January 1971 15 UDC 651. M)s. 313:551.611.2 NUMERICAL INTEGRATION OF THE PRIMITIVE EQUATIONS WITH A FLOATING SET OF COMPUTATION POINTS: EXPERIMENTS WITH A BAROTROPIC GLOBAL MODEL FEDOR MESINGER 2 Department of Mefeorology, University of California, Los Angeles, Calif. ABSTRACT A Lagrangean-type numerical forecasting method is developed in which the computational (grid) points are ad- vected by the wind and the necessary space derivatives (in the pressure gradient terms, for example) are computed using the values of the variables a t all the computation points that at the particular moment are within a prescribed distance of the point for which the computation is done. In this way, the forecasting problem reduces t o solving the ordinary differential equations of motion and thermodynamics for each computation point, instead of solving the partial differential equations in the Eulerian or classical Lagrangean way. The method has some advantages over the conventional Eulerian scheme: simplicity (there are no advection terms), lack of computational dispersion in the, advection terms and therefore better simulation of atmospheric advection and deformation effects, very little incon- venience due to the spherical shape of the earth, and the possibility for a variable space resolution if desired. On the other hand, some artificial smoothing may be necessary, and it may be difficult (or impossible) to conserve the global integrals of certain quantities. A more detailed discussion of the differencing scheme used for the time integration is included in a separate sec- tion. This is the scheme obtained by linear extrapolation of computed time derivatives to a time value of to + aAt where to is the value of time at the beginning of the considered time step At and where a is a parameter that can be used to control the properties of the scheme. When choosing a value of a between % and 1, a scheme is obtained that damps the high-frequency motions, in a similar way as the Matsuno scheme, but needs somewhat less computer time and, with the same damping intensity, has a higher accuracy for low-frequency meteorologically significant motions. Using the described method, a 4-day experimental forecast has been made, starting with a stationary Haurwita- Neamtan solution, for a primitive equation, global, and homogeneous model. The final geopotential height map showed no visible phase errors and only a modest accumulation of truncation errors and effects of numerical smoothing mechanisms. Two shorter experiments have also been made to analyze the effects of space resolution and damping in the process of time differencing. It is felt that the experimental results strongly encourage further testing and inves- tigation of the proposed method. 1. 2. 3. 4. 5. 6. 7. CONTENTS 15 16 18 21 21 22 22 22 24 27 28 28 1. INTRODUCTION AND FORMULATION OF THE METHOD Ever since the beginning of the development of modern meteorology, Lagrangean methods have had a strong appeal to theoretical meteorologists. The laws of motion and thermodynamics relate the forces and heating to the individual derivatives of velocity components and po- 1 UCLA Department of Meteorology Contribution No. 180 2 Present affiliation: Department of Meteorology, University of Belgmde, Yugoslsvh tential temperature; and in this way, the Lagrangean approach is the most fundamental one. Nevertheless, due to practical reasons, during the last two decades of extensive research in numerical prediction and atmos- pheric general circulation experimentation, the Eulerian method of formulating and solving the relevant atmos- pheric equations has been used almost exclusively. There has been only a modest number of attempfs to use a Lagrangean or a quasi-Lagrangean technique. Graphical forecasting methods are essentially Lagrangean (Fjgrtoft 1952, 1955); however, their use has diminished with the easier availability of fast digital computers. Anf example in the paper by Welander (1955) illustrates the major difficulty to be expected in trying to perform a numerical forecast with the use of the classical Lagrangean appradch : a chessboard set of fluid elements soon deforms into a highly irregular pattern, unsuitable for space differentia- tion along the originally straight material lines. Thus, frequent redefinition of the material coordinate lines would be necessary. Besides, the computational complexity of this approach appears to be considerable; see, for example, the analysis by Wiin-Nielsen (1959). To avoid these problems or for other reasons, a number of investigators have used the Lagrangean formulation of 16 MONTHLY WEATHER REVIEW Vol. 99, No. 1 the advection terms only (qkland 1962, Krishnamurti 1962, Leith 1965). In this way, an advective change at a space-fixed grid point and in a particular time step is essentially obtained by finding the location of the parcel that mill in this time step be carried by the flow to the grid point and by computing the corresponding individual value through space interpolation. Since by this procedure a space interpolated value is advected a t each grid point and time step, i t is not clear whether the method has any Lagrangean-type advantages, It does represent a simple possibility to avoid the “nonlinear” computational instability; however, it is questionable whether the gain in simplicity, as compared t o the more involved application of the Arakawa (1966) concept of maintaining the integral constraints of physical im- portance, will often be of more value than the advantages of the latter alternative. In fact, it has likely been to a rather large estent due to the wide application of the Arakama conservation schemes that the remarka,ble achievements in atmospheric simulation experiments have been accomplished in recent years. Besides, other less refined possibilities to overcome t’he nonlinear instability problem also exist. Despite this outstanding performance of the present day Eulerian methods, they do contain a number of intrinsic deficiencies. It appears convenient to mention them briefly: No attempt can be made to conserve, within advection terms, individual properties following the motion of air particles. That is, n false dispersion of the wave components of the flow is produced t )y the finite-difference formulation of the advection terms (Wurtelc 1961, Matsuno 1966b). The error in phase speed of individual waves increases cata5trophicallg when the wavelength approaches the smallest possible wavelength resolvable by the grid. This precludes the formation of sharp gradients of an advected quantity, such as the gradients observed in frontal and intertropical convergence zones, and smooths out these zones in case that some initially exist. Another consequencc of the Eulerian formulation of advection terms is that the nicntioned process of elongation, tangling, and subse- quent mixing of fluid elements is not realistically sirnulated (DjuriE 1966). It seems futile in Eulerian methods to try alleviating this wave dispersion difficulty by having a higher grid resolution in zones of sharp gradients. These zones are usually moving and after some time would so leave the regions of high resolution. Besides, a space discontinuity in grid resolution results in false reflection of waves at the boundary separating the two regions. The task of defining the locations of grid points on spherical earth so as to accomplish a quasi-homogeneous space resolution, though feasible in a number of straight-forward ways, is an uncom- fortable one, leading to laborious computational schemes (Kurihara and Holloway 1967, Sadourny- et al. 1968). A Eulerian model accepts information a t predetermined grid- point locations only, while the observations are and will be made at relatively randomly spaced points. This restriction may be of some disadvantage, especially when, as expected, a crucial part of the atmospheric observations becomes continuous in time. In trying to remedy the leading of the mentioned shortcomings, i t was proposed by Charney (1966) to let the grid points move, as if imbedded in some hypothetical medium, in such a way that the grid spacing constantly adjusts to the gradient of the quantity one wishes to . resolve. In this way, one would have two simultaneous flows: the physical flow of the fluid and the flow of this hypo thetical medium. Charney rejected the possibility of allowing the two fluids to coincide since, as described already, the grid lines moving with the physical flow would soon become highly distorted and useless for the com- putation of space derivatives. It is suggested here to try the other alternative: let the computation points move with the physical fluid and, instead, reject the requirement for the continuity of their relative position. The necessary space derivatives have then to be computed by using the values of the variables at all the computation points that a t the particular moment happen to be in some way surrounding the point for which the computation is done. Even if they initially are not, after some time, say a number of days, these surrounding points will be dis- tributed essentially a t random; it is conceivable that this may result in a significant increase in the number of computation points necessary to achieve some required accuracy in space differencing. But, in return, the prognostic equations reduce to a set of ordinary, instead of partial, differential equations, with a considerable gain in simplicity. Besides, no errors are produced by the finite- difference treatment of the advection terins since there are no advection terms. The proposed method has been tested using a global primitive equations model for a homogeneous and incom- pressible atmosphere. Following sections describe the technical details of the computational procedure as well as the results of a number of performed experiments. Finally, a comparison of the properties of the method with those of the conventional Euleriaii schemes will be given. 2. SPACE DIFFERENCING Throughout this study, we shall deal with a homo- geneous, incompressible, and frictionless atmospheric model. Although simple, such a niodel contains a sig- nificant part of the essential features of the large-scale atmospheric motions ; see, for example, the discussion in the paper by Arakawa (1970). For these reasons, i t has been used many times to test the performance of the numerical finite-diff erence schemes. The most recent reference describing such a use is probably the one by Grammeltvedt (1969) ; it contains a number of the previous ones. As is customary, the pressure change with height is assumed in the present study to be given by the hydro- static equation. The pressure force is then not a function of height; and if the initial wind is independent of height, i t will have to remain so. We shall assume this to be the case. We shall further consider earth to be a perfect sphere and choose a A, 0, r spherical coordinate system with, for later convenience, an arbitrary orientation of the O= f 7r/2 axis. Having a shallow atmosphere, we replace the radius T with a constant value a, and accordingly (Phillips 1966) neglect the vertical velocity terms in the equations of horizontal motion. The governing equations Fedor Mesinger 17 January 1971 then become and Here, U and V are the horizontal velocity components in the directions of constant 0 and constant A, respectively; + is the geopotential of the free surface; t is time; time derivatives are individual rates of change, following the motion of the fluid; fi is the magnitude of the earth’s angular velocity; and p is latitude. We shall further use h for longitude and denote by u and v the eastward and northward components of the horizontal velocity, respectively . The problem of space differencing now consists of computing the four space derivatives in eq (1). We want to do this by making use of the known values of u, v, and + a t the point for which the computation should be per- formed, which we shall call the “reference point,” and of the values of these variables a t a number of surrounding computation points, which we shall call “neighbors.” The space distribution of these points is considered to be arbitrary, except for assuming that they are not or- ganized in some special inconvenient way (along a line, for instance). This distribution is given by the known values of their geographical coordinates h , cp. For a de- fined orientation of the A, 8 system, the components u, v and coordinates h , cp can be transformed readily, of course, into components U, V and coordinates A, e, if this definition is such as to make this necessary. Now, we may expect that the property of being a neighbor has to be a reciprocal one. If it were not, we would have a situation in which the values of dependent variables a t a computation point affect the time change of those a t another one, without the opposite being true. It would appear that this should lead to an instability of the computation; indeed, some test runs support this con- clusion. The simplest way of making the property of being a neighbor reciprocal is to take as neighbors all the computation points which a t that particular moment happen to be within some prescribed distance p of the considered reference point ; this possibility has been used in the present study. Now, a method for the computation of space derivatives will always require a certain minimal number of neighbors; to make sure that we always have a t least that many neighbors in a computation with a randomlike distribution of points, we then have to make p such that the average number of neighbors is consider- ably greater than this minimal one. I n this situation, the method of least squares offers a straightforward way to compute the space derivatives, and it has been used here. The accuracy of the least-squares fitting, if performed in terms of the geographical coordinates A, cp, would ob- viously break down when approaching the two Poles. For avoiding this, the least-squares fitting was done in the coordinate system A, 0, which had its origin always in the considered reference point, and e= &~/2 axis in its meridional plane, with 0 increasing northward. This choice of origin has an additional benefit of simplifying eq (l ), in that the three tan 0 terms disappear, if computed forward in space. The necessary coordinate transformation equations can be written in the form A = arctan (Y/X) e=arctan(Z/JX2+ y2) X= cos cpo cos cp cos (h - h,) + sin cpo sin cp, and where Y=coscp sin(h--Xo), Z= -sin cpo cos cp cos(X--Xo) + cos cpo sin cp, 0 and ho, cpo are the geographical coordinates of the refer- ence point. Except along the meridian through the refer- ence point, the coordinate lines of the A, 8 system, in general, make an angle with respect to those of the h, cp system. Denoting this angle by P and having it increase in the positive direction, we obtain from eq (2) sin p=-sin cpo sin(h-Xo)/JX2+Y2 and (3 1 cos p=[cos cpo cos cp + sin cpo sin cp COS(~-A~)~/JX~+Y~. The velocity components U, V can then be computed making use of the relations U=u cos p+v sin P and . (4 1 V=-u sin B+v cos 0. After having in this way performed the transformation of coordinates A, cp and components u, v into coordinates A, e and components U, V, the least-squares fitting was done in the experiments reported here by defining Here, x stands for any of the variables, U, V, and 4, and 18 MONTHLY WEATHER REVIEW Vol. 99, No. 1 xo is the value of x at the reference point; the five coeffi- cients are computed so as to make 2 differ as little as possi- ble, in the least-squares sense, from the values of x a t the individual neighbors. Having, as described, more than five neighbors, the two will have to be, in general, different. If we wish, we can prescribe some weights w to the squared differences a t the individual neighbors and compute the coefficients so as to minimize the sum s=2 Wt(X,-;;,)2. i =l The subscripts here refer to the values of the variables a t particular neighbors, the total number of neighbors being n. In the experiments that will be described here, the same weight has been prescribed always to all the neigh- bors. It might be possible, however, to optimize the weights of individual neighbors in some way, say as a function of their space distribution or the reliability of the x values a t particular neighbors. NOW, obviously, we assume (7) a) a) au av an + ae- --&A , --B+, aK=A,, and - ae --=€ I,; the computation of space derivatives in eq (1) is in this way defined. A few words could be added about the choice of the neighbor-defining distance p and the search for neighbors. Preliminary experiments with the present method have shown that a tendency for the appearance of computa- tional noise exists, this tendency becoming catastrophic with the decrease in p. This noise may be partly a result of the constant change of neighbors; this change must be associated with some disruption of the balance in the fields. Another source of the noise should be expected in the errors in space differencing that are produced by in- convenient space distribution of the neighbors. For keeping this noise tendency a t an acceptably low level, it was found necessary to have a fairly large average number of neighbors, Z. The experiments reported here have all been done with the value of p=a, arccos 1-2 - ( 3 such as to make E equal to 30; here, N i s the total number of computation points. With such a large value for E , no need was felt to check whether the particular values of n are, a t each reference point and time step, always greater than or equal to 5. For enabling a fast search for the neighbors, the globe was divided into spherical grid boxes; and a t the beginning of the computation and after each displacement of the computation points, lists of addresses have been made of all the points found in particular boxes. For a specific reference point, then only the points found in boxes that are within the distance p of the reference point have been checked for the possibility of being neighbors. This procedure appeared to be fairly fast, in the sense that it did not consume a major portion of the total computation time. 3. TIME DIFFERENCING: THREE-LEVEL EXTRAPOLATION SCHEME It has earlier been pointed out by Matsuno (1966a) that a discontinuity in the grid size of a Eulerian grid can result in a false generation of high-frequency waves. As pointed out, this apparently takes place also in our case of an irregular distribution of computation points. For such situations, Matsuno recommended use of a computational scheme that filters out high-frequency oscillations during the process of time integration. His first-order accuracy scheme (Matsuno scheme) has since been used extensively in some of the general circulation studies (Mintz 1965, Arakawa et al. 1969). It was pointed out by Lilly (1965) that this scheme may possibly cause a considerable loss in the energy of the system; on the other hand, for the purpose of initialization for the primitive equations model, some authors felt the need for a scheme with even more powerful damping of the high-frequency motions (Nitta 1969). Use of such high-frequency damping schemes is likely to become much more extensive with the expected advent of the time-continuous observations: we may then wish to use these schemes to speed the adjustment of the imbalances introduced by continuous input of the data into a numerical €orecast. In view of this general interest and possible advantage in having a variable-damping scheme, a scheme that was analyzed for use in the present model and can have a variable damping will be described here in more detail. This is the scheme obtained by per- forming a linear time extrapolation of computed time derivatives to a time value of to + aAt where to is the value of time at the beginning of the considered time step At and a is a parameter that we can use to control the properties of the scheme. We shall write a particular one out of eq (1) in the general form ax -=j( a t x, t ), x =x( t ) . In writing this, we have disregarded the coupling within the system (1) in exchange for the possibility of considering x to be complex and thus having eq (9) stand for a pair of eq (1). If we now use an integer subscript to denote the number of time steps elapsed, we can write the considered finite-difference version of eq (9) as For brevity, we shall sometimes call approximation (10) January 1971 Fedor Mesinger 19 TLE (three-level extrapolation) scheme. It reduces to the Euler (forward) scheme when a=O and to the simplified Adams-Bashforth scheme when a=jh. When a= 1, it represents a simulation of the backward difference scheme. If we define as the truncation error e of the considered scheme, the remainder on the right side of eq (10) when xr+, are substituted by their true values obtained through Taylor series expansion, we have ~=-(~-~)f:(At)"-(g+a)f:'(At)~+ . . . . Thus, approximation (10) is of first-order accuracy, unless a=%, when its accuracy is of the second order. To analyze the behavior of the error over a period of time, however, we have to prescribe first the function f(x, t ). It is of particular interest to consider the case of the oscillation equation ax . z =z w x since it describes the linearized gravity-inertial motions present within eq (1). We shall do so throughout the remainder of this section. We tentatively write our finite-difference solution of eq (11) in the form X,=&XT. (12) Combining eq (10-12), we then see that eq (12) is the required finite-difference solution of eq (1 1) , provided that (-A+B)'/'I X=+[1 f +(A +B y 2 ] +i*[ (1 +a )p i ~ 211-Ul 1-a (13) where p=wAt, A=3[1- (l+~)'p'], and B=$[1+2( 1 -6a+az)p2+( 1 +a) 4 p 4 ] l/Z* In eq (13), the upper signs correspond to the "physical," and the lower correspond to the spurious "computational mode" that appears through the use of the three time levels differencing scheme. Now, we can also write h=Iileio. (14) Since the true solution of eq (11) can be written as we see it appropriate to refer, as customary, to 1x1 as the amplification factor and to O/wAt as the relative phase change (per time step) of the finite-difference solution. For stability, it is now necessary to have both values of 1x1 5 1; for accuracy, it is desirable to have 1x1 and O/wAt of the physical mode close to unity and to have 1x1 of the computational mode as small as possible. We shall first investigate the stability and damping properties of the considered TLE scheme. From eq (13), we obtain ' I X I = &[ 1 + ( 1 + a) 'p2 + B f (A + B) '1' 1--a2 5 --~p(-A+B)"']'~'. (15) 11-4 The right side of eq (15) is equal to unity when w A t =-(- 1 1-2a ---) "' . a l+% .. The TLE scheme is, therefore, stable for values of wAt equal to or smaller than the right side of eq (16). This right side has its maximum value (wAt) ,,,az=42(5&- 11) = 0.6006 when a= (1 +dg) =0.8090. (17) Thus, as far as the extent of the stability range is con- cerned, eq (17) gives the optimum value of the parameter a. Amplification factors given by eq (15) for this optimum stability and three other particular values of a are plotted against wAt in figure 1. It demonstrates a favorable damping of the computational mode for values of al , maximum damping occurs at the intersection of the two 1x1 curves. This happens at wAt=l/(l+a). (20) 3 - 2~ & 23/2( 1 - a ) '/' 1 Both expressions (19) and (20) are of course valid in the limiting case when a= 1. A summary of the dcpendence of considered wAt values on a is given in figure 2. Thc figurc shox-s the extent of the stability range, eq (16), and wAt values at which the considered TLE scheme has a maximum damping, eq (19) with upper signs for arcssioi~ as to the amount of noise present in the final gcopotential field is to some extcut deceiving. A fair amount of noise can be seen along the Equator \\-here thc field is also flat; but the remaining contour lines in regions of stronger gradients are often remarkably smooth. In figure 5, the initial (npper map) and final (lower map) space distribution of computation points is shown on an cyual-area cyliiiclrical projection so as to reflect their number per unit area of the globc. One can notice that, in the 4-day map, the points exhibit a tendcncy to organize dong irregular lines; this must be a consequence of the fact that, except for somc iioisc, 110 turbulence \vas pre- sent in the computed flo\v fields. A more irregular final distribution of points should of course be expected when the points arc advccted by a less orgunizcd flow. Starting with an initial distribution of points that is better than random, onc must expect i t to deteriorate with time and eventually become random or possibly 11-orse than random--“bctter” and “v-orsc” meaning that the points are more and less c\-enly dispersccl than in a random distribution , rcspcctivcly. In fact, somc earlier cxperimcnts (Mesinger 1965) have sho\\-ii that the iiondivcrgcnt part of an observed t\\-o-dimciisioiial atmospheric flo\\- had thc property of producing as well as maintaining a raiidom distribution of initially organized and floating particles while divergent flows maintained distributions worse than random in the above sciisc. The time rate of this process is of‘ interest and d l be discussed here along with its effect on the accuracy of the forecast. For giving a statistical description of the distribution of our computation points, it is convenient to consider the mean distance from each point to its nearest neighbor. It has been shown by Hertz (1909) that, when a very large number of points is distributed at random on a plane and the unit of length chosen so as to make their average concentration equal to 1, this mean distance will be equal to 0.5. If, however, the points are distributed at random only relative to one another but with a concentration that is a function of space, this mean distance has to be less than the random distribution value of 0.5 (Mesinger fn zs n I 2 3 4 TIME (DAYS1 FIGURE 7.-Root-mean-squcwe individual change per time step in the geopotential height as a function of time. The values refer to the 4-dag- experiment. 1965). Finally, if the points are organized relative to one another, this mean distance can of course be greater; it is, for example, equal to &/4.&=1.0746 when the points are arranged into a hexagonal grid. Apparently when the mean nearest neighbor distance is greater, me might expect a better performance of a particular distri- bution of computation points. This mean distance was computed at 1-hr intervals in the 4-day experiment: its initial value, corresponding to the distribution shoxn in the upper map of figure 5, was equal to about 0.7555; i t decreased rather slonrly later on, reaching a value of about 0.6319 at the elid of the experiment. The obtained mean distances as a function of time are shown in figure 6. It can be compared with figure 7 in the aforementioned paper by the author; in that previous experiment, the iiondivergent part of an observed flow needed much less time, only about half a day, to accomplish the same de- crease in the mean nearest neighbor distance. The dif- ference, again, is due to the lack of turbulence in the present experiment. For obtaining a more quantitative idea of the amount of various kinds of noise present in the 4-day experiment, a record of the root-mean-square (rms) individual change per time step in the geopotential heights is shown in figure ’i-also based on computations at 1-hr intervals. Computed per unit time and divided by the rms value of the geopotential height, this quantity would give the mass-weighted rms value of div (divergence) v. Thus, i t is equal to zero in the analytic solution and does not depend on time. I n the experiment, the initial value of the rms individual geopotential change was equal to about 3.38 m2 s-’/12 min; this corresponds to the value of about 0.5X10-’ s-l for the mentioned value of div v. Compared with the Pole-to-Equator change in the free-surface height of somewhat more than 1600 gpm, this initial error of about % gpm/12 min appears to be moderate; in fact, some caution should be exercised here since these numbers may already be influenced by the round-off error of the performed IBM single precision computations. Since there is no noise in the initial fields, this error-assuming no round-off influence-reflects only the difference be- 26 MONTHLY WEATHER REVIEW Vol. 99, No. 1 zi I 2 3 4 TIME IDAYS) F~GURE 8.-Average height of the free surface as a function of time for the &day experiment. tween the analytic space derivatives of velocity com- ponents and those computed by the least-squares fitting of a second-degree polynomial to the values a t available neighbors. Thus, it shows a combined effect of the incon- venient space distribution and number of neighbors and of the truncation of higher order terms in the polynomial. Some feeling for the relative significance of these two factors can of course be obtained through suitable experi- ments. For instance, the 3000-point run, with points arranged into a 120x25 pattern, had this initial error equal to 3.97 m2 r2/12 min; this must have been mostly a consequence of the increased importance of the neglected higher order terms. On the other hand, a different organi- zation of the 4000 points into a 100x40 instead of 160 X25 pattern with a 0.7997 mean nearest neighbor distance was found to result in a reduction of this error to the value of 2.75 of the same units. During the first day or so, the considered quantity is seen to undergo vigorous gravity-type oscillations with a period of very nearly 3 hr. These oscillations must be set off by small systematic differences between the analytic and computed space derivatives. By the end of about 1 day, the oscillations are damped out through the process of time differencing. The parallel run, differing from the 4-day run only by being computed with the Adams-Bashforth time-diff erencing scheme, did not ex- hibit such damping; and the oscillations persisted without noticeable change in amplitude throughout the l-day period of the experiment. At times after about 1 day, the rms individual geo- potential change seen in figure 7 settles at the value of about twice the initial one, showing the contribution of the random space noise in the velocity fields. After about 2 days, a slow rise in this rms change can be noticed, interrupted by three irregular sudden increases. The gradual rise in this quantity should probably be attributed to the slow deterioration of the space dispersion of compu- tation points, and the three peaks to occasional occur- rences of very inconvenient configurations of neighbors. Performance of the model can further be analyzed by examination of the behavior of global integrals of a number of relevant physical quantities. Conservation of none of these integrals is formally guaranteed in the present method, and a change in conservative quantities thus reflects the effect of numerical errors or d the artificial smoothing mechanisms. A record of one of such quantities, the average height of the free surface, is shown in figure 8, again based on values computed a t l-hr intervals. This average was obtained by using the geopotential values interpolated to points of the spherical 72x46 point grid, by interpolation described earlier in this section. The interpolation protects the computed average values against the effect of possible systematic variations in the space density of computation points; this protection seems necessary since such variations would be correlated with those in the height of the free surface. The time changes in the average free-surface height shown in figure 8 appear to be random and of a modest amplitude. The observed amplitude of less than 2.1 gpm represents about 0.02 percent of the total average height. The considered average is of course proportional to the total mass of the fluid. It may be noted that, using eq (32), one can obtain (34) where is the global average value of 4. With the values of constants given in section 5, eq (34) gives 9057.33 gpm for the average free-surface height. The initial value of 9057.02 gpm entered in figure 8 happens to be slightly less than this "exact" value, the difference showing the amount of the sampling error. This exact value could also have been entered t o start the curve in figure 8, what would have further reduced its amplitude. Finally, a record of the average values of total, kinetic, and available potential energy computed also at 1-hr intervals is shown in figure 9. Denoting the total kinetic energy averaged per unit mass by K, and the available potential energy averaged in the same way by A,, we have and (35) (36) where S represents the total area of the globe. I n computing the values of K, and A,, shown in figure 9, these intergals and global average of 9 in eq (36) were substituted by corresponding sums and by the average of based on values of 4 and v2 interpolated again to points of the spherical 72x46 point grid. Thus, the values obtained for Kl and A, were also protected against systematic January 1971 Fedor Mesinger 21 TOTAL 650 ,n r AVAILAELE POTENTIAL 1 2 3 4 TIME (DAYS1 FIGURE 9.-Average values per unit mass of the total, kinetic, and available potential energy as functions of time. The solid curves refer to the Cday experiment; the dashed curves refer to the parallel 3000-point experiment, differing from the +day one only in the number of computation points; and the dotted curves refer to the parallel a= 54 (Adams-Bashforth scheme) experiment, dif- fering from the &day one only in the value of this parameter of the timedifferencing scheme. variations in the space density of computation points. Such a protection seems desirable since it appears that there should exist a definite correlation between these variations and those in the contributions to total energy values. However, it is this space interpolation that gives rise to the time noise seen in the kinetic (and thus also total) energy curves in figure 9; no time noise appears when the integrals in eq (35) are computed by taking simple sums of corresponding quantities over all the corn- putation points. Of the energy curves, the solid ones obviously refer to the 4-day experiment; the dashed curves refer to the “coarse resolution’’ 3000-point run, and the dotted ones to the run performed by using the nondamp- ing Ad ams-Bashf orth time-diff erencing scheme. An outstanding characteristic of the energy curves is a periodic exchange between the kinetic and available potential energy. These oscillations are apparently brought forth by the truncation of the higher order terms in the least-squares polynomial. Namely, it would seem that in this way one computes pressure gradients that are pre- dominantly smaller, in their absolute values, than the analytic ones. Since the initial Coriolis forces are not affected by the space-differencing method, an imbalance is obtained such as to set off inertia-gravity oscillations in which the fluid particles initially move predominantly toward higher free-surface heights; this is associated with an increase in available potential and a decrease in kinetic energy. The amplitude of these oscillations should then increase with an increase in the neighbor-defining distance due to the increased importance of the neglected higher order terms; this is confirmed by the dashed kinetic energy curve. Namely, since the average number of neighbors E was the same in all of the three experi- ments, the neighbor-defining distance of the 3000-point experiment was greater than that of the two 4000-point ones; accordingly, this 3000-point experiment exhibits a greater amplitude of the discussed oxcillations. I n the 4000-point 4-day experiment, this amplitude initially amounts to about 1.1 percent of the total kinetic energy and decreases later on. Damping in the time differencing and physical dispersion of the waves are responsible for this decrease in amplitude. The period of the oscillations is less than the pure inertial one, in accordance with the usual linear theory; see, for example, the paper by Arakawa (1970). The energy curves in figure 9 exhibit a general de- creasing trend, showing the loss of energy due to the artificial space smoothing and damping in the time differencing. During the 4-day experiment, the total kinetic energy decreases by about 3.71 percent of its initial value. One part of this energy is transformed into the available potential energy through the described inertia-gravity oscillations and subsequent geostrophic adjustment; it somewhat more than compensates for the loss in the latter to the existing space-time smoothing mechanisms, leaving the potential energy a t a value slightly higher than the initial one. The remaining part is dissipated through the numerical smoothing processes. It is tempting to try estimating the relative contribu- tions of space smoothing and time differencing to the observed kinetic energy loss. One may attempt to do this with the help of the results of the parallel run computed using the Adams-Bashforth time-differencing scheme. This run was terminated at the l-day time; a t that moment, its loss in kinetic energy amounts only to some- what less than one-third of that demonstrated by the solid kinetic energy curve in figure 9. However, some caution should be exercised in interpreting this difference since there is a possibility that it appears partly as a consequence of the weak instability of the Adams-Bash- forth scheme. No visual sign of such instability, though, could be noticed by careful comparison of the two 1-day geopotential height maps; thus, a value for the parameter a smaller than the one used for the 4-day experiment appears preferable. 7. CONCLUDING REMARKS The proposed Lagrangean method offers a number of distinct advantages over the conventional Eulerian tech- nique. We shall summarize them here: Formulation of the problem involves no advection terms; this makes the governing equations simple, free of nonlinear instability, and guarantees, except for truncation errors, the conservation of MONTHLY WEATHER REVIEW Vol. 99, No. 1 individual properties following the motion of fluid particles. In other words, there is no computational dispersion in the advection terms. This should improve the simulation of mesoscale phenomena, such as fronts, jets, and intertropical convergence zones, associated with sharp gradients of an advected quantity. Further, this enables a realistic simulation of the deformation-dependent process of lateral diffusion. The simulated flow is not subjected t o systematic space changes in grid resolution and properties. With computation points moving with the fluid, a quasi-homogeneous resolution is accomplished without any geometrical considerations. The two mentioned properties significantly reduce the amount of work needed to code a computer program for the model. Absence of considerations related to the geometry of a particular grid and to the treatment of advection terms relieves the programmer of a major effort that these considerations require in a Eulerian global model. It would appear easy to accomplish a variable resolution, such as t o have a higher density of computation points in regions of rapid spatial change in appropriate quantities. Such regions exhibit a strong tendency to move with the fluid and would thus maintain this increased resolution for an extended time. A density dependent neighbor-defining distance could then be used, while still observing the principle of property of being a neighbor having to be a reciprocal one. The model can accept information at arbitrary points in space, and they can be prescribed different weights, depending on their reliability. This seems ideally suited to the expected future global observation system (ICSU/IUGG 1967). On the other hand, a number of objections.could be raised : The present experiments showed a tendency for the appearance of computational noise, and artificial smoothing was needed t o suppress this tendency. However, the situation should be better when a simulation of the real atmospheric diffusion is incorporated into the model. It may be difficult or impossible to construct schemes that would conserve global integrals of relevant quantities. It seems that more computation time and more points are needed than in a comparable Eulerian model. The 4000-point experiments described here required about 50 min of the IBM 360/91 computer time per 1 day of simulated time. Finally, some uncertainty remains as to horn severe the performed test was; it would appear that a task of main- taining a stationary state, with computation points floating through it, is fairly demanding. Still, one feels a need for a more nonlinear two-dimensional test and of course a three-dimensional one. The amount of effort invested into the present method is exceedingly small compared to that which brought the Eulerian methods to the present state; this, coupled with the results of the present study, mould seem to highly encourage further investigation of the proposed method. ACKNOWLEDGMENTS The author is grateful t o Profs. Yale Mintz and Akio Arakawa for their encouragement of this research and for some helpful suggestions. M&ssrs. Wayne Schubert and Robert Haney kindly helped t o prepare the manuscript by proposing numerous im- provements in its text. Computing assistance was obtained from the Campus Computing Network, University of California, Los Angeles. REFERENCES Arakawa, Akio, “Computational Design for Long-Term Numerical Integration of the Equations of Fluid Motion: Two-Dimensional Incompressible Flow, Part I,” Journal of Computational Physics, Vol. 1, No. 1, J ~l y 1966, pp. 119-143. Arakawa, Akio, “Nunierical Simulation of Large-Scale Atmospheric Motions,” Numerical Solution of Field Problems in Continuum Physics, American Mathematical Society, Providence, R.I., 19710, Arakawa, Akio, Katayama, Akira, and Mintz, Yale, “Numerical Simulation of the General Circulation of the Atmospherc,’’ Proceedings of the WMOlIUGG Symposium on Numerical Weather Prediction, Tokyo, November 26-December 4, 1968, Japan Mcte- orological Agency, Tokyo, Mar. 1969, pp. IV-7-IV-12. Charney, Jule G., “The Use of the Primitive Equations of Motion in Numerical Prediction,” Tellus, Vol. 7, No. 1, Feb. 1955, pp. Charney, Jule G., “Some Remaining Problems in Numerical Weather Prediction,” Advances in Numerical Weather Prediction, Travelers Research Center, Inc., Hartford, Conn., 1966, pp. 61-70. DjuriC, DuBan, “On the Accuracy of Air Trajectory Computations,” Journal of Meteorology, Vol. 18, No. .5, Oct. 1961, pp. 597-605. DjuriC, DuSan, “The Role of Deformation in the Large-Scale Dis- persion of Clusters,” Quarterly Journal of the Royal Meteorological Society, Vol. 92, No. 392, Apr. 1966, pp. 231-238. Fj$rtoft, Ragnar, “On a Numerical Method of Integrating the Baro- tropic Vorticity Equation,” Tellus, Vol. 4, No. 3, Aug. 1952, FjGrtoft, Ragnar, “On the Use of Space-Smoothing in Physical Weather Forecasting,” Tellus, Vol. 7, No. 4, Nov. 1955, pp. 462- 480. Grammeltvedt, Arne, “A Survey of Finite-Diff erence Schemes for the Primitive Equations for a Barotropic Fluid,” Monthly Weather Review, Vol. 97, No. 5, May 1969, pp. 384-404. Haurwitz, Bernhard, “The Motion of Atmospheric Disturbances on the Spherical Earth,” Journal of Marine Research, Vol. 3, 1940, Hertz, Paul, “Uber den gegenseitigen durchschnittlichen Abstand von Punkten, die mit bekannter mittlerer Dichte im Raume angeordnet sind” (On the Average Mutual Distance of Points Distributed in Space With a Known Average Density), Mathe- matische Annalen, Vol. 67, Springer-Verlag, Berlin, 1909, pp. 387- 398. ICSU/IUGG-Committee on Atmospheric Sciences, Report of the Study Conference Held at Stockholm, 28 June-11 July 1967, on the Global Atmospheric Research Programme (G A R P ), World Mete- orological Organization, Geneva, 1967, 144 pp. Krishnamurti, Tiruvalam N., “Numerical Integration of Primitive Equations by a Quasi-Lagrangian Advective Scheme,” Journal of Applied Meteorology, Vol. 1, No. 4, Dec. 1962, pp. 508-521. Kurihara, Yoshio, and Holloway, J. Leith, Jr., “Numerical Integra- tion of a Nine-Level Global Primitive Equations Model Formu- lated by the Box Method,” Monthly Weather Review, vel. 95, Leith, Cecil E., “Lagrangian Advection in an Atmospheric Model,” World Meteorological Organization Technical Note No. 66, Geneva, Lilly, Douglas K., “On the Computational Stability of Numerical Solutions of Time-Dependent Non-Linear Geophysical Fluid Dynamics Problems,” Monthly Weather Review, Vol. 93, NO. 1, Jan. 1965, pp. 11-26. Matsuno, Taroh, “Numerical Integrations of the Primitive Equa- tions by a Simulated Backward Difference Method,” Journal of the Meteorological Society of Japan, Ser. 11, Vol. 44, NO. 1, Feb. 1966a, pp. 76-84. pp. 24-40. 22-26. pp. 179-194. pp. 254-267. NO. 8, Aug. 1967, pp. 509-530. 1965, pp. 168-176. January 1971 Fedor Mesinger 29 Matsnno, Taroh, “False Reflection of Waves at the Boundary Duc to the Use of Finite Differcnces,” Journal of the Meteorological Society of Japan, Ser. 11, Vol. 44, No. 2, Apr. 1966b, pp. 143-157. Mesinger, Fedor, “Behavior of a Very Large Number of Constant- Volume Trajectories,” Journal of the Atmospheric Sciences, Vol. 22, No. 5 , Sept. 1965, pp. 479-492. Minte, Yale, “Very Long-Term Global Integration of the Primitive Equations of Atmospheric Motion,” World Meteorological Organi- zation Technical Note No. 66, Geneva, 1965, pp. 141-167. Neamtan, S. M., “The Motion of Harmonic Waves in the Atmos- phere,” Journal of Meteorology, Vol. 3, No. 2, June 1946, pp. 53-56. Nitta, Takashi, “Initialization and Analysis for the Primitive Equation Model,” Proceedings of the WMOII UGG Symposium on Numerical Weather Prediction, Tokyo, November 26-December 4, 1968, Japan Meteorological Agency, Tokyo, Mar. 1969, pp. VI-1 1-VI-20. qkland, Hans, “An Experiment in Numerical Integration of the Barotropic Equation by a Quasi-Lagrangian Method,” Geo- jysiske Publikasjoner, Vol. 22, No. 5, Oslo, Jan. 1962, 10 pp. Phillips, Norman A., “Numerical Integration of the Primitive Equations on the Hemisphere,” Monthly Weather Review, Vol. 87, Phillips, Norman A., “The Equations of Motion for a Shallow Rotating Atmosphere and the Traditional Approximation,” Journal of the Atmospheric Sciences, Vol. 23, No. 5, Scpt. 1966, Sadourny, Robert, Arakawa, Akio, and Minte, Yale, “Integration of the Nondivergent Barotropic Vorticity Equation With an Icosahedral-Hexagonal Grid for the Sphere,” Monthly Weather Review, Vol. 96, No. 6, June 1968, 351-356. Welander, Pierre, “Studies on the General Development of Motion in a Two-Dimensional, Ideal Fluid,” Tellus, Vol. 7, No. 2, May 1955, pp. 141-156. Wiin-Nielsen, Aksel, “On the Application of Trajectory Methods in Numerical Forecasting,” Tellus, Vol. 11, No. 2, May 1959, Wurtele, Morton G., “On the Problem of Truncation Error,’’ Tellus, NO. 9, Scpt. 1959, pp. 333-345. pp. 626-628. pp. 180-196. Vol. 13, NO. 3, Aug. 1961, pp. 379-391. [Received March 5, 1970; revised May 6, 19701 I CORRECTION NOTICE Vol. 98, No. 7, July 1970: p. 531, photo of fig. 3 should be interchanged with photo of fig. 4. I