number
of
stacks

1
1
1
1
1
1
1

2
2
2
2
2
2

8
3
2
1
1
2
1
2
Us,
ft

200
200

246
410
288
493

250
250
425
425
492
492

250
300
400
500
500
500
600
600
D,
ft

11.3
11.3
6.5
7.5
11.5
20.5
19.5

12.0
12.0
23.0
23.0
19.7
19.7

14.0
16.5
14.0
20.8
25.0
25.0
26.0
26.0
w0,
ft/ sec

32.6
32.6
31.9
15.7
28.0
12.0
65.0

18.3
56.0
40.9
54.7
46.3
70.0

48.7
42.9
94.8
71.5
52.4
23.7
51.3
57.2
Range of u,
ft/ sec
2 to 14
14 to 30
17 to 38
14 to 33
16 to 25
15 to 29
20 to 23
25 to 49

14 to 35
14 to 35
10 to 26
10 to 35
13 to 52
13 to 52

8 to 29
10 to 17
6 to 22
8 to 21
7 to 34
5 to 39
6 to 55
12 to 34
Qn/stack,
10 cal/sec
0.0096
1.10
1.10
1.54
0.855
1.88
2.93
11.6

1.54
4.72
11.95
16.0
7.9
11.95

5.45
6.74
10.8
16.8
16.9
8.55
20.2
21.9
x* •(•
ft
14§
370
370
485H
380
705
680
1630

485
760
1510
1700
1400
1660

805
975
1400
1910
1920
1460
2300
2380
x,
ft
60
2950
1900
600
820
1150
1000
3250

4800
4800
4800
4800
5900
5900

2500
1000
2500
2500
3000
2000
4500
4500
u Ah,
ft- ft/ sec
112
4,430
3,980
2,450
2,150
3,400
5,500
22,100

5,580
8,150
14,800
18,600
10,900
11,150

6,210
7,200
10,100
8,000
14,250
7,850
21,200
20,000
•("Calculated from Eq. 4.35.
±Not included in selected data.
§ Height chosen for computing x* = 20 ft.
51 Height chosen for computing x* = 250 ft.

-------
Table 5.1 (Continued)

Code
B
HA
HB
BO
DS
DB
T
L

E
L
CD
CD
N
N

S
c
J
we
G
G
P
F



Source
Ball
Harwell A
Harwell B
Bosanquet
Darmstadt
Duisburg
Tallawarra
Lakeview
CEGB plants
Earley
Earley
Castle Donington
Castle Donington
Northflcet
North fleet
TVA plants
Shawnee
Colbert
Johnsonville
Widows Creek
Gallatin
Gallatin
Paradise
Paradise
Median for
Median for

Reference
76
75
75
73
80
80
87
88

91
91
91
91
93
93
99








all data
selected data

Moses and
,, 110
Carson
1.59
0.43
0.48
0.92
0.78
0.74
0.57
0.28

0.40
0.48
0.43
0.39
0.47
0.56

0.68
0.66
0.59
0.94
0.53
0.68
0.39
0.42
0.54 ± 34%
0.48 ± 19%

Stumke108

0.74
0.83
0.75
1.04
1.12
1.53
0.41

0.72
0.57
0.74
0.62
0.79
0.84

0.90
0.96
0.66
1.32
0.90
1.53
0.65
0.70
0.79 ± 11%
0.72 ± 247,
Ratio
Holland6
0.04
0.23
0.25
0.44
0.25
0.38
0.30
0.31

0.18
0.37
0.44
0.47
0.44
0.65

0.54
0.55
0.66
1.18
0.64
0.58
0.51
0.58
0.44 ±37%
0.47 ± 26%
af calculated to
n • a 73,87
Priestley
(first phase)
1.31
2.00
1.60
1.19
1.49
1.47
0.91
0.78

2.49
2.26
1.57
1.34
2.24
2.43

1.88
0.86
1.37
1.94
1.35
1.41
1.19
1.28
1.44 ±26%.
1 .4 1 ± 1 8%
Dbserved values (
Lucas, Moore,
and Spurr91
1.51
1.70
1.59
1.38
1.69
1.62
1.02
0.62

1.59
1.44
1.01
0.86
1.24
1.35

1.69
0.96
1.24
1.76
1.05
1.37
0.80
0.85
1.36 ±21%
1.24 ±22%
jf u Ah
Lucas109
0.78
1.27
1.19
1.12
1.36
1.60
0.87
0.68

1.30
1.16
1.01
0.86
1.35
1.47

1.36
0.82
1.21
1.92
1.15
1.50
0.96
1.03
1.18 ± 20'";
1.16 ± 14%

Eq. 4.32
("2/3 law")
0.86
1.40
1.17
0.98
1.13
1.17
0.76
0.66

.73
.71
.29
.13
.75
.96

1.53
0.77
1.19
1.73
1.10
1.21
1.03
1.11
1.17 ±23%
1 . 1 7 ± 1 2%.

Eq. 4.34
0.72
0.95
0.93
0.98
1.09
1.16
0.75
0.64

1.05
1.25
1.17
1.05
1.46
1.73

1.40
0.77
1.17
1.72
1.09
1.20
1.00
1.09
1.09 ± 19"',,
1.09 ±7%

-------
46                                      CALCULATED AND OBSERVED PLUME BEHAVIOR

      data; Tallawara and Lakeview, much higher rise than comparable sources in Fig. 5.2,
      possibly due to lakeshore effect; Widows Creek, down wash, possibly due to a 1000-ft
      plateau nearby, shown in Figs. 5.1  and 5.2; Northfleet, terrain downwash reported by
      Hamilton93 and rise much lower than at Castle Donington at same emission;Colbert
      and Shawnee, many stacks. The results in Table 5.1 help justify the exclusion of these
      data,  since  with  the  selected  data the average  deviation  from  the  median is
      considerably reduced for seven of the eight formulas.
         The  first three formulas tested in Table 5.1 are  completely empirical  and do not
      allow  for the effect of distance of measurement on plume  rise as the remaining five
      formulas do; consequently, these three formulas give poorer agreement with  data. The
      Holland6 formula  (Eq. 4.1) in  particular shows a  high  percentage  of scatter.  The
      formula of Stiimke108 (Eq. 4.4) is perhaps  slightly preferable to that of Moses and
      Carson110  (Eq. 4.8), although  the latter shows less scatter in comparison  with the
      selected  data.  All  three  of these  formulas underestimate  plume  rise,  but  this
      shortcoming can be  corrected  by  multiplying the  formulas  by  a constant  that
      optimizes the agreement.
         The  next three formulas are based  on  the Priestley73  theory.  The  first  is the
      asymptotic formula for the  first-phase theory87 (Eq. 4.36), which predicts  a  rise
      proportional to x3/(. Even  though this is a transitional-rise  formula, which  does not
      apply  to  a leveling off stage of plume rise, it shows less scatter  compared with
      observations than the three empirical final-rise formulas. The next formula (Eq. 4.37),
      by Lucas, Moore, and Spurr,91  includes both a transitional- and a final-rise  stage and
      gives a little better agreement with data. When Eq. 4.37 is multiplied by the  empirical
      stack-height factor suggested  by Lucas,109 i.e., 0.52 + 0.00116 hs,  the agreement is
      considerably better. However, one should be cautious about applying this formula to
      plants with heat emission less than  10 Mw, because it predicts continued plume  rise to
      almost 1 km  downwind regardless  of source  size.  For instance,  for the very small
      source used by Ball,76 the predicted  final rise is 12 times the rise measured at 60 ft
      downwind; it seems unlikely that such a weakly buoyant plume so close to the ground,
      where turbulence is stronger, will continue to rise over such a long distance.
         The  last two formulas are based  on the  simplified theory given in the section,
      "Basic Theory  Simplified" in Chapter 4. The "2/3  law"  (Eq. 4.32), another transi-
      tional-rise formula, agrees about as  well with these data as the Lucas109 formula just
      discussed. Equation 4.34, which includes both a transitional-rise  and a final-rise stage,
      gives both improved numerical agreement and much  less percentage of scatter. Clearly
      it is the  best of the eight formulas tested in Table 5.1 and is the one recommended for
      buoyant plumes in neutral conditions  (for optimized fit it should be divided  by  1.09).
         Eq. 4.34 should not be applied beyond x = 5x*, because so few data go beyond
      this distance. In some  cases the  maximum ground concentration occurs closer to the
      source than this, and Eq. 4.34 applied at the distance of the maximum gives the  best
      measure of effective stack height. (Beyond this distance plumes diffuse upward,  and
      the interaction  of diffusion with plume rise cannot be neglected.) One conservative
      approach is to set x = 10 hs, which is about the minimum distance downwind at  which
      maximum  ground  concentration  occurs.  For the   fossil-fuel plants  of the Central

-------
NEUTRAL CONDITIONS                                                              47

Electricity Generating Board (CEGB) and TVA in Table 5.1, at full load this distance
turns out to be in  the range 2.5 < (x/x*) < 3.3. At x/x* = 3.3, Eq. 4.34 gives a plume
rise only  10%  lower than Eq. 4.32, but at  twice  this distance the  plume  rise is
increased  by  only  27%.  This  suggests  a rule  of thumb that Eq.  4.34  can be
approximated by Eq. 4.32, the "2/3 law," up to a distance of 10 stack heights, beyond
which further plume rise is neglected, i.e.,

                       Ah= 1.8 FHu~'x%    (x<10hs)

                     Ah= 1.8 FV1(10hs)Ss  (x>!0hs)

For other sources a conservative approximation to Eq. 4.34 is  to use Eq. 4.32 up to a
distance of x =  3x* and then to consider the rise at this distance  to be the final rise.
Surprisingly,  Eq. 5.1  compares  even better  with  the data   in  Table 5.1 than  the
recommended  Eq.  4.34. Excluding  Ball's  data, which  were for a ground source, the
median  ratio of calculated  to observed plume rises is about  1.13,  and  the  average
deviations are ±17% for all data and ±4%  for the selected data. Because of the nature
of the approximation used in Eq.  5.1 and the scarcity of data beyond x = 5x*, Eq. 5.1
is recommended as an alternative to Eq.  4.34 only  for fossil-fuel plants  with a heat
emission of at least 20 Mw at full load.
    For multiple stacks the data show little or no enhancement of  prume rise over that
from comparable  single stacks in neutral conditions. Observations at  the Paradise
Steam Plant were  about  equally split  between one-stack operation and  two-stack
operation  with  about the  same heat emission  from the second stack. In  Fig. 5.1 the
plume rises  in  these  two conditions can be  seen  to be  virtually indistinguishable.
However, the same  figure  shows a clear  loss in plume rise at Gallatin for  the cases in
which the  same heat emission was split between  two stacks. In Table 5.1  average
plume rises for plants with two stacks are somewhat less than those for plants with one
stack, at least in comparison with Eq. 4.34. Colbert, with three stacks, seems to have
an enhanced rise,  but Shawnce, with eight or nine  stacks operating, has  a lower rise
than would be expected for a  single stack. This may be due to down wash, as noted in
the discussion  of Fig. 5.1. In  summary, die observations do not  clearly support any
additional allowance  for  plume  rise when more than one stack is operating.  It is
beneficial  to combine as  much of the effluent as possible into one stack to get the
maximum heat emission and the maximum thermal plume rise. This has been the trend
for large power  plants both in England and in the United States.
    Few data are available to  evaluate plume  rise in unstable  conditions. Slawson88
found a just slightly higher average rise  in unstable than in  neutral conditions, as well
as  more  scatter, as  might be expected owing to convective  turbulence. The same
general  features are  evident  in  the TVA data. The  buoyancy  flux of  the  plume
increases as it rises in unstable air, but there is also increased atmospheric turbulence:
it is not clear which influence has the greater effect on the plume. However, because of
lack  of empirical evidence,  it is possible only  to recommend for  unstable conditions
the same formulas  that apply in neutral conditions, specifically Eq. 5.2.

-------
48
                                          CALCULATED AND OBSERVED PLUME BEHAVIOR
      Jets in Neutral Conditions

          Most data for jets in a crosswind do not extend very far downwind; so in Fig. 5.4
      they are compared with the bending-over plume model in "Basic Theory Simplified,"
      Chapter 4; Ah/D is plotted as a function of R=w0/u for two different  distances
          60

          50
       a:  20
           10
       C£>
       a:
       UJ
       o
       o
                                     5            10            20
                          RATIO  OF EFFLUX VELOCITY TO CROSSWIND VELOCITY
                             50 60
        Fig. 5.4  Plume rise of jets in crosswind compared with values for bending-over plume model.
             B and C, Bryant and Cowdry
             C and R, Callaghan and
               Ruggeri62
             F.Fan71
             J, Jordinson
             K and B, Keffer and
                    63
                                     61
N and C, Norster and Chapman6
P-C, concentration profiles,
 Patrick65
P-S, Schlieren photographs,
 Patrick65
P-V, velocity profiles,
 Patrick65

-------
NEUTRAL CONDITIONS
                                                                                   49
downwind, x = 2D and x = 15D. The two families of curves group together rather well,
considering the variety of experiments and measurement techniques, which include the
photographic center lines by Bryant and Cowdry67 (B and C), the temperature survey
by Norster and Chapman65 (N and C), the velocity survey by Keffer and Baines63  (K
and  B),  the total pressure   measurements  by  Jordinson65   (J), the top  of  the
temperature profile measured by  Callaghan  and Ruggeri62 (C and R), the photo-
graphic  measurements by Fan71  (F), and the three  different  sets of measurements
made by Patrick,65  i.e., concentration profiles (P-C),  velocity profiles (P-V), and
Schlieren photographs (P-S).  The  data  are  fit rather  well  by the dashed line that
represents the formula given by the bending-over plume model (Eqs. 4.14 and 4.24);
the resultant formula is probably not of practical value since it applies  only near the
source and, being unwieldy, is not written out. This is just a test of the entrainment
assumption. Only the Callaghan and Ruggeri data do not fit the pattern. A number of
reasons are possible, one being that the jet velocities were near supersonic and another
being that  this jet was more  nearly horizontal, the distance downwind being about
twice the rise. The main reason this curve is higher is probably that it represents the
top of the jet rather than the center line.
    A comparison of values from Eq. 4.33 with the few sets of data that go  as far as
100 or 200 stack diameters downwind is  shown in Fig. 5.5. Equation 4.33 does fairly

         80
      
-------
50                                      CALCULATED AND OBSERVED PLUME BEHAVIOR

      well even when the plume is more vertical than horizontal (Ah > x) and works quite
      well when the plume is more horizontal. The exception is that it overestimates the rise
      measured by Fan at the lower value of R = w0/u, specifically at R = 4. This lends some
      credence to the suggestion made by Hoult, Fay, and Forney68 that the entrainment
      constant 7 may be a  function of R although the particular function that they suggest
      works poorly in the present model. It should be noted that Fan's plumes were partially
      buoyant, but these effects are minimized by rejecting  data for which  Fx/Fmu, the
      ratio of buoyancy-induced momentum flux to initial momentum  flux, is greater than
      0.5.
          As for the final rise of a jet, again it appears that none has been measured, but the
      asterisks in Fig. 5.5 at Ah/D = 3.0R (see  Table 4.1) indicate  a reasonable value for
      maximum observed rise; i.e.,

                                       Ah =3 ^2 D                             (5.2)


      This is  twice the value given by Eq. 4.9, the often-cited formula  of Rupp and his
STABLE CONDITIONS

      Penetration of Elevated Inversions

          A hot plume will penetrate an inversion and continue to rise if at that elevation the
      plume is warmer than the air above the inversion, i.e., if its temperature excess exceeds
      ATj. A jet, on the other hand, must have enough momentum to force its way through
      an inversion, and then it must eventually subside back  to the level of the inversion
      since  it is cooler than  the air above. For the case of no  wind, the simplified vertical
      model with boundary conditions implies that penetration  ability is a function of bj, zi;
      Fm , and F  Then conventional dimensional analysis predicts penetration when

                                 zibP-6F-°-4
-------
STABLE CONDITIONS
51
   g uj o
   if) -1-  'u_
    s oS ° -
    5 tt:   -°
    2 uj
    O >   "









0


0

7.




.2



, . r
6 h







?
14










b,








-n









R









c































O








A_




















o




















^-
.








^-









.9
u

i





^1
^**^ o
- / u *V3
m/b;)

0





>-^








^




4(
                                NONDIMENSIONAL MOMENTUM FLUX
                                            ,  0.8 ,--1.2,
Fig. S.b  Maximum nondimensional inversion height for penetration by plume vs. nondimensional
momentum flux (based on data from Vadot  ).
proportionality is roughly 1.6, as given in Eq. 4.31. As a simple, conservative criterion
for a vertical plume, Vadot's experiments suggest penetration when
                                           K-0.6
                                                                             (5.4)
    A bent-over buoyant plume rising through neutrally stratified air should penetrate
 an inversion at height Z; if, as expressed by Eq. 4.30,
 This equation  (Eq. 4.30) was  derived  from the simplified  bent-over plume model,
 which gives a characteristic temperature excess of the plume of
                                          g uz
                                                                             (5.5)
 for a plume rising through neutral air.  Eq. 5.5 is easier  to apply to cases where there
 are two or more inversions separated by neutral stratification.  Initially Fz = F, and 6'
 decreases with  the  inverse square  of  the height above the source  until the plume
 reaches the  first  inversion.  As  the plume  rises  through the inversion, its potential
 temperature  is unaffected, but the potential temperature of the ambient air increases
 by AT,; thus 6' is reduced by AT;. If 6' remains positive, the plume is buoyant and
 continues to  rise  with 6' proportional to z~2 until it reaches  the height of the next
 inversion. The  same  procedure  is  repeated  until the plume reaches an inversion  it
 cannot penetrate, i.e., until 6' <  AT,.

-------
52                                      CALCULATED AND OBSERVED PLUME BEHAVIOR

          The results  obtained  by applying this procedure to  the  data of Simon and
      Proudfit103 from the Ravenswood  plume in New York City, which include plume
      penetrations  of multiple  inversions,  are   shown  in Table 5.2, along  with the
      temperature excesses of the plume relative to the air above the inversion as calculated
      by subtracting AT; from Eq. 5.5 applied at the top of the inversion. It can be seen that
      every one of the eight nonpenetrations is predicted by a negative calculated 6'. In one
      case penetration is questionable  because  the plume center  line  ascended only 10m
      higher than the inversion;  so the  lower part of the plume was undoubtedly below the
      inversion.  Only one of the five penetrations was not predicted, and that was with a
      negative  6' of only  0.2°C, near  the limits  of  the  accuracy  of  temperature
      measurements. The procedure given in the discussion following Eq. 5.5 appears to be a
      good predictor but, perhaps, just slightly conservative.
      Rise Through Uniform Temperature Gradient

          Also of particular interest is the case in which the plume rises through air with a
      fairly uniform temperature gradient. In this case we can approximate s as a constant.
      For the calm case the simple vertical model predicts that the buoyancy of a hot plume
      decays to zero according to Eq. 4.25. This formula was derived by M,T,&T58  from
      virtually the same model, and a similar formula was derived by Priestley and Ball.11 7
      The  ability of Eq. 4.25 to predict the final height of the tops of plumes is shown in
      Fig. 5.7. Data are plotted from the modeling experiment in stratified salt solution by
      M,T,&T,58 from the modeling experiment in air near  the floor  of an ice rink of
      Crawford  and  Leonard,59  from  the experiments of Vehrencamp, Ambrosio, and
      Romie100 on the Mojave Desert, and from the observation by Davies1 01'102 of the
      plume from  a  large  oil fire. Equation 4.25 correctly approximates the top of the
      massive smoke plume that billowed out of the  Surtsey volcano in 1963.14°  The rate
      of thermal emission was estimated to be of the order of  100,000 Mw,141 or about a
      thousand times greater than the heat emission from a large stack. For the average lapse
      rate  observed in the  troposphere (6.5°C/km), Eq. 4.25 gives a rise of 5 km,  or about
      16,000 ft; the observed cloud top ranged from 3 to 8 km.
          As the nondimensional  momentum  flux  is  increased,  Morton's119 numerical
      solution indicates lessened plume rise, just as inversion penetration ability was seen to
      decline in Fig. 5.6. There are no data to show this, but three experiments with vertical
      plumes by Fan71 indicate gradual  enhancement  of rise  over that given  by  Eq. 4.25
      when Fm s^/F > 1.8.  Dimensional analysis of the vertical model indicates that

                                       Ah = CF*s-*                             (5.6)

      for a pure jet,  where C is a constant. The values of C that correctly describe  Fan's
      plumes, which were momentum dominated but not pure jets, are 4.53, 4.43, and 4.18.
      A value of C = 4 is suggested as an approximation, as in Eq. 4.27.

-------
                                                                                                                              H
Table 5.2
INVERSION PENETRATION AT THE RAVENSWOOD PLANTt
Date
May 25

July 20


July 21


September 8

September 9



Time
1825

0552-0559

0617-0820
0600-0724
0828

0648-0930
1000-1020
0640-0705
0747-0850

0930-1000
QH,
107 cal/sec
1.97

0.98

1.11
1.13
1.64

1.66
1.77
1.20
1.54

2.13
u,
m/sec
9.0

10.5

7.3
4.3
2.7

7.5
5.4
9.6
9.1

9.6
Plume
height,
m
295

350

360
360
510

410
560
350
370

390
Inversion height, m
Bottom
145
325
255
365
540
410
240
360
360
620
360
260
370
420
Top
180
475
275
395
580
450
280
410
400
650
400
300
410
530
°C
0.2
0.7
0.3
2.0
1.9
0.6
0.6
0.4
0.8
0.4
2.1
0.7
1.6
1.8
Calculated
F
6 ,
°c
15
-0.5
0.05
-2.0
-1.9
-0.45
1.7
0.0
-0.6
-0.3
-2.0
-0.2
-1.6
-1.7
m
n
o
o
H
O
z
Penetration
Yes
No
Yes
No
No
No
Yes
Yes
t
No
No
Yes
No
No
fStack height, 155 m.

-------
54
                                        CALCULATED AND OBSERVED PLUME BEHAVIOR
                  10,000
                   (000
                -  100
                     (0
                    O.I





/•




,rff^RA
/'*
MORTON, T
(

0

/
iVFORD AND
(ICE RIN
AYLOR.TUR
TANK)
AVIES (LON
/
A
/
/ VEHRE
/i AMBRC
(MOJAV
LEONARD
<)
NER
3 BEACHlT
NCAMP,
SIO, ROMIE
E DESERT)


y





                       O.I
                                          10        100
                                          5F'/4s-V8(ft)
                                                           1000
10,000
                        Fig. 5.7  Rise of buoyant plumes in calm, stable air.
         For  the case of a bent-over plume rising through stable air with constant s, the
      quasi-horizontal model can be applied both to a buoyant plume and to a jet to yield
      Eqs. 4.26 and 4.28, respectively. There are no data to test Eq. 4.28, but Eq. 4.26 and
      several other formulas can be compared with data from buoyant plumes released in
      stable air. These data include nine runs made at Brookhaven86 with 15-sec ignitions of
      rocket fuel, six runs by TV A99 with large  single stacks,  and seven runs by Van Vleck
      and Boone79  with 60-sec firings of horizontal rocket motors. Admittedly the plumes
      were not continuous in two of these experiments, and the plume rises were defined
      somewhat differently in each case. In each case the ratios  of the calculated to observed
      rises  were  computed.  The resulting  median values of this  ratio and mean deviation
      from the median are

                               Holland6        0.44 ±131%
                               Priestley73      0.42 ± 43%
                               Bosanquet20     1.22 ±26%
                               Briggs, Eq. 4.26  0.82 ±13%

      Holland6 suggested that Eq. 4.1 be reduced by 20% to predict rise in stable conditions,
      but this may  be  seen to  work poorly. The Priestley73  and Bosanquet20 theoretical
      formulations are both  complex; so  they were simplified to the  case  for a buoyant

-------
STABLU CONDITIONS

     5
 in
 or
 d
 2
 O
 a
 z
 o
                                                  55
                                                 	PLUME  TOP
                                                           PLUME  CENTER  LINE
                           23456
                      x/(us-'/2), NONDIMENSIONAL  DISTANCE DOWNWIND

l<"ig. 5.8  Rise of buoyant plumes in stable  air in crosswind at the TVA Paradise and Gallatin
plants.
point source. Clearly Eq. 4.26 gives the most consistent agreement, and on the average
it slightly underestimates rise. A constant of 2.4/0.82 = 2.9 works best, i.e.,
                                  Ah
        /F Y*
   = 291 — 1
        Vus/
(5.7)
    A further  test  of the simplified theory for bent-over plumes is shown in Fig. 5.8
for six periods of TVA data, which include the complete  trajectories of the plume
center lines and plume tops in stable air. The  center lines follow the "2/3 law" in the
first stage of rise with a fairly typical amount of scatter and reach a maximum in the
neighborhood  of x = 7rus~^  as is  predicted  by  theory. There  is less scatter in the
final-rise  stage, where four of the  six trajectories  almost coincide. The actual final
heights range from 450 to 1500 ft. The plume tops level out at
Ah = 4.0  —
                                                                            (5.8
   When  two  or  three  slacks were operating  at  the  TVA plants, there was  some
evidence of enhanced final rise in stable conditions. The maximum enhancement that

-------
56                                     CALCULATED AND OBSERVED PLUME BEHAVIOR

      could be expected according to Eq. 5.7 would be 26 and 44% for two and three stacks,
      respectively, if the total heat emission could simply be lumped together in computing
      F. The averaged observed enhancement relative to Eq. 5.7 was +20% with two stacks
      operating and +30% with three  stacks operating  except  that when  the wind was
      blowing along the  line  of three stacks  at Colbert  the  enhancement was +40%.
      Enhancement also depends on stack spacing since the plumes  can hardly be expected
      to interact with  each  other  if  they  are too far apart,  especially  if the  wind is
      perpendicular to the line of stacks. In the  preceding cases the  stacks were spaced less
      than 0.9(F/us)^, or about one-fourth of the plume rise apart.

-------
6
CONCLUS/ONS
AND  RECOMMENDATIONS
     There is no lack of plume -rise formulas in the literature, and selection is complicated
     by the fact that no one formula applies to all conditions. For a given situation many
     different predictions emerge, as  is shown  in Table 5.1. The  variety  of theoretical
     predictions follows  from the great variety of assumptions used in  the models; the
     disagreement  among empirical formulas is due to the different weighting of data used
     in their formulations and to variability among the data. Another factor is the frequent
     disregard of the dependence of plume rise on distance downwind of the stack. In the
     formulas recommended in the following paragraphs, all symbols are given in Appendix
     B, and the constants in the  formulas are optimized for the best fit to data covered by
     this survey. Readjustment of the  constants in previously cited equations is indicated
     by primes on the equation numbers.
        An important result of this study is that buoyant plumes are  found to follow the
     "2/3 law" for transitional rise for a considerable distance downwind when there is a
     wind, regardless of stratification; i.e.,

                                    Ah=1.6FV'x%                       (4.32')

     The bulk of plume-rise data are fit by this formula.
        In  neutral stratification Eq. 4.32' is  valid up to  the distance x/x* = 1, beyond
     which the plume center line is the  most accurately described by


                                                                          (434',
                                          57

-------
58                                              CONCLUSIONS AND RECOMMENDATIONS

      where

                                           F2/5h|!    (hs<1000ft)

                                                                                (4.35)

                                           F%      (hs>1000ft)
      Equation 4.35  is the best  approximation of x* at present for sources 50 ft or more
      above the ground; for ground sources an estimated plume height can be used in place
      of hs. Equation 4.34' applies to any distance such that x/x* > 1, but owing to lack of
      data at great distances  downwind x/x* = 5 is suggested as the maximum distance at
      which  it be applied at present. Even though Eq. 4.34' is the  best of the dozen or so
      formulas considered, the average plume  rise at a given plant may  deviate from the
      value given by Eq. 4.34'  by ±10% if the site is flat and uniform and by ±40% if a
      substantial terrain  step or  a  large body of water is nearby.  Furthermore, normal
      variations in the intensity of turbulence at plume heights at a typical site cause x* to
      vary by about ±20% on  the average,  with corresponding variations in  Ah. For
      fossil-fuel  plants  with  a   heat  emission  of 20 Mw  or more,  a  good  working
      approximation to Eq. 4.34' is given by

                              Ah= 1.6 F^u"1 x%      (x<10hs)                    ,
                           Ah= 1.6 FV1 (10hs)*>    (x>10hs)                 (    '

      For other sources, a conservative  approximation to  Eq. 4.34'is to use Eq. 4.32'up to
      a distance of x = 3x*, then  to consider the rise at this distance to be the final rise.
          Equations 4.34' and 5,1' are also  recommended for  the mean rise in unstable
      conditions although  larger fluctuations about  the mean should  be  expected (see
      Fig. 2.4).
          In  stable stratification  Eq. 4.32' holds approximately  to a distance x = 2.4us~^,
      beyond which the plume levels off at about
                                                                                 (5.7)
      as illustrated in Fig. 5.8. The top of the stratified plume is about 38% higher than that
      predicted by Eq. 5.7, which describes the plume center line. Although no significant
      increase  in  transitional rise is found when  more than  one  stack is operating, some
      enhancement of the final rise in stable  conditions is observed provided the stacks are
      close enough. If  the wind  is so light that the  plume rises vertically, the final rise is
      given accurately by
                                        Ah=5.0FV%                           (4.25)

-------
CONCLUSIONS AND RECOMMENDATIONS                                              59

In computing s for Eqs. 4.25 and 5.7, an average potential temperature gradient is
calculated for the stable layer or for the layer expected to be traversed by the plume.
    A buoyant plume  will penetrate a ground inversion if both Eq. 5.7 and Eq. 4.25
give a height higher than the top of the inversion. The plume will penetrate an elevated
inversion if the top of the inversion lies below both Eq. 5.4 and Eq. 4.30, i.e.,

                           zi<4F°-4b^°-6     (calm)                        (5.4)
    All the preceding formulas apply to buoyant plumes, which include most plumes
from industrial sources, and they are fairly well confirmed by observations. Because of
a relative lack of data, it is more difficult to make firm recommendations of formulas
for jets. It appears that in neutral, windy conditions the jet center line is given by
                                                                          (4.33)

at least up to the point that

                                  Ah=3^D                              (5.2)

as long as w0/u > 4. It can be  only  tentatively stated that in windless conditions the
jet rises to
                                  Ah = 4  HH
                                                                           (5.6)

where  4 is used as the value of C. This is on  the basis of only three experiments. If
there is some wind and the air is stable, the minimum expected theoretical rise is
                               Ah = 1.5 flll   s"%                        (4.28)
Unfortunately there are no published data for this case, and it would be presumptuous
to recommend any formula without testing it. However, since Eq. 4.28 is based on the
same  model, we  should not use Eq.  5.6  or  Eq. 5.2 if  it  gives  a higher rise than
Eq. 4.28  does. The most conservative of the three  formulas is the one that best
applies to a given situation. The same can be said of Eqs. 4.34',  5.7, and 4.25  for a
buoyant plume.
    Obviously  more experiments are needed  to complete  our basic understanding of
plume rise.  In particular they  are needed for jets at large  distances downwind for all

-------
60                                              CONCLUSIONS AND RECOMMENDATIONS

      stability conditions and for buoyant plumes at distances greater than ten stack heights
      downwind in neutral conditions. Once the fundamental results are complete, it will be
      worthwhile to study in detail the effect of the finite source diameter, the bending-over
      stage  of plume rise, the effect of wind shear and arbitrary temperature profiles, the
      interaction of plumes from  more than one stack,  and the interaction of plume-rise
      dynamics with diffusion processes.

-------
A
APPENDIX:
EFFECT  OF  A7WIOSPHER/C
TURBULENCE
ON   PLUME RISE
      As  discussed in "Basic Theory Simplified," in Chapter 4, entrainment of ambient air
      into the  plume by atmospheric  turbulence  is  due mostly to eddies in the inertial
      subrange; so, for a bent-over plume or a puffin a neutral atmosphere, the entrainment
      velocity, or velocity of growth, is given by

                                   dr/dt =j3el'rl>                          (A.I)

      where p1 is a dimensionless entrainment constant, e is the eddy energy dissipation rate,
      and r is a characteristic radius defined as (V/u)^ for a bent-over plume. To apply this
      entrainment assumption, some simple method of estimating e at plume heights is
      needed, and p1 must be determined.
         Ideally e would be related in some simple way to wind speed (u) and height above
      the ground (z). In  the neutral surface layer, e.g., the lowest 50 ft or so, such a relation
      is well described by the  expression1! s e  = u*3/0.4z, where u* is the friction velocity
      and is proportional to the wind speed at some fixed height. Unfortunately, at typical
      plume heights no such simple relation is found to exist. The turbulence becomes more
      intermittent and is affected more by departures from neutral  stability and  by terrain
      irregularities over a wide area. Still, enough  data exist to estimate mean values  ol e
      along with the amount of variability that should be expected.
         Recent  estimates of e were made by Hanna,14: who used  vertical-velocity spectra
      measured in a variety of experiments, and by Pasquill,'43 who used high-frequency
      standard deviations of wind inclination measured with a lightweight vane mounted on
      captive balloons at Cardington, England.  Hanna used data from towers at Round  Hill,

                                        61

-------
62                                                                       APPENDIX A

      Mass.,144  and Cedar  Hill, Tex.,145  from aircraft measurements made over a great
      variety of terrain by the Boeing Company,146 and from several low-level installations
      (below 50 ft). These values  of e are used in Table A.I to test the relation1^ oc um
      by computing the median value of e^> u"m and the average deviation from the median
      value for m = 0, V3, 2/3, and 1 at each height of each experiment. Because e is sensitive
      to atmospheric  stability, only  runs in which -1.0  < Ri < 0.15 were used from the
      Round Hill and Cedar Hill data, where Ri is the local Richardson number; the Boeing
      runs during very stable conditions and PasquilPs measurements above inversions were
      omitted. Also omitted were the few runs made during very low wind speeds, i.e., less
      than 2m/sec.
         Table A.I shows that the excellence of the fit is rather insensitive to increasing the
      value of m, especially  at Round Hill and Cedar Hill. The best overall fit is with m = V3;
      the average percentage deviation from the median is lowest with m = l/3 for four of the
      eight sets of data and, on the average, is only 9% greater than the  minimum value of
      percentage  deviation (indicated by  f in Table A.I). This is fortunate because the
      expression for x*,  the distance at which atmospheric turbulence begins to dominate
      entrainment,  turns  out  to  be independent  of  wind  speed  when  e^u^ (see
      Eq. 4.35  and the preceding discussion in Chap. 4).  It is therefore very desirable to
      adopt this approximation, keeping in mind the scatter about the median values shown
      in the table.
         It  is evident  in Table A.I  that e^/u^ decreases with height.  With a power law
      relation of e^/u^ <* z~", the  optimum value of n depends on which data are used. The
      best least-squares fit to log e* /u* =  constant-n  log z is n = 0.29 for all the  data but
      n = 0.37 if the Pasquill data at 4000 ft are omitted. At Round Hill n  = 0.31 between 50
      ft and 300 ft, and at Cedar Hill n = 0.39 between 150 ft and 450 ft, but in Pasquill's
      data n is  only 0.15 between  1000 ft and 4000 ft. These values are  roughly consistent
      with the  following  three published  conclusions: (1) Hanna142'147 confirmed the
      relation e* = 1.5 aw A^ for a wide variety of data, where ^w  is the variance of
      vertical  velocity and  Xm  is  the  wavelength of maximum specific  energy  in the
      vertical-velocity spectra; (2) data compiled in a note by Moore1 31  indicate almost no
      dependence  of CTW on height from about  100  to 4000 ft except for very high wind
      speeds (u >  10 m/sec); (3)  Busch  and Panofsky148 conclude that Xm oc z near the
      ground and reaches a maximum or  a  constant value somewhere above z = 200 m. The
      simplest expression consistent with all of the preceding evidence is e^/u^  « z~H up
      to a height of the order  of 1000 ft and then becomes constant with height. In the last
      column of Table A.I,  an expression of this type is compared with  the  data. The best
      estimate of energy dissipation appears to be

                        e* =0.9 [ft%/sec%]uH z~ *    (z< 1000 ft)
                                                                                (A.2)
                        e1* = 0.09  [ft* /sec%] u*      (z > 1000 ft)

         There remains the problem of how to determine the value of the dimensionless
      constant 0, particularly when no observations of plume, puff, or cluster growth include

-------
EFFECT OF ATMOSPHERIC TURBULENCE
63
                                    Table A.I

                  ENERGY DISSIPATION VS. WIND SPEED AND HEIGHT
Source
Round Hill
Round Hill
Cedar Hill
Round Hill
Cedar Hill
Boeing
Pasquill
Pasquill
Height,
ft
SO
150
150
300
450
750
1000
4000
Number
of runs
8
11
9
4
6
22
31
10
e*.
ftH/sec
0.636 ± 17%
0.495 ± 11%
0.457 ±20%
0.470 ±11%
0.331 ± 9%t
0.256 + 20%+
0.269 + 38%|
0.172 ±49%
elS/uii,
ft W /sec $4
0.266 ± 14%t
0.177 ± 10%
0.159 ± 18%
0.151 ± 7%t
0.104± 9%f
0.083 ± 24%
0.097 ±44%
0.079 ±4274
eS/u?s,
sec" ^
0.103 + 16%
0.063 ± 874
0.057 ± 16%t
0.049 ± 7%f
0.034 + 11%
0.028 ± 34%
0.042 ± 46%
0.030 + 47%'
ftt/u,
ft-H
0.042 + 18%
0.022 + 10%.
0.020 ±19%.
0.017 ±774
0.010 + 17%
0.009 ±53%,
0.01 8 ±53%
0.01 1 +59%
(fZ/u)H,
ft *s /sec*'
0.98
0.94
0.84
1.01
0.80
0.75
0.97
0.79]:
  "[Minimum value of percentage deviation.
  ±z = 1000 ft.
simultaneous, independent measurements of e. The approach used in this review is to
assume the validity of Eq. A.2 at the time and place of diffusion experiments and to
compare the  results with Eq. A.I.
    Frenkiel  and Katz149 used two motion-picture  cameras to photograph smoke
puffs released above  an island in  the Chesapeake Bay. The puffs were produced by
small detonations of gunpowder from an apparatus on the cable of a tethered balloon.
The radii of the puffs were calculated from their visible areas at 1-sec intervals. The
values of /3e^> shown  in Table A. 2 were calculated from the first 2 sec of puff growth
by using Eq. A.I as a  finite difference equation, i.e., by setting dr/dt = Ar/A t. Smith
and Hay150  published some  data  from several  experiments  on the expansion of
clusters  of  particles.  In their  short-range experiments,  Lycopodium spores were
released at  a  height  of 2 m and were  collected on  adhesive  cylinders lined up
perpendicular to the wind at 100 m downwind, yielding a lateral standard deviation of
particle  distribution (ay). In their medium-range experiments, fluorescent particles
were released from an airplane at heights of 1500 to 2500 ft several miles upwind of a
sampling apparatus mounted on  the cable of a captive  balloon, yielding a vertical
standard deviation of particle distribution (az). The values of /3ew shown in Table  A. 2
for the Smith and Hay experiments were  calculated from the integral of Eq. A.I for a
point source, namely,
 Interpreting the effective radius of a rising plume in terms of ay or az is difficult, but
 in this case it was assumed that CTZ = ay and that r = 2^ ay, as is true in the "top hat"
 model equivalent to a Gaussian plume in the Morton, Taylor, and Turner5 8 theory.
    The last column of Table A.2 shows the value of (3e^ inferred from the diffusion
 data divided by the value of e* calculated from Eq. A.2. The values of 0 inferred from
 this calculation range  from 0.62 to 0.82, a remarkably small range considering the

-------
64                                                                        APPENDIX A

                                          Table A.2
                      GROWTH RATE OF PUFFS AND PARTICLE CLUSTERS
Source
Smith and Hay
Runs 1-5
Runs 7-10
Frenkiel and Katz
z = 15 to 22 m
z= 39 to 61 m
Smith and Hay
(May 7, 1959)
Number
of runs

5
4

6
7

4
z, ft

14 = 0y
13 = Oy

58
164

2500
u, ft/sec

18
30

19
52

16
(3e^, ft % /sec

0.60 ± 7%
0.96 ± 18%

0.40 ± 7%
0.48 ± 23%

0.17± 17%
calculated

0.62
0.82

0.64
0.78

0.74
      indirectness of this approach and the wide range of variables involved. Note that the
      short-range experiments  of Smith and Hay  were probably  carried out within  the
      surface layer, where Eq. A.2 is not actually valid; nevertheless, the error in estimating e
      is not large for moderate wind speeds at these heights. Table A.2 suggests that |3 ~ 0.7,
      but, considering the small  number of data and the  indirectness of this analysis, the
      more conservative value of/3 = 1.0 is recommended.
          It should  be  cautioned that the characteristic plume radius, r, that  appears in
      Eq. A.I is not necessarily the same as the visible radius or other measures of size of a
      passive puff or plume,  and  so the evaluation of /3 made in Appendix A is not directly
      applicable to diffusion problems other than plume rise.

-------
B
APPENDIX:
NOMENCLATURE
     Dimensions of each term are given in brackets: / = length, t = time, T = temperature,
     m = mass.

         bj   Inversion parameter = g AT;/T [l/t2 ]
        CD   Drag coefficient [dimensionless]
         D   Internal stack diameter [/]
         F   Buoyancy flux parameter [/4/t3] ; see Eqs. 4.19c and 4.20
        Fm   Momentum flux parameter [I4 /t2 ] ; see Eq. 4.19b
        Fz   Vertical flux of buoyant force  in plume divided by Tip [I4/t3]; see Eq. 4.17
         Fr   Froude number = Wo/[g(AT/T)D] [dimensionless]
          g   Gravitational acceleration [//t2 ]
         h   Effective stack height = hs + Ah [/]
         hs   Stack height [/]
        Ah   Plume rise above top of stack [/]
         k   Unit vector  in the vertical direction [dimensionless]
         L   Characteristic length for buoyant plume in crosswind = F/u3  [/]
         Q   Emission rate of a gaseous effluent [m/t]
        QH   Heat emission due to efflux of stack gases [m/2/t3]
         R   Ratio of efflux velocity to average windspeed = w0/u [dimensionless]
          r   Characteristic radius of plume or puff, defined as (V/u) ^2  for a bent-over
                plume [/]
         r0   Internal stack radius  [/]

                                          65

-------
66                                                                        APPENDIX B

           s   Restoring acceleration per unit vertical displacement for adiabatic motion in
                 atmosphere [t"2]; see Eq. 4.16
           T   Average absolute temperature of ambient air [r]
          Ts   Average absolute temperature of gases emitted from stack [r]
         AT   Temperature excess of stack gases = Ts — T[T]
        AT;   Temperature difference between top and bottom of an elevated inversion
                 M
               Vertical temperature gradient of atmosphere [r/l]
           t   Time [t]
           u   Average wind speed at stack level [l/t]
          u*   Friction velocity in neutral surface layer [l/t] ; see Ref. 115
           V   Vertical volume flux of plume divided by TT [/3 /t]; see Eq. 4.15
           v   Average velocity of plume gases [l/t]; see Eq. 4.18
           v"   Velocity excess of plume gases = \?p — ve [l/t]
           ve   Average velocity of ambient air [l/t]
          vp   Average local velocity of gases in plume [l/t]
           w   Vertical component of v = k • \f [l/t]
          w'   Vertical component of Vp = k • vp [l/t]
          w0   Efflux speed of gases from stack [/ /t]
           x   Horizontal distance downwind of stack [/]
          x*   Distance at which atmospheric turbulence begins to dominate entrainment
                 [/]; see Eq. 4.34.
           y   Horizontal distance crosswind of stack [/]
           z   Vertical distance above stack [I]
           z   Height above the ground [/]
           Z;   Height of penetratable elevated inversion above stack [/]
           a   Entrainment constant for vertical plume [dimensionless] ; see Eq. 4.22
           /3   Entrainment constant for mixing by atmospheric turbulence [dimensionless];
                 see Eq. A.2
           r   Adiabatic lapse rate of atmosphere = 5.4°F/1000 ft [r/l]
           7   Entrainment constant for bent-over plume [dimensionless] ; see Eq. 4.23
           e   Eddy  energy  dissipation  rate  for atmospheric  turbulence [/2/t3] ;  see
                 Ref. 115
           6   Average potential temperature of ambient air [T]
          6'   Potential temperature excess of plume gases = 6p — 0  [T]
          dp   Average potential temperature of gases in plume [T]
               Vertical potential temperature gradient of atmosphere  [r/l]; see Eq. 2.1
           p   Average density of ambient air [m//3 ]
          Po   Density of gases emitted from stack [m//3 ]
          pp   Average density of gases in plume [m//3 ]
       az/ay   Ratio of vertical dispersion to horizontal dispersion [dimensionless]
           X   Concentration of a gaseous effluent [m//3 ]

-------
c
APPENDIX:
GLOSSARY  OF TERMS
      Adiabatic lapse rate  The rate at which air lifted adiabatically cools owing to the drop
         of pressure with increasing height, 5.4°F/1000 ft in the earth's atmosphere.
      Advection The  transport of a fluid property by the mean velocity field of the fluid.
      Buoyant plume  A plume initially of lower density than the ambient fluid after the
         pressure  is adiabatically brought  to  equilibrium.  Usually,  the  term  "buoyant
         plume" refers to a plume in which the effect of the initial momentum is  small, and
         the term  "forced plume" refers to  a plume with buoyancy in which the effect of
         the initial momentum is also  important.
      Convection   Mixing motions in a fluid arising from the conversion of potential energy
         of hydrostatic instability into kinetic energy. It is more precise to term this motion
         "free  convection" to distinguish it from "forced convection," which arises from
         external forces.
      Critical wind speed  In the context of this critical review, the wind speed at  the height
         of an elevated  plume for which the maximum ground concentration is  highest in
         neutral conditions.
      Diffusion The mixing of a fluid property by turbulent and molecular motions within
         the fluid.
      Downwash The downward motion of part or all of a plume due to the lower pressure
         in the wake of the stack or building or due to a  downward step of the terrain.
      Effective stack height Variously defined. The three most common definitions are: (1)
         the height at which a plume levels off,  which has been observed only in stable
         conditions; (2)  the  height  of a plume above the point of maximum ground
         concentration; (3) the virtual height of plume origin based on the diffusion pattern

                                          67

-------
68                                                                         APPENDIX C

          at large distances downwind  of the stack.  Definition 1 is the easiest to apply in
          stable  conditions; definition 2 is  the most practical  in  neutral  and unstable
          conditions; definition 3 is comprehensive but difficult to apply.
      Efflux velocity  The mean speed of exiting stack gases.
      Entrainment  The dilution of plume properties due to mixing with the ambient fluid.
      Final rise  The total plume rise after leveling off, if this occurs, especially as opposed
          to the term "transitional rise."
      Froude  number  The ratio of pressure forces to buoyant forces. The efflux Froude
          number of a stack may be defined as Wo/[g(AT/T)D].
      Fumigation The  downward  diffusion  of  pollutants  due  to  convective  mixing
          underneath an inversion that prevents upward diffusion.
      Inversion  A layer of air in which temperature increases with height. Such a layer is
          also stable.
      Jet  A nonbuoyant plume.
      Lapse rate  The rate at which temperature drops with increasing altitude; the negative
          of the vertical temperature gradient.
      Neutral   In hydrostatic equilibrium.  A  neutral atmosphere is characterized by an
          adiabatic lapse rate, i.e., by potential temperature constant with height.
      Plume  rise  The rise of a plume  center line  or center of mass above its point of origin
          due  to initial vertical momentum or buoyancy, or both.
      Potential temperature The temperature that a gas would obtain if it were adiabati-
          cally compressed to some standard pressure, usually 1000 mb in meteorological
          literature.
      Stable   Possessing hydrostatic stability. A stable atmosphere has a positive potential
          temperature gradient.
      Stratification  The variation  of potential temperature with height. Usually the term
          "stratified fluid" refers to a fluid possessing hydrostatic stability, as does the
          atmosphere when the potential temperature gradient is positive.
      Temperature gradient  In meteorology, usually the vertical gradient of mean tempera-
          ture.
      Transitional rise  The rise of a plume  under the influence of the mean  wind and the
          properties of the plume itself; i.e.,  the rise before atmospheric turbulence or
          stratification has a significant effect.
      Turbulence Three-dimensional diffusive motions in  a fluid on a macroscopic scale.
          According to Lumley and Panofsky,115 turbulence is also rotational, dissipative,
          nonlinear, and stochastic.
      Unstable  Possessing  hydrostatic  instability. An unstable atmosphere has a negative
          potential temperature gradient.

-------
                                                                    REFERENCES
 1. A. E. Wells, Results of Recent Investigations of the Smelter Smoke Problem, Ind. Eng. Chem.,
    9: 640-646 (1917).
 2. C. H. Bosanquet and J. L. Pearson, The Spread  of Smoke and Gases from Chimneys, Trans.
    Faraday Soc., 32:  1249-1264(1936).
 3. G. R. Hill, M. D. Thomas, and J. N. Abersold,  Effectiveness of High Stacks in Overcoming
    Objectionable Concentration of Gases at Ground  Level, 9th  Annual Meeting, Industrial
    Hygiene  Foundation of  America,  Nov.  15-16, 1944,  Pittsburg,  Pa., Industrial  Hygiene
    Foundation,  Pittsburgh, Pa.
 4. Symposium on Plume Behavior,//;?. /. Air  Water Pollut.,  10: 343-409 (1966).
 5. P. R.  Slawson and G. T. Csanady, On the Mean Path of Buoyant, Bent-Over Chimney Plumes,
   J. FluidMech., 28: 311-322 (1967).
 6. U. S. Weather Bureau, A Meteorological Survey of the Oak Ridge Area: Final Report Covering
    the Period 1948-1952, USAEC Report ORO-99, pp. 554-559, Weather Bureau, 1953.
 7. J. E.  Hawkins and  G. Nonhcbel, Chimneys  and the Dispersal  of Smoke, J. Inst. Fuel,
    28: 530-545  (1955).
 8. O. G. Sutlon,Micrometeorology,  McGraw-Hill Book Company, Inc., New York, 1953.
 9. F. Pasquill, Atmospheric Diffusion, D. Van Nostrand Company, Inc., Princeton, N. J., 1962.
10. U. S. Weather Bureau, Meteorology and Atomic  Energy, USAEC Report AECU-3066, 1955.
11. Environmental Science  Services  Administration, Meteorology  and  Atomic Energy —1968,
   USAEC Report TID-24190, 1968.
12. G. H. Strom, Atmospheric Dispersion of Stajk Effluents, Air Pollution, Vol. 1, pp. 118-195,
   A. C. Stem (Ed.), Academic Press, Inc.,  New York, 1962.
13. M. E. Smith,  Atmospheric Diffusion Formulae and Practical Pollution Problems,/. Air Pollut.
   Contr. Ass., 6: 11-13 (1956).
14. M. E. Smith,  Reduction  of  Ambient Air  Concentrations of Pollutants by Dispersion from
   High Stacks,  USAEC  Report BNL-10763, Brookhaven National Laboratory, 1966.
15. K W. Thomas, TVA Air Pollution Studies Program, Air Repair, 4: 59-64 (1954).

                                         69

-------
70                                                                              REFERENCES

        16. F. E. Gartrell, Transport  of S02 in  the Atmosphere from a Single Source, in Atmospheric
           Chemistry  of Chlorine and Sulfur  Compounds,  pp. 63-67, American  Geophysical  Union
           Monograph 3, 1959.
        17. F. W. Thomas, S. B. Carpenter, and F. E. Gartrell, Stacks—How High?,  /. AirPollut.  Contr.
           Ass., 13: 198-204 (1963).
        18. G. N. Stone and A. J. Clarke, British Experience with Tall Stacks for Air Pollution Control on
           Large Fossil-Fueled Power Plants, in Proceedings of the American Power Conference, Vol. 29,
           pp. 540-556, Illinois Institute of Technology, Chicago, 111., 1967.
        19. C. H. Bosanquet, W. F. Carey, and E.  M. Halton, Dust Deposition from Chimney Stacks, Proc.
           Inst. Mech. Eng., 162: 355-367  (1950).
        20. C. H. Bosanquet, The Rise of a Hot Waste Gas Plume,/. Inst. Fuel, 30: 322-328 (1957).
        21. G. Nonhebel, Recommendations on Heights for New Industrial Chimneys  (with discussion),/.
           Inst. Fuel, 33: 479-511(1960).
        22. Ministry of Housing and Local Government, Clean Air Act, 1956: Memorandum on Chimney
           Heights, S. O. Code No. 75-115, Her Majesty's Stationery Office, London.
        23. G. Nonhebel, British Charts for Heights of New Industrial Chimneys, Int. J. Air Water Pollut.,
           10: 183-189 (1966).
        24. R. S. Scorer and C. F.  Barrett, Gaseous Pollution from Chimneys, Int. J. Air Water Pollut.,
           6: 49-63 (1962).
        25. CONCAWE, The Calculation of Dispersion from a Stack,  Stichting CONCAWE,  The Hague,
           The Netherlands, 1966.
        26. K. G. Brummage, The Calculation of Atmospheric Dispersion from  a Stack, Atmos. Environ.,
           2: 197-224(1968).
        27. M. E. Smith (Ed.), Recommended Guide for the  Prediction  of the Dispersion of Airborne
           Effluents, 1st ed., American Society of Mechanical Engineers, New York, May 1968.
        28. G. A. Briggs, CONCAWE Meeting:  Discussion of the Comparative Consequences of Different
           Plume Rise Formulas, Atmos. Environ., 2: 228-232 (1968).
        29. Discussion of Paper on Chimneys and  the Dispersal of Smoke (see Ref. 7), /. Inst. Fuel,
           29: 140-148(1956).
        30. C. H. Bosanquet and A. C. Best, Discussion,/. Inst.  Fuel,  30: 333-338 (1957).
        31. Symposium on the Dispersion of Chimney Gases, Dec. 7, 1961, Int. /. Air Water Pollut.,
           6: 85-100(1962).
        32. Round  Table on Plume Rise and Atmospheric  Dispersion, Atmos. Environ., 2: 193-196 and
           2: 225-250(1968).
        33. Symposium  on Chimney  Plume  Rise  and  Dispersion: Discussions,   Atmos.  Environ.,
           1: 425-440(1967).
        34. R. H. Sherlock  and E. A.  Stalker, A Study of Flow Phenomena in the Wake of Smoke Stacks,
           Engineering Research Bulletin 29, University of Michigan, Ann Arbor, Mich.,  1941.
        35. S. Goldstein (Ed.), Modern Developments  in  Fluid Dynamics, p.  430,  Dover Publications,
           Inc., New York, 1965.
        36. R. S. Scorer, The Behavior of Chimney Plumes, Int. J. Air Pollut., 1: 198-220 (1959).
        37. Committee  Appointed  by the Electricity Commissioners, The Measures Which Have Been
           Taken in This Country  and in Others to Obviate the Emission of Soot, Ash, Grit and Gritty
           Particles from  the Chimneys of Electric  Power Stations, Her Majesty's Stationery Office,
           London, 1932.
        38. J. Halitsky, Gas Diffusion Near Buildings, Meteorology and Atomic Energy—1968, USAEC
           Report TID-24190, pp. 221-255, Environmental Science Services Administration,  1968.
       39. D. H. Lucas,  Symposium  on the Dispersion of Chimney Gases, Royal Meteorology Society,
           Dec. 7, 1961,Int. J. AirWater Pollut., 6:  94-95  (1962).
       40. F. E. Ireland, Compliance with the  Clean Air Act: Technical Background  Leading to the
           Ministry's Memorandum on Chimney Heights,/. Inst. Fuel,  36: 272-274 (1963).

-------
REFERENCES                                                                               71

 41. P. J. Barry, Estimates of Downwind Concentration of Airborne Effluents Discharged in the
    Neighborhood of Buildings, Canadian Report AECL-2043, 1964.
 42. \V. M.  Culkowski,  Estimating the Effect of Buildings on Plumes from  Short Stacks, Nud.
    Safety, 8(3): 257-259(1967).
 43. P. O. Davies and D. J. Moore, Experiments on the Behavior of Effluent Emitted from Stacks
    at or near the Roof of Tall Reactor Buildings, Int. J. Air Water Pollut., 8:  515-533 (1964).
 44. H. St'umke, Consideration  of Simplified Terrain Types  in the Calculation of the Turbulent
    Propagation of Chimney Gases,  Staub, 24: 175-182 (1964); translated in  USAEC Report
    ORNL-tr-981, Oak Ridge National Laboratory.
 45. H. Stu'mke, Correction  of the Chimney Height Due to an Influence  of the Terrain, Staub,
    24: 525-528  (1964);  translated  in USAEC  Report  ORNL-ti-997,  Oak  Ridge  National
    Laboratory.
 46. R. S. Scorer, Natural Aerodynamics, pp. 143-217, Pergamon Press, Inc., New York, 1958.
 47. R. A. Scriven, On  the Breakdown of Chimney Plumes into Discrete Puffs, Int. J. Air Water
    Pollut., 10: 419-425 (1966).
 48. G. K. Batchelor, Heat Convection and Buoyant Effects in Fluids, Quart.  J.  Roy. Meteorol.
    Soc., 80: 339-358 (1954).
 49. R. Serpolay, Penetration d'une couche d'inversion par un cumulus industriel, Int. J. Air Water
    Pollut., 8: 151-157 (1964).
 50. M. E.  Smith  and  I.  A.  Singer,  An Improved  Method for Estimating  Concentrations and
    Related Phenomena from a Point Source Emission,/. Appl. Meteorol., 5:  631-639 (1966).
 51. C. R. Hosier,  Climatological Estimates of Diffusion Conditions in the  U. S., Nud. Safety,
    5(2): 184-192 (1963).
 52. E. W. Hewson, E. W.  Bierly,  and G. C. Gill, Topographic Influences on the Behavior of Stack
    Effluents, in Proceedings of the American Power Conference, Vol. 23, pp. 358-370, 1961.
 53. H. Schlichting, Boundary Layer Ttieory, pp. 607-613, McGraw-Hill Book Company, Inc., New
    York, 1960.
 54. S. I. Pai, Fluid Dynamics of Jets, D. Van Nostrand Company, Inc., New York,  1954.
 55. W. Schmidt, Turbulente Ausbreitung eines Stromes Erhitzter  Luft,  Z. Angew. Math. Mech.,
    21: 265-278 and 21: 351-363 (1941).
 56. C. S. Yih,  Free  Convection  Due to a Point  Source of Heat, in Proceedings of First  U. S.
    National  Congress of Applied Mechanics, pp. 941-947, American Society of Mechanical
    Engineers, 1951.
 57. H. Rouse, C. S.  Yih,  and H. W. Humphreys, Gravitational Convection from  a Boundary
    Source, Tellus, 4: 201-210 (1952).
 58. B. R.  Morton,  G.  I.  Taylor,  and J. S. Turner, Turbulent Gravitational  Convection  from
    Maintained and Instantaneous Sources, Proc. Roy. Soc.  (London], Ser. A,  234: 1-23 (1956).
 59. T. V. Crawford and A. S. Leonard, Observations of Buoyant Plumes in Calm Stably Stratified
    Air, /. Appl. Meteorol,  1: 251-256 (1962).
 60. L. Vadot, Study of Diffusion of Smoke Plumes into the Atmosphere  (in French), Centre
    Interprofessionnel Technique d'Etudes de la Pollution Atmospherique, Paris, 1965.
 61. A. F. Rupp, S. E. Beall, L. P. Bornwasser, and D. F.  Johnson, Dilution of Stack Gases in Cross
    Winds,  USAEC Report AECD-1811 (CE-1620), Clinton Laboratories, 1948.
 62. E. E. Callaghan and R. S.  Ruggeri,  Investigation of the Penetration of  an Ail Jet  Directed
    Perpendicularly to  an Air Stream, Report NACA-TN-1615, National Advisor,' Committee for
    Aeronautics, 1948.
 63. T. F. Keffer and W. D.  Baines, The Round Turbulent Jet in a Crosswind, /. Fluid Mech.,
    15: 481-497 (1963).
 64. J. Halitsky, A Method  for Estimating Concentration in Transverse Jet Plumes, Int. J. Air
    Water Pollut.,  10: 821-843 (1966).
 65. M. A.  Patrick,  Experimental Investigation of the Mixing  and Penetration  of  a Round
    Turbulent Jet Injected  Perpendicularly into  a Traverse  Stream,  Trans. Inst.  Chem.  Eng.
    (London), 45:  16-31 (1967).

-------
72                                                                              REFERENCES

       66. L. W. Bryant, The Effects of Velocity and Temperature of Discharge on the Shape of Smoke
           Plumes from  a Funnel or Chimney: Experiments in a  Wind Tunnel, National Physical
           Laboratory, Great Britain, Adm. 66, January 1949.
       67. L. W. Bryant and C. F. Cowdrey, The Effects of Velocity and Temperature of Discharge on
           the Shape of Smoke Plumes from a Tunnel or Chimney: Experiments in a Wind Tunnel, Proc,
           Inst. Mech. Eng. (London), 169: 371-400(1955).
       68. D. P. Hoult, J. A.  Fay,  and L. J.  Forney, Turbulent Plume  in  a Laminar Cross Wind,
           Massachusetts Institute of Technology, Department  of Mechanical Engineering, 1967.
       69. R. S.  Richards, Experiment  on  the  Motions  of Isolated  Cylindrical Thermals through
           Unstratified Surroundings, Int. J. Air Water Pollut., 1: 17-34 (1963).
       70. D. K.  Lilly, Numerical  Solutions  for  the  Shape-Preserving  Two-Dimensional  Thermal
           Convection Element, J.Atmos. Scl, 21: 83-98 (1964).
       71. L. Fan, Turbulent  Buoyant Jets into  Stratified  or  Flowing  Ambient Fluids,  California
           Institute of Technology, Report KH-R-15, 1967.
       72. F. T.  Bodurtha,  Jr.,  The  Behavior of Dense  Stack Gases, /. Air Pollut. Contr.  Ass.,
           11: 431-437 (1961).
       73. C. H. B. Priestley, A Working Theory of the Bent-Over Plume  of Hot  Gas, Quart. J. Roy.
           Meteorol. Soc.,  82:  165-176 (1956).
       74. N. G. Stewart, H. J. Gale, and R. N.  Crooks, The Atmospheric Diffusion of Gases Discharged
           from the Chimney of the Harwell Pile (BEPO), British Report AERE HP/R-1452, 1954.
       75. N. G. Stewart, H. J. Gale, and R. N.  Crooks, The Atmospheric Diffusion of Gases Discharged
           from the Chimney of the Harwell Reactor BEPO,/«r. /. Air Pollut., 1: 87-102 (1958).
       76. F. K. Ball, Some Observations  of Bent  Plumes, Quart. J. Roy. Meteorol. Soc.,  84:61-65
           (1958).
       77. H. Moses and G. H. Strom, A Comparison of  Observed Plume  Rises with Values Obtained
           from Well-Known Formulas, J. Air Pollut. Contr. Ass., 11: 455-466 (1961).
       78. A. M. Danovich  and S. G. Zeyger, Determining the  Altitude of Rise of a Heated Contaminant
           in  the Atmosphere,   JPRS-28,188,  pp. 52-66;   translated  from  Trudy, Leningradskii
           Gidrometeorologischeskii   Institut,  Vypusk 18 (Proc.  Leningrad   Hydrometeorol.   Inst.,
           Vol. 18).
       79. L. D. Van Vleck and F. W. Boone, Rocket Exhaust  Cloud Rise and Size Studies, Hot Volume
           Sources, paper  presented at 225th National Meeting of the American Meteorology Society,
           January 29-31, 1964, Los Angeles, Calif.
       80. H. Rauch, Zu'r Schornstein-Uberhohung, Beitr. Phys. Atmos., 37: 132-158 (1964); translated
           in USAEC Report ORNL-tr-1209, Oak Ridge National Laboratory.
       81. B. Bringfelt, Plume  Rise Measurements at Industrial Chimneys, Atmos. Environ.,  2: 575-598
           (1968).
       82. V. Hbgstrom, A Statistical  Approach to the Air Pollution Problem of Chimney Emission,
           Atmos. Environ., 2:  251-271 (1968).
       83. S. Sakuraba, M. Moriguchi,  I. Yamazi,  and J. Sato, The Field Experiment of Atmospheric
           Diffusion  at Onahama Area (in Japanese with  English figure headings), /. Meteorol.  Res.,
           19(9): 449-490(1967).
       84. Diffusion  Subcommittee,  Report on Observations and Experiments  on Smoke Rise Patterns
           During Formation of Inversion Layer (in Japanese), Tech. Lab.,  Cent. Res. Inst. Elec. Power
           Ind., (Japan), Mar. 12, 1965.
       85. I. A.  Singer, J. A. Frizzola, and M. E. Smith, The Prediction of the Rise  of a Hot Cloud from
           Field Experiments,/. Air Pollut.  Contr. Ass.,  14: 455-458 (1964).
       86. J. A. Frizzola, I. A. Singer, and M. E. Smith, Measurements of the Rise of Buoyant Clouds,
           USAEC Report BNL-10524, Brookhaven National Laboratory,  April  1966.
       87. G. T. Csanady,  Some Observations  on  Smoke  Plumes, Int. J.  Air  Water Pollut 4-47-51
           (1961).

-------
 REFERENCES                                                                              73

 88. P. R.  Slawson, Observations of Plume Rise from a Large Industrial Stack, Research Report 1,
    USAKC Report NYO-3685-7, University of Waterloo, Department of Mechanical Engineering,
    1966.
 89. G. T.  Csanady (Project Supervisor), Diffusion in  Buoyant Chimney  Plumes, Annual Report,
    1966,  USA EC  Report NYO-3685-10, University of  Waterloo, Department of Mechanical
    Engineering, January 1967.
 90. G. T.  Csanady  (Project  Supervisor),  Research  on Buoyant Plumes, Annual Report, 1967,
    USAKC Report NYO-3685-13, University of Waterloo, Department  of Mechanical Engineer-
    ing, January 1968.
 91. D. 11. Lucas, D. J. Moore, and G. Spurr, The Rise of Hot Plumes from Chimneys, Int J. Air
    Water Pollut., 1: 473-500 (1963).
 92. D. H.  Lucas, G. Spurr,  and  !•'. Williams, The Use of Balloons  in Atmospheric Pollution
    Research, Quart. J. Roy. Meteorol. Soc., 83:  508-516 (1957).
 93. P. M.   Hamilton,  Plume Height  Measurements  at  Two  Power Stations, Atnios. Environ.,
    I: 379-387 (1967).
 94. P. M.   Hamilton,  The Use  of Lidar in  Air Pollution Studies, Int. J. Air Water Pollut..
    10: 427-434 (1966).
 95. P. M.  Hamilton, K. W. James, and D. J. Moore, Observations of Power Station Plumes Using a
    Pulsed Ruby Laser Rangcfmder, Nature. 210: 723-724  (1966).
 96. D. H.  Lucas,  K. W. James, and I. Davies, The Measurement of Plume Rise and Dispersion at
    Tilbury Power Station, ,4tnios. Environ.. I: 353-365  (1967).
 97. I1'. E.  Gartrell, I'. W. Thomas, and S. B. Carpenter, An Interim Report on Full Scale Study of
    Dispersion of Stack Gases,/. A ir Pollut. Con tr. Ass..  11: 60-65(1961).
 98. !•'. E.  Gartrell, E. W. Thomas, and  S. B.  Carpenter,  Ful]  Scale Study of Dispersion of Stack
    Gases, Tennessee Valley Authority, Chattanooga, 1964.
 99. S. B.  Carpenter  et al.. Report  on a  Full  Scale Study  of  Plume  Rise at Large  Electric
    Generating Stations,  Paper 67-82,  60th Annual  Meeting  of the Air  Pollution  Control
    Association, Cleveland, Ohio,  1967.
100. J.E. Vehrencamp, A. Ambrosio, and F. E.  Romie,  Convection from Heated Sources in  an
    Inversion  Layer,   Report 55-27,  University  of  California,  Los  Angeles,  Department  of
    Engineering, 1955.
101. R. W.  Davies,  Large-Scale  Diffusion  from an Oil Fire, in Advances in Geophysics.  Vol. 6,
    pp. 413-415,  F. N.  Frenkiel  and  P. A.  Sheppard (Eds.), Academic Press, Inc.,  New York,
    1959,
102. R. W. Davies, Jet Propulsion Laboratory. Pasadena, Calif., personal communication, 1966.
103. C. Simon  and W. Proud fit.  Some  Observations  of Plume Rise and Plume Concentration
    Distributed Over  N.Y.C., Paper 67-83,  60th Annual  Meeting of  the Air Pollution  Control
    Association, Cleveland, Ohio, June 11 — 16, 1967.
104. 11. St'u'mke, Zur  Berechnung  Der Aufstiegshohe von Rauchfahnen,  VDI Forschungsh., 483,
    Ausg. B. 27: 38-48,  1961.
105. W. 1-'.   Davidson, The Dispersion  and Spreading of  Gases arid Dusts from Chimneys, in
    Transactions of  Conference  on Industrial Wastes.  14th  Annual Meeting of the Industrial
    Hvgicnc Foundation of America, pp. 38-55, Industrial Hygiene Foundation, Pittsburgh, Pa.,
    1949.
106. M. Ye. Berlyand, Ye. L. Genikhovich, and R. 1. Onikul. On Computing Atmospheric Pollution
    by Discharge  from the Stacks of Power Plants, in Problems of Atmospheric Diffusion and Air
    Pollution. JPRS-28, 343: 1-27 (1964); translated from  Tr.  Gl. Geofi:. Obser., No.  158.
107. C. A.  Briggs,  A Plume Rise  Model Compared with Observations,/ Air Pollut.  Contr. Ass.,
    15: 433-438 (1965).
108. H. St'u'mke, Suggestions for an Empirical Formula for Chimney Elevation, Staub, 23:  549-556
    (1963); translated in USAEC Report ORNL-tr-977, Oak Ridge National Laboratory.
109. 1). 11.  Lucas, Application and Evaluation of Results  of the Tilbury  Plume Rise and Dispersion
    Experiment, Atnws. Environ.. 1: 421-424 (1967).

-------
74                                                                              REFERENCES

       110. H.Moses and  J. E.  Carson, Stack Design Parameters Influencing Plume  Rise, Paper 67-84,
           60th Annual Meeting of the Air Pollution Control Association, Cleveland, Ohio, 1967.
       111. G. N. Abramovich, The Theory of Turbulent Jets, The M.I.T. Press, Cambridge, Mass., 1963.
       112. G. I. Taylor, Dynamics of a Mass of Hot Gas Rising in the Air, USAEC Report MDDC-919
           (LADC-276), Los Alamos Scientific Laboratory, 1945,
       113. G. A. Briggs, Penetration of Inversions by Plumes, paper presented at 48th Annual Meeting of
           the American Meteorological Society, San Francisco, 1968.
       114. F. P. Ricou and D. B. Spalding, Measurements of Entrainments by Axisymmetrical Turbulent
           Jets,/. FluidMech.,  11: 21-32 (1961).
       115. J. L. Lumley and H. A.  Panofsky, The Structure of Atmospheric Turbulence, pp. 3-5,  John
           Wiley & Sons, Inc., New York, 1964.
       116. W. Tollmien, Berechnung der Turbulenten Ausbreitungsvorgange, Z. Angew. Math. Mech.,
           4: 468-478 (1926).
       117. C. H. B. Priestley and F. K. Ball, Continuous Convection from  an Isolated Source of Heat,
           Quart. J. Roy. Meteorol Soc., 81: 144-157 (1955).
       118. M. A. Estoque, Venting of Hot Gases Through Temperature Inversions, Geophysical Research
           Directorate Research Note 3, Report AFCRC-TN-58-623 (AD-160756), Air Force Cambridge
           Research Center, 1958.
       119. B. R. Morton,  The Ascent of Turbulent  Forced Plumes in a Calm Atmosphere, Int. J. Air
           Pollut.,  1: 184-197 (1959).
       120. B. R. Morton, Buoyant Plumes in a Moist Atmosphere,/ Fluid Mech.,  2: 127-144 (1957).
       121. M. Hino, Ascent of Smoke in  a Calm  Inversion  Layer of Atmosphere; Effects of Discharge
           Velocity and Temperature  of  Stack Gases, Tech. Lab., Cent. Res.  Inst. Elec. Power Ind.
           (Japan), Report TH-6201, May  1962.
       122. M. Hino, Limit of Smoke Ascent in a Calm Inversion Layer of Atmosphere, Tech. Lab., Cent.
           Res, Inst. Elec.  Power Ind. Rep. (Japan), (text in Japanese; figure headings in English),
           14(1): 9-43 (1963).
       123. J. S. Turner,  The 'Starting Plume' in  Neutral  Surroundings,  /.  Fluid Mech., 13:356-368
           (1962).
       124. A. Okubo,  Fourth Report  on the "Rising Plume" Problem in the Sea,  USAEC  Report
           NYO-3109-31, Johns Hopkins University,  1968.
       125.1. V. Vasil'chenko, On the Problem of a Steady-State  Convection  Flow,  Tr.  Gl.  Geofiz.
           Observ., No. 93, 1959.
       126. J. W. Telford, The Convective Mechanism in Clear Air,/. Atmos. Sci., 23: 652-665 (1966).
       127. B. R. Morton,  On Telford's Model for Clear Air Convection (with  reply), /.  Atmos.  Sci,
           25: 135-139(1968).
       128. S. Lee,  Axisymmetrical  Turbulent Swirling  Natural-Convection  Plume,  /. Appl.  Mech.,
           33:647-661(1966).
       129. O. G. Sutton,  The Dispersion  of  Hot  Gases  in  the Atmosphere, /.  Meteorol., 1: 307-312
           (1950).
       130. R. S. Scorer, The Rise of a Bent-Over Plume, Advances in Geophysics,  Vol. 6, pp. 399-411,
           F. N. Frenkiel and P. A. Sheppard (Eds.), Academic Press, Inc., New York, 1959.
       131. D. J. Moore, Discussion  of Paper: Variation of Turbulence with Height, Atmos. Environ.,
           1:521-522(1967).
       132. F. H.  Schmidt, On the Rise of Hot Plumes in  the Atmosphere, Int. J. Air Water Pollut
           9:175-198(1965).
       133. D. J. Moore, On the Rise of Hot Plumes in the Atmosphere, Int. J. Air Water Pollut
           9:233-237(1965).
       134. F. A.  Gifford,  The Rise of Strongly  Radioactive Plumes, /.  Appl  Meteorol  6- 644-649
           (1967).
       135. J. S. Turner, Model Experiments  Relating to Thermals with Increasing Buoyancy, Quart J
           Roy. Meteorol. Soc., 89: 62-74 (1963).

-------
REFERENCES                                                                              75

136. G. T. Csanady, The Buoyant Motion Within a Hot Gas Plume in a Horizontal Wind,/ Fluid
    Mech., 22:  225-239 (1965).
137. D. P. Hoult,  J. A. Fay, and  L. J. Forney, A  Theory of Plume Rise Compared with Field
    Observations, Paper 68-77, 61st Annual Meeting of the Air Pollution Control Association,
    June 23-28,  1968, St. Paul, Minn.
138. H. Moses, G.  H.  Strom, and J. E. Carson, Effects of Meteorological and Engineering Factors
    on Stack Plume Rise.M/ci Safety, 6(1): 1-19 (1964).
139. H. Moses and G. H. Strom, A Comparison of Observed Plume Rises with Values Obtained
    from Weil-Known Formulas,/. AirPollut. Contr. Ass., 11: 455-466  (1961).
140. R. Anderson et al., Electricity in Volcanic Clouds, Science, 148(3674): 1179-1189 (1965).
141. S. Thorarinsson and B. Vonnegut, Whirlwinds Produced by the Eruption of Surtsey Volcano,
    Bull. Amer. Meteorol. Soc., 45: 440-443 (1964).
142. S. Hanna, A  Model of Vertical  Turbulent Transport in  the Atmosphere, Ph. D. Thesis, The
    Pennsylvania  State University, 1967.
143. F. Pasquill, The Vertical  Component  of Atmospheric Turbulence at Heights up  to 1200
    Metres, Atmos. Environ., 1: 441-450(1967).
144. F. Record and H. Cramer,  Turbulent Energy Dissipation Rates and Exchange Processes Above
    a Non-homogeneous Surface, Quart. J. Roy. Meteorol. Soc., 92: 519-532 (1966).
145. J. Kaimal,  An Analysis of Sonic  Anemometer  Measurements from  the Cedar Hill Tower,
    Report AFCRL-66-542, Air Force Cambridge Research Laboratory, 1966.
146. Boeing Company, Low Level Critical Air Turbulence, Technical  Progress-Monthly  Report,
    Contract No. AF33(615)-3724, Doc. No. 83-7087-11 and 83-7087-16 (1967).
147. S. Hanna, A Method of Estimating Vertical Eddy Transport in the Planetary Boundary Layer
    Using  Characteristics of  the Vertical  Velocity Spectrum, J. Atmos. Sci., 25: 1026-1033
    (1968).
148. N. E, Busch and H. A. Panofsky, Recent Spectra of Atmospheric Turbulence, Quart. J. Roy.
    Meteorol. Soc., 94:  132-148 (1968).
149. F. N. Frenkiel and I. Katz, Studies of Small-Scale Turbulent Diffusion in the Atmosphere,/.
    Meteorol.,  13: 388-394 (1956).
150. F. B. Smith and J. S. Hay,  The Expansion of Clusters of Particles in  the Atmosphere, Quart. J.
    Roy. Meteorol. Soc., 87: 82-101 (1961).

-------
                                                   AUTHOR  INDEX
Abersold, J. N., 2
Abramovich, G. N., 25
Ambrosio, A., 21, 34, 52
American Society of Mechanical
  Engineers (ASME), 4, 23
Anderson, R.,  52
Baines,W. D.,  17,37,48,49
Ball,F. K., 19,34,39,44,46,52
Barrett, C. F.,4
Barry, P. J., 7
Batchelor,G. K.,8,34, 37
Beall.S. E., 17,22,24,25,50
Berlyand,M. Ye.,23,24
Best, A. C., 4
Bierly,E. W., 15
Bodwitha, F. T., 18
Boeing Company, 62
Boone,F.W., 19,54
Bornwasser, L. P., 17, 22, 24, 25, 50
Bosanquet, C. H.,  2, 4, 18, 35, 38, 39,
  43,54
Briggs, G. A., 23, 29, 32, 37, 39, 43, 54
Bnngfelt,B.,20
Brummage, K. G., 4
Bryant, L. W., 18, 23, 35, 39, 48, 49
Busch, N. E., 62
Callaghan, E. F.,  17, 25, 48, 49
Carey, W. F.,4, 18,35,38
Carpenter, S. B.,  3, 15, 21, 30, 44, 54
Carson,J. E., 24, 38, 45, 46
Chapman, C. S., 48, 49
Clark, A. J., 3
C ONCAWE (see footnote, page 4),
  4,24,38,39
Cowdrey.C. F., 18,48,49
Cramer, H.,  62
Crawford,!. V.,  17,52
Crooks, R.N., 19,23,38,44
Csanady, G. !., 3, 20, 23, 36, 37,

Culkowski.W. M.,7
Danovich, A. M., 19,37
Davidson, W. F.,  23, 39
Da vies, I., 20
Davies,P. 0., 7
Davies,R. W.,21,52
                                   77

-------
78
                                                                  AUTHOR INDEX
      Estoque, M. A., 35
      Fan,L., 18,35,48,49,52
      Fay,J. A., 18,37,50
      Forney, L. J., 18,37,50
      Frenkiel, F. N., 63
      Frizzola, J. A.,20, 54
      Gale,H. J., 19,23,38,44
      Gartrell, F. E.,3, 15,21
      Genikhovich, Ye.I.,23,24
      Gifford, F. A., 37
      Gill, G. C., 15
      Goldstein, S., 6
      Halitsky,J.,7, 17
      Halton,E.M.,4, 18,35,38
      Hamilton, P. M., 20, 24, 44, 46
      Hanna, S. R.,61,62
      Hawkins, J. E., 4, 18
      Hay,J.S.,63
      Hewson,E. W., 15
      Hill.G. R.,2
      Hino,M.,35
      HOgstro'm, V., 20
      Holland, J. Z., 3, 18, 22, 24, 38, 39,
       45,46,54
      Hosier, C. R., 14
      Hoult,D. P., 18,37,50
      Humphreys, H. W., 17,34
      Ireland, F. E., 7
      James, K. W., 20
      Johnson, D. F., 17, 22, 24, 25, 50
      Jordinson,R.,48,49
      Kaimal, J., 62
      Katz,I.,63
      Keffer,T. F., 17,37,48,49
      Lee, S., 35
      Leonard, A. S., 17,52
      Lilly, O.K., 18,37
      Lucas, D. H.,  2, 3, 7, 20, 23, 24, 36,
       38,44,45,46
      Lumley, J. L., 61
      Moore, D. J.,  7, 20, 23, 36, 37, 38, 44,
       45,46,52,62
      Moriguchi, M., 20
      Morton,  B. R., 17, 29, 34, 35, 52, 63
Moses, H., 19,23,24,38,45,46
Nonhebel,G.,4, 7, 18
Norster,E. R.,48,49
Okubo,A.,35
Onikul, R. I., 23, 24
Pai,S. I., 17
Panofsky,H. A.,61,62
Pasquill, F., 3, 61
Patrick, M. A., 17,25,48,49
Pearson, J. L., 2
Priestley, C. H. B., 18, 23, 34, 36, 44,
  45,46,52,54
Proudfit, W., 2, 52
Rauch.H., 19,'23, 24, 38,44
Record,  F., 62
Richards, R. S., 18,29
Ricou, F. P., 29
Romie.F. E.,21,34, 52
Rouse, H., 17,34
Ruggeri, R. S., 17,25,48,49
Rupp, A. F., 17,22,24,25,50
Sakuraba, S., 20
Sato, J., 20
Schlichting, H., 17
Schmidt, F. H., 37
Schmidt, W., 17,34,36
Scorer, R. S., 4, 7, 8, 36, 37
Scriven,  R. A., 8
Serpolay.R., 10
Sherlock, R. H., 6
Simon, C., 21, 52
Singer,!. A., 13,20, 54
Slawson, P. R., 3, 20, 37, 44, 47
Smith, F B.,63
Smith, M. E., 3, 4, 13, 20, 23, 54
Spalding, D. B., 29
Spurr, G., 2, 20, 23, 36, 38, 44, 45, 46
Stalker, E. A., 6
Stewart, N. G., 19, 23, 38, 44
Stone, G. N., 3
Strom, G. H., 3, 19,23,38
StUmke, H., 8, 23, 24, 38, 45, 46
Sutton,0. G.,3,36
Taylor, G. I., 17,29, 34, 52,63

-------
AUTHOR INDEX

Telford, J. W., 35                          Vasil'chenko, I. V., 35
Thomas, F. W., 3, 15,21,23                Vehrencamp, J. E., 21, 34, 52
Thomas, M. D., 2                          Vonnegut, B., 52
Thorarinsson, S., 52                        Wells, A. E., 2
Tollmien, W., 34                          Williams, F., 20
Turner, J. S., 17, 29, 34, 35, 37, 52, 63       Yamazi, I., 20
Vadot, L., 17, 18,50,51                    Yih,C. S., 17,34
Van Vleck, L. D., 19,54                    Zeyger.S.G., 19,37
                                                                              79

-------
SUBJECT  INDEX
      Bifurcation, 8
      Brookhaven National Laboratory,
       20,54
      Building effects, 7
      Buoyancy, 6,8-9, 10, 11,18,22-24,
       26,31-33,36,50,52
      Buoyancy flux, 23, 27-28, 47
      Central Electricity Research
       Laboratories, 20, 38
      Condensation of plume, 10, 35
      Coning, 12-13
      Diffusion, 2-4, 11-15, 37, 46, 63-64
       effect of temperature profile on, 12
      Dispersion (see Diffusion)
      Down wash, 5-8, 39
      Drag force on plume, 27, 28, 29, 31,
       35
      Efflux velocity, 5-7,8,35
      Entrainment, 8, 28-31, 33, 34, 37, 49
      Entrainment velocity, 28, 31, 35, 37
      Fanning, 12-13
      Froude number, 6, 8, 17
      Fumigation, 12-15
Inversions, 9, 13, 14-15, 17,21,37,
  50-53,59
Jets, 17-18, 24-25, 29, 37, 48-50,
  52,59
Lofting, 12-13
Looping, 12-13
Modeling studies, 16-18
Momentum, 6, 8, 26, 27, 31, 33, 35,
  36,50
Momentum flux, 27, 50, 52
Multiple stacks, 47, 55-56, 58
Plume radius, growth of, 8, 30, 34, 36
Plume rise, aerodynamic effects on,5-8
  definition of, 3, 39, 46-47
  effect on diffusion, 2, 13-15
  fluctuations in, 10, 11, 58
  measurement of, 18-21
  modeling of, 16-18
  in neutral air, 10, 17-21, 33, 38-50,
    51,57-58
  qualitative description of, 8-11
  in stable air, 10, 17, 18, 19,21,29,
    31-32,50-56,58-59
                                         80

-------
SUBJECT INDEX
                                                                                    81
  near stack (first stage), 32, 36, 55,
    57,59
  in unstable air, 10, 47, 58
Plume rise formulas, empirical, 22-25
  recommended, 57-59
  theoretical, 31-33, 36
Plume rise model, bending-over
    plume, 31
  bent-over plume, 29-31
  vertical plume, 28-29, 34
Plumes, dense, 17, 18
  downwash  of, 5-8
  inclined,  18, 35
  looping of, 12-13
  puffing of, 8, 12
Potential temperature, 9, 26, 51
Potential temperature gradient,
    9-10,59
  (See also Stability)
Radiation, thermal, 11
Reynolds number, 16-17, 26
Stability, effect on plume, 9-10, 13
  measurement of, 19-21
Stack height, determination of, 3-4, 7,
   13-15
  effect on plume rise, 24, 34, 46-47
Stratification (see Stability)
Taylor entrainment hypothesis, 28-30,
  34
Temperature gradient, 9, 52
  (See also Stability)
Temperature inside plume, 8, 17, 26
Tennessee Valley Authority (TVA),
  3,14,21,30,38,54
Terrain effects, 8, 15, 46
Turbulence, atmospheric, 9, 30-31,
   33,35,61-64
  inertial subrange, 31, 37
  self-induced, 8, 17, 28, 29, 35, 37
Two-thirds law of rise, 32, 37, 42-47,
  55,57
Velocity inside plume, 8, 9, 17, 26-27,
  29
Volume flux of plume, 27, 28, 29-30
Wind speed, effect  on plume, 8, 17-18,
   29,35,36,39-42,61-64
  measurement of, 19-21
                                     NOTICE
       This book was prepared under the sponsorship of the United States Government.
       Neither the United States nor the United States Atomic Energy Commission, nor
       any  of  their  employees,  nor any of  their contractors, subcontractors, or their
       employees, makes any warranty, express or implied, or assumes any legal liability
       or responsibility for the accuracy, completeness or usefulness of any information,
       apparatus, product  or process  disclosed, or represents that its use would not
       infringe privately owned rights.

-------
NUCLEAR SAFETY INFORMATION CENTER


Plume Rise was originally prepared for the Nuclear Safety Information Center,
      one of the U. S.  Atomic Energy  Commission's specialized  information
      analysis  centers.  Established  in 1963 at the Oak Ridge National Labo-
      ratory, the Nuclear Safety  Information Center serves as a focal point for
      the  collection, storage,  evaluation, and  dissemination  of nuclear safety
      information. The subject coverage,  which is comprehensive in the nuclear
      safety  field, includes such primary subject areas as
            General Safely Considerations       Fission-Product Transport
            Plant Safety Features              Reactor Operating Experiences
            Consequences of Activity Release    Instrumentation,  Control,
            Accident Analysis                    and Safety Systems

      The Nuclear Safety Information Center  publishes periodic staff studies,
      bibliographies, and state-of-the-art  reports; disseminates selected informa-
      tion on  a biweekly basis; answers  technical  inquiries as time is available;
      provides counsel and guidance on nuclear safety problems; and cooperates
      in the  preparation  of Nuclear Safety, a bimonthly technical progress review
     -sponsored by the AEC.

      Services of the Nuclear Safety Information Center are available to govern-
      ment  agencies, research and  educational institutions, and  the  nuclear
      industry. Inquiries are welcomed.


                                          J.  R. Buchanan,  Assistant Director
                                          Nuclear Safety Information Center
                                          Oak Ridge National Laboratory

-------
                                                                              AEC
                                                                      CRITICAL
                                                                        REVIEW
                                                                          SERIES
As a continuing  series of  state-of-the-art studies published  by the AEC Office of
Information Services,  the  AEC Critical Reviews  are designed to evaluate the existing
state of knowledge  in a specific  and limited field of interest, to identify significant
developments, both published and unpublished, and to synthesize new concepts out of
the contributions of  many.
SOURCES OF TRITIUM AND ITS
BEHAVIOR UPON RELEASE
TO THE ENVIRONMENT
   December 1968 (TID-24635)
   D. G.Jacobs
   Oak Ridge National Laboratory

REACTOR-NOISE ANALYSIS
IN THE TIME DOMAIN
   April 1969  (TID-24512)
   Nicola Pacilio
   Argonne National Laboratory
   and Comitato Nazionale
   per I'Energia Nucleare

PLUME RISE
   November 1969  (TID-25075)
   G. A. Briggs
   Environmental Science Services
   Administration

ATMOSPHERIC
TRANSPORT PROCESSES
   Elmar R. Reiter
   Colorado State University
   Part 1 : Energy Transfers
   and Transformations
   December 1969  (TID-24868)
   Part 2: Chemical Tracers
   January 1971  (TID-25314)
   Part 3: Hydrodynamic Tracers
   May 1972   (TID-25731)
THE ANALYSIS OF
ELEMENTAL BORON
    November 1970  (TID-25190)
    Morris W. Lerner
    New Brunswick Laboratory

AERODYNAMIC CHARACTERISTICS
OF  ATMOSPHERIC BOUNDARY LAYERS
    May 1971  (TID-25465)
    Erich J. Plate
    Argonne National Laboratory
    and Karlsruhe University
NUCLEAR-EXPLOSION
SEISMOLOGY
   September 1971   (TID-25572)
   Howard C. Rodean
   Lawrence Livermore Laboratory

BOILING CRISIS AND CRITICAL
HEAT FLUX
   August 1972  (TID-25887)
   L. S. Tong
   Westinghouse Electric Corporation

NEPTUNIUM-237
PRODUCTION AND RECOVERY
   October 1972  (TID-25955)
   Wallace W. Schulz and Glen E. Benedict
   Atlantic Richfield Hanford Company
Available from the National  Technical Information Service,  U.  S.  Department of Commerce,
Springfield, Virginia 22151.

-------
                                                                         m m
                                                                         
-------
   ALL  Problem  Sheets and ANY additional calculations

   m¥ust be  returned to APTI to receive credit.
                                                              c
               EFFECTIVE STACK HEIGHT PROBLEMS
                                    SET ONE
       Submit  all  calculations leading to and involving final answer(s)  to each

       problem.
•H
60
H
cd


01
                         u
                                       f,y
                A power plant with an 80-meter stack 3.5 meters in diameter,

                emits  effluent gases at 93°C with an exit velocity of 15

                meters/sec.  What is the effective stack height when the

                wind speed is 4 meters/sec, using the Bryant-Davidson stack

                rise equation?  Assume air temperature is 20°C.
                    3.:
5.
                                                                 ex
                                                                 •H
) (/. z
                      2^.7 /».
                                                   26-7
CO
•H

a)
oo
cd
tx
                                                               co
                                                               co
                                                               a) cu
                                                             cu M 4->
                                                             0 id
                                                             Cd T3
                                                             B 
-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
 28

-------
   ALL Problem Sheets and  ANY  additional calculations

   must be returned  to APTI  to receive credit.
                                                           Problem Set One       (C)
<*-<


C
•H

M
IB





4-1


C
T3
0)
4J
rt
M
o
14-1
O.

U)
01
•H
           Using the above" conditions and atmospheric pressure of 1010 mb,

           what is the effective stack height  calculated from the Holland

           equation for  neutral stability?
                         x*
                         (/.s
                                                                                  ui
                                                                                  U)
                                                                                  a) a)
                                                                              29

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
30

-------
   ALL Problem  Sheets and ANY additional calculations
   must be returned to APTI to receive credit.
f-J
O
•H
M
t-i
a
CO
•H

H
                                                           Problem Set One       (C)
       3.  Briggs has  published generalized plume rise  equations  which EPA

           is  incorporating into dispersion calculations  involving elevated

           emission sources.   A simplified working equation is  given by:


i—i
cd
>                                     1/3 _i   ?/3
g                         Ah  =  1.6 F ' u   x  '  ;   (x  <  10 h )
3                                                              s


•^                                     1/3-1        ?/3
S                         Ah  =  1.6 F ' u   (10 h )  '  •   (x > 10  h )
tu                                                 s                 s
                     Where:
                          Ah  =  plume rise  (m)
cu                                                AT       ?   4    3
•£                          F  =  buoyancy flux = —  gvsrz (m /sec )


•H                             =  £^(?'%'"(j^) /Ss*i&<>^ {s ?~: •'<-• 1<"  •  d

lu                          u  =  wind speed  (m/sec) •  '
4-J
0)
o                          x  =  downwind distance (m)
                          h   =  physical stack height  (m) - <> O
                           S


                          v_  =  exit velocity (m/sec)
                                                          Pu
                                                          •H
                                                          C-J
                                                        CO
                                                        CO
                                                        QJ QJ
"                                                     (U J_l 4-1
                             _                        g T3 W

r  =  stack radius  (m)     3-^> ^/^   / 7^-/r,             ^,~3\ ^
       For the power plant  in  Problems #1 and #2, assuming ambient air  temperature

       of 20°C, what is  the plume rise at:  a) 350 m and b) 1750 m downwind?
                                                                               31

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
32

-------
  ALL Problem Sheets  and  ANY  additional calculations

  -must be returned  to APTI to receive credit.
rt

o

0)
oo
M
rt


0)
0)
4J
m
M
o
M-l
M
•H


H
                                                          Problem Set One      (C)
      Discuss very  briefly whether or not this simplification of the Briggs

      equation  should  be used for this power plant.
                                                                                  en
                                                                                  w
                                                                                  m m
                                                                                cu
                                                                                g -a
                                                                                a T3
                                                                              33

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
34

-------
   ALL Problem Sheets and ANY additional calculations
   giust be returned  to APTI to receive credit.
V
e.
                                                                              F
              EFFECTIVE STACK HEIGHT  PROBLEMS
                                   SET TWO
      Submit all calculations leading to and including final answer(s)  to each
      problem.
      Using the Colbert  power plant data in Table  5.1, p. 44, of Plume Rise
      by Briggs, calculate  the expected plume rise under the following
      stability conditions:
          1.  Neutral and Unstable   u = 5 m/s,  Eq. 4.20 for F,  Eq.  5.1' for Ah
              a. Ah at  800 m                /^ *  3OO -F+  =  9V.VJ/5-,

                                          J =   /o'.SfV -  5 03 m
                                          ^
              b. Ah at  8000 m              ^j  A   iff q  ft'yic. -  /3. o^*
                   37x/o-£/r'»%*.3 7^
                            ^ ^-/>fec J P«
"         C3()^tre-   X *  '' ®


-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
48

-------
   ALL Problem Sheets  and ANY additional calculations

        be returned  to APTI to receive  credit.
rt
QJ
t

a)
a
10
•H
                                                             Problem Set Two       (F)
                                                           SQ          —2
       2.  Stable                  u = 2 m/s,  T = 280°K, — =  2  x 10
                                                           a Z
                                                -1/2
           a.   At what distance is x = 2.4  us    ;  why is  this  calculation

                important?
SJ          b.   Ah at 800 m

o
                                                        / ,-,   (--_>>, t>  C /" /* ,T r ' "f"  '«':>,-,
                   l^   T *i / J~  /7-y C-* / /n     -w=—,-,—*»-rw   -^   >   -•( o ^  ,)  /  ^
                                                            2-0 A
       2
                                                                                    49

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
50

-------
    ALL Problem Sheets and ANY additional calculations
    >r.ust be returned to APTI  to receive credit.
                                                            Problem Set Two       (F)
       3.  "No wind' Ah from  Eq.  4.25; Assume top  of  surface-based inversion at
           500 m.
tH
ed
to
rt
'rl
60
13
(U
o.
CO
(U
t>0
CO
•H
£1
                                     *
                                                                            5,00


                        t ft    r\*~f
                                                                                  51

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
52

-------
  ALL Problem Sheets and ANY additional calculations
  must be returned to APTI to receive credit.


              EFFECTIVE  STACK  HEIGHT PROBLEMS
                                                                 I
                                 SET THREE
       Submit  all  calculations leading to and including  final answer(s) to each
       problem.
       Refer to  Part  H of  this package, which reflects  current usage of the
       Briggs'  equations in the Meteorology Laboratory, EPA.
       Under Unstable  or Neutral Conditions:


           1.  Using the Colbert plant data,  what is  x*?
                                                                 •H
                                                                 NJ
60
            ,
           //
                                     >^}( i~c> 3 '",'
0)
Ml
ctf
to
•H
-77
                                        3.
                                                                  \
                                                                  }
                                                                 /
                                                                              0)

                                                                              I
                                                                             81

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
82

-------
   ALL Problem Sheets and ANY  additional calculations

   must be  returned to APTI  to receive credit.
                                                           Problem Set Three      (I)
       Under Unstable or  Neutral Conditions:

          2.  What  is the distance of the  final plume rise, xf?
ra
cfl
a)
•H
00
(-1
n)
B
a
•H

T3
a)
4-1
n)
M
o
M-l
l-l
01
a

                                                                                     CO
                                                                                     (n
                                                                                     QJ a)

bo                                                                                  B T3 a)
rt                                                                                  I n) "3 -"-1
                                                                                83

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
84

-------
    ALL Problem-sheets and AJM£ additional calculations
    must be returned to APTI to receive credit.
                                                           Problem Set Three      (I)
       Under Unstable  or Neutral Conditions:

            3.  What  is the plume rise  Ah,  that can be  expected a mile  (1500 m)
                from  the plant if the wind  speed is 5 m/sec?
 o

 0)
 M

 >.
 ra
                                                                                         o.
 0)                                                                                        -H
                                                                                         M
 H
 o
M-l

 C
•H
 60
 M
 oj
a
•H

T3
0)
4-J
B)
M
O
>W
M
0)

                                                                                       CO
                                                                                       a) a)
•H                                                                                    a) M -u
                                                                                     g x) n)
a)                                                                                    n) 73 *J
oo                                                                                    2; <; c/3
td
a.

to
•H

H
                                                                                   85

-------
USE THIS SIDE FOR CALCULATIONS BEFORE USING ADDITIONAL SHEETS
86

-------
      EFFECTIVE  STACK HEIGHT/PLUME RISE
                     STUDENT  CRITIQUE
PLEASE CIRCLE A RESPONSE.

 1.   How many hours  did you spend completing this package?

     <2    2-4    4-6    6-8    8-10  •  >10

 2.   In terms' of  coverage of the topic, how would you rate  this  package?

     Too Narrow       123-4-567    Too  Broad
Too Elementary
Too Little
Material
1
1
2
2
3
3
.4
4
5
5
6
6
7
7
Too Advanc
Too Much
Material
 3.   How would  you  rate  this package in terms of overall  value  to you?

     Not Worth         1     2    3    4 '   5    6    7.    Significantly
      the Time                                            Improve My Work

 4.   What responsibility do you have'-for calculating or reviewing effective
     stack height estimates?

     None at     Currently    Will Assume   None Planned
     Present     Involved     Shortly       (Within next two years)

 5.   Additional Comments:

     A.   Package Contents


     B.   Administrative Aspects (grading, Certificate,  etc.)


     C.   General Suggestions
                                                                          if,
                                                                         1 .'D 1)
                                                                        II i-i u
                                                                        e -a a
 A CRITIQUE  MUST  1TE  SUBMITTED BEFORE A CERTIFICATE CAN  j
 BE AWARDED  BY  THE AIR POLLUTION TRAINING INSTITUTE.    \
                                                                       89

-------