EPA-600/4-76-007
January 1976
Environmental Monitoring Series
                                SPECTRAL  MODELING OF
 ATMOSPHERIC FLOWS  AND  TURBULENT  DIFFUSION
                                    Environmental Sciences Research Laboratory
                                         Office of Research and Development
                                        U.S. Environmental Protection Agency
                                  Research Triangle Park, North Carolina  27711

-------
                     RESEARCH REPORTING SERIES
Research reports of the Office of Research and Development,
U.S. Environmental Protection Agency, have been grouped into
five series.  These five broad categories were established to
facilitate further development and application of environmental
technology.  Elimination of traditional grouping was consciously
planned to foster technology transfer and a maximum interface in
related fields.  The five series are:

          1.  Environmental Health Effects Research
          2.  Environmental Protection Technology
          3.  Ecological Research
          4.  Environmental Monitoring
          5.  Socioeconomic Environmental Studies

This report has been assigned to the ENVIRONMENTAL MONITORING series.
This series describes research conducted to develop new or improved
methods and instrumentation for the identification .and quantification
of environmental pollutants at the lowest conceivably significant
concentrations.  It also includes studies to determine the ambient
concentrations of pollutants in the environment and/or the variance
of pollutants as a function of time or meteorological  factors.
This document is available to the public through the National
Technical Information Service, Springfield, Virginia  22161.

-------
                                             EPA-600/4-76-007
                                             January  1976
   SPECTRAL MODELING OF ATMOSPHERIC FLOWS

           AND TURBULENT DIFFUSION
                    by
       Arthur Bass, Steven A. Orszag
    Flow Research, Inc.  (N.E. Division)
                 1 Broadway
            Cambridge, MA 02142
          Contract No. 68-02-1297
              Project Officer

             Kenneth L. Calder
    Meteorology and Assessment Division
 Environmental Sciences Research Laboratory
Research Triangle Park, North Carolina 27711
    U.S. ENVIRONMENTAL PROTECTION AGENCY
     OFFICE OF RESEARCH AND DEVELOPMENT
 ENVIRONMENTAL SCIENCES RESEARCH LABORATORY
RESEARCH TRIANGLE PARK, NORTH CAROLINA 27711

-------
                              DISCLAIMER
     This report has been reviewed  by the Environmental  Sciences Research
Laboratory, U.S. Environmental  Protection Agency,  and  approved for
publication.   Approval  does not signify that the contents  necessarily
reflect the views and policies  of the U.S.  Environmental  Protection
Agency, nor does mention of trade names or commercial  products constitute
endorsement or recommendation for use.
                                 11

-------
                           Preface








     In order to provide  a strong quantitative basis for air



quality management through selective emission controls, there



have been increasing demands in recent years for air quality



simulation models of considerable sophistication and complexity.



In these models a major component is that describing the atmos-



pheric transport and turbulent dispersion of the pollutants, and



many efforts are in progress to improve the specification of these



effects.  For the turbulent dispersion this is normally done through



the use of diffusion coefficients (eddy diffusivities) or by more



sophisticated turbulence  parameterizations that are developed



through various "closure" schemes for the governing equations of



flow.  These closure approximations are, of course, necessary



because of the well-known mathematical intractability of the



general non-linear time-dependent Navier-stokes equations that



govern the turbulent flow of a viscous fluid.  Closure approxi-



mations all involve the use, at some stage in their development,



of empirically based constants of parameters, that in some cases



may require ad hoc field  experiments or observations.



     During the past few  years, and largely stemming from the



pioneering efforts of Professor Steven A. Orszag and his associates,



a major breakthrough has  been made in developing powerful new mathe-



matical-computational approaches for the accurate numerical simu-




lation of turbulent flow.  It has been shown that many aspects of






                           -iii-

-------
 turbulent motion  may  be  numerically simulated with impressive




 accuracy, directly  from  the  equations  of  motion,  and without




 the  somewhat  arbitrary closure  approximations involved in normal




 parameterizations.   It was these  remarkable  results that stimulated




 the  present contract.  It was hoped that  it  might provide the ini-




 tial step towards providing  a basic technique for atmospheric tur-




 bulence  and dispersion estimates  by accurate non-empirical methods.




 In due  course  such  results might  provide  a set of diffusion data




 against  which  various approximations might be tested.




      The recent methods  of Professor S. A. Orszag involve so-called




 spectral techniques,  along with sophisticated computer programming.




 These techniques  require expansion  of  the dependent variable into




 a Fourier (or  other orthogonal) series of smooth  functions.   The




 calculation proceeds by performing  certain of the mathematical




 operations in  Fourier  (or other)  space and others in physical con-




 figuration space, according  to the  mathematical  form of the opera-




 tion  and its relative efficiency.   Transformations between the two




 spaces are made by Fast Fourier Transform (FFT) methods and the




 use  of special computational techniques.  In addition to having




 very powerful  applications to the problems of turbulent flow and




 dispersion, the spectral technique  has a  wide potential for numer-




 ical simulation of many atmospheric flows.   The present report




 attempts  to provide an overview of  such applications.   It also com-




pares the accuracy and advantages of spectral techniques,  with the




more conventional finite-difference numerical techniques for the




solution of partial differential  equations.   In some places  a tu-




torial-type presentation results, from the attempt to  provide ade-
                            -iv-

-------
quate  references  and to make the potential of spectral methods
more familiar  to  the developers of meteorological air quality
simulation models.  In addition, it provides an account of an
initial  research  attempt to develop direct solutions, without
parameterization  approximations, for turbulent dispersion of
an inert pollutant  in a thermally  stratified atmospheric layer
that is  governed  by the Boussinesq equation.  Unfortunately,
this research  failed to reach its desired goal during the rela-
tively short period of the contract, although in anticipation of
renewed  efforts along similar lines in the future, a fairly
detailed account  is desirable.
     As  an example  of the power of spectral techniques, some new
results  in turbulence dynamics are presented,  concerning the rate of
relaxation of  anisotropic flows to asymptotic isotropy.
     Finally,  general recommendations are made concerning the
potential utility of spectral methods in a number of specific
problem  areas, and  specific suggestions are given concerning the
prospective improvement of several existing air quality simulation
models, using  these spectral methods.
Environmental Portection Agency               Kenneth L. Calder
    Research Triangle Park                    Project Officer
      North Carolina
      October, 1975
                            -v-

-------
                          CONTENTS
                                                            Page

Title Page                                                    i

Disclaimer Page                                              ii

Preface                                                     iii

Contents                                                    vii

List of Figures                                              ix

List of Tables                                                x

Acknowledgements                                             xi

1.0  Introduction                                             1

2.0  Overview of Numerical Techniques for Meteorological
     Flow Simulations                                         3

2.1  Comparative Accuracy of Finite Difference and Finite
     Spectral Methods                                        H

3.0  Spectral Methods                                        25

3.1  Boundary Conditions and Geometries                      31

3.2  Treatment of Boundary Layers                            33

3.3  Pseudospectral Methods                                  39

4.0  Applications of Numerical Spectral Methods              45

5.0  Comparison of Methods of Spectral and Finite
     Difference Methods for Two Examples                     48

5.1  Comparison of Methods for the Diffusive Color
     Problem                                                 48

5.2  Comparison of Methods for Neuringer's Problem           65

6.0  Spectral Modeling of the Turbulent Diffusion of a
     Passive Scalar                                          77

6.1  The Numerical Model                                     79

6.2  Physical Parameters                                     86

6.3  An Unsuccessful Experiment                              88
                             -vi*-

-------
                        CONTENTS  (CONT'D)

                                                            Page


7.0  Relaxation of Anisotropic Turbulence to Isotropy       107

8.0  Recommendations for Future Research                    113

9.0  References                                             115

Appendix 1  The Data Content of Discrete-Grid
            Representations                                 118

Appendix 2  Discrete Fourier Transform Representations      121

Appendix 3  Elimination of Aliasing in Discrete Real-
            Space Multiplications                           124

Appendix 4  The Fast Fourier Transform                      132

Appendix 5  Construction of Initial Perturbation
            Flow Field                                      138
                               -vii-

-------
LIST OF FIGURES
Number
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11
6.1A
6. IB
6. 1C
6.2A
6.2B
6.2C
6. 3A
6.3B
6.3C

Initial Point-Source Distribution (t = 0) . . .
Finite-Difference Scheme, 32 x 32 point resolu-
Spectral Scheme, 32 x 32 point resolution,
Finite-Difference Scheme, 32 x 32 point resolu-
tion v = 05, t = 1/2 revolution 	
Spectral Scheme 32 x 32 point resolution,
v = .05, t ~ 1/2 revolution 	
Finite-Difference Scheme, 32 x 32 point resolu-
Spectral Scheme, 32 x 32 point resolution,
Finite-Difference Model, 32 x 32 point resolu-
Spectral Model, 32 x 32 point resolution,
v = .005, t = 1/2 revolution 	
Correct Neuringer FCN for y = 1 > 't = 0 . 5 . .
Neuringer Problem (y = 1.0, t = 0.5)
Analytic Value of F (x,y) /F (0 , 0) 	
U Field (y— z projection) ,t=0 	
U Field (y-z projection) , t = 10 sec ....
U Field (y-z projection) , t = 2 x 10 sec . .
U Field (x-z projection) , t = 0 	
U Field (x-z projection) , t = 10 sec ....
U Field (x-z projection) , t = 2 x 10 sec . .
U Field (x— y projection) ,t=0 	
U Field (x-y projection) , t = 10 sec . . - .
3
U Field (x-y projection) , t = 2 x 10 sec . .
Page
52
53
54
55
56
57
58
59
60
67
68
90
91
92
93
94
95
96
97
98
     -viii-

-------
                      LIST OF FIGURES  (CONT)

Number                                                          Page


 6.4A      V Field  (y-z projection), t = 0	     99

 6.4B      V Field  (y-z projection), t = 10  sec  ......    100

 6.4C      V Field  (y-z projection), t = 2 x 103  sec  ....    101

 6.5A      W Field  (y-z projection), t=0.........    102

 6.5B      W Field  (y-z projection) e t = 10  sec	    103

 6.5C      W Field  (y-z projection), t = 2 x 10   sec  ....    104

 7.1       Time History of R (t) for E,(0)  = 0 and En (0) =
             1/2E2(0)   .. .a.  ...............    110
                                -ix-

-------
                          LIST OF TABLES


Number                                                         Page

  2.1     Phase Errors for  a = 0.1	     13

  5A      Analytic Evaluation of F(x,y)/F(0,0)  for  y =  1.0,
            t = 0.5	     69

  5B      Analytic Value of F (x,y)/F (0,0), y =  4.0, t = 0.5      71

  5C      Absolute Errors in Spectral Run	     72

  5D      Absolute Errors in 48*48 Point Finite Difference
            Run	     73

  5E      Absolute Errors in 96*96 Point Finite Difference
            Run	     74


-------
                       Acknowledgements






     We acknowledge with thanks the constructive suggestions of




K. L. Calder, R.  Eskridge,  and K.  L.  Dermerjian, who have contri




buted greatly to the effectiveness of this report.  We are




especially greatful to K.  L.  Calder for suggesting to us the




spectral simulation of Neuringer's problem.
                             -xi-

-------
1.0  Introduction



     Advances in computer technology, together with important



achievements in the design of efficient numerical algorithms,



have begun to make practical the use of direct spectral methods



for numerical simulations of many fluid flow problems.  In a



spectral method, the dependence of the flow variables upon one



or more of the spatial coordinates is represented by expansions



in a discrete series of smooth orthogonal functions (Fourier



series expansions being the most familiar).  The governing



dynamic equations are then reformulated as equations relating



the spectral coefficients of the fields so analyzed.  To permit



numerical computation, the expansions are truncated to some



finite upper mode.  The resolution and accuracy properties of



spectral methods are very different from those of the more



familiar finite-difference methods, although both classes of



methods are related to discrete spatial grids.



     Although rapidly growing in popularity, these spectral



methods are as yet relatively unfamiliar to large segments of



the meteorological communities.  We have attempted, therefore,



to provide a brief review of some important finite spectral



methods for a variety of flow simulations relevant to air quality



modeling.   The report discusses at some length the relative ad-



vantages and disadvantages of finite-difference and finite spec-



tral techniques, especially in regard to ease of implementation,



accuracy,  and physical utility.  Implicit in this description



is  a clear understanding of the considerations which influence



the choice of a particular spectral approach, and determine its



range of utility.  We have included in Appendix 1, a discussion

-------
of the resolution implications of discrete-grid methods,  from




the point of view of both finite-difference and finite spectral




representations.



     Next, a review is made of the current state of spectral




modeling in a variety of geophysical fluid flow problem  areas




for which spectral methods have proved fruitful and/or hold




much prospective promise:   (a) numerical weather prediction,




 (b) turbulence modeling,  (c) methods for advective-diffusive




simulations.



     In the course of the current contract effort, we have




developed a spectral model for the direct simulation of  fully




three-dimensional, non-linear, time-dependent, turbulent,




incompressible flows, and a corresponding computer program




for the solution of the Navier-Stokes equations using high-




resolution Fourier spectral methods.  The model has been de-




signed to treat,  within the Boussinesq approximation, the influ-




ence of differing stratification profiles upon the spectral evolu-




tion of turbulent flow fields.  We have attempted, using this  pro-




gram, to study the evolution of a passive scalar field in a




sheared, turbulent velocity field, but unforeseen limitations




in computer core and running time precluded the attainment of




physically realistic results.  More successfully, the program




was used to look at an important problem in turbulence transport




theory, that is the rate at which an initially anistropic flow



will relax to isotropy; the relaxation time constant is  an impor-




tant term in some turbulence closure models.




     A two-dimensional version of these finite spectral  methods




has been used for the accurate simulation of two advective-diffu-



                             -2-

-------
sive point-source problems, which may have useful .implications

for air quality modeling efforts in which discrete point sources

are to be represented.  It is shown that, in the two cases,

finite spectral methods offer significant advantages, both with

regard to accuracy and computer storage requirements, compared

with finite-difference representations of the same problems.

     Finally, recommendations are made concerning the practioc'il

uses to which such finite-spectral methods can be put, in the

present state of air quality modeling, and in the light of more

realistic assessment of available computational resources.



2.0  Overview of Numerical Techniques for Meteorological Flow
     Simulations

     The historical development of numerical methods for the

simulation of meteorological and other fluid flows can be traced

back to the early part of the century.  Especially noteworthy is

the pioneering manual forecast effort of L.F. Richardson, during

the first World War.  Although unsuccessful, this effort provided

later generations of numericists with much insight into the limi-

tations imposed by small-scale phenomena which are meteorologically

irrelevant, but whose implicit presence in a numerical model leads

to computational catastrophe.  The impetus of technological pro-

gress during the second World War spurred greatly the development

of numerical methods for Shockwave and other supersonic model

calculations, and resulted in the first systematic examinations

of finite-difference numerical models for such calculations.

With the advent of high-speed calculational devices by the early

fifties, it became practical to attempt the first numerical weather

                              -3-

-------
predictions  (p-jortof t , Von  Neumann,  Charney),  using simple models



which  isolated,  by  suitable scaling  arguments,  the meteorologi-



cally-relevant  large  scales of  motion  and eliminated the small-



scale  structural  behavior which had  frustrated  Richardson's



efforts  thirty  years  before.  The  first  such  numerical  forecasts



were performed  by constructing  finite-difference  approximations



for the  evolution of  the two-dimensional inviscid barotropic



vorticity  equation
 (2.1)     (   + v-V) (c + f)  =  0





where  V  (x,t) •=  u(x,t)l  +  v(x,t)j   is  the  horizontal  velocity



field,  f  is the Coriolis  parameter,   and   r,   =  1Z _  —  is the
                                                  3x   9y


geostrophic vorticity.  Considerably more ambitious simulations



were soon attempted in a  variety of  fluid-mechanical contexts,



beginning from the viscous , incompressible  Navier-Stokes equa-



tions :





(2.2)     (3- + V-V)V = -  ¥1 +  WV2V
          3 1              p      —




(2. 3a)    V-V = 0





where  P(x,t)  is the pressure,  p   the  (constant)  density, and



v  the kinematic viscosity.  Using the  incompressibility condi-



tion (2.3a)  one obtains a Poisson equation  for the  pressure field,





(2.3b)    yV2P = - V- (V-V)V  .





Extensions of such models to slightly compressible  flows (in the



Boussinesq sense) provide the  needed coupling  of  flow  kinematics



and thermodynamics.



     It was  soon recognized that the proper formulation of  boundary


                              -4-

-------
conditions  is  among  the most  crucial  difficulties  encountered



in numerical flow  simulations.  Appropriate  boundary  conditions



may include   (a) the specification  of the  velocity V  at  no-slip

                                                <-\

boundaries;  (b) the  specification of   V    and  — V_   at no-stress
                    .,                  n       3 n  T


boundaries   (n and T are  respectively, the normal  and tangential



directions  to  the  boundary);  (c) the  specification of boundary-



layer momentum, flux  when  such fluxes  cannot  be computed explicitly;



 (d) thermal boundary conditions,  (e)  the choice  of inflow  and out-



flow boundary  conditions  for  systems  whose natural boundaries lie



outside the designated computational  boundaries,  (f)  the incorpora-



tion of topographic  influences  and  differences in  land-sea surface



properties.  The problems  associated  with  imposition  of boundary



conditions  remain, by and  large, the  principal impediments to



truly accurate numerical  flow evolution, and continue to be  areas



of intensive investigation.



     The most  widely applicable and popular  numerical methods



for flow simulations  are based  upon finite-difference approxi-



mations, in one or more geometries, to various versions or exten-



sions of (2.1  - 2.3); comprehensive reviews  of finite-difference



methods have been  provided by Fromm (1970),  Kreiss and Oliger



(1973), Roache  (1972), to name but  a  few.  Difference methods

                                         >

fall into several  general classes.  We may distinguish, first,



between those  difference schemes developed for Eulerian, and



those developed for  Lagrangian, versions of  the  equations  of



motion.   In the Lagrangian formulation, the  coordinate system



is embedded in the moving fluid.  For flows  which  are relatively



uncomplicated, and not subject to extreme  distortions,  such



methods can lead to  accurate  results,  while  accurately preserving


                              — 5—

-------
advection properties and real discontinuities.  However,  for  suffi-




ciently disorganized flows, the Lagrangian mesh becomes badly dis-




torted, and the calculational accuracy quickly deteriorates.




Lagrangian methods have been especially popular in the simulation




of strongly compressible flows.



     Eulerian systems, in which the coordinate grid is fixed,




are much less susceptible to violent computational breakdown  as




the fluid undergoes large distortion.  The penalty paid is  the




corresponding loss of accuracy in the representation of advection,




diffusion, and flow discontinuities.  We shall discuss some of




these difficulties shortly.



     Difference methods may be also distinguished with respect




to the choice of variables which are computed prognostically,




and those obtained diagnostically-  A variable is said to be




computed prognosfcically if the equation used to govern its




temporal evolution explicitly  includes its time-dependence.




For example, the velocity field in  (2.2) is a prognostic  variable.




A variable is said to be diagnostic when the equation used  to




govern its temporal evolution makes no explicit reference to  its




time dependence.   For example, the pressure field in  (2.3)  is a




diagnostic variable, since its change with time is entirely




implicit in the time variation of the velocity field.  These




choices are far from trivial to make and govern the physical




realism which will inhere to the computation.   For example,  in




two-dimensional versions of (2.1 - 2.3) one may solve directly




for the variables  V  and  P,  or alternatively, convert  to  the




vorticity-stream function formulation





                              -6-

-------
 (2.4)        +  J(4»,C)  =  W2?
where  i|>   is  the  stream function defined by
 (.2.5b)   c  =  V
-------
However,  I ho  vorticity-streamfunction  idea  is  still very useful




for  the  determination of boundary  conditions  in primitive vari-




able  simulations   (Orszag  and Israeli  1974).



      The  ability of  an  inviscid numerical sgheme to conserve




energy and enstrophy (square vorticity)  is  an  important consider-




ation from the point of view of physical realism,  and long-term




computational stability.   Let us consider briefly  how well such




properties are preserved by finite-difference  schemes.   If the




Jacobian  in  (2.4)  is evaluated using second-order  centered differ-




ences on  a uniform grid the resulting  scheme violates energy and




enstrophy conservation.  Therefore, Arakawa (1966,  1970) proposed




that  the  Jacobian  be approximated  by second-order  centered differ-




ences when it is first  written in  the  equivalent form






 (2.6)     J = ^[(CT|I )  -  (Sty )  + \\) l,   -  4> 5 + (\K )   - (iK )  ]
              3   Yx  y      y x   rx y    y  x       y x      x y





If  (2.6)  is approximated by second-order centered  differences,




the  discrete analogs of energy and enstrophy conservation






 (2. 7)     //*J dx  dy =  0;  {/ C  J dx dy = 0






respectively, are  preserved.  Arakawa1s  energy-conserving methods




have  been very popular  in  recent years,  apparently because it has




been  believed that energy  conservation is necessary to avoid non-




linear computational instability (Phillips  1959),  usually called




"aliasing" instability.  However,  energy-conserving finite-differ-




ence  schemes may not be as useful  as expected.   The Arakawa schemes




"semi-conserve"  energy; that is, they  conserve energy only in the




absence of time-differencing errors, but Kreiss &  Oliger (1972)



have given examples  of schemes that semi-conserve  energy but




that are  unstable  when  leapfrog time-differencing  is  used.   in

-------
general, though, Arakawa-type schemes  are  less  susceptible  to




numerical instability than straightforward centered schemes




(GrammeItvedt  1969).  They are, however, considerably  less  effi-




cient and yet  not noticeably more  accurate.   It is now understood




that aliasing  instability will appear  only in simulations with




inadequate grid resolution)  (Orszag  1971a,  1972, Fornberg 1972,




Kreiss & Oliger 1973),  and the deliberate  use of an energy-con-




serving scheme to avoid aliasing instability may lead  to a  grossly




inaccurate simulation.




     Among the most  attractive features of the  spectral methods




to be discussed is the  ease with which in  the inviscid case




energy and enstrophy  are preserved in  the  absense of time-differencing




errors.  This  is mainly due to the accuracy with which phase and




amplitude information is retained  using a  spectral form of  the




Jacobian term.




     An essential part  of the vorticity-streamfunction method is




the inversion  of the  Poisson equation  (2.5b).   In a finite  differ-




ence scheme, the inversion is performed by a  standard  relaxation




method.  In a  spectral  scheme, the inversion  is trivial, and




direct spectral methods, rather than relaxation methods, should




be used whenever possible  (Hockney 1970, Dorr 1970, Swartztrauber




1972).




     Primitive-variable methods are  increasingly widespread,




especially in three-dimensional NWP  simulations.  In the staggered-




mesh second-order scheme (Fromm 1963,  Lilly 1965, Harlow and Welch




1965,  Orszag 1969, Williams 1969,  Deardorff 1970), velocities are




defined at cell boundaries and pressures at cell centers.   It has




the attractive features of efficiency, small  truncation error and



                             -9-

-------
energy semiconservation.  The  Poisson  equation for the pressure




is most conveniently solved by  direct  spectral methods.   1-Iarlow




and Welch (1965) and Piacsek and Williams  (1970) have  developed




techniques to ensure compliance with the incompressibility con-




dition; these are especially useful when relaxation  methods are



used to solve (2.2, 2.3).
                            -10-

-------
 2 . 1   Comparative Accuracy of Finite Difference and Finite

      Spectral Methods



      In this section we compare finite difference and finite



 spectral methods with regard to several important classes of



 numerical error.  These may be identified as follows:





 (2.1.1)  First-Differencing Phase Errors



      Consider the usual approximation to the first derivative



 3A/9x'   where  A  is a scalar function of the spatial variable



 x :
 (2.8
                            2Ax
Suppose we  are  attempting  to  represent  in  this  manner  the  first


                                              i KX
derivative  of the  sinusoid function   A(x)  =  e   .   Then  the



ratio of  the approximation (2.8)  to  the exact result iK  A(x)



is given  by sin (KAx) /KAx-  This  ratio  approaches  unity



only in the limit   KAx->-0.   Note,  that for  large K and  finite



Ax, the approximation  (2.8) becomes  progressivly poorer.   In



particular  for  K-HT/AX ,  i.e.,  as  we  approach the limit of  spec-



tral resolution of  the finite  mesh,  the ratio approaches zero,



which is  hardly acceptable if  there  is  significant information



contained at wavenumbers near  the truncation  limit N/2.   Con-



sider now the effect of  such  first-differencing errors upon the



solution  of the one-dimensional advection  equation (N  is the



number of mesh intervals in the grid)
(2.9)       -- A(x,t) + U   A(x,t) = 0
                              -11-

-------
where   (J   is  a  constant,  uniform advecting velocity.   The solu-



 tions  to  (2.9)  can  be  synthesized by linear combinations of



waves  of  the  form






 (2.10)      A (x,t)  =  eiKn(x-Ut).
              n




 In  fact,  if the values of  A(x,t)   are given on  N  discrete



 points  equally  spaced  by   Ax,   an  arbitrary excitation on these



 N   points may be approximately  represented as a linear combina-



 tion of the   N   solutions (2.10)  with  KAx - 2rrm/N  (-^ f_ m< - ) ,



 m   integral.



     The  leapfrog centered "2Ax difference" approximation to



 (2.9)  is
 (2.11)                                 _





where  A. = A(jAx, nAt) ,   At   is  the  time  increment, and



a= UAt/Ax.  The  finite-difference equation (2.11)  is satisfied



by





 (2.12)      An = ei(KjAX -  n9)
             D





where  sin0 =otsin  (KAx).   If  |a|  <  1  then   6   is real for  all



K  and hence the scheme is  neutrally  stable.   In each time step,



the exact change of phase of   A(x,t)   given by (2.10) is  -KUAt



= -KaAx  while the change of  phase  of (2.12)  = -0 = - sin~



[asin KAx].   Since  |9/(KaAx)|
-------
     Let us compare the  ratio  of   
-------
     The values of  6  listed in Table 2.1 show that the fourth-


order scheme  has smaller phase errors than the second-order


scheme, and that the truncated Fourier scheme has phase errors


of less than  1%  for all waves satisfying  | KAx | <-y,  when
      It should be noticed that the second-order scheme has lagging


phase error  (6<1) , while the fourth-order scheme has mostly lagging


errors.   [There is a region of leading phase error   6>1  for very


small  K Ax   for the fourth-order scheme] .   The lagging phase


errors provide an explanation for the wakes of bad numbers given


by  finite-difference approximations  (see also the discussion in


Chapter 6) .


      The  truncated Fourier scheme has only leading phase errors


for  finite   At,  and even these disappear in the limit  At^-0 ,


if   KAx0


(if  the exact solution that is being approximated is infinitely


dif ferentiable) .   Finite-difference  approximations have non-zero


phase errors in the limit  At+0.





2.1.2     Second-differencing phase errors


     These errors are introduced in  evaluating the Laplacian

           2
operator  V   by the standard finite-difference operation


(2.13) (v2A)x     ,. A(x£+i'V   + A(*^ym) + A(x£--ym-i} + ^
            X£'Ym ~ ----- ' - - - ' - 2
                                              Ax




     Consider the two-dimensional sinusoid  A(x,y) = e   ix     2


Then the  ratio of the approximate expression (2.13) to the exact


                             -14-

-------
 result   -(kj   +  k22  )   is  just


                   •  2  klA*  .   .  2
 (2.14)        4    	__fL:	: ^  '
             Ax2            2  ,  .  2
                        kl   +  k2


which approaches  unity  only  in the  limit   k  Ax-»0   and   k  Ax->0.


Note again  that for  large  k]   or   k,,   the second-order finite


difference  scheme seriously  misrepresents  the  Laplacian operator

                                                           2
as, for example,  in  the evaluation  of  the  viscous  term  vV V.


Reduction of these errors  is possible  by use of a  fourth-order

                                     0
finite-difference approximation to  V V, due  to the higher order


of approximation,  but little benefit is gained.


     There  are no second-differencing  errors in the truncated-

                                                 2
Fourier expansion method,  as the transform of  vV  V is exactly

   p                              O
-vk tJ(fi).   The algebraic form   -vk  U(k)  makes it  a simple matter


to treat viscous  dissipation implicitly in time in the  Fourier


method.  However,  this  advantage is slight, because in  fully


turbulent flows,  it  is  the advective terms which are most impor-


tant, and the numerical simulations are thereby usually limited


by advective rather  than diffusive  stability criteria.


     It should be  noted here that,  for systems where the  pressure


field (or the streamfunction field) is recovered diagnostically


by solution  of a  Poisson equation (cf.  (2.3b)   or (2.5b), for


example), the second-differencing phase errors associated with


the representations  of  the Laplacian of these  fields are  consider-


ably more detrimental to the evolution of  the  system than those


introduced into the  viscous  terms.



                             -15-

-------
 2.1.3     I _n compressibility Errors

     The  imposition of the imcoinpressi bill ty condition is ,1

 trouble-some problem in many f inite-di f feronce schemes and com-

 putationally expensive iterative procedures are often required

 to maintain incompressibility to within a specified precision.

 Incompressibility can be imposed exactly in a truncated Fourier

 scheme  by the following simple procedure.

     Suppose that the three-dimensional velocity field is repre-

 sented  in Fourier space by  U.(k_) , where  U(k) is the three-dimen-

 sional  Fourier transform of the real-space velocity field  V(x) .

 The divergence of  U(k)  is given in Fourier space by  ik-U(k_).

 Let .us  subtract out vectorially the divergent component of  U(k)

 by replacing  U(k)   with


 (2.5)        U' (k)  = U(k)  - -(-'-(k)) .
                              k-k


 Note that if we now take the divergence of this entire expression

 (that is, by performing the operation ik-  upon both terms, we

 get the identity


             ik.u-(k)  = ik-u(k)  - JQi-kXk'lHiO)  =  o.
                                       (k-k)


Hence, incompressibility  is  maintained exactly if, at the con-

clusion of each time  step, we  replace  U(k_)   by  U' (k) .   We have

thus removed the divergent component of the  velocity field.  (Note

how trivially such a  constraint can be imposed in the Fourier

spectral scheme).  What is truly remarkable  about this  method

is that, if the Navier-Stokes  equations are  written in  rotational
                             -16-

-------
form,  i.e.

                                                1            i~\
 (2.16)        -— V(x,t)  -  V(x,t)* u>(x,t)- V(P + y V-V) + vV  V(x,t)
              dt —        — —     — —           ^   —


where   u  =  VxV   is  the  vorticity,  then this system may be  evolved

in  time without any reference to the pressure head term, V (P+^-  V-V)

so  long as  incompressibility is  imposed at each time step  in  the

manner indicated above.   This is because, in an incompressible

system, the pressure field serves  only to enforce incompressibility,

by  dynamic  readjustment of the velocity fields.
 2.1.4   Coupling  errors

      Among  the  crucial  difficulties with finite-difference approaches

 are  the coupling errors  introduced when non-linear differential opera-

 tors are  represented.  Consider again for example the simple two-

 dimensional  barotropic vorticity equation;


 (2.17)       ^L   v2(/;= _v-v (V2^)

                  ^\
             V = k  x Vip


 where   if;  is the geostrophic  streamfunction.   If we represent  the

 Fourier transform of  ^ (x)  by  0 (k)   as defined in Appendix  2,

 then it can  be shown  (Lilly  1965)  that the wave-space version

 of (2.17) is just
 (2.18)        -^ {k.k<|>(k)}   =   II*1  x *_"  (k"-k" - k1 -k1 KMk" )$ (kj1) dk
                                k'+k"=k .
For the continuous system,  the interaction between components

with wavenumbers  k'   and   k"   must vanish identically if either

k' is parallel to k  or   k1 |  =  k"|.   However, the Arakawa  J
                                                                _L

                               -17-

-------
algorithm  leads  instead to a non- vanishing,  spurious coupling



in this  case  (Lilly,  op.cit.)-   (Other finite-difference schemes



lead to  different  spurious coupling relationships.)   Moreover,



the misrepresentation of spectral  interactions occurs every-



where throughout the  spectral range,  and cannot be surmounted by



ignoring a particular subrange,  as discussed below.







2.1.5  Aliasing



     Another  fundamental difficulty with finite-difference tech-



niques is  "aliasing"  associated  with a finite grid interval.



The phenomenon of  aliasing is intrinsic to the representation



of non-linear interactions defined over a discrete physical-



space grid.   It  can be illustrated in the following simple manner.



Consider the  spatial  interval of length  L,   over which is defined



an N-ippirit discrete grid  • x.  ,  x.=jAx,   j = l?  ...  N  (L=NAX)



Let us now represent  the discrete  analogues  of the fundamental


                      2fTX
sinusoid  f (x) = sin  — — ,  and  the first two harmonics
                       LI

            4 71 }£                   6 TT 5C
g(x) = sin — —   and   h(x)  = sin  — - — .   The discrete version of
            LI                   LI


the fundamental  sinusoid,  defined  over this  grid,  is just



f,(x_;) = sin  — ^-  and in like manner   g , (x . )  = sin
  ,_;                                   ,   .
 d   j         N                         a   ]          N


h , (x . ) = sin  7r-'  .  Consider  now  the  discrete-grid representation
 d  j         N


of the two product  fields





             P(x) = f (x)g(x)  = i(cos  lj£  -  cos
             Q(x) = g(x)h(x) = |(cos      -  cos
We might then suppose that
             P  (x  ) = -Tcos ^Li -  cos
                ^   ;     ^
                             N          N
                              -18-

-------
and that
     To make matters concrete, choose  N=8.   If we  were  to tabulate


P  (x)  and  Q, (x) ,  we would  find them to be  numerically identical


over the discrete-grid  j = l,  ... 8,  whereas  the  continuous-space


functions certainly differ with regard to the  terms  cos -j— ,


cos 107TX .  what has occurred?  Trivially, of  course, we have  im-
      L

posed an identity, since, for  N=8, cos -^- =  cos ~~-  •   But


the physical consequence of this little example is  that,  over


a  discrete grid of length eight points, the sixth harmonic of


the fundamental spatial waveform is indistinguishable from the


second spatial harmonic, and  that, in fact, the information


contained in the sixth harmonic, if it occurs  as  a  result of


non-linear interaction among, say, two lower modes, would be


misrepresented by, or aliased down upon, the  second harmonic.


In like manner, the seventh harmonic would appear aliased as


the first harmonic, and so on.  We have illustrated the  notion


of a "folding" frequency (the Nyquist frequency N/2)  associated


with a discrete spatial grid  of finite length  N points.


     Let us examine the aliasing problem in more  formal  mathe-


matical manner.  The values of the arbitrary  functions   f(x),  g (x)


at  N  equally spaced grid points  xn = -— -  (n=0 , 1,  ...  N-l)  may


be expanded in the discrete Fourier series




            f(x ) = f  -   I   U(k)elkXn

(2.19)          n     n   |-k.|
-------
where   k   is  an integer satisfying  -K^k
-------
approximation to the Navier-Stokes  equations  may  lead to difficulty,




because of aliasing errors in the discrete-grid approximation of




the  (scalar) product v-V v-  Phillips  (1959)  gives  a simple example




where aliasing terms in the representation  of a product induce an




instability not present without aliasing.   This aliasing property




of  finite-difference representations has  several  other consequences




in  the numerical simulation of non-linear fluid flows.   We must




recognize,- first, that  the physical process of non-linear turbulent




spectral cascade of energy  (or vorticity) to  small  scales is falsi-




fied, inasmuch as the folding frequency acts,  inthe absense of




dissipation, as a reflective boundary  in  wavespace.   That is,




since no modes above the Nyquist frequency  exist  to receive energy,




the energy transferred  by non-linear interaction  must be redistri-




buted among the available modes which  lie below the Nyquist cutoff,




in  order that energy remain a conservative  quantity.   [This is




true, as well, of vorticity, or enstrophy].   Accordingly, it might




suggest itself to ignore the small-scale  structure,  in the expec-




tation that the largest scales are  at  least adequately represented.




But, in fact, all the scales of the system  are contaminated by




downward-aliased spectral information.




     One may ask, therefore, why conventional finite-difference




methods are reasonably  successful for  certain classes of geophysica]




fluid flow simulations, most notably,  for example,  in large-scale




operational numerical weather prediction.   The answer is rather




subtle, and concerns the spectral structure of the  flow fields




being simulated.   If the spectrum falls off sufficiently rapidly




with increasing wavenumber, then the aliasing-induced contamination




of the lowest modes, while present, may be  so small as to be ignorabl




                             -21-

-------
It is precisely this property of  large, geostrophic-scale atmos-




pheric flow systems  (which seem to  follow minus-third or minus-



fourth spectral power  laws) that  permits physically credible



numerical weather  forecasts on these scales.  In general, how-



ever, the spectral power  law of the simulated physical system does



not  fall off with  sufficient rapidity, so as to permit the ignoring



of aliasing contaminations.  For  such problems, therefore, it is



especially important that the numerical technique employed can



eliminate aliasing errors or strongly limit their extent, if



such errors are deemed important.  Appendix 3 presents the method



developed by Orszag  for efficient alias-free multiplication of



fields defined over  discrete spatial grids.








2.1.6  Treatment of  Singularities and Local Discontinuities



     Among the most  troublesome problems encountered in numerical



flow simulation is the  treatment  of flow structures with singular-



ities and local discontinuities embedded in such flows, as for



example, point sources  of pollution.  Eulerian finite-difference



methods permit the numerical evolution of strongly discontinuous



flows, but only at the  cost of smearing out, or smoothing over,



these discontinuities,  often in an entirely unrealistic manner.



These effects have been demonstrated by Orszag (1971b),who compared



second and fourth order finite-difference methods with truncated



spectral methods, for  the problem of the two-dimensional advection




of a passive scalar field, whose  initial distribution is circular



(hence referred to as  a scalar "cone")/ by a uniform rotation



velocity field.  (An extension of this study to viscous flows



is included in Chapter  5 of the present report.)  One of the most





                             -22-

-------
remarkable features of the truncated spectral method that  is




demonstrated by this problem is the result that the cone is




better localized in space, using these truncated spectral  repre-




sentations, than by finite-difference schemes formulated directly




in physical space.  That is, the cone location is preserved more




accurately by expanding the initial excitation in discrete Fourier




series, permitting these Fourier coefficients to evolve in time




in accordance with the spectral form of the governing dynamical




equations, and finally, reconstructing the discrete real-space




field from the Fourier coefficients, than by a direct evolution




of the real-space field using a finite-difference method.  In




other words, flow structure with intermittencies in physical




space seem to be more accurately followed in terms of their Fourier




components than directly in physical space.  To explain this rather




counter-intuitive result, we discuss in Appendix 1 the data content




of discrete real-space representations of continuous spatial func-




tions.  The principal consequence, in the present context, is that




spectral representations  (as by Fourier expansions), allows much




better interpolation between discrete real-space grid points than




is permitted by the structural data preserved by the local grid-




point values alone.  That is, the global information preserved in




the spectral decomposition of a spatial function permits more pre-




cise reconstruction of local details than does the grid-point




decomposition of corresponding resolution.  This may be better




understood by reference to the previous discussion of the  phase




and aliasing errors which underlie finite-difference methods; in




particular the point made that finite-difference schemes have




mostly lagging phase errors, so that some of the data needed to




reconstruct a spatially local disturbance at a given instant in




                             -23-

-------
time is erroneously phase-lagged and so is as yet unavailable




at the appropriate local spatial region.
                              -24-

-------
3. 0  Spectral Methods



     The utility of spectral transformations in the analytic




solution of many partial differential equations of classical




physics has been long recognized.  It is only relatively recently,




however, that advances in available computational power and algor-




ithmic breakthroughs have made possible the wide use of numerical




spectral methods for the solution of PDE's, in particular, for




the solution of a variety of geophysical and other fluid flow




problems.



     The essential principles which govern the use of spectral




methods in numerical simulations may be stated quite simply;




instead of solving multidimensional finite-difference equations,




one applies semianalytic methods in which the dependence of the




flow on one or more independent variables (generally, but not




exclusively, the spatial dimensions) is expanded into a series




of smooth, and usually orthogonal, functions.  The potential




choice of orthogonal expansion functions is quite broad; e.g.,




Fourier, Chebyshev, Lengendre, and Hermite series, among others,




are possible candidates in various different flow problems.  The




specific choice made depends upon the following kind of consider-



ations :




     (a) the nature of the boundary conditions to be imposed,




         and the suitability of the expansion functions chosen




         in regard to the imposition of these boundary conditions;






     (b) the ease with which the governing differential equation




         can be formulated and numerically evolved with time, in




         terms of the expansion functions;
                             -25-

-------
      (c). the  relative  computational effort involved in the



          numerical  evaluation of the expansion coefficients;





      (d)  the  accuracy  achieved in the expansion,  in particular,



          the  rate of convergence of the expansion coefficients;





      (e)  the  relative  generalizability which a given class of



          orthogonal polynomials offers in a variety of flow



          problems.







      A  few  simple examples  will suffice to illustrate these



points.   For  example,  in  flow in cylindrical geometries the



velocity field  must be periodic with respect to the zonal coor-



dinate      so that  the dependence on  cj>  is expandible in a



Fourier series



                          oo

          V (r, 
-------
(3.1)           +     =  0          (
             d t   d X




with the following mixed initial-boundary value conditions:





(3.2a)       U(x,0) = 0




(3.2b)       U(-l,t) = sin irni t.




The  exact solution is just




         U(x,t) = 0;  (tx+l).




(The index  m  designates the number of full wavelengths within



the  interval   |x|<
-------
 the  boundary conditions.   Chebyshev polynomials (Orszag 1971d)

 are  particularly appropriate.   The nth-degree Chebyshev polynomial

 Tn(x)  is  defined by  T (cos8)=cos n9 .   If  u(x)  has  p  piecewise

 continuous  derivatives on the  interval [-1, 1] , then  u(x)   may be

 expanded  as (Fox and Parker 1968, Orszag 1971e)
                      00
 (3.3)         u(x)  =  I  AnT (x)
                     n=0  n n
 (3.4)
 In  particular,  if  u(x)   is infinitely dif ferentiable ,  then the

 Chebyshev coefficients  go to zero faster than any finite power

 of   1/n ;   we  say that the approximation obtained by truncating

 (3.4)  at   N   terms  is of infinite order,  since the error goes

 to  zero faster  than any  finite  power of  1/N.   Also,  the latter

 Chebyshev series may be  differentiated termwise any number of

 times.  Notice  that the  convergence  rate of the expansions is

 governed  by the smoothness  of the function  u(x),  irrespective

 of  boundary conditions.

     At present,  there are  three  principal classes of discrete-

 spectral  methods  for the representation of continuous systems

 such as (3.1).   These are the Galerkin approach (Collatz,  1960),

 the tau method  (Lanczos,  1956), and  the collocation method

 (Collatz,   1960,  Lanczos  1956).  The  three methods can be use-

 fully -compared  in regard to  equations  (3.1-3.2).   The Galerkin

methods employ  a new dependent  variable  v (x, t) =u (x, t) -sin mTrt,

so that   v(x,t)  satisfies  the  homogeneous  boundary condition

v(-l,t)=0  and  the  inhomogeneous  wave  equations,  and equations


                              -28-

-------
for the time-dependent  coefficients  are obtained by requiring

that an appropriate  inner  product  of the dynamical equation

(3.1) with each of the  retained  Tn(x)   be zero.  When rewritten

in terms of  u(x,t),  the  semidiscrete  Chebyshev-Galerkin equa-

tions become  (Orszag 1971d)


                        N
(3.5)        u(x,t)  =   I   a  (t)Tn(x)
                      n=0

                da         N
(3.6)        c  -J-P-  = -2   V   pa  + (-l)nb(t)      (0l) .  Here  b(t)   is  a "boundary" term

that is determined by compliance with (3.7).

     The tau method  is  based on using the  finite expansion

(3.5) to solve exactly  the approximate  differential equation

u +u  = T  (t)T  (x).  The resulting equations  for  a (t)  are
 L  x    j i    n                                     ti

precisely  (3.6) and  (3.7)  but with the  boundary term  b(t) set

to zero for  n=0,  ...,  N-l and  b(t) = TM(t)   for  n=N.

     Finally, collocation  uses the spectral representation (3.5)

to interpolate values of u(x,t)  at   (N+l)  "selected points"

x  (p=0, ..., N) .  With  Chebyshev polynomials,  it is most natural

to choose  x =cosTp/N,  so  that  (3.5)  is just  a discrete Fourier

transform.   The polynomial interpolation (3.5)  is used merely to

evaluate  3u/9x  at  the points  x    by  termwise differentiation.

In terms of  a , collocation gives (Orszag 1972) equations (3.5-3.7)

with  b(t)   replaced by b(t)/C    ;  the only  change is in  (3.6)
                             -29-

-------
for  n=N  where   (-l)Nb(t)  is replaced by  1/2 (-l)Nb(t).




Collocation  (called pseudospectral approximation in a later




section) is particularly convenient when there are no-neonstant




coefficients or nonlinear terms.  In these latter cases, pseudo-




spectral approximation is quite different from, and much simpler




than, spectral  (Galerkin or tau) approximation because of the




appearance of aliasing terms in pseudospectral approximation.




     Notice that  solution of 3.6, 3.7 requires only  0(N)




arithmetic operations per time step if care is taken to accumu-




late required sums, and that the transformation  (3.5) to the




"physical space"  u(x,t)  can be accomplished in  0(N log N)




operations when   N  is a power of  2,  because  (3.5) is just a




Fourier cosine series for which fast Fourier transforms are




applicable.
                              -30-

-------
3 . 1  Boundary Conditions  and Geometries;







3.1.1  Periodic Boundary  Conditions



     Physical intuition leads us  to expect  that  for planar



geometries  and/or periodic boundary conditions,  the use of



discrete Fourier series is appropriate.   In this  case,  our



intuition is indeed  correct; finite Fourier series  are  the



most natural and efficient expansion method.







3.1.2  Free-slip Boundary Conditions



     Free-slip--free-slip boundary conditions  such  as





          (a) U =V =0       z=0, IT
              z  z



          (b) W=0           Z=O,TT





are mathematically equivalent to  symmetry boundary  conditions,



inasmuch as conditions  (a) are most naturally  implemented in



cosine series  and conditions  (b) in sine series  representations



of   (U,V),  and  W   respectively.  Hence  Fourier-series expansions



are the appropriate  choice for free-slip--free-slip boundaries.



For mixed f ree-slip--np_-slip conditions,  the use  of Fourier



series would, in general, be ill-advised.







3.1.3  No-Slip Boundary Conditions



     By contrast with free-slip conditions,  no-slip boundary



conditions impose stringent constraints upon the  dynamical fields,



which,  in general do not  lend themselves  readily  to computationally



efficient or accurate implementation in Fourier-series  expansion





                             -31-

-------
due to Gibbs phenomena  at  the boundaries.   Rather,  as has  been




demonstrated by  Orszag  (1972) Chebyshev  expansion is the tech-




nique of  choice, principally because  the rate  of convergence  of



such .an expansion  can be specified  a  priori, without reference



to the boundary  conditions, and  because  Chebyshev methods  are



efficiently implemented by skillful use  of  Fast Fourier  (cosine)



transforms.  Chebyshev  series are not restricted to planar geo-



metries;  they may  profitably be  applied,  for example, to handle



the radial expansion in flows in cylindrical or spherical  geo-



metries,  or in any geometry that can  be  smoothly mapped, by con-



venient transformation, into either rectangular or  circular



domains.  Metcalfe (1973) has employed Chebyshev expansions in



the analysis of  a  variety of viscous  incompressible flows  in



various cylindrical and spherical geometries.








3.1.4  Inflow and  Outflow Boundary Conditions



     Spectral methods are capable of  high-order accuracy at



inflow and outflow boundaries, for example, as in a limited



forecast  area model (LFM), without the need for special methods



to impose the boundary  conditions.  By contrast, high-order



finite-difference  methods deteriorate, or may break down com-




pletely,  near boundaries because grid points outside the physical



domain must be invoked.  The necessary modifications to maintain



accuracy near the  boundary can get quite complicated (Kreiss  and




Oliger 1973).
                            -32-

-------
3.2  Treatment of Boundary^Layers



     Among the most valuable properties  of Galerkin methods



using Chebyshev expansions, is  the  much  greater fidelity with



which boundary layer structure  can  be  resolved using Chebyshev



expansions, compared to a conventional finite-difference repre-



sentation with comparable number of grid intervals.  This follows



naturally from the definition of the Chebyshev polynomials, namely:




 (3.8)        T  (0) = cos n 8 ,  0<8 0

i e = .
n


1, nN
In particular, for  n=N-2, we have
(3.15)
              _  = 4N(N-l)a
                           N
                             -35-

-------
 Substituting (3.13)  and (3.15)  into (3.14)  we may eliminate



 the  terms   b ,   and obtain a linear tridiagonal matrix system



 of the  form
where   a   is  the  vector
                          L a
                           'NJ
                                                   rao
a  is the vector
This  system may be inverted by any of a number of standard



matrix  inversion algorithms, the standard LUD algorithm being



especially  useful.
      Suppose   Up(z)   is  of the form  1-
                                         cos  26
               (z=cos
Then  the  only  non-zero modes  in the  Chebyshev representation



of  this function  are   a =1,   and  a  =-1/2.
                        u           £


      Let  us  now compare the Chebyshev solutions  to  (3.12)



(using (3.13)  - (3.15))  as we vary parameters  y  and  N,



the number of  retained modes.   As  before  we  designate the  Cheby-
shev modes of   IF   by   a   (n=0„  ...,  N),   and  the  modes  of   U



will be designated  by   a  :
                                                             p+1
Case 1: Y =10 '

n=0
2
4
6
8
10
12

j
N=8
.9385
-.6093
-.0827
-.0451
-.2013



a
N=16
.927039
-.63308
-.10976
-.07966
-.05092
-.02863
-.01398
-36-

N=32
.927056
-. 63306
-.10976
-.07971
-.05102
-.02885
-.01443

                                                                n






                                                              1.0000



                                                              -.5000



                                                                0

-------
 14


 16


 18


 20


 22


 24


 26


 28


 30


 32
N=16


00538


00561
  N=32


-.00641


-. 00253


-8.9 xlO


-2. 8 xlO


-8. 0 xlO


-2.0 xlO


-4.7 xlO


-9. 8 xlO


-1.8 xlO


-4.6 xlO
                     -4
                     -4
                     -5
                     -5
                     -6
                     -7
                     -7
                     -8
Note that for  N=8  retained modes, no convergence   (a +0 for n->-™)


is apparent.  In the case  N=16, there is  just  marginal conver-


gence, but for  N-32  modes the convergence  is  extremely strong.


This indicates that there is just marginally enough resolution,


using 16 modes, to resolve the boundary  layer whose characteristic


thickness is  6 ~—  ~ .03,  but that merely by doubling the number
                /v

of modes, the resolution is about as perfect as we  could ever


hope to achieve (another doubling would  give no better resolution),
Case 2:  Y =10
              -2
                                n

n=0
2
4
6
8
N=8
. 857279
-.698663
-. 106009
-. 036378
-. 016238
N=16
. 857280
--698875
-.106911
--0391647
-.0101173
                                                 N=32


                                                 same as N=16
                            -37-

-------
                  N=8             N=16           N=32


  10                            -.00190785       -.00190785




  12                           -2.71303 xlO~4    -2.71333 xlO~4



  14                           -2.966 xlO~5      -2.991 xlO~5



  16                           -3.089 xlO~6      -2.619 xlO~6



  18                                             -1.86 xlO~7



  20                                             -1.09 xlO~8



  22                                             -5 xlO~10



  24                                             -2 xlO'11



  26                                             -5 xlO~13



  28                                             -2 xlO~14



  30                                             -6 xlO~16



  32                                             -2 xlO~17
 In  this  case,  for the characteristic boundary layer thickness



 6 ~  .1,   N=8   modes  is just marginally adequate,  while 16 modes



 already  displays  rapid convergence,  and an additional doubling



 gives  essentially nothing more.



     Orszag has  found an approximate'rule which relates the



 boundary  layer thickness implicit in the computation,  to the



 number of modes just necessary to adequately  resolve such



 a layer:



          ..     ^  ^-1/2

          NCRIT   36




 Note that for  Case 1,  NCRIT ~ 16.8>,.  .and for Case2f   NCRIT ~  9.5,



which is  entirely consistent with the  results shown.



     Stated in more  familiar terms,  the number of Chebyshev



modes  required for an accurate spectral simulation of a laminar


                                                        1/4
 flow with a viscous  boundary layer increases  only as  R '   (where





                             -38-

-------
R is the Reynolds number of the  flow),  whereas  the resolution


requirement of a finite difference model  varies as  R


For the high Reynolds numbers characteristic of realistic mete-


orological flow problems, the computer  storage  advantage of


Chebyshev methods is enormous.   Such methods thus offer the


potential of extremely accurate  numerical simulation of boundary



layer structure.


     If, in conjunction with these Chebyshev methods, appropriate


coordinate-stretching techniques  (Israeli 1971, Ph.D. Thesis) are


employed to assist in the resolution of a boundary layer of thick-


ness  S ,  then the spectral errors are  decreased faster than any


finite power of  6  as  S+0,  whereas the errors in finite differ-


ence methods, due to such coordinate transformations, decrease by


finite power of  6 .






3.3  Pseudospectral Methods


     Our emphasis thus far has been focused  upon the purely spec-


tral (Galerkin) methods.  We have seen  that, within limitations


imposed by geometry and boundary conditions, these methods cire verv


powerful.  We have suggested also that  Galerkin methods are rela-


tively easy to implement for certain important  classes of non-linear


interactions, as for example the rotational  term V x U) j_n the Wavier


Stokes equations.  The accuracy with which phase information is pre-


served, and with which aliasing  can be  controlled, has been discussed


and the elimination of aliasing  is described in detail in Appendix  3,
(1) in a finite-difference representation  of vj  ,  say

 v                                               zz               -I/?1
—j~ (U. ,  - 2U . + u . _, )  i.t is evident  that   Az  must vary wilh N

 z                        1/2
   or eq;.ivalently with  R




                              -39-

-------
     However, there are many  cases  of  interest  for which aliasing

errors can be expected to be  small,  and/or  tolerable, and we would

not wish to pay the computational price  required  to eliminate  alias-

ing in such cases.  Secondly,  it is  sometimes necessary to evaluate

highly non-linear  terms within a spectral framework, and such  terms

would be extremely difficult  to formulate directly in wavespace.

Thirdly, if we wish to represent physical processes that are local

in real space  (such as, for example, the onset  of condensation in

a numerical model  of  cloud dynamics) such a  local-space condition

would be extremely awkward to  formulate  in  a Galerkin framework

because a single point in real space is  represented in wavespaco

as a superposition of many modes, and  a  product term in real space

corresponds to a convolution  sum in  wavespace.  Accordingly, Orszag

has introduced a modified Galerkin-like  approach, which he has termed

"pseudospectral."  This approach is  computationally very economical,

retains much of the power of  the purely  spectral  Galerkin method,

yet permits us to handle more  general  non-linear  terms, complicated

boundary  conditions, and processes  which   are   local   in

real space.  The idea behind  the pseudospectral method is to repre-

sent certain terms in physical  space when it would be computationally

prohibitive to do so spectrally, but to  switch  back to wavespace

when it is very important, and computationally  easy, to preserve

accurate phase and amplitude  information.

     Consider, for example, the  temperature equation written  in

an incompressible Boussinesq  system


(3.16)          ^~  f£ = ~ ^~ V-(V T)  -  N2(z)W  +  -3-  KV2T
               Too  at     Too   ~                oo
                              -40-

-------
where  T  is the variable  temperature  field,  TQO is a standard


reference temperature, V is  the  three-dimensional velocity vector,


W  is the vertical velocity  component,   N(z)   is the Brunt-VaLsala


frequency, which depends upon  altitude   z,   g  is  the acceleration


of gravity,   K  is the  (constant)  thermal  diffusion coefficient.


An appropriate pseudospectral  procedure  for evaluation of the


right hand side might be as  follows:


      (a )  Represent each  component  of   V,   and the temperature


           field  T,  in real  apace, and form the vector quantity


            (VT) at each physical  grid point,  that is


           VT = U(x,y,z)T(x,y,z)i  +  V(x,y,z)T(x,y,z)j
                                 /-\
              + W (x , y , z ) T (x , y , z ) k ;



      (a )  transform the vector  quantity (VT)  to a vector term


           in wavespace, denoted by   (VT).



      (a.,)  Evaluate the divergence of the  quantity directly in


           wavespace by multiplying  by the  wavespace divergence


           operator  ik-  (where  k_ is the  wave vector) ;  this has


           the virtue of preserving  the  exact phase relationships


           in the divergence expression, without the first-differ-


           encing errors introduced  by a finite-difference approxi-


           mation to the divergence  term in real space.


      (b)   Assuming an explicit  evaluation of the thermal diffusion


           term is adequate, transform   T   to wavespace, denoting


           its transform by  T,   and evaluate the Laplacian term


           directly in wavespace  by  multiplying  T  by the scalar

                               2
           operator - k -k  =  -k  .   This procedure avoids the second-


           differencing errors  associated  with the finite-difference


           form of the Laplacian  operator.  [The diffusion term  can

-------
            be treated implicitly in a spectral formulation very


            simply, if, for example, (3.16) is written in the


            transformed form
                                             (VT) -  (N2(z)W),

                                    T
                  00                 00



            where, as, before, the tilda  (~ )  designates the wave-



            space transform of the real -space field.]





      (c)    Represent the vertical velocity  W  in real space, and



            at each point of the physical grid form the product


             2                                       2
            N (z)W;  then transform the expression  N (z)W  to



            wavespace.  By first evaluating the product in real



            space, we avoid a nasty convolution sum, but accept



            the inevitable aliasing penalty,  which may not be a



            severe limitation, especially if  the term is small



            or if  N  is a slowly varying function of  z.





      (d)    Since each term on the right hand side of (3.16)  is



            now available in its spectral form,  we may evaluate

                                            j\ m

            each of the spectral components TTT.    Finally,  then,
                                            d t

                                  r>"T
            we inverse transform  v~-  to real space to obtain
                                  o t


            the  local rate of change of temperature.
     Thus,  the  pseudospectral method makes effective use of finite



spectral operations when  it  is convenient and/or crucial to do so,


but returns to  finite  real space  collocation operations  when the


equivalent  spectral operation is  computatively  prohibitive,  or


aliasing errors are tolerable.  A more  elegant  statement of the


method, which has perhaps been adequately motivated by the example




                               -42--

-------
above, has been  provided by  Orszag  (1971a).   That is, the



philosophy of  the  pseudospectra.l  method is to calculate in



either spectral  or physical  space according t.o whichever repre-



sentation is more  natural.   In  particular, one chooses a suitable



orthogonal expansion  as  an interpolatory tool, in order to be



able to differentiate without phase  error.  Derivatives are evalu-



ated entirely  by explicit multiplications in spectral space  (rather



than by finite differences in real space).  Operations like differ-



entiations in  finite-difference schemes and convolutions in spec-



tral schemes are avoided.



     Let us apply  the pseudospectral  idea for efficient evaluation



of the non- linear  terms  of the  equations of motion.   The principal



error of finite-difference schemes is first-differencing or phase



error  (cf. section 2.1).  These errors  are avoided so long as spec-



tral methods are used to compute  derivatives.   For example,  9v/9x



at  x=x . ,  where  x. = 2Trj/N  (j = l,  ..., N) ,  may be computed without



phase error by first Fourier analyzing   v-=v(x.)   into its discrete



Fourier components  v(k)  (|k| < N/2)  and then setting
A term like  v9v/3x  in  an equation  of  motion can then be evaluated



at the grid point  x.  as the  local  physical-space product of  v(x-)



with   Ov/9x)    .    If derivatives are  computed in this manner,
             .A. ^C «


the resulting scheme is not subject  to  phase  error, but it is "fully



aliased" in the sense that spectral  terms modulo  +2k  appear in



the spectral analysis of the scheme  (see  section 2.1.5).
                              -43-

-------
      If  a pseudospectral approximation is  made to the Navier-

 Stokes equations  written in the form (2.2)  by evaluating deriva-

 tives as  in  (3.18),  the resulting fully-aliased equations may  be  sub-

 ject  to  unconditional instability (Orszag,  1971b).   However, if

 the Navier-Stokes equations are first expressed in  rotation form




 (3-19)          j£ V(x,t)  = V(x,t)x
-------
4.0  Applications of Numerical Spectral Methods




     In this section we discuss briefly some  of  the  principal




areas of fluid research, relevant to meteorological  problems,




in which numerical spectral methods have  been found  especially




fruitful, namely:



     (a) the simulation of atmospheric dynamics;




     (b) the simulation of turbulence and tests  of turbulence




         theories;



     (c) advective-diffusive modeling.



     Recent years have witnessed an avalanche of papers concerning




spectral simulations of various problems  in atmospheric flows and




numerical weather prediction.  For example, Eliasen, et al. (1970),




have considered spectral harmonic methods for integration of the




primitive equations on the sphere; Merilees  (1973, 1974)  has de-




scribed pseudospectral methods for such problems;  Bourke (1972),




Gordon and Stern  (1974) have described experiments with multilevel




baroclinic global general circulation models  constructed in spec-




tral manner.  An especially important spectral model  is the strato-




spheric dynamics model of the MIT group  (Cunnold, et al.  1975) which




incorporates elements of both the purely  Galerkin and the pseudo-




spectral approaches for the simulation of the dispersion of chemical




pollutants in the upper atmosphere.  Many of  the recenb advances




in the field of atmospheric spectral modeling are described in




the Copenhagen symposium volume  (GARP, 1974), to which the reader's




attention is directed.




     The employment of spectral methods for direct numerical simula




Lion of turbulent flows is of very recent origin.   Spectral simula-
                              -45-

-------
tions of this  sort, which  require  explicit  retention  of  1-3x10




modes, have become practical  only  with  the  introduction  of  effi-




cient transform techniques (Orszag 1969,  Patterson  and Orszag




1971).  Simulations have been performed for homogeneous  two-




dimensional turbulence  (Herring,et al.  1974)  and homogeneous




isotropic  three-dimensional turbulence  (Orszag  and  Patterson




1972), at moderate Reynolds numbers.  Extensions of  these studies




to inhomogeneous flows, turbulent  wakes and jets, etc.,  are being




actively pursued  (Herring 1974,  Orszag  and  Pao  1974).




     It would  not be  appropriate here to  attempt to review  the




salient conclusions regarding the  structure of  turbulence,  and




the validation of turbulence  theories,  which  have emerged from




such studies.   One point,  however,  is certainly worth emphasizing




here.  That is,  throughout the course of  such spectral simulations,




it has become  empirically  evident  that  the  large-scale flow fea-




tures exhibit  a  remarkable degree  of Reynolds-number  independence.




As a consequence, it  may not  be necessary to  simulate turbulence




at the huge Reynolds  numbers  appropriate  to the atmosphere, if




our principal  concern is with synoptic-scale  features; that is,




much valuable  information  can be obtained from  simulations whose




Reynolds numbers are  such  that all  dynamically  relevant  scales




of motion  can  be included  explicitly in the computation, without




recourse to closure arguments  or subgrid-scalemodeling.  It appears




that, once a certain  critical  Reynolds  number is attained  (and such




Re  .  are remarkably low!) the large-scale features  are sensibly




independent of the detailed mechanisms  by which energy  (or  ens trophy)




are cascaded to  the smallest  scales, nrovided that  the initial flow




conditions are Reynolds-number independent.





                             -46-

-------
     Of more immediate importance, with  regard  to  the  improvement




of air-quality simulation modeling efforts,  are  the insights which




are emerging concerning  the utility of spectral methods in the




numerical simulation of  advective-diffusive problems.   In a pio-




neering effort, Orszag  (1971b) considered  the Crowley  color pro-




blem (Crowley, 1968), namely the two-dimensional convection of




a passive scalar field by a uniform rotation velocity-   This




problem has been considered as well by Molenkamp (1968) ,  from




the finite-difference point of view.  Orszag compared  solutions




to this problem using second-order and fourth-order Arakawa




finite-differencing schemes, and the Galerkin Fourier  approach




discussed before.  He demonstrated that  results obtained using




the truncated Fourier-expansion scheme on  a 32x32  space grid




(cutoff wavenumber K=16) are at least as good as those obtained




by the fourth-order scheme on a 64x64 grid,  and significantly




better than those obtained by the fourth-order  scheme  on  a 32x32




grid, or by the second-order scheme on 32x32 and 96x96 grids.




It appears that, for those problems for  which the  "color" problem




and the Neuringer problem discussed in Chapter 5  are typical exam-




ples, second-order schemes require at least four times as many




grid intervals in each spatial direction,  and fourth-order schemes




require at least two times as many grid  intervals, as  the truncated




Fourier expansion scheme.




     We have recently extended this study  to :i ncorporate diffusive




terms.   These new results are presented  in Chapter 5 of this report




As will be seen, the conclusions which emerge support  strongly the




argument for the use of  spectral techniques where  feasible, espe-




cially  in regard to high Reynolds number flow simulations.





                             -47-

-------
5 - 0  A compg£i£gn__of  Spectral  and Finite Difference  Methods
     for Two Examples

     To date,  little  effort  has  been  directed toxvard exploiting

spectral techniques  for  numerical air-pollution  simulations.   This

section illustrates  some of  the  computational advantages  of the

spectral technique  for the simulation of advective-dif f usive  prob-

lems, by reference  to (a) a  diffusive generalization of the Crowley

 "color " problem  (Crowley 1968,  Molenkamp  1968,  Orszag   1971a),

that is, the advection-dif fusion of an instantaneous point-source

scalar in  a two-dimensional  field of  solid rotation;  and  (b)  a prob-

lem proposed by  Neuringer (Neuringer,  1968) ,  concerning the advection

diffusion  of a point  source  in a sheared wind field.  We  compare

numerical  solutions  using the  spectral (truncated Fourier)  method

with those using the  conventional second-order finite-difference

method.  It is shown  that, for simulations of comparable  accuracy,

the spectral method permits  very substantial  reductions in  required

core storage,  and less computational  time as  well.

     We also present  a rebuttal  to some criticisms  offered  in a

private communication by Chan  and Stuhmiller  (denoted by  C-S)  to

a preliminary  version of these results.



5 . 1  Comparison  of Methods for the Diffusive  Color  Problem

     We follow the evolution of  the two-dimensional  system
                                          6
-------
rotation;   \;  is  a  diffusion coefficient; and  the  
-------
where  k^'k2   represent the  two  horizontal wavenumbers ^
and  the  tilda   (~)   designates  the  two-dimensional  discrete  Fourier



transform  of the  indicated field.



       In particular,  the  term
            k   k
            K1'K2




designates  the (k  ,k  )  component  of  the  alias-free  product of  the



velocity  component U  with the-real-space field  C   at  time  step   n.



The  achievement of alias-free  non-linear products is  performed as



described in  Appendix 3.   The  time-differencing  used  for  the advec-



tive terms  is the  Adams-Bashforth scheme,  while  the  viscous  term is



represented with Crank-Nicholson  time  differencing.   The  expression



(5.3) makes use of the  fact  that  the constant  advecting velocity



field is  non-divergent.



      To  emphasize the  power of the  spectral approach,  the finite-



difference  computations  are  performed  with both  32x32 and 96x96 points



grid resolution, whereas  the spectral  calculations  are  performed only



on a 32x32  point grid.   In each case we  used a true point singularity



as the initial source term,  of non-dimensional magnitude  1.0 units.



Thus the  initial conditions  are identical  in each case, which  satis-



fies an objection  raised  by  (C-S)  to the preliminary  results.   Pre-



vious experience with spectral calculations of the  color  equation



show that the  spectral  approach is remarkably  accurate  for a distrib-



uted  scalar   field   (Orszag, 1971a) but should-be  appreciably less



accurate  for  a true point source.      In every case the numerical



time step was  chosen sufficiently small  that time-differencing errors





                               -50-

-------
were negligible.



      For each system, the physical-space grid  is  square,  and the



center of rotation coincides with the center of the  grid.   The



initial point-source distribution is located midway  between the



center of rotation and the left boundary of the grid.   The fre-



quency of rotation is such that one full rotation  requires 100




time steps.



      We present results for the following cases:



       (a) Finite-Difference Scheme with 32x32 point  resolution,



          viscosity v = .05     w



       (b) Spectral Scheme with 32x32 point resolution  (i.e.  wave-



          number cutoff K=16), viscosity  v = .05



       (c) Finite-Difference Scheme with 32x32 resolution,   v = .005



       (d) Spectral Scherfte with 32x32 point resolution,  v =.005.



      Figure 5.1 displays the initial point-source distribution,



that has been enticofLy machine-constructed with ten  equal  contour



intervals, dashed lines representing negative-valued contour lines.



The contour intervals were generated automatically with contour



spacing

           C    - C
      .  _  max	min

      AL ~      10




where  C   , C  .   are respectively, the maximum and minimum values



of the scalar field  C  encountered anywhere on the  grid at the given



time step.  Thus, no two figures show the same  contour  spacing, a



point which will become important shortly.  However, the choice of



contour interval is now made in entirely objective manner, as was



not the case in the preliminary results seen by (C-S).
                               -51-

-------
Figure 5.1   Initial Point-Source Distribution  (t=0)
           '','."('

            Contour from 0. to l.OOE+00
                          -52-

-------
        Figure  5.2  Finite-Difference  Scheme
            (v  =  .05,  t =  1/4  revolution)
             32  x  32  point  resolution
              F.D.  Scheme  T = 25.0000
Contour from -1.42E-02 to 5.31E-02 Interval 6.73E-03
                         -53-

-------
              Figure 5.3  Spectral  Scheme
               32  x 32  point resolution
                      v  = .05
                  t = 1/4 revolution
             Spectral Scheme T = 25.0000
Contour from -1.19E-03 to 6.21E-02 Interval 7.91E-03
                         -54-

-------
        Figure 5.4  Finite-Difference Scheme
              32 x 32 point resolution
                      v  = .05
                 t = 1/2 revolution
               F.D. Scheme  T = 50.0000
Contour from -4.58E-03 to 2.84E-02 Interval 3.30E-03
                         -55-

-------
              Figure  5.5   Spectral  Scheme
               32  x 32  point  resolution
                       v   =  .05
                  t = 1/2  revolution
             Spectral Scheme T = 50.0000
Contour from -5.92E-04 to 3.15E-02 Interval 4.01E-03
                         -56-

-------
        Figure 5.6  Finite-Difference Scheme
              32 x 32 point resolution
                      v =  .005
                 t = 1/4 revolution
               F.D. Scheme T = 25.0000
Contour from -1.09-01 to 1.13E-01 Interval 2.22E-02
                         -57-

-------
             Figure 5.7  Spectral Scheme
              32 x 32 point resolution
                     v  = .005
                 t = 1/4 revolution
                                                  C>
                              o                      o
             Spectral Scheme T = 25.0000
Contour from -6.38E-02 to 3.70E-01 Interval 4.34E-02
                         -58-

-------
        Figure 5.8  Finite-Difference Model
             32 x 32 point resolution
                     v = .005
                t = 1/2 revolution
             F.D. Scheme  T = 50.0000
Contour from -5.02E-02 to 7.58E-02 Interval 1.26E-02
                        -59-

-------
            Figure 5.9  Spectral Model
            32 x 32 point resolution
                   v  = .005
               t = 1/2 revolution
                                          o,
                                           •  o
                                                 o
Contour from -5.62E-2 to 2.55E-1  Interval  3.11E-1
                       -60-

-------
     In figures 5.2, 5.3 we compare, respectively,  cases  (a)  and




 (b) after one-quarter resolution.  Note that  the  contour  interval




in the spectral case (5.3) is about 17% larger than that  in  the




finite-difference case.  It is observed that  case  (b) more accu-




rately reproduces the circular symmetry to be expected, while case




 (a) displays a small phase lag, relative to the expected  center




of symmetry, axial distortion, and a spurious numerical noise




manifested as characteristic negative "trailing wake"  (the explan-




ation for which we have considered in section 2.1).   Proceeding




with the temporal evolution of both systems out to   t=l/2 revolu-




tion, figures 5.4, 5.5 show that the spectral run continues  to




yield a distribution which is quite circularly symmetric, with no




wake structure, and preserves the expected center of mass of the




distribution much more closely than does the  finite-difference




run, wherein the axial distortion and phase lag are becoming quite




pronounced.  The differences between these two cases (for v  = .05)




are not yet striking, however.  The situation changes greatly as




v->0.




     Let us now compare the two methods for a ten-fold reduction in




"iscosity,  v = .005, i.e., cases  (c) and  (d)  as shown in  figures




5.6, 5.7, for  t=l/4 resolution.  With a reduction  in  v,  we




v;ould expect that the effect of the strong discontinuity  in  the




source distribution would become more apparent, since the smoothing




effect of the diffusion term is correspondingly reduced.  Now the




spectral scheme shows to considerably better  advantage with  regard




to preservation of phase, symmetry of distribution, and  suppression




of superfluous trailing wake structure.  In fact, it would  already




be difficult to argue that the finite-difference  model corresponds





                               -61-

-------
to physical  reality  in  any  useful  sense.   Let  us  look  as well  at




i-he results  for   t=l/2   resolution (figures  5.8,  5.9).  The  contin-




ued high  fidelity of the spectral  model  to the expected continuum




solution  is  truly impressive.   Note  that  the phase  information is




still preserved  remarkably  well, and the  spurious numerical  noise




is of small  amplitude (< 10%  of the  peak).




     The  (32x32)  spectral runs  required  about  eight times  longer




cpu execution  time than the (32x32)  finite-difference  runs.  How-




ever, it  should  be emphasized that these  runs  did not  use  optimally-




coded Fast Fourier Transform  routines  (which would  save about  50%




execution time).   Moreover, we  employed  for  these runs a fully




alias-free spectral  algorithm,  and Orszag has  shown that the partly




aliased pseudospectral  method offers,  in  this  case,  a  four-fold re-




duction in required  computation time,  without  appreciable  degrada-




tion of results.   Together, therefore, these modifications would




effect at least  an eight-fold savings  in  computation time.   These




results suggest  therefore that  spectral methods are cost-effective for




simulation of  the  diffusive color  problem and  offer significant core




storage savings  compared to finite-difference  methods, for equiva-




l3nt accuracy.




     Finally,  it  is  necessary to correct  a misconception entertained




by Chan and  Stuhmiller,  as  evidenced by a comment of theirs  concerning




the use of spectral  techniques  for the color problem:




     "The example  used  in the comparison  is  a  linear problem,  i.e.,




the velocity field is given explicitly ..."In  spectral method, there




are two general approaches  to nonlinear problems in which  the  non-




linearity is of the product type.  The first approach is to  do the




entire calculation in the Fourier  space;  the nonlinear products






                              -62-

-------
lead to evaluation of  'convolution sum' which  is  in  general  very




time consuming.  The second approach is to transform from Fourier




space into physical space and perform  the multiplication  there  for




the nonlinear terms.  When transforming back into Fourier space,




 "aliasing" has to be controlled to avoid the buildup of errors.




Bass was probably referring to the second approach as  the "spectral




method" in his memo.  The first approach is not susceptible  to  phase




errors since the amplitude and phase of each Fourier mode are dealt




with directly as dependent variables.  But the evaluation of con-




volution sums  for nonlinear problems  makes this  approach rather




expensive.  On the other hand, the second approach is  not free  of




phase error because of the calculation performed  in  the physical




space."




     In response to these comments we  should note that, first,  the




color problem is "linear" in the sense that the velocity  fields do




not enter quadratically, but inasmuch  as it is necessary  to  evaluate




the advective product  V-VC  where  V=V(x,y)   is  a spatially variable




field, the computation is nonlinear in the usual  sense; and  we  require




an accurate efficient technique for the alias-free representation of




multiplicative fields.




     Second, it is not true that the "spectral method" to which we




referred is implicitly aliased, or is  phase-error contaminated.   The




spectral method (a truncated Galerkin  Fourier  method)  which  we  used




in these calculations is




     (a)  entirely free of phase error




     (b)  entirely free of aliasing misrepresentations, since we




         used the alias-renewal technique described  in Appendix  3.




     Apparently Chan and Stuhmiller have confused spectral with




pseudospectr; 1 methods.  The alias-free evaluation of  nonlinear




                               -63-

-------
products does not require the explicit evaluation of convolution



sums.  Perhaps the most single advantage of the methods introduced




by Orszag is that we  can achieve the same results as would be ob-



tained by computationally expensive convolution, with much greater



efficiency via the "shifted-grid"  algorithms (Orszag 1971a,  1971b).
                               -64-

-------
5.2  Comparison of Methods  for Neuringer's  Problem



     In this section we compare  spectral  and  conventional second-



order finite-difference numerical  solutions to the problem solved



analytically by Neuringer  (1968),  namely, the advection-diffusion



of an instantaneous point source in  a  two-dimensional linearly



sheared wind field.



     Neuringer presents an  analytic  solution  for  the  evolution



of the two-dimensional system
 (5.4)
CTT = d
 3y
                                (S(X_X', y-y',  t)
where the   6-function represents an  instantaneous  source and



a, c, d are constants.  He shows that with  the  transformations




         t = at


         	                  2

         x = (a/c)(x-aty-act /2)



         y = (a/c) (y+ct)





and with the  5-function source at the origin,  x' = y'  =0,



the analytic solution is
(5.5)
where
F(x,y,t) =-
                                exp
                            12
                -y
(x/2-t.y/4)


t(l+t2/12)
          2    2 .  ,
         Y  = c /ad  .
Figs. 2-5 in Neuringer's paper attempt  to  show  the  variation of



the dirnensionless distribution function F.   It was observed that



these figures are entirely incorrect, except  for the special para-
                             -65-

-------
etric choices  x =  0  or  y =  0.  We have  therefore reevaluated


numerically Neuringer's analytic expression  for   F(x,y~,t),  equa-


tion  (5.5) above.   Fig. 5.10 exhibits  some correct representative


curves of  F  vs.   x  for the  case  y  =  1.0,   t  = 0.5.   (Note,  for


example, that by comparison with fig.  5.10,  Neuringer's  curve  for


y = -5  is wrong by more than  six orders of  magnitude, while the


curve for  y = 0  is correct.)


     Superimposed on fig. 5.10 are  some points sampled from a  spec-


tral simulation of  equation 5.4 with 32x32 point resolution, and


the same choices of parameters  y,  t,  and  (x,y)  dependence.  On


the scale of the figure the spectral results  are indistinguishable


from the analytic results.  Table 5A presents  some sampled  values


for the ratio  F (x,y )/F(0, 0) spaced  at  unit intervals  of   x  and  y,


obtained from direct numerical evaluation  of  equation 5.4.  In this


and subsequent tables the notation  2 (-8)  is  to be understood as

      — P
2 x 10  , and so forth.  In Fig. 5.11  we display a two-dimensional


contour plot of this ratio, constructed from more detailed  numerical


tables.  The corresponding contour  plots constructed  from the  spec-


tral and finite difference runs cannot be  distinguished  by  eye from


those shown in the  figure and  so are not included.


     To permit more exact comparison of the  spectral  and finite-


difference computations for Neuringer's problem,  we have performed


a second series of  three runs, with parameter  y now chosen for


convenience as  4.0,  and the  computation  carried out to time


t = 0.5.  For each  run, the time increment was chosen small enough


so that time-differencing errors were  negligible.  The three runs


compared were:


      (a) spectral computation on a  32x32 point space  grid



          (hereafter called run  (a)) ;


                              -66-

-------
                           Figure  5.10
    Legend
—•  Y  »  o
	Y  • -2
	V  « -3
-•- y  . .5
».... Y  • -» 2
	Y  ' +3
                                                  Y » + 5 curve is
                                                  off -scole to  lower
                                                  right ,
       Figure 5.10  Correct Neuringer FCN for j =1,  t = 0.5
                    A  = Points  Obtained from Spectral Run
                         for  Same Combinations   (X,Y)
                                -67-

-------
                    Fiqure  5.11
                                                --2
Figure 5.11  Neuringer Problem  (y  = 1.0,  t  = 0.5)
             Analytic Value  of  F(x,y)/F(0,0)
                         -68-

-------
                                                    TABLE 5A

                         Analytic  Evaluation of  F(x,y)/F(0,0)  for  y = 1-0»  t  = 0.5
                  -4        -3        -2       -1         0          1          2          3          4



      4             -          2(-8)   5.5  (-7)  6.40(-6)   2.7S(-5)   4.54(-5)  2.78(-5)   6.40(-6)   5.5 (-7)


      3             4(-3)   2.53(-6)   6.10(-5)  5.53(-4)   1.88(-3)   2.40(-3)  1.15(-3)   2.03(-4)   1.40(-5)
I

^     2          2.45(-6)   1.23(-4)   2.33(-3)  1.65(-2)   4.40(-2)   4.40(-2)  1.65(-2)   2.33(-3)   1.23(-4)


      I          5. 29 (-5)   2.OS (-3)   3.08 (-2)  1.71(-1)   3.57(-l)   2.79(-l)  8. 21 (-2)   9.06 (-3)   3.75 (-4)


  Y   G          3.95(-4)   1.22(-2)   1.41(-1)  6.13(-1)   1.00( 0)   6.13(-1)  1.41(-1)   1.22(-2)   3.95(-4)


    -1          1.02C-3)   2.46(-2)   2.23C-1)  7.59(-l)   9.70C-1-)   4.65(-l)  3.38(-2)   5.66(-3)   1.44 (-4)


    -2          9.12(-4)   1.72(-2)   1.22(-1)  3.25(-l)   3.25(-l.)   1.22(-1)  1.72(-2)   9.12(-4)   1.81(-5)


    -3          2.S2(-4)   4.17(-3)   2.32(-2)  4.33(-2)   3.73(-2)   l.ll(-2)  1.22(-3)   5.0S(-5)   7.9 (-7)


    -4          3.02(-5)   3.49(-4)   1.52(-3)  2.48(-3)   1.5K-3)   3.49(-4)  3.02(-5)   9.8  (-7)      l{-8)

-------
      (b)  second-order finite-difference computation on a




          48x48 point grid (called run  (b)),with one and




          one-half times greater horizontal resolution



          than run (a);





      (c)  Second-order finite-difference computation on a




          96x96 point grid (called run  (c)),with three times




          greater horizontal  resolution than run (a).






      The  equations used in the finite-difference and spectral models




 to  simulate  (5.4)  are virtually identical  to those described for the




 color problem, e.g.,equations (5.2)  and (5.3),  the only differences




 occurring in  the representation of the advective terms.  That is,




 the  velocity  field  U(x,y)   is replaced by  +(ay), and the velocity




 field V(x,y)   is  replaced by the constant  -c.  The time-stepping




 schemes used  here  are identical to those described for the color




 problem.




      Table 5B displays  the values of the ratio  F (x,y~) /F (0 ,0)




 obtained  by tabulating  expression (5.5),  for equally-spaced points




 (x,y)  spanning the  region of maximum concentration amplitude.




 Table 5C  shows  the absolute  errors  made by the spectral run, i.e.,




 the  differences between the  analytic values of the ratio  F(x,y)/




 F(0,0)    and  those obtained  from the spectral  computation.  In like




 manner , Tables  5D,  5E display the absolute errors of the 48x48 point




 and  the 96x96 point  finite difference runs,  respectively.




     Upon  examination of these  tables,  we  may  draw several signi-




 ficant conclusions.  First,  we  observe that the errors  made by the




 run  (b) are consistently about  four times  greater than, and uni-




 formly in  the  same direction as, the errors  made by run  (c).  Second,




we see that,  in the  region of maximum amplitude,  both the  (a) and






                              -70-

-------
           TABLE 5B
Analytic Value of F(x,y)/F(0,0)



       (y = 4.0,   f =  0.5)

t

-
-
-
_
- 1
1/2
0
1/2
1
3/2
2
.019
,141
.368
,332
.104
,011
- 1/2
,104
.613
1.252
.885
.216
.018
X
0
,216
1.000
1,600
. 885
..169
.011
+ 1/2
.16-9
,613
,767
,332
.050
.003
+ 1
.050
.141
.138
,047
.005
<.001
               -71-

-------
           TABLE 5C
Absolute Errors in Spectral Run
     -  1
- 1/2
+ 1/2
                                         + 1
t

-
-
-
-
1/2
0
1/2
1
3/2
2
<.001
<,001
+ .004
+ .007
+ ,003
<,001
-.001
<.001
+ .007
+ .013
+ .008
+ .003
-.002
-
+ .011
+ .01,9
+ .001
+ .005
-.003
-.001
+ .012
+ .016
+ .007
+ .002
-,004
-,004
+ .004
+ .009
+ .004
+ .001
              -72-

-------
                     TABLE 5D
Absolute Errors in 48*48 Point Finite Difference Run
              - 1
- 1/2
1/2
1/2
0
1/2
1
3/2
2
-.003
-.006
-.023
-.039
-.007
+ .004
-.013
-.017
-.056
-.085
-.006
+ .007
+ .020
-
-.028
-.063
<.001
+ .005
-.015
-.017
-.036
-.034
-.003
<.001
-.007
.-,006
-.009
-,006
<.001
<.001
                        -73-

-------
                      TABLE  5E
Absolute Errors in  96*96  Point Finite Difference Run
                                 x
              -  1
- 1/2
-I- 1/2
t 1
1/2
0
1/2
1
3/2
2
+ .001
-.002
-,006
-,010
-.002
t.OOl
-.003
-.004
-.014
-.022
-.001
+ .002
-.005
-
-.007
-.016
<,001
+ ,001
-.004
-.004
-.008
-.009
-.001
<.001
-.002
-.002
-.002
-.002
<.001
<.001
                        -74-

-------
  (c) computations are of about equal accuracy  and  capture  equally




well the spatially local peak structure  (although the  finite-differ-




ence approach is "local"  and the spectral approach is  "global'1   in




nature).   Third, with some reference to  the more extensive  tabula-




tions from which these tables have been  extracted,  it  is  seen  that,




in the wings of the distribution, the analytic result  and the  finite-




difference calculations decay to zero very rapidly.  By  contrast,




the spectral computation makes a small oscillatory  error  of absolute




magnitude less than 3*10~  that of the peak amplitude.   This small-




amplitude residual oscillatory error is  typical of  spectral repre-




sentations for strongly peaked fields.   However, as we  have noted




previously, the error made by truncating a spectral representation




at  N  terms decreases more rapidly as   N  is increased  than any




power of  1/N.  Hence, for only modest additional increase  in  spec-




tral resolution, the magnitude of this characteristic  oscillation




can be reduced to negligible proportions.




     It is important to note that, for this two-dimensional compu-




tation, therefore, the core storage advantage of the spectral  method,




in this case, is almost thirty-fold.  This makes it practicable,  from




the point of view of available core, to  pursue a much  larger class




of calculations than are presently possible with finite-difference




methods.




     Again, as regards execution time, the spectral method  is  more




than competitive.  Run (a) required about 36% more  computation




time than did run (c).  However, as before, the spectral calcula-




tions were made in fully alias-free manner, and an  immediate  four-




fold reduction in computation time is possible if we relax  the




requirement for fully alias-free spectral interactions,  by  per-




forming the calculations is pseudospectral manner.   Thus, for




                              -75-

-------
comparable accuracy, the spectral approach is to be preferred, on



economic grounds, for solution of advection-diffusion problems




typified by Neuringer's problem.  We believe, therefore, that



spectral methods merit serious attention for prospective air-




quality simulation efforts.
                             -76-

-------
 6.0   Spectral Modeling  of  the Turbulent  Diffusion  of  a
      Passive Scalar

      This  chapter  describes  a modeling effort,  as  yet unsuccessful,

 using state-of-the-art  spectral methods  to  study the  turbulent dif-

 fusion of  a passive  scalar source  in  a fully  three-dimensional,

 turbulent, sheared velocity  field, without  recourse to  closure

 arguments.  The impetus  for  this effort  stemmed from  the  expecta-

 tion  that  such a model  would provide  an  idealized  set of  turbulent

 diffusion  data, free of  closure parameterization assumptions,  against

 which various air-quality  turbulent diffusion models  might  be  tested.

 Moreover,  it was anticipated that  the model might  permit  the compila-

 ation of model turbulent diffusion data, suitable  for the evaluation

 of eddy Austauch coefficients, for a variety of cases for which observa-

tional data  are lacking,  or for which existing data  are  unreliable.

      These expectations  seemed well justified,  at  the initiation

 of this study effort, by the very  real successes achieved by earlier

 spectral numerical models  of this  type in reproducing the essential

 dynamics of three-dimensional turbulent  flows in a number of cases

 (Orszag and Patterson 1972,  Orszag and Pao  1974),  and by  the demon-

 strated utility of the numerical results so obtained  in testing the

validity of various formal turbulence theories.  The  model  experi-

ments  contemplated for the current effort presupposed, therefore,

the availability of computational resources (core  storage and  CPU

time)  comparable to that employed in the previous  successful efforts.

It was not until well along  in the current contract study,  and only

after the various computer codes had been written  and checked  out,

that  the computational resources available to us were thenceforth

severely restricted.   The  critical limitation was  that of core

storage.  Th^ experiments had been planned on the premise  that  the


                              -77-

-------
codes could be  run with  (32)J modal resolution, as had previously



been the case at the Lawrence Berkeley CDC  facility.  Unfortunately,



new and unforeseen user  demands upon that facility had the practical



effect of restricting our model experiments to only  (16)  modes.



Previous experience with homogeneous turbulence codes suggested



that (16)  modal resolution wais but marginally adequate for the



evolution of physically  realistic turbulent flows, and we had but



little hope for the prospective success of  a  (16)  model for strongly



sheared flows.  Nevertheless because of our problems with Lawrence



Berkeley we had no choice but to proceed with our experiments at



the reduced resolution.  Initially we hold out some hope for partial



success because the largest scales of motion  (which are the most



important for turbulent  eddy transport) could still be adequately



resolved.  However, after the fact, the difficulty of achieving



physically credible small-scale equilibrium statistics in the pres-



ense of a strongly sheared mean flow at the (16)  resolution pre-



cluded our getting satisfactory results.  (The attainment of equili-



brium is necessary to make possible a meaningful comparison of our



numerical model with the corresponding predictions of a turbulence



closure model based upon equilibrium turbulent conditions.)



     The remainder of this chapter is devoted, first, to a thorough



description of the model we have developed.   Then, a discussion is



given of the difficulties encountered, by reference to figures which



depict the evolution of  the velocity and vorticity fields during



the course of one of our unsuccessful experiments.
                            -78-

-------
6.1  The Numerical Model

     In this section we define the Bouss'inesq equations describing

the physical system to be simulated, and develop the spectral

forms of these equations which are actually implemented  in the

computer model which has been constructed.

     We begin from the familiar Boussinesq system of equations

for shallow layer convection in an effectively incompressible

fluid.   The scaling arguments underlying the derivation  of these

equations are presented in Spiegel and Veronis  (1960)  and Calder

(1967), and will not be restated here.  The Boussinesq equations

may be written as
(6.1)
        at
           = -JJL
 (6.2) «  V-V = 0
 (6.3)
         oo at           ioo
where
        A  is a unit vector in the vertical direction

        V  is the•three-dimensional velocity field

        p  is the deviation of pressure from its equilibrium

           value P

        T  is the deviation of temperature from its equilibrium

           value T ,  which is a linear function of height  z,


           Te = T00(1-Z/H)

        T00'  P00  are the surface equilibrium temperature  and
           density, respectively,

                             -79-

-------
          2
         N ,   the square of the Brunt-Vaisala frequency, =
                   ST
               cr     el
              —^— (  • •  + —), assumed constant.
              P00   3Z    cp
         v,   the kinematic viscosity, = y/Pnn/  where  y  is the

             coefficient of viscosity

         a,   the thermal viscosity =  cr/P  , where  a  is the ther-

             mal conductivity



Equation (6.1)  is the momentum equation, incorporating the con-

tinuity  condition (6.2).   Equation  (6.3) is the heat equation.

For  convenience,  we will define the variables  P = p/p  ,

 02?      -U
 • =  T"^ •   ,  whence
      00


 (6.4)    |=s + v-VV = - VP + 0A +vV2V
         ;\cy,            o      9
 (6.5)    •£•  +  we + irw = KV .
The term    6A   in (6.4)  provides the convective coupling of tem-
                                                         2
perature to  the vertical velocity field, and the term  N W  in

(6.3) represents  the  effect o.f basic-state stratification upon the

evolution of the thermal field.

Equation  (6.4)  may  be  written in the rotational form



(6.6)   % - V  x w  = - V(P + ~ V-V)  + 0X + VV2V
       i dt   —   —            4 — —



where  w = Vx v  is the  vorticity field. .

Equation  (6.5)  may  be  written, using the incompressibility condi-

tion, as
(6.7)    + V-(V0)  + N2W =
                             -80-

-------
Equations  (6.6),  (6.7) and  (6.2), together with  appropriate ini-

tial and boundary conditions, define the  system  to  be  simulated

spectrally.

     To integrate this system with time,  we  use  truncated Fourier

spectral methods.  That is, we represent  the velocity  and tempera-

ture fields in truncated discrete Fourier series  (Appendix 2)  of

the form
(6.8)   V(x,t) =   7  U(k,t)e1-'-
                 |k|
-------
where  a,3,y  may each range over  the  three  indices  1,  2,  3

(i.e., the three orthogonal directions  in wavespace).  The  symbol

6    is the Kronecker delta, and summation over  repeated Greek

indices is implied.  The pressure  term has been  eliminated by

use of the incompressibility condition (see  also the discussion

in section 2.1.3}.  Note, however,  that the form  (6.10)  requires

that a convolution sum be evaluated  for each pair of indices

(3,y).  Such an approach is quite  costly computationally,

especially for a fully three-dimensional system.  Instead, we

have employed the pseudospectral approach to integration of the

momentum and heat equations.  Returning to equation  (6.6), we

see that the source of the convolution term  is the nonlinear

vector product term  VXUK  To avoid the convolution requirement

we use the following procedure:

      (a) Beginning from the Fourier  transform of. the velocity

         field, U(k,t), create the Fourier transform of the

         vorticity field,n(k,t) by performing the wavespace

         operation  J2(k,t) = i k x U(k,t).

This is the direct analogue of  to  =  VxV.  Note that  because

the vorticity field is evaluated in  wavespace, it is  free  of

first-differencing phase and amplitude errors.

      (b) Recover the ir^al-space vorticity field   u(x,t) by  inverse

         discrete Fourier transformation upon  ft(k,  t), namely

            u)(x,t)  = I    g^tje1^'-
                    |k|
-------
      (c) Evaluate,  in  discrete real-space,the vector cross



         product  term   V(x,t) *w(x,t)  - ,  say, £(x,t).  This



         product  is, of course,  fully  aliased unless the pro-



         cedure outlined in  Appendix  3 is  used.



      (d) Generate the  discrete Fourier transform of  ^(Xft), de-



         noted by   V_(k't)



            l(k,t)  - I  ^(Xftje"1-*- •

                     x


                                          2
      (e) Represent  the  diffusion term   vV  V(x,t)   in wavespace as



                2
            - vk  U(k,t) -



         Then equation  (6.6) may be written in prognostic spectral



         form as




 (6.11)       (~ + vk2)U(k,t) - nk,t)  - 6(k,t)A.
             at       — —     — —        —




The spectral form of the  pressure head term,  -V (P + -^ V-V) , need



not be included in  (6.11), since the only  effect of this term is



to guarantee that the velocity field remains  incompressible.  In-



stead, incompressibility  can be  enforced,  at  each successive time



step, by requiring  the  replacement  of   U(k_,t)  by







 (6.12)      u' (Jc,t) =   U(k-,t) -  -=*-(k-U(k,t))

                                 It




whence  ik-U'(k,t)  = 0,  thus maintaining the  incompressibility



 (see section 2.1.3) .   The nonlinear term   V-  (V0)  in the ther-



mal equation  (6.7)  may  be treated in a similar pseudospectral



manner, by the following  procedure:



      Ca) From the  real-space  velocity field  V(x,t), and -the



        real-space  temperature field   6(x,t),,  from the real-space



        vector   V(x,t)6(x,t) =,  say,   Z_(x,t).




                             -33-

-------
      (b) Create  the  Fourier  transform of  Z^(x,t),  which will

         be denoted  by    C_(k,t):


            Uk,t) - I  Z^xjtje""1^  .
                      x

      (c) Evaluate the divergence  term  V- (V9 ) ,  directly in wave-

         space,  as   ik'C_(k.»t).   (The divergence operation is ik-

         in wavespace).



     Thus,  (6.7) may be written in spectral form as


(6.13)       IE  + Kk2)8(k,t)  -  -  ikk?(k,t)  - N2U3(k,t)


where  U_(k,t)   designates the  Fourier  transform of the vertical

velocity field   w.

     Leapfrog time differencing was used for the advective terms

and Crank-Nicolson   time  differencing for  the  diffusion terms.

Thus, if the superscript  ()n  designates the value of a  field at

the n   time step  (t =nAt) the  time evolution  of the system can

be explicitly represented as


(6.14)      (1 +

                                            en(k,t)X]
(6.15)
                            -  At [ik'
(6.16)       Un+1(k,t)  -  Un+1(k,t)  -
                             -84-
                                       k

-------
     Finally, we have included  in  our model  a passive scalar

field, whose turbulent evolution is  governed by  the velocity

field   v(x,t).  The scalar conservation  equation may be written

as


(6.17)      || = vcV2c + s(X,t)



where  c(x,t) is the scalar field,   v   is the scalar diffusion

coefficient, and  s(x,t)  is a  source term for the scalar field.

If we let  C(k,t) denote the Fourier transform of  c(x,t),

S(k_, t) denote the Fourier transform  of  s(x, t),  and  M(k,t)

denote the Fourier transform of the  product   V(x,t)c(x,t),

then the temporally discretized version of  (6.17), by analogy

with  (6.15) is just


                 v k2At    .              v k2At
(6.18)       (1 + _£_—)Cn+1(k,t)  =  (1	^-=	)Cn~i(k,t)
                                  - At[ik-Mn(k,t) ]
                                  +  At  Sn(k,t).
The prognostic equations  (6.14, 6.15,  6.18),  together with the

replacement  (6.16) define the numerical model.   Initial and

boundary conditions will be discussed  shortly.
                             -85-

-------
6-2  Physical Parameters



     An attempt was made to design an experiment to simulate the



turbulent diffusion of a passive scalar  field in a sheared, tur-



bulent flow, in such manner that reasonable comparison could be



made with the results of one or more turbulence closure models for



the same physical problem.  To this end, the geometrical dimensions



of our numerical model were fixed as 9.6 km in each of the two



horizontal directions  and 0.96 km in the vertical direction.  The



horizontal and vertical grid intervals were 600 m  and 60 m  res-



pectively.  All fields were taken to be  periodic in the two hori-



zontal directions.  At the upper and lower boundaries of the sys-



tem, free-slip boundary conditions were  imposed on the horizontal



velocity components  (that is  U  = V  =0), while the vertical
                               Z    Z


velocity  W  and the scalar field  C  were constrained to be zero.



     For the initial experiments, we chose a neutrally stratified



flow, with no convective coupling between the thermal and vertical



velocity fields, because it was felt that the inclusion of the



stratification and Boussinesq forcing terms from the outset would



impose an unwanted degree of complexity  in the interpretation of



the results.  Accordingly, the terms  At9n    in  (6.14)  and  N U3



in (6.15)  were suppressed.  Since the temperature field no longer



had any dynamical significance, equation  (6.15) was in fact ignored.



A linearly sheared mean-velocity profile  U(z)  was chosen of the



form:
                             -86-

-------
     The mean flow field was perturbed initially by a small-



amplitude, random three-dimensional field, the construction of



which is explained in Appendix 5.  The mean shear profile  U(z)



was maintained constant throughout the evolution of the  flow  by



fixing in time the corresponding spectral components  0^(0,0,K  ).



Of course, fixing the mean wind field implies a continuous source



of kinetic energy for the perturbation wind field  (by Reynolds-



stress conversions of the form (u'w'U ).  It was hoped that the
                                     z


turbulent perturbation wind field would come to approximate energy



equilibrium, the Reynolds-stress excitation being matched by suf-



ficiently vigorous viscous dissipation at the smallest scales.



 (In fact, during each of a number of experiments over a  range of



Reynolds numbers, the system never tended toward equilibrium in



any physically meaningful sense.   At large Reynolds numbers, there



would result a strong cascade of energy toward high wavenumber,



which could not be dissipated sufficiently rapidly.  If  the Rey-



nolds number was reduced to the point where energy dissipation



could keep up with the generation of kinetic energy, the pertur-



bation flow structure typically degenerated into a set of large-



scale vortices ,iri.-a totally unphysical manner.  Some examples of



these characteristic difficulties will be shown in a later sec-



tion.   In retrospect, this behavior may be understood as a conse-



quence of the very limited modal resolution with which the experi-



ments were attempted, together with the strong mean^shear profile



adopted. )
                            -87-

-------
 6* 3   An Unsuccessful Experiment


      In this  section we follow the  evolution of  the  turbulent


 flow  fields in  one experiment.   The objective was, as previously


 discussed, to create a turbulent flow structure  for  which  the


 generation of kinetic energy  (by conversion  of the mean  shear  U  )


 would come to equilibrium with the  viscous dissipation of  kinetic


 energy.  The  difficulties encountered,  which proved  insurmountable


 at present, are  illustrated.


      To aid in  visualization,  all velocity fields  are pictured


 in the  form of  contour plots.   Positive-valued contour  lines are


 shown as solid  lines,  negative-valued contour lines  are  shown  as


 dashed  lines.   The minimum and maximum contour values and the


 contour interval  are  indicated for each  figure.   These  contours


 have  been constructed by  an automatic contouring package.


      In order to depict the three-dimensional spatial variation


 of the  velocity  fields u,  v,  and w,   the fields are displayed


 in each of three orthogonal planes.   That is,  for  example, we


 show  the variation of  the  u   velocity  component in  the  y-z


 plane  for some  fixed x-coordinate,  say x ,  the  variation  of u


 in the  x-z   plane for fixed   yn,   and the variation of  u  in


 the   x-y  plane  for fixed  z  .   The  choices   XQ, y   z   are


 arbitrary; because the fields  are turbulent,  the individual  con-


 tours are to  be  understood only  as  representative  of the charac-


 teristic spatial structure  in  the given plane.


      In the experiment to be discussed, the  perturbation kinetic


energy  spectrum  was  chosen  to  have  the  spectral  power law  depen-

                1               222'1/2
dence   E'(k)
-------
wavenumber.   The  rms magnitudes  of  the  initial perturbation



velocity  fields,  compared  to  the rms  magnitude of tho mean



velocity  field  U(z), were
              =  2.2  x  10~2;   (U')rms  ~  0.2  m/sec
        ?'  i /•>
        '
    < (\7 ' } > '            — 9
      _J -, p. - =  2.4 x  10   ;  (V')rms  -  0.2  m/sec

          '
       '
                         -3
               =  5.6 x  10   ;  (W')rms  -  0.05  m/sec.
The ratio of total perturbation energy  to  mean  energy was initially


about 5 x 10   .  The viscosity coefficient used in the vertical


direction was  fixed as  1/100  that  of  the viscosity coefficients


used in the two horizontal directions, consistent with the (square


of the)  ratio  of vertical to  horizontal spatial intervals.   The

                                                                 2
particular choice made  in this experiment  was vtrrvmAT = 10 m /sec.
VVERTICAL = °"1 m /sec-   (In other experiments  we took values


of  v  which were tenfold higher  and  lower) .  The equivalent large-



scale Reynolds number, based upon a characteristic velocity  U=10



m/sec. , a horizontal length scale L  =  -^— km, and the value of
                                        2rt


VHORIZONTAL' was thus  ~ 160°-  Tne characteristic eddy circulation


time  L/U  was  ~ 10  sec.  The time step,  as prescribed by the



Courant advection condition for a spectral scheme,
            UAt  <

             AX  -
was taken as 10 sec., which is  a  slightly  conservative choice


                            -89--

-------
Figure 6.1A  U Field (y-z projection), t =0
               Umin = 5'4 m/S6C
               U    =10.2 m/sec
                max         '
        Contour Interval = 0.3 m/sec
                     -90-

-------
Figure 6.IB  U Field (y-z projection), t = 10  sec
              Umin = 5'4  m/sec
              Umax = 10"2 m/sec
         Contour Interval 0.3 m/sec
                     -91-

-------
Figure 6.1c  U Field  (y-z projection), t = 2 x 10  sec
               U .   =4.8 m/s'ec
                mm


               U    =10.8 m/sec
                max


          Contour Interval 0.3 m/sec
                      -92-

-------
Figure 6.2A  U Field (x-z projection),  t = 0
°
               min
                         m/sec
              Umax = 10'2  m/sec
        Contour Interval =0.3 m/sec
                     -93-

-------
Figure 6.2B  U Field  (x-z projection), t = 103 sec
               Umin =  5'
               Umax = 10'2 ra/sec
         Contour Interval =0.3 m/sec
                       -94-

-------
Figure 6.2C  U Field (x-z projection),  t = 2 x 10  sec
                 U .   =5.1 m/sec
                  mm        '


                 U    =10.8 m/sec
                  max         '


            Contour Interval 0.3 m/sec
                         -95-

-------
Figure 6.3A  U Field (x-y projection), t = 0
              U .   =7.5 m/sec
               mm        '


              U    =8.35 m/sec
               max


        Contour Interval 0.05 m/sec
                     -96-

-------
Figure 6.3B  U Field (x-y projection), t = 10  sec
                 U .  =7.1 m/sec
                  mm        '


                 U    = 9.2 m/sec
                  max        '


            Contour Interval = 0.1 m/sec
                        -97-

-------
Figure 6.3C  U Field (x-y projection), t = 2 x 103 sec
                  U  .  =5.4 m/sec
                   mm


                  U    =10.2 m/sec
                   max         '


             Contour Interval 0.3 m/sec
                        -98-

-------
    Figure 6.4A  V Field  (y-z projection), t = 0
>/"	A.    ""  \
                  V .   = -.4  m/sec
                   mm        '


                  V    = .8 m/sec
                   max


            Contour Interval  = 0.08 m/sec
                         -99-

-------
Figure 6.4B  v Field (y-z projection), t = 103 sec
                Vmin = -'48 m/sec
                Vmax = '80 m/sec
           Contour Interval = 0.08 m/sec

-------
Figure  6.4C   V Field  (y-z projection), t =  2  x 10  sec
 _.•"	i;-	'.'.'.I'   ;•""  i""   •:[   ••"
                 V  .   =  -.64 m/sec
                  mm
V
                  max
                             m/sec
            Contour Interval 0.08 m/sec
                         -101-

-------
Figure 6.5A  W Field (y-z projection), t = 0
Wmin = ~'
Wmax = -1
                         m/sec
                        m/sec
        Contour  Interval  =  .01 m/sec
                    -102

-------
Figure 6.5B  W Field (y-z projection), t = 10  sec
     Wmin = ~'24
     Wmax - -12
Contour Interval
                            =  .02 m/sec
                       -103

-------
Figure 6.5C  W Field (y-z projection), t = 2 x 10  sec
                 W .   = -.30 m/sec
                  mm         '

                 W    = .27 m/sec
                  max        '


              Contour Interval .03 m/sec

-------
     The typical structure of the initial velocity fields is


represented in figures 6.1A, 6.2A, 6.3A, 6.4A, 6.5A.  Note in


particular that the mean vertically sheared wind field  U(z)

is prominent in 6. 1A and 6.2A, although perturbed slightly by


the initial turbulent  u'  components.  In figure 6.3A, which

represents an x~y slice of the initial  u  field, we see that the


horizontal structure of the  u  field is, indeed, initially quite

random, with a  p1   fluctuating component of magnitude   0.4 m/sec
                                             /
(the mean  U  value is about  8 m/sec).  Figures 6.4A, 6.5A.show,


respectively, typical slices in the y-z plane, of the small-ampli-

tude velocity components  V   and  w1   respectively.  The structure


of the  v>  field is quite random spatially, while the  w*  field

is rather more organized into vertical rolls.  (This is a conse-

quence of the specific procedure by which the perturbation fields


were synthesized; it reflects the fact that the vertical extent

of the system is one-tenth that of the horizontal extent.)

     In figures 6.IB ... 6.5B we display the same fields after the

system was permitted to evolve for 100 steps  (or 10  seconds),

corresponding roughly to the large-eddy circulation time. " The

mean shear structure of  U(z)  is still strongly evident, although


the  u1  perturbation velocity field has more or less doubled in


amplitude,  as pan be inferred from the range  U     -U .    in

figure 6.3B.   However, by comparing figure 6.4B with figure 6.4A,

and figure 6.5B with figure 6.5A, we see that the small-scale


structure of the perturbation velocity fields  v1  and  w'  has

been damped appreciably, which suggests that, in this experiment,

turbulent energy is being dissipated at the higher wavenumbers


more rapidly than it is being cascaded downscale by nonlinear


                            -105-

-------
exchange with low-wavenumber modes.



     After another 100 times steps of evolution, the situation



is worse yet.  Although the perturbation velocity fields have



roughly  doubled again in magnitude, this growth is almost en-



tirely, at the very lowest wavenumbers as may best be seen in



figures 6.1C and 6.5C.



     Since, after two eddy "circulation times, the turbulent vel-



ocity fields were clearly further  from equilibrium .than before,



it was pointless to continue the time evolution of this experi-



ment.  We therefore started anew,  in a second experiment, this



time with ten-fold, les.ser viscosity.  Although the decrease in



viscosity tended to preserve relatively more of the small-scale



structure, the growth of total perturbation energy was unrealisti-



cally rapid, and. the contour plots showed the mean wind field was



soon masked by the magnitude of the perturbation velocity fields.



Alternatively, we also tried a ten-fold increase in viscosity,



which resulted in a rather more constant magnitude of perturbation



kinetic energy, but at the price of even stronger suppression of



all but the very lowest harmonic components.  After further exper-



iments with somewhat difficult initial conditions, it was reluc-



tantly concluded that, with the limited resolution and cpu time



available, it would not be possible at this time'to evolve a



turbulent flow in reasonable equilibrium with the desired mean



shear profile.
                           -106-

-------
1'®  Relaxation of Anisotropic Turbulence to Jsdtropy


     One of the critical parameters of turbulent transport models


is the rate at which an initially non-isotropic flow decays to


its asymptotic isotropic state.  Return-to-isotropy terms appear


exnlicitly, for example, in the turbulence closure model of


Donaldson  (1973)  and that Lumley and Khajeh-Nouri  (1973), in


the form of the pressure-strain correlation expression
               3u' .
            pi I _ i. i
              <
This represents> for incompressible flow, a redistribution of tur


bulent kinetic energy among the three velocity components.  A


variety of closure prescriptions have been given for this term.


For example, Donaldson has chosen the Rotta model  (Rotta, 1951)

             -                                  .    2

               3uT   3u        p^  _     ikq
                                    /„ i n i   _  - - ^ - \
                                    (u  u             '
            " X3x,  -ax. '      A T  V i   k      3
                 k      i         1



            	  .  1/2
where  q = (u .u1.)      is a scalar measure of the turbulent


velocity field,   Xi   is a length scale introduced for dimensional


consistency  (and requires independent specification), and  p..   is


the density.


     Instead of specifying the rate of return to isotropy in ad-hoc


fashion, 'one may have recourse to numerical simulation  for an ex-


perimental measure of this parameter.  It is important  to use


direct spectral methods to ensure high accuracy.  To perform such


a simulation, we require a fully three-dimensional closure-free


Navier-Stokes model.  The basic model described in section 6.1
                            -107-

-------
serves  especially well with the following minor modifications:



      (a)  the^ thermal and scalar concentration equations are



          ignored;



      (b)  the  vertical boundary conditions are modified to per-



          mit  general periodic symmetry conditions in the vertical



          (rather than the particular cosine-like or sine-like



          conditions previously specified).



     To best  examine the relaxation-ta-isotropy process, it is



useful to start  from an axisymmetric turbulent flow, i.e., a tur-



bulent three-dimensional flow field with one axis of symmetry.



The initial flow conditions for such a relaxation problem may



be constructed in the following manner.   We begin from a velocity



field  V(x,t)  which is statistically homogeneous and axially



symmetric about  an axis whose direction is  described by the



vector  n.  Suppose we define the second-order two-point moments



of  V, namely:



        <\7  I v +- ^ V I v'  +• ' \ ~>  —  V  I Y v1  +• -h M
        <-v.v.xfujv.ix_,t. ) >  —  T . . ix/ x_ ,u,t  ;.





     The  Fourier transform  of  ¥   is given  by





          ¥. . (K,t,f )  = / exp (IK- (x-x1 ) Y . . (x,x' , t, t' )d(x-x' ) }.
          13 —          '        —  —     i j





For the axisymmetric  case,  this may be  decomposed into






          ^ (K,t,t')  = $(1) (K,t/t')e[1)(K)e.f1) (K)





                      + $(2) (K,t,t')e(2)(K)e.|2) (K)
                             —       i  —  j   —
                            -108-

-------
whore the two vectors   e      and  e_     are





        e(1)(K) =  (K x  n)/| K  x n|






        e(2) (K) =  K x(K x  n)/|K x (K x n) | .






Here,  e     is the direction transverse  to the axis of symmetry,



and  e      is orthogonal to both  K  and  e/ '.  The functions



<£>(1)(K)  and  $(2)(K) are  given by





        $ (1) (K) =  <| (K  x n) -U(K) |2>/|K x  n|2






        $(2)(K) -  <|n-U(K)|2>/|K x n|2






where  U(K)   is the Fourier  transform of  V(x).   The function



$    (K)  represents the modal kinetic energy in the direction


                                            (2)
transverse  to the  symmetry axis  n,   and  $    (K)   represents



the modal kinetic  energy in the direction orthogonal both to



wave vector K  and to   n.   If the  statistical properties of



the initial velocity field are invariant  to rotation about the



symmetry axis, then  $    (K)   and   $'2'(K)   are invariant as



well.



     A useful 'experimental'  measure of the rate of return to



isotropy is provided by the non-dimensional parameter
                            E2(t)  -  E1(t)
where
        E, (t) =  2ir/   dK K2$(1) (K,t)

          1         0
                             -109-

-------
                         JL  ln{E2(t)-  E^l)}
Figure 7.1    Time History of  R  (t)   for  E.. (0) = 0 and
              E^O)  = 1/2 E2(0)
                            -110-

-------
                    ,-<»      2  (7 )
        E  (t) = 2-n  I  dk K $v  '(K,t)

                    0



represent the kinetic energies  in  the  orthogonal directions



e(1)(K),  e(2)(K),  respectively,   E(t)  =  E-^t)  + E2(t)   is



the total kinetic energy,  and






        e(t) - 4TT2   /°°dk K4[${1)(K,t)  + $(2)(K,t)]

                     0




is the total viscous energy  dissipation rate.   The  ratio



 E(t)/e(t)     describes the rate  of fractional decay of total
energy, while the ratio  -r—
E2 (t)-E1(t) |/[E2(t)-E1(t)|  describes
the time-dependent fractional  rate of  return  to isotropy.   R (t)
                                                             a.



may be considered equivalent to  the Rotta  constant.



     To study the time-dependence of this  parameter upon the extent




of anisotropy initially present  in a turbulent  flow,   we performed




two independent experiments.   In the first case,  the  axisymmetric



turbulent velocity spectrum was  chosen so  that  initially all the


                                             (2)
modal kinetic energy was in the  direction   e_    (K) ,  thus E (0)=0,




E  (0)^0, and the initial flow  field was totally anisotropic.   In




the second case, the initial conditions were  chosen as E  (0)=yE2(0)



so that the flow was fairly isotropic  to begin  with,




     Rotta1s phenomenological  argument suggests that   R (t)  should
                                                        ci


be constant in time with a magnitude ~  2.   However,  this argument




is predicated upon small departures from isotropic  conditions.  It




is interesting, therefore, to  see how  well this prediction holds




for flows which are initially  quite anisotropic.   Figure 7.1 dis-




plays the time history of the  expression  R (t)   for each of the
                                            cl



two cases studied (the time  t  is normalized by the  large-eddy




circulation time).  Somewhat surprisingly, in view of Rotta's
                            -111-

-------
argument,  we sec that for the second case  E1(0) = ^ E2(0),



which represents fairly isotropic initial conditions, the rate of



return to isotropy is considerably less constant than that observed



in the first case  E, (0)=0, which began with completely anisotropic



initial structure.  This is indicative of large statistical fluc-




tuations in the first case, and suggests that even fairly iso-



tropic turbulent flows remain in strong disequilibrium for long



periods of time.  That is, isotropic relaxation of turbulence is



a very slow process.  Thus, turbulence closure models which suppose



self-similar asymptotic conditions are strongly suspect in this



regard.



     We may note, finally, that these numerical results are roughly



consistent with calculations based upon the Direct Interaction Ap-



proximation turbulence theory of Kraichnan  (Herring  1974).
                            -112-

-------
8.0  Recommendations for Future Research




     Although the primary research objective of direct turbulence




simulations was not achieved in the current contract  because of




limited computer resources, we remain confident that the use of




spectral methods holds considerable promise for operational




improvement of air quality numerical modeling in a number of




specific applications areas.  These include:




     (a) more accurate treatment of advection and viscous terms




         by means of spectral representations which preserve




         better than do finite-difference schemes the true phase




         and amplitude structure of such terms;




     (b) accurate representation of the advection-diffusion of




         point-source and line-source pollutants, as suggested




         by the discussion in chapter 5>




     (c) efficient spectral schemes for accurate representation




         of vertical structure and boundary layers, in particular




         the use of Chebyshev spectral methods to provide effec-




         tively increased resolution at viscous boundaries;




     (d) the use of spectral methods, in conjunction with mapping



         techniques, for accurate treatment of topography, as,




         for example, the modeling of pollutant advection-diffusion




         from a highway situated in a valley;




     (e) the use of spectral methods for the accurate simulation




         of lateral inflow and outflow boundary conditions, with-




         out requiring the imposition of ad-hoc one-sided differ-




         encing schemes.





     We recommend, therefore, that careful investigation be made




of the  prospective benefits to be gained by implementation of




these spectral methods within existing air-quality models of




                               -113-

-------
importance  to  EPA,  in  particular,  the  photochemical  air-pollution



model  developed at  Systems Applications,  Inc.,  and the  air-pollu-



tion model  developed at  IBM.




     As  to  the problem of simulating the  turbulent diffusion of



passive  contaminants in  fully  three-dimensional turbulent sheared



flows  by the  direct spectral  approach of the current effort, we



do not recommend further pursuit of this  study  at present for two



reasons.




     The first of these  concerns the timeliness  of application of



this research.   It  does  not  appear likely that  extension of the



present  study  would provide  to EPA, in the near term, the capability



for better  estimation  of eddy  Austauch coefficients  for turbulent



diffusion of pollutants  in the place of field observations.  We



do believe, however, that the  prospects for success  of a longer-



term research  effort remain  very good.  We do not know of any



fundamental theoretical  impediments to the successful numerical



simulation  of  such  turbulent flows by  direct spectral methods,



especially  in  the light  of previous recognized  achievements in



this area by Orszag and  collaborators.



     The second reason concerns the prohibitive  economics of the



required calculations.   Neither the computer core time necessary



for adequate spectral  resolution,  nor  the computer time required



to establish realistic turbulent dynamics, will  be available to



us in  the near future.   At present, such  calculational efforts



can possibly only be undertaken by university-affiliated groups



with access to the  NCAR  computer facility or a comparable installa-



tion.  When the  computing-resource situation improves, as may be



anticipated, it would be most  worthwhile, we believe, to pursue




this study anew.




                            -114-

-------
9.0  References

Arakawa, A. (1966) J. Comp. Phys. _1, 119-143.

Arakawa, A. (1970) Numerical Solution of Field Problems  in
     Continuum Physics 24,,  American Mathematical Society,
     Providence, R.I.

Bourke, W. (1972) Mon. Weath. Rev.  100, 683-689.

Calder, K. L.  (1967) QJRMS, 88-92.

Collatz, L. (1960) Numerical Treatment of Differential Equations,
     Springer-Verlag, Berlin.

Cooley, J. W.  and Tukey, J. W.  (1965) Math. Comp. 19, 297-301.

Crowley, W. P.  (1968) Mon. Weath. Rev. 96, 1-11.

Cunnold, D.,  Alyea, F., Phillips, N. and Prinn, R.  (1975) J.A.S.
     32^ 123.

Deardorff, J.  W.  (1970) J. Fluid Mech. 41, 453.

Donaldson, C.  duP.  (1973) "Atmospheric Turbulence and the Dis-
     persal of Atmospheric Pollutants," EPA report EPA-R4-73-016a.

Dorr, F. W. (1970) SIAM Rev. 12, 248.

Eliassen, E., Machaver, B. and Rasmussen, E.  (1970) "On a Numer-
     ical Method for Integration of the Hydrodynamical Equations
     with a Spectral Representation of the Horizontal Fields,"
     Institute for Teoretisic Meteorologi, Univ. of Copenhagen,
     Report No. 2.

Fornberg, B. (1972) "On High Order Approximations of Hyperbolic
     Partial Differential Equations by a Fourier Method." Dept.
     of Computer Sciences, Uppsala Univ. Sweden, Report No.  39.

Fox, L., and Parker, J. B.  (1968) Chebyshev Polynomials  in Num-
     erical Analysis, Oxford Univ. Press, Oxford

Fromm, J. E.   (1963) "A Method for Computing Nonsteady Incom-
     pressible Viscous Fluid Flows," Los Alamos Scientific
     Laboratory, Report LA-2910.

Fromm, J. E.   (1970) An Introduction to Computer Simulation in
     Applied Science, ed., F. F. Abraham, W. A. Tiller.  IBM
     Data Processing Division, Palo, Alto, Calif.

GARP (1974) "The GARP Programme on Numerical Experimentation,"
     Report of the Int. Symp. on Spectral Methods in
     Report No. 7, Copenhagen.
                             -115-

-------
Gentleman, W.  M.  and Sande,  G.  (1966)  Proc.  Fall Joint Computer
     Conference,  1966,  563-578.

Gordon, A. and Stern, W.  (1974)  "Spectral Modelling at GFDL,"
     on GARP  (1974)  (see  above).

Grammeltvedt,  A.  (1969) Mon. Weath.Rev. 97,  384.

Harlow, F. H.  and Welch,  J.  E.  (1965)  Phys.  Fluids 8,  2182.

Herring, j. R.  (1974) Phys.  Fluids 17.,  859-872.

Herring, J. R., Orszag, S. A. Kraichnan,  R.  H.  and Fox,  D.  G. ,
     (1974) J.  Fluid Mech. ^6_,  417-444.

Hockney, R. W.  (1970) Methods in  Computational  Physics 9,  135.

Israeli, M.  (1971)  "Time  Dependent Motions of Confined Rotating
     Fluids,"  PhD. Thesis,  Dept. of Applied Mathematics,  M.I.T.,
     Cambridge, MA.

Kreiss, H. O.  and Oliger,  J, (1972)  Te&lus 24,  199.

Kreiss, H. O.  and Oliger,  S. (1973), Methods for the Approximate
     Solution  of  Time Dependent Problems,  GARP  Publication Series
     No. 10, WMO,  Geneva.

Lanczos, C.  (1956) Applied Analysis, Prentice Hall Pub.  Co.,
     Englewood, N.J.

Lilly, D. K.  (1965)  Mon. Weath. Rev. 93, 11.

Lumley, J. L.  and Khajeh-Nouri, B.  (1973)  "Computational Model-
     ing of Turbulent Transport," Proc.  Second  IUGG-IUTAM
     Syrap. on  Atmoshperic  Diffusion on  Environmental Pollution,
     Academic  Press, N.Y.

Merilees, P. E.  (1973) Atmosphere rL,  13-20, (1974)  Mon. Weath.
     Rev. 102,  82-84.

Metcalfe, R. W.  (1973)  "Spectral  Methods  for Boundary  Value
     Problems in  Fluid Dynamics," PhD.  Thesis,  M.I.T.  Cambridge
     MA.

Molenkamp, C.  R.  (1968) J.A.M.  T_, 160-167.

Neuringer, J.  L.  (1968) SI AM J. Appl. Math 16_,  4.

Orszag, S. A.  (1969) Phys. Fluids (Suppl.  2) 12_,  250.
                              -116-

-------
Orszag, S. A.  (1971a)  Stud,  in Applied  Math 50,  293.

Orszag, S. A.  (1971b)  J. Fluid Mech.  49,  75

Orszag, S. A.  (1971c)  Stud,  in Applied  Math 50,  395.

Orszag, S. A.  (1971d)  Phys.  Rev. Letters  26,  1100.

Orszag, S. A.  (1971e)  J. Fluid Mech.  50,  689.

Orszag, S. A.  (1972) Stud,  in Applied Math  51,  253.

Orszag, S. A. and Israeli, M.  (1974)  Ann. Rev.  Fluid  Mech. _6_,  281,

Orszag, S. A. and Pao, Y. H.  (1971)  "Numerical  Computation  of
    Turbulent Shear Flows,"  Advances  on Geophysics  ISA,  Academic
    Press, Inc., NYC 225-236.

Orszag, S. A.  and Patterson, G. S.  (1972)  Statistical Models
     and Turbulence, ed. Rosenblatt,  Springer-Verlag,Berlin,  127.

Patterson, G. S.  and  Orszag, S. A.  (1971)  Phys.  Fluids  14,
     2538-2541.

Phillips, N. A,  (1959) The Atmosphere and the Sea in  Motion.
     Rockefeller Institute Press, NYC.

Piacsefc, S. A. and Williams, G. P-  (1970) J. Comp.  Phys.  6,  392.

Roache, P. J.  (1972) Computational Fluid Dynamics,  Hermosa
     Publishers, Albuquerque, NM.

Rotta, J. C.  (1951) Z. Phys. 129, 547.

Spiegel, E. A. and Veronis,  G.  (1960) Astrophys.  J. 131,  442-447.

Swartztrauber, P. (1972) A Direct Method for the  Discrete Solu-
     tion of Separable Elliptic Equations  (unpublished).

Williams, G. P.,  (1969) J. Fluid Mech.  37,  727.
                              -117-

-------
                          Appendix 1






       The  Data  Content of  Discrete-Grid Representations







     It is conventional practice / in the numerical solution of



many fluid motion problems, to perform such calculations with



respect to a discrete spatial grid  and  suitably spatially dis-



cretized versions of the governing differential equations whether



by finite-difference of finite-spectral  methods.  In this section,



we discuss the underlying assumptions and limitations which inhere



to the discretization method.



     Imagine that we wish to estimate the frequency spectrum of a



continuous spatial waveform,  f(x), which is periodic with wave-



length  L.  A harmonic  analysis  of this  waveform might be performed



as follows :



     (1)  partition the  interval  L  into  N  equal segments by



         the discrete points  x  =pAx  (p=0, ... N-l) ;  Ax=L/N,



         where for convenience,  N is a power of 2);




     (2)  tabulate the string of  values   f (,x )  at each of



         these points;




     (3)  employ the Discrete Fourier Transform algorithm



          (Appendix 2) to compute the Fourier modes  9(^q)



         corresponding  to the discrete wavenumbers







                        9-°' i-  ••• N/2  •
Note that there is an upper wavenumber cutoff  (the  "folding" or



Nyquist, frequency N/2) so that we possess no  data whatever about
                             -118-

-------
harmonic components higher than  q = - N/2.  If we now attempt  to




reconstruct the waveform  f(x)  by inverse discrete Fourier  trans-




form upon the truncated set of modes  g(A  ), q=0, ... 2'  the only




data recoverable about the spatial behavior of  f(x)  is  precisely




its values at just the discrete points  x   (p=0,  ... N-l).   That




is, the data content of the truncated discrete spectral represen-




tation of the waveform is precisely equivalent to that contained




in the  N  distinct functional values  f(x ); P=0, ... N-l.  There-




fore, the spectral representation of this data contains precisely




N  distinct degrees of freedom.  To see why this  is so, consider




the Fourier series representation of a real function periodic with




fundamental wavelength  L, sampled at the uniformly spaced points




x  (p=0, ... N-l), viz:





         f V = aO + q^ Uq C03^ + bq Sin^) '






If this system of equations is to be uniquely soluble for the set




of Fourier coefficients  a's  and b's,  then we can only  specify




a total of  N  Fourier coefficients.  The most general choice of




coefficients which are determinable for the data  string   f (x ) ,




... f(xNt_-i)  in the absense of any additional information about




the function  f,  is precisely the set
         V V V a2' b2 ••••

Note that the discrete Fourier representation of the  function




incorporates exactly as much data about the spatial structure




of the function  f (x)  as does the discrete string  f (x  )  ...  .




     How then do we interpret the correspondence between a real




physical variable which is a continuous spatial function  and
                             -119-

-------
its point-wise representation in a discrete-grid numerical model?



One approach is to assume that the physical variable is sufficiently



smoothly varying; that is, its assigned value at a discrete point



x   is representative of its behavior over some neighborhood  K-XO



of the point in question, although the notion of neighborhood is



not made precise.  In fact, for highly discontinuous processes,



as, for example, in the propagation of shocks, or small-scale tur-



bulent cascade processes such an assumption is extremely dangerous.



A more satisfactory approach is to consider the discrete-point rep-



resentation as only a " best-fit" to the data contained within the



numerically resolvable scales of the variable.  The finite cutoff



wavelength forces us to surrender any detailed data about small-



scale structure which cannot be determined entirely from the re-



solvable larger scales.  This interpretation does not require a



notion of smoothness.
                             -120-

-------
                           Appendix 2
                /
           Discrete Fourier Transform Representations


      The  bulk of the computational effort in the model we have

'developed is  expended in performing discrete Fourier transforms

 from  discrete three-dimensional real space to discrete three-

 dimensional wave space,  and vice versa, using the fast Fourier

 Transform (FFT)  as discussed in Appendix 4.   In this appendix

 the discussion is limited only to the discrete three-dimensional

 Fourier transform.  The  discrete three-dimensional Fourier trans-

 form  pair is  defined as  follows:
Let   f             be a three-dimensional field defined over the
      Jl'  J2'  J3
N, x  N_  x  N   real-space grid:
                                               J2=l,  ...  N2

                                               J.,=l ,  ...  N_ .
Let  F              be  a three-dimensional field defined over the
         T'   2 '   i
NI x N   x N_   discrete wave-space grid
          (kn, k_,  k_)  =  (K,Ak  K^Ak  K Ak)     K =1,  ...  N
            L    £•    J       L -  I   £•  i   3        1-         -L
                                               K =1,  ...  N .



Then  f            and  IT  -       „    are  a discrete  Fourier  trans-
       Jl" J2' J3        Kl'  K2'  3
form pair if  (with one  choice of  normalization)


                        .        N,    N0    N,,    9
    1}   FR.   R      =  i	r   y1    y2    y3  e"27T
                                                       Nl
                                                            (cont'd)

                              -121-

-------
(A2.1)



fcont'd)
(J2-

1)(K2-1)
N2
(J -1) (K -1)
+
J
                                                                J
(A2.2)

J-i , J,, , J
(
+

N N
yl y2
I L
3 K^=l K2=l
Kl-1) (J -I)
+
N2
?3 e+2,,1
z
K'=l
(K^-l) (J1-l)
Nl
(Kl-1) (Jl-1)
3 3
N !<[, K' K'
     The basis  functions are orthogonal  in  the  sense
 Na   +0  .(J -1)(K -K1)
 y    ± 2 TT i	a     a   a


j' = l
 a
                   N
                    a
                               N 6M (K ,K'
                                a N   a'  or
                                   a
                                                  (a=l,2,3)
                                                  (3=1,2,3)
wh e re


                 'l,  for  modulo[(K-K1),N]=0




                 0,  otherwise




delta function, modulo   N.



     Writing the functions  in  vectorial form
                                             ,   i.e., the Kronecker
         J  =  (J,,  J2,  J3)



         K  =  (K;L,  K2,  K ) ,
                            -122-

-------
wo have
             2-iri  j. (K-K1 )
                                Nl N2
If  f
      1, 2,U3
is a real array, then  F,
is a conjugate-
symmetric array  (the  conjugate symmetry being denoted by a star)
         FK1 K  K, =   |FK,   K,   K,  1    ,   where
           1,2,3     I  ^1'   2'   3
         K!=l  if   K.=l
                        if
Noting that the wave-space coordinates   K^,   (i = l,2,3)   are  indexed
relative to one rather than  zero   and  folded relative  to the  Nyquist
frequency  N/2,  the  following  table provides the  correspondence
between the computational wave-space coordinate   K.  of the algorithm
and the associated non-dimensional  physical wavenumber:
      K.   of DFT algorithm
                 1
                 2
                      corresponding physical
                            wavenumber
                                0
                                1
                N/2
                N/2 + 1
                N/2 + 2
                               N/2-1
                               N/2
                              -(N/2  -1)
                 N
                               -1
                             -123-

-------
                          Appendix  3

Elimination of Aliasing in Discrete Real-Space  Multiplications


     This appendix describes, in a  simplified two-dimensional con-

text, the algorithm developed by Orszag  (1971a)  for the efficient

alias-free evaluation of real-space products.   In  order to help

the reader become fully familiar with such  techniques,  we have

included all relevant mathematical  details.

     Let us define the two-dimensional real-space  grid  (Xj ,  Yj )
                                                           -L    «£
for  0 _< J , J  <_N-1;  that is,  X   = J^^x,  y    =  J  Ay,   Ay = Ax.

For compactness we will use  the vector notation J=(J1,J?).

Similarly, we define the two-dimensional wave-space grid  K=(k ,k ),

where - — _< k , k < j.  These are equivalent to the notation  of

Appendix 2, but for a shift  in coordinate origins.   Consider  now

two real functions of the variables  X   , Y      These will be
                                      Jl    J2°
denoted as


         gT = g(X  , YT )
          —       1    9


         hj = h(Xj , Yj )



defined over the real-space grid.  Their product


         W(XT   YT ) = w, =  q  h
             1'   2     —    3  ^Z

is fully aliased in the sense that  all spectral components | KJ >'•_- appec

falsely, within the spectral domain   |K| £  j.   The problem is to

construct a scheme for evaluation of  W  such that only those

spectral components for which   |K|  < l^-  are  properly included.
                              -124--

-------
      Define, as in Appendix 2, the  two-dimensional discrete


  ourier transform pair(1'2).
 (A3.1)    f  = £ p  e——- K-J =  say,  DFT,
           «^   „  i\   JM  — ~-~           J\



           |K|  <_ N/2





where   £  , |K|  ^ N/2  is taken to mean   -  ^ <  K. ,  K <  ~
        K                                    2-122



     The  inverse DFT is




                ,      _ 2-rri K-J

 (A3. 2)    F  = -L- Jf e  ~N      = o'1  fT
           K    2 '±. J                  J
           —   N  J —                  —
where  7  is  taken to mean  0 < Jn , J_ < N-l  .
       j                       -12-



     Then the DFT  of the fully aliased local  product  W,.   is
                                                        j


                                      27TJ K-J

                             g  h e

                        N  j
          W  = D   W  = -^  I g  h e    N

                         2
                               2iri K1 -J          ,  2iri K"-J   _ 2TTJ K-J

 (A3. 3)    =  -~  J   I   G_ e +  N        I  H  „ ie    N       e    N

            N2  J   K-    K              K"  ^




By rearrangement of order of summations we have
 ,-,„.    w     lVVr.   „   v  -2TTi(K-(K'+'K")) -J
 (A3. 4)    W   =  -^ I  I G ,  H „ I e -^- -  -  -    -

           -   N   K'  K" -   -  J
where
         GK =
         HK =  °    hJ '
(1)  The notation K-J  is  understood to mean the scalar  dot  product


    K1J1+ K2J2 +   K3J3-

(2)  The notation D is shorthand for the inverse Fourier  transformation

    operation applied to PK-
                              -125

-------
     The orthogonality  relationships can be written  as
          N-l   2-ni  (K  -(K'  + K")) J

                ~~
(A3.5)
         J1=0
     Since
                N     -1-    -1-     •*•   -1- = N, if K = (K'  +  K")  modulo N



                                     = 0, otherwise .



                 < ^-,  IK" I  <  5L  then   IK'  + K: I  <  N   and the
              j. • — 2   '  1 '  — I           i    z '  —


summation   (A3.5) - N,  if   K± = (K| + Kj) or  (KJ  +  KJ + N),  or



(K' + K" - N).  Similar orthogonal relationships  hold for summa-



tion over index J?.   Therefore, the summation over  index J^.


Therefore, the  summation



         \   ^L It J_  * T_    ,  .     ii % \ 	
e N v- v-
2 r
N , if
H

Ki
Ki
Ki
-r j\ /

+ K" =
+ KJ =
+ K" =
( • J

Kl
K -N
K +N


K2
and K'
K2


+ K2 -
+ K'2 =
+ K2 =

,
K2
K2-N
K +N
^-- >
and
         = 0,  otherwise
(A3.4) may therefore be written out fully as
                                                +

                        K2

                                   K2
                              -126-

-------
whence






 (A3. 6)   WT,   T,   =  W,
                                              -KUK +N
where
         WK   K   =                      G       H

           a' b    K
The first term on  the right hand side of (A3. 6),  W „  „ ,   is the
                                                    K-i- t K.p

                                                            — 1
desired alias-free convolution of  GT,  and  H^,  i.e., the  D
                                     K       — K


transform of  the alias-free product of  g_  and  h .   Each of
                 - -              j        j^


the other terms in A3. 6  implies the presence of aliased contri-



bution of the form    K  ±N ,   etc.   The task now  is  to isolate



the first term from the  other eight terms on the right  hand side



of  (A3. 6).  To achieve this in economical fashion, Orszag has



introduced the "shif ted-grids" procedure.  There are a number of



variants of this method,  but  we will choose the "quartergrid"



scheme as particularly convenient.   Let us define the four shifted




grids



                      s          S

                                            (p=lf "4)
where
^ ='(J]
-P
+ %
' J2 '
S2
•• -/p)
                   ,


                 P    P
         S  =  (1,1);
                           -127-

-------
 (T'hc shifted  grids are introduced for the purpose  of  developing


expressions which permit the isolation of the  alias-free term.)


     For each shifted grid (p=l,--4), define


                          2TTJ K-J

 (A3.7)   g     )  =   GKe  N     Sp
                         2TJ K-J
 (A3.8)    h,T     = I  He  N     -p
           ^q  '    V   *
            O      XV
            -p    -


 and  define



 (A3. 9)    W^       =  D'1  g     -h(   )
            K1'K2         (-S )    -S
                             -P     -P




 =  -^ I   I  I   I   GK,   K,  H     „  I  I  exp    <^i)   [(KJ+K--K )
   N^ K|  K^  K^  K2'  Kl' K2   Kl'   2 J1 J2           N        111
Again using the  orthogonality relationships




         I  I   exp
                                                 S!
         J, J.         ^ N
          1 "2
                        S2                    Sl        S?
                         ^—      T    ^  -        -L_        £•
         .- ^ -2  -2,  ,_2-  4  , _,   „   c  N



where
         K* , K*  = 0, ± N   only
Thus,  evaluating the quadruple sums  ^   ...  £   for each  of  the

                                      Ki
shifted grids in turn, we  find:



                              -120-
                                               V II

                                               K2

-------
  (A3.10a)
           §1
           KlfK2
-1  W
     KlfK2-N
                + i W.
                +  W
 (A3.10b)
           W
            K ,
                  1W
                     K1-N,K2
                                  iW,
 (A3..10C)
          WT
                        ,  «
               - w
(A3.10d)
           -4
                                          + i W.
                                -129-

-------
(A3.10d)  (Cont'd)
- i  W
                  +  i  W
                                         — W
                                            K1+N,K2+N
Thus :


           1, 2
                '^2
                              S         _

                           +  w~4      = 4v\r

                              K1'K2       Kl, K2
whence we have the  remarkable  result
 (A3.11)
         W
          K1>K2
       = T  I
            W,
                  p= 1 , • • 4
Finally, by performing  one  further DFT
(A3.12)
W.
                = DW
                     K1'K2
we recover the alias-free  product of  g   and  h  .
                                        J        J


     As outlined above,  the  grid-shifting procedure requires



twelve DFT operations, namely
          K
                -P
         H
          K
          -P
                        (p=l,  ••  4)
                         W
                          ~p
                             -130-

-------
Orszag  shows  that it is possible to save half of these operations



by a simple modification to the above scheme, as follows:



      (a)  use  only one diagonal pair of shifted grids,  say
                 J     and  J  ;

                 -1        -2



      (b) perform evaluations  (A3*7,  3.8);



      (c) replace (A3.11)  with


                                 2     s


                 W1        = -   I    W~p

                   K1'K2     2  p=l      K1'K2






where the prime  signifies that   W1        is not  yet  the  alias-

                                    1'  2

free convolution of  G^,   and  H ;
                       K         K


      (d) truncate  the  field  W  to  the wave-space region




                 0  £ (K2  4- K2





and let
                W       =   fw1      )
                 KlfK2      V  KX,K2'TRUNCATED'





      (e) obtain the alias-free  real-space  product  field by



performing operation  (A3.12).



     The corresponding algorithm for  three-dimensional systems



is considerably more  complex in implementation, but  is concep-



tually identical.
                             -131-

-------
                           Appendix 4
                    The Fast Fourier Transform


     The basic  algorithms  underlying the Fast Fourier Transform
are well documented in the literature (Gentleman and Sande  1966,
Coo ley and  Tukey  1965) .
     The definition of the discrete Fourier transform  (DFT) de-
fined over  the  one-dimensional string of  N  data points  D. (j^O,
. . . , N-l) can be  written as

               i  ^~ ^       ' v        ?  /N
 (A4.1)   T^ = —    I    D.  WDK  ,   W=e^ 17 ,   k=0 ,  ..., N-l.
          K  /N" j = 0    J

 (where we have  incorporated  /N   normalization in each direction,
in contrast to  the  one-way  1/N  normalization used in Appendix 2)
     Explicitly for  N=8,

        /8  TQ = D W° +  D-j^W0 +  D2W° + . . .
        /8  T, = D.W0 +  D..W1 +  DW2 + ...
            ,     .      ..      0
(A4.2)       -1    °      1      2
                   A      «
                 Q

                 Q
                    A       «       »
         /8 T  = DW  +  D-j^W  +  D2W

         /8 T  = DW°  +  DW3  4-  DW6
     Naively, to perform  the  evaluation of  T ,   one would compute
                                              K
directly from (A4.2).   This  would require 64 complex multiplications
and 56 complex additions.   However,  the trigonometric function
T,jk    2irijk/N        2-irjk ,  .   .   2-rrjk            ^   .     ,_
WJ  = e   J    - cos  — rj* — + i sin — ^ —  possesses two important
properties namely:
      (1) Periodicity  ...  WN - e2TTiN/N  = 1^  SQ thafc fQr  N=g^

         W8 = W° =  1,  W9  =  W, W10 -  W2, etc.
      (2) Symmetry  ...  For N a power  of two, we may utilize the
                              -132-

-------
          reflections  of the cosine and sine functions in the

                           2345
          eight  octants:   W  = i,  W  - iW, W  = -1, W  = -W, etc.
     Hence, expressions  (A4.2)  reduce to 32 multiplications and


56 additions:
 (A4.3)



          /8 TQ =  DQ  +  DI  + D2 + D3 + D4 + D5 + DS + D?





          /8 T, =  D_  +  D,W + D^i + D_iW - D. - D._W - Dri - D_iW
              10123      4567




          /8 T2 =  DQ  +  DI± - D2 - D3i + D4 + D5i - D6 - D?i





          /8 T- =  D   +  D iW - D,i + D3W - D  - D iW + D i - D_W





          /8 T4 =  DQ  -  DI  + D2 - D3 + D4 - D5 + D6 - D?





          /8 T  =  D   -  D-j^W + D i - D3iW - D^ + D^ - DQi - D iW






          ^ T6 =  °0  -  °li - °2 + °3i + °4 - V * °6 + °7i




          /8 T  =  D   -  D iW - D i - D_W - D. + D-iW + D i + D W  .
              /    U     J-       ^-•J^tJ      O     /





          By adroit grouping of terms in (A4.3), it is possible to



reduce the amount of computation still further.  Proceeding sys-



tematically, define:





         A0 = DQ//8



         Ax = D4//8



         A2 = D2//8



         A3 = D6//8



         A4 - D1//8



         A5 = D5//8




                              -133-

-------
 (A4.4)  (Cont'd)
     A6 =  D3//8
        -  D?//8
Next:
(A4.5!
     Bo = Ao  +  Ai
     Bi = Ao  -  Ai
     B3 =  (A2-A3)i

     B4 = A4  + A5
     B5 = A4  - ^
     B6 = A6  + A7
     B? =  (A  -A?)i
     C0 = B0
     C2 = B0 - B2
     C3 = Bl - B3
     C4 = B4 + B6
     C6 = 'W1
     C? = (B5-B?)iW
        = C  + C
      0     0    4
        = r  + c
      1    Cl   °5
     T3  = C3 + C7
     T4  = CQ * C4
                              -134

-------
(A4.5)  (Cont'd)
T  —
 6


T  =
             - C
               C
             — C
               C
     Excluding the normalization  of  (A4.4),  we  have  accomplished


the DFT in only 5 complex multiplications  and 24  complex  additions.


In general, when  N  is  a power of two,  the  Fast  Fourier  Transform


(FFT) performs a DFT in  Only   j   log2N   complex multiplications


(for N >_ 32)  and only   N log0N   complex additions.  Comparatively,
                              .£      ~
                                         2                         2
the naive approach to  (A4.2)  requires  N  multiplications and   N


additions.


     The flow diagram below displays the FFT technique  of (A4.4)


and  (A4.5).  A circle of the  form  It/ y  means to add  (subtract)


all the values flowing in from the left  and  then  to  multiply by


w".  An empty circle means do nothing to the incoming value.

-------
     Notice that the first stage of the calculations,  (A4.4),

D'Tforms a peculiar permutation of the data.

                Index before        Index after

                     0                   0
                     1                   4
                     2                   2
                     3                   6
                     4                   1
                     5                   5
                     6                   3
                     7                   7


     Writing  these  indices in binary notation, we see why this

process has received the name "bit reversal":

                Index before        Index after

                    000            '     000
                    001                 100
                    010                 010
                    Oil                 110
                    100                 001
                    101                 101
                    110                 Oil
                    111                 111



     This sorting is required for all the major variations of

the Fast Fourier Transform algorithm.   It can be absorbed into

the computational stages of the FFT, but only at the cost of

doubling the amount of computer storage.  The time required for

this permutation is generally about 25% of the computational

stage time.   Thus, it is normally better to trade time for

storage and perform bit reversal explicitly.   (Notice that

no additional storage is needed to perform bit reversal, since

all data are exchanged in pairs.)

     It is easy to generate the sequence of integers 0.  1, 2,

etc., each from the previous one.  In binary notation, the process
                             -136-

-------
may be stated  as:   add  one  to  the  low  order bit,  and carry upward




if it overflows  (becomes  larger  than one), repeating with higher




order bits  until no further overflows  occur.   Similarly, beginning




with 0, which  is certainly  the bit reversal of the integer 0, we




generate the reversal of  the successive  integers  by an analogous




process:  add  1 to  the  highest bit of  the  reversed index; if it




overflows,  carry one to the next-lower bit, repeating with lower




order bits  until no further overflow occurs.




     E.g., for  N=8,  we start  with 000 and add one to the high-




order bit:




     100 =  4.




Generate the next reversed  integer by  adding  a high order one:




     200+010 = 2.




Again add one:




     110 =  6.




And again




     210^020-^001 =  1.




And so forth.



     In a high-level language  like Fortran, an integer may be




tested for  the value  (0 or  1)  of the high-order bit by comparing




it to N/2;  the high bit is   1  if  the  integer is  greater than N/2.




If the high bit is   0,  then we  must test  the second highest bit




by comparing to N/4, and  so on.
                             -137-

-------
                          Appendix 5



        Construction of Initial Perturbation Flow  Field







      The initial flow fields specified in this model, as  described



in chapter 6, are of two types:



      (a) the mean velocity field, with a suitable shear structure



          U(z) ,



      (b) the perturbation velocity field, which is comprised  of



          a large ensemble of small-amplitude incompressible modes,



          distributed in statistically homogeneous and isotropic



          manner, and obeying a designated spectral power  law   E(k).



          This Appendix describes the process by which such a  per-



          turbation field is constructed in wave space.



      Let us  define the two random processes   = <8 .  (k ) B .(k1' )>=0
                                       i    3  —        i —  ] —
 (kyk1)   where  <••>  designates ensemble averaging, and  k  is the



 discretized wavenumber vector.   The random processes  a  and  3



 can  be  obtained from any of a variety of pseudo-random number



 generation  algorithms.  Next, define the vector potential field



 in wave space as





         A(k)  = { [E(k)]1/2/k2} +  | 2 SLn a (k) |1/2 e27TlS (-} .





     Then,  we construct the initial perturbation velocity field



 in wave space  U(k)   as the curl, of  A(k)





         U(k)  = ik  x ,A(k)
                            -138-

-------
which automatically guarantees incompressibility-  Finally, the



real-space perturbation velocity  field  V(x)  is obtained from



U(k)  by discrete Fourier transformation.
                            -139-

-------
                                    TECHNICAL REPORT DATA
                             (Please read Instructions on the reverse before completing)
  REPORT NO!~
    EPA-600/4-76-007
                              2.
 4. TITLE AND SUBTITLE

   SPECTRAL MODELING OF  ATMOSPHERIC FLOWS AND
   TURBULENT DIFFUSION
                                                           8. PERFORMING ORGANIZA IION CODE
7. AUTHOR(S)      "	"	'	

   Arthur Bass, Steven A.  Orszag
                                                            8. PERFORMING ORGANIZATION REPORT NO.
                                                           3. RECIPIENT'S ACCESSION-NO.
                                                           6. RFt'OHT DATE
                                                              January 1976
 9. PERFORMING ORGANIZATION NAME AND ADDRESS
   Flow  Research, Inc.   (N.E.  Division)
   1  Broadway
   Cambridge, MA  02142
                                                           10. PROGRAM ELEMENT NO.
                                                             1AA009
                                                           11. CONTRACT/GRANT NO.
                                                              EPA 68-02-1297
 12. SPONSORING AGENCY NAME AND ADDRESS
   Environmental Sciences Research Laboratory
   Office of Research  and Development
   U.S.  Environmental  Protection Agency
   Research Triangle Park,  North Carolina  27711
                                                           13. TYPE OF REPORT AND PERIOD COVERED
                                                             Final
                                                           14. SPONSORING AGENCY CODE
                                                             EPA-ORD
 15. SUPPLEMENTARY NOTES
 16. ABSTRACT
         This report presents a survey of discrete  spectral  and pseudospectral
   numerical methods to  simulate atmospheric flow  and turbulent diffusion.
   Some  applications of  these methods to air quality simulation modeling are
   presented.  A three-dimensional  spectral incompressible  numerical model  is
   described in detail.  Computational resource  limitations precluded successful
   evaluation of eddy Austauch coefficients.  Some numerical  results are presented
   for the rate of relaxation of anisotropic flows.

         Recommendations  and  suggestions for further research  are made concerning
   the prospective utility of these spectral methods for air  quality simulation
   modeling.
17.
                                KEY WORDS AND DOCUMENT ANALYSIS
                  DESCRIPTORS
                                               b.lDENTIFIERS/OPEN ENDED TERMS
                                                                        c.  cos AT I  Field/Group
   * Numerical  analysis
     Turbulence
   * Turbulent  diffusion
     Air  flow
     Air  pollution
     Environmental simulation
     Mathematical  models
                                                                               12A
                                                                               20D
                                                                               04A
                                                                               13B
                                                                               14B
 8. DISTRIBUTION STATEMENT

           RELEASE TO PUBLIC
                                              19. SECURITY CLASS (This Report)
                                                   UNCLASSIFIED
21. NO. OF PAGES
     151
                                               20. SECURITY CLASS (This page)
                                                   UNCLASSIFIED
                                                                         22. PRICE
EPA Form 2220-1 (9-73)

-------