-1 1963 SEP MESOSPHERIC HEATING and SIMPLE MODELS OF THERMALLY DRIVEN CIRCULATION by Conway Leovy A.B., University of Southern California (1954) SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY July, 1963 . Signature redacted Signature of Author . Department a Meteoilogy, July 24, 1963 Accepted by . . . ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chairman, Departmental Committee on Graduate Students I . Certified by . . Signature redacted Mesospheric Heating and Simple models of Thermally Driven Circulation by Conway Leovy Submitted to the Department of Meteorology on July 24, in partial fulfillment of the requirement for the degree of Doctor of Philosophy. 1963 ABSTRACT Possible thermally driven symmetric circulations of the mesosphere and upper stratosphere are considered, subject to the following major assumptions: the motions are in hydrostatic and geostrophic balance, the motions are small perturbations about a state of rest which is defined by the horizontal average of the radiative equilibrium temperature, the eddy fluxes of momentum and heat act only in a diffusive manner on the mean zonal momentum and temperature fields. The assumptions of geostrophy and small perturbations are justified by means of a scale analysis and by a detailed numerical investigation of photochemical and radiative processes. The derived meridional circulations have a principle flow branch from summer pole to winter pole with maximum amplitude less than a meter per second. Good agreement with the observed amplitude, distribution, and seasonal phase of the mean zonal winds is obtained if eddy viscosity and eddy conductivity parameters are assumed to have values between 500 m 2 /Is and 1000 m2s Radiative and photochemical processes are found to have an important damping effect on motions of this type; a radiative-photochemical damping parameter can be derived having the order of magnitude 10-6 per second. Thesis Supervisor: Title: Jule G. Charney Professor of Meteorology ACKNOWLEDGEMENTS It is a pleasure to acknowledge the encouragement, support and valuable advice given by Professor Jule G. Charney. The author is indebted to Dr. Lewis D. Kaplan for providing a copy of the Curtis CO 2 cooling matrices and to Dr. R. Tousey for providing the recent solar spectrum used. He also wishes to thank Pro- fessor R. E. Newell for making figure 13 on the Rocket Network data available. Valuable discussions with Dr. R. R. Rapp, Mr. E. S. Batten, Dr. R. M. Goody, and Dr. H. I. Schiff are also greatly appreciated. Preparation of a manuscript under conditions requiring both speed and precision is a trying task, and the author wishes to thank Miss Marie Guillot, Mrs. Theresa Blumen, Mrs. Berit Larsen, and Miss Ann Ordway for this work, particularly Miss Guillot on whom the brunt of the task fell. The numerous figures were drafted by Miss Isabel Cole and Mrs. Janet Leovy. Much of the numerical work on the radiative and photochemical problem was done at the M.I.T. Computation Center, Cambridge, Massachusetts. TABLE OF CONTENTS Page Chapter 1. INTRODUCTION 1 Chapter 2. BACKGROUND OF THE PROBLEM 4 Chapter 3. FORMULATION OF THE PROBLEM A. The General Equations B. Scale Analysis and Formulation of the Linearized Problem C. Energy Equations 11 11 RADIATIVE EQUILIBRIUM AND HEATING A. General Remarks B. Photochemical Considerations C. The Equilibrium Computations D. The Time Dependent Problem E. The External Heating F. Discussion 29 29 35 44 SOME LINEAR SOLUTIONS 69 69 72 Chapter 4. Chapter 5. A. B. C. D. E. F. G. Chapter 6. DISCUSSION OF RESULTS A. B. C. D. Chapter 7. The Equations Boundary Conditions on the Stream Function Solutions in the Case of Rayleigh Viscosity and Conductivity Solutions for Constant and Exponentially Increasing Viscosity and Conductivity Boundary Conditions on , and an Example with Eddy Viscosity and Eddy Conductivity Eddy Viscosity, Eddy Conductivity and Radiative Coupling A Solution in the Case of a Bounded Atmosphere 48 60 63 75 98 101 110 124 133 Comparison with Observations The Evidence from Radioactive Tracers Eddy Viscosity and Eddy Conductivity Some Implications of the Deduced Circulations CONCLUSIONS AND SUGGESTIONS 15 25 FOR FURTHER RESEARCH 133 139 143 149 154 Appendix 1. DETAILS OF THE PHOTOCHEMICAL & RADIATIVE CALCULATIONS A-1 A. Photochemical Data A-1 B. Calculation of CO2 Emission A-10 C. REFERENCES j Calculation of Ozone Emission A-12 U LIST OF FIGURES Page Figure Figure Figure Figure Figure 1. 2. 3. 4. 5. 32 Characteristic times for equilibrium in the ozone problem. 39 Equilibrium temperatures, *K, computed by matrix method. 46 Log (base 10) ozone concentration, molecules per cc, computed by matrix method. 47 Diurnal variation of 03 Lat. 0, Dec. 21. tial conditions: equilibrium. 54 Ini- Figure 6. Time integrations of temperature at the equator. 55 Figure 7. Variations of ozone concentration on the equator. 56a Figure 8. Equilibrium temperatures, *K, computed by marching method. 57 Log (base 10) ozone concentration, molecules per cc, computed by marching method. 58 Figure 10. Diurnal temperature range, *K. 59 Figure 11. The basic statical equilibrium temperature distribution. 61 The "external heating" of an atmosphere whose temperature is T0(z), *K/day. 62 Figure 12. Figure 13. Figure 14. Figure 15. Figure 16. Rocket network zonal wind observations, taken from Newell (1963). 79 The idealized "external heating" used in the dynamical computations. 88 Vertical velocity (cm/s) model I, Rayleigh friction Newtonian conductivity. 89 - 9. Meridional velocity (m/s), model I, Rayleigh friction Newtonian conductivity. - Figure ii Cooling rate due to CO2 versus Planck function at 15 microns. 90 mm Figure 18. Figure 19. Figure 20. Figure 21. Figure 22. Figure 23. Zonal velocity (m/s), model I, Rayleigh friction Newtonian conductivity. 91 Vertical velocity (m/s), model II, Rayleigh friction - Newtonian conductivity. 93 Meridional circulation (m/s), model II, Rayleigh friction - Newtonian conductivity. 94 Zonal velocity (m/s), model II, Rayleigh friction Newtonian conductivity. 95 Phase relationships between v&, the external heating, P = 0.3. -' A A ur, - Figure 17. - Page u ,T and 96 Comparison of "well-behaved" solution, and solution matched to molecular dissipation layer. No radiative coupling. 106 Zonal wind (m/s), model I with eddy viscosity and conductivity. 107 A Figure 24. Figure 25. Figure 26. Figure 27. Figure 28. Phase lag of L conductivity, P , = model I with eddy viscosity and eddy 1. 109 Roots of the equation A values of GP in parentheses. 114 Well-behaved solutions (dashed), and solutions matched to eddy dissipation layer (solid), radiative coupling included. 116 Vertical velocity (cm/s), model III, eddy viscosity and conductivity with radiative coupling. 119 Meridional velocity (m/s), model III, eddy viscosity and eddy conductivity with radiative coupling. 120 Figure 29. Zonal winds (m/s), model III, eddy viscosity, eddy conductivity and radiative coupling. 121 Figure 30. Phase relationships between model III. Figure 31. A V, /I er , A A A , 7 and 122 Heating used by Murgatroyd and Singleton (1961), *K/day.127 Page Figure 32. Truncated series approximation to the heating of Figure 31, *K/day. 128 Figure 33. Vertical velocity (cm/s), model IV. 129 Figure 34. Meridional velocity (m/s), model IV. 130 Figure 35. Zonal Winds (m/s), model IV. 131 Figure 36. Observed zonal winds (m/s), after Batten (1961). 134 Figure 37. Zonal wind (m/s) for model IV with eddy viscosity and eddy conductivity, showing the effect of various boundary conditions on 144 Temperature distribution corresponding to model I (OK)- 153 Figure 38. Appendix 1. Figure Al. Solar spectrum from 1750 to 2100A. A- 5 Figure A2. Cooling rate due to ozone infrared emission versus the 9.6 micron Planck function. A-13 LIST OF TABLES Page Table Table Table Table Table 1. 2. 3. 4. 5. 6. Reactions determining ozone and atomic oxygen concentrations. 36 Errors due to overestimating the absorption at sunrise. 67 Zonal wind data deduced from rocket network observations. 80 Solutions and separation constants for the equation. 86 Coefficients in representation of Murgatroyd Singleton heating. - Table Comparison of deduced and observed zonal winds. 132 135 Appendix 1. Table Al. Spectral data used in photochemical calculations. A- 7 Table A2. Calculation of time constants used in Figure 2. A- 9 Table A3. Constants used in calculating cooling due to ozone infrared emission. A-14 1. INTRODUCTION The possibility and desirability of understanding the dynamics of large-scale motions in the upper atmosphere arise as detailed information, particularly information on variations in space and time becomes available. In this regard, the upper stratosphere (which in this work is taken to be the region from 20 kilometers height to the mesopeak) and mesosphere, where the absorption of solar radiation by ozone is an important process, play a unique role. Most of this region lies beyond the range of balloons, and it is too low for probing by meteor trail or radio reflection techniques. The only direct observations from this region must be obtained by means of rockets. Above 55 kilometers, temperature cannot be measured directly except by acoustical methods (see for example Stroud et al, 1956, 1960, Nordberg and Stroud, 1961), and the alternative procedure of meas- uring the vertical distribution of density is more difficult than in regions accessible to satellites. Nevertheless, wind observations by the rocket network since 1958 have given us a large store of information about these regions, at least in the North American region. The variation of zonal wind with height, latitude and season is now fairly well known (Murgatroyd, 1957; Batten, 1961), and it has even been possible to construct crude synoptic charts at mesospheric levels (Teweles and Finger, 1962). An attempt has been made to assess the importance of eddy processes up to 60 kilometers (Newell, 1963). -1- In addition, the mesosphere and the lower thermospheric region just above it play unique roles as atmospheric chemical laboratories. Nowhere else are large-scale dynamics and molecular scale processes so directly and intricately related. The detailed distributions of trace constituents must be known before the dynamics can be well understood, and conversely a knowledge of large-scale circulations is important for the understanding of a variety of chemical processes (Kellogg, 1961; Wallace, 1962). In developing a theory of large-scale circulations, a good procedure is often to start with highly simplified models, and perhaps the simplest model of all is one in which deviations from zonal symmetry are either ignored or else are treated parametrically. attempt to apply such a model to the mesosphere. This paper is an It is recognized that this approach leaves much to be desired; it has proved wholly inadequate to deal with the dynamics of the troposphere and lower stratosphere (see for example Starr, 1954). Nevertheless, there is no a priori theoretical reason for rejecting such a dynamical scheme in the mesosphere. The theory of baroclinic instability which seems to be generally applicable to the troposphere, as well as to the dishpan experiments, depends crucially on the temperature gradient along the lower boundary, as Charney and Stern (1962) have shown. It is not reasonable to expect that the mesospheric stability properties depend critically on the surface temperature gradient, and the latter authors have thus far only been able to demonstrate a necessary condition for instability in the absence of a boundary temperature -2- gradient, regardless of temperature gradients or lateral wind shears within a region. This necessary condition for instability is probably exceeded in the mesosphere during the winter, and in fact there is considerable observational evidence that instabilities do arise in the winter mesosphere. On the other hand at other seasons the mesosphere may well be stable according to this criterion, and indeed, large scalelarge amplitude eddies are notably absent from the summer wind observations. It is probable, then, that a symmetric circulation model may be applicable to the summer mesosphere, and possibly to the transition seasons as well. It has also been shown (Charney and Drazin, 1960) that it is unlikely that the large-scale quasi-geostrophic motions of the troposphere penetrate into the mesosphere, except perhaps for brief periods during the spring and fall. Another reason for applying a symmetric model to the mesosphere is that it may lead to conclusions regarding the zonal circulation components which are incompatible with observations. For example, we are able to derive a mean dissipation rate which would be required to give the observed magnitude of the zonal circulation, furthermore an estimate of the dissipation rate can be obtained from the observed phase relation between the zonal flow and the annual component of solar heating. If these prove to be incompatible with each other or with what we know about the physical mechanism of dissipation, we may reject the symmetric circulation concept, at least in the particular latitude, height and season range where the symmetrically derived zonal circulation is incompatible with that observed. -3- 2. BACKGROUND OF THE PROBLEM The idea of axially symmetric convection driven by latitudinal variations in heating is perhaps the oldest in theoretical meteorology. Attempts to apply this concept quantitatively have ranged from that of Oberbeck (1888) to Holl (1961). In another type of approach, Kuo (1954) has investigated the free axially symmetric motions which may arise in a rotating fluid when a horizontal temperature gradient exists. Such motions are a form of instability and will not be treated here; instead it will be assumed that horizontal temperature gradients always lie well below this instability threshold. Aside from the instability of the mean state of the troposphere with respect to quasi-horizontal disturbances (baroclinic instability), another difficulty, perhaps even more fundamental, attaches to quantitative attempts to deduce the basic state of the troposphere from theoretical considerationsalone. This is the fact that the state of radiative equilibrium is one of negative static stability (Manabe and M6hler, 1961); the observed positive static stability of the troposphere must be a consequence of the motions, yet the static stability exerts a controlling influence on both the magnitude and qualitative properties of the motions themselves. Hence the theoretical problem is an essentially non-linear one in this sense. This problem has been discussed in detail by Lorenz, (1953). -4- The upper stratosphere on the other hand is statically stable in the radiative equilibrium state, and the same may be true even up through the meoodecline. In Chapter 4 an attempt to ascertain the radiative equilibrium state of the mesosphere will be described, and it is found that the radiative equilibrium state does indeed appear to be statically stable. This result is of fundamental importance, since it means that theoretical studies of these regions, whether based on symmetric or non-symmetric models may be able, in a first approximation at least, to treat the vertical temperature distribution as given mainly by the radiation field with a relatively small component superimposed by the motions themselves. This is the underlying philosophy of the present treatment. Wilckens (1962) has reviewed our knowledge of mesospheric circula- tions emphasizing the attempts to explain the observations in terms of axially symmetric models; these last may be divided into two classes: those which emphasize an explanation of dynamical properties, and those which emphasize an explanation of the distribution of trace substances. In the first group, Kellogg and Schilling (1951) proposed that a meridional cross-isobar flow occurs from summer pole to winter pole. They utilized wind and temperature measurements available at that time to deduce the average slopes of pressure surfaces as a function of latitude and height for the summer and winter seasons. Assuming that the southern Hemisphere resembles the northern hemisphere in the corresponding season, they extended their isobaric surfaces from pole to pole. Using these cross-sections they surmised the existence of a frictionally induced -5- cross-isobar component from the summer high pressure region to the winter low pressure region. Such a pole to pole circulation would be thermally direct, and would differ radically from the lower atmosphere circulation where transport across the equator is generally regarded as of less significance than transfer within hemispheres. Murgatroyd and Goody (1958) have made a calculation of the distribution of radiative heating in the mesosphere. Their results clearly show that the principle zonally averaged diabatic heating has a maximum at the summer pole and a minimum at the winter pole. The underlying reason for this difference between the distribution of heating in the mesosphere and in the troposphere can be explained by the difference between volume and surface absorption of solar radiation. The principle heating agent for the lower atmosphere is absorption of solar radiation at the ground. This is roughly proportional to the cosine of the zenith angle multiplied by the length of day. In the summer these two factors oppose and nearly compensate each other with increasing latitude, normal pole to equator temperature gradient is never reversed. and the On the other hand, the heating produced by absorption in a volume is proportional only to the length of the day, at least at reasonably small optical depths. This is why the heating gradient reverses between summer and winter within a hemisphere. Murgatroyd and Goody's estimates of the radiative heating were used by Murgatroyd and Singleton (1961) to deduce the magnitude and distribution -6- of meridional wind components which would be required to balance the heating. They did not consider momentum balance requirements, and in fact their circulation implies a particular distribution of momentum sources and sinks to give the observed zonal momentum distribution. Their calculation agreed in essence with the proposal of Kellogg and Schilling and gave meridional velocity components up to about four meters per second in the upper mesosphere. They also had a less intense direct circulation from equator to pole in the stratosphere. The complimentary point of view was taken by Haurwitz (1961), who calculated the meiional components which would be required to balance the loss of relative angular momentum, eddy viscosity. assuming that this loss is due to Implied by this circulation is a distribution of eddy sources and sinks of heat. Haurwitz's circulation applies only to middle latitudes and is somewhat more complex appearing than that of Murgatroyd and Singleton, but in general shows a direct circulation in the mid- mesosphere, with indirect circulations at higher and lower levels. The dryness of mid-latitude stratospheric air led Brewer (1949) to postulate a direct stratospheric circulation with rising air over the equator and descent at middle and high latitudes. Dobson (1956) pointed out that such a circulation model could also account for the observed latitudinal distribution of ozone. Machta (1959) also suggested that such a circulation could account for the distribution of fallout of radioactive materials initially deposited in the stratosphere. Libby and Palmer (1960) invoked a somewhat similar scheme to account for the distri- -7- bution of radioactive material. More recently, however, Newell (1961) has shown that the ozone and radioactivity observations can be explained by quasi-horizontal eddies in which downward and northward velocities are positively correlated. An important paper by Eliassen (1950) has a very considerable bearing on the present problem. Eliassen showed that in a rotating circular vortex which is in hydrostatic and geostrophic balance, every distribution of momentum and heat sources taken together with the distributions of angular momentum and specific entropy in the vortex implies a particular meridional circulation, and furthermore the meridional circulation responds instantaneously to changes in sources or in vortex structure provided these changes are slow enough that the vortex always remains balanced. This is because the equations for a balanced symmetric vortex are just a special case of the general quasi-geostrophic equations discussed originally by Charney (1948) and Eliassen (1949). The meridional component in the symmetric model arises from the divergence required to maintain the vortex in geostrophic and hydrostatic balance. momentum sources and heat sources is quite analogous. The role of A point source of heat produces a dipole-like flow toward higher values of specific entropy at the source point. A point source of angular momentum produces a dipole- like flow toward values of higher specific angular momentum at the source point, and away from such values at other points on the same line of constant specific angular momentum. Distortions of the dipole fields arise from the vortex structure. In the present problem, the assumptions of hydrostatic and geostrophic -8- balance are made, however Eliassen's approach is more suited as a diagnostic tool when actual sources and sinks of momentum and heat as well as the zonally averaged vortex structure are known. To determine both zonal and meridional components of motion using Eliassen's theory it is necessary to use an iterative procedure in time and to adapt his equations to realistic boundary conditions. The present work is thus an attempt to derive meridional and zonal circulation components under conditions resembling those in the mesosphere*, under the following main assumptions: ) The vortex is in hydrostatic and geostrophic balance. ii) The motions are a small perturbation on a basic state of statical equilibrium, which is defined by the lat- itudinal and zonal mean of the radiative equilibrium temperature field. iii) Sources of angular momentum and entropy due to both large and small scale eddies result only in effects resembling eddy viscosity and eddy conductivity. The validity of assumptions i) and ii) will be explored to some extent in Chapters 3 and 4. Assumption iii) can only be verified or * rejected after more observational and theoretical work has been done. Hereafter "tmesosphere" will be used to denote the entire region from 25 to 85 kilometers height. -9- The approach to be followed is similar in some respects to that of Kuo (1956), but differs in the treatment of dissipation, in the consideration of time variations, and in the treatment of static stability as a constant or slowly varying function of height, an exponentially increasing function. -10- rather than 3. A. FORMULATION OF THE PROBLEM The General Equations Since hydrostatic equilibrium will be assumed throughout, pressure but pressure varies by several orders can be used as vertical coordinate, of magnitude through the mesosphere so that it is more convenient to use TF(P) instead the quantity H IT(?) =- P5 where (3.1) is the ratio of pressure to surface pressure, H and H is can be expressed in terms of the average ~r.0 H where /p n the average scale height. temperature defined by R TOO (3.2) is the gas constant for dry air and the acceleration of gravity. IT where T has the following useful properties: in atmospheric regions is close to , -F0 TT differences will approximate height differences; if temperature differences are small everywhere, iT resembles the height, and the main advantage of pressure coordinates in such an atmosphere - the elimination of the need to deal with more than one inde- pendent thermodynamic variable - dP dT P H _ e is preserved. We also have -r/ (3.3) H -11- Dt p where W The PDt P nt -- mt is the density, so that as long as (3.4) ~T~ is close to closely resembles the vertical particle velocity, TT' D 7/ system was introduced by Eliassen (1949). The equations of motion, hydrostatic equilibrium, continuity, and energy may now be written L ;T. "-sP Dt (z + D t.r -k (3.5) r cosCOS )s LL51 -L P COS d _ (3.6) RT T H It + 4F - (3.7) W -- re cos a (r cos cp) + e TT/H .r/H a- (3.8) (3.9) - c3u r COS T with, as usual 3t at- Le (3f L4 recos?- -12- a J C9 Tr (3.10) i In these equations " is time, longitude, and its rotation rate, and is the earth's radius, SL- r. X f latitude, Cp is the ,M specific heat of dry air at constant pressure. LL dependent variables are the zonal velocity Lr , the geopotential height of a the temperature T . Tr -W~ In addition to , the the meridional velocity (or pressure) surface , and is diabatic heating per unit mass and per unit time. We now take an ensemble average of each term in these equations where the particular type of ensemble may be thought of as a large collection of observations taken at the same point in ( X , and at the same time of year but at all times of day. f, Tr ) space In this way the slow annual variation is retained while the diurnal, semidiurnal and irregular variations all become fluctuations about the mean. aging operation is identified by the over-bar (_), from it by the prime ( )'. This aver- and fluctuations We also introduce the zonal average, iden- tified by the curly over-bar f( )dA (3.11) * S- and indicate its fluctuations by a star ( ) The successive application of the ensemble and zonal averages will be denoted by means of the caret ( ). The zonal average automatically satisfies the Reynolds postulates (Kampe de Feriet, 1951); the ensemble average is assumed to do so. quently, if we first take the ensemble average of equations (3.9), and then the zonal average, we obtain -13- Conse- (3.5) through r A A u ~Jrr 4- Ll C) AS n - a -* +CO (3.12) dH e Tr/ A ( 4jw -* GL A' ' ) A + cr a - diV 4+ ( 2. n.+ R* (3.13) -- +- A Lt r. Cas9 +o Ctr T e - F(p, rr ft) S4-r 1-)CsJ r-e c 0 4 -- e J wf w tU e -A/ e-r .' - E~ 7rr t RH ~TT H r. Cos -- ( U,cos T) a<p + eTr/H c) T ( - (3.14) (3.15) -14- L A A re C Cos CP (3.16) A J Le CP C ( P ) T.,) The problem of the symmetric circulation is to regard these equations as a closed set to be solved with suitable boundary conditions for the doubly averaged quantities appearing on the left-hand sides. The eddy fluxes of heat and momentum as well as the diabatic heating act as forcing functions for the symmetric components (Eliassen and Kleinschmidt, 1957, p. B. 144; Saltzman, 1961). Scale Analysis and Formulation of the Linearized Problem. We assume that the averaged temperature field can be decomposed in the following way - )~, TT, --(3.17) T~_ where T , q ( Tr~ which is uniform in all pressure surfaces, is the temperature distribution of the basic statical equilibrium state. time dependent heating which varies within to (3.17), we split into -15- 1t~ surfaces. arises from Corresponding A + pT.) (3.18) is defined by where RTO __ (3.19) H c)Tr and we assume that (TT=o) = Const It is convenient to introduce the following nondimensional variables: ;z 7T/ Y~ SiCT Cos A 2 -ft 1-e A LUT- HodL' (3.20) T- R D1e - - 7 P, Cr -16- L cos p F coscp r. er cr and we also define the static stability by the relation New quantities appearing on the right-hand sides of these equations are (Y , the annual frequency, of the heating and motions. and D the characteristic vertical scale Two height scales D have been H and introduced in order to contrast certain features of mesospheric motions with those of the troposphere. in the mesosphere D Murgatroyd and Goody, " 3H 1958). In the latter region D '\ H , whereas (see, for example, Batten, 1961; Equations (3.12)-(3.16) may now be put into the non-dimensional forms C) T' 1 '/2.c)2(3. --t + n L0 (3.22) + LD2_ ro -17- (3.23) -fl + nr - R where 7Z. -(3.24) *, 7-T W-- (3.25) /r-nJ(rje can be assigned orders of mag- and The operators nitude unity since we are dealing with processes of annual and global scales; also has order of magnitude one by definition. D > H , Because equation (3.24) then shows that ja) Ur (3.26) We note also that 7r where U l(u U R ~~0 (3.27) is a characteristic (dimensional) zonal velocity, and is the Rossby number for the problem. -18- R. It is possible to place upper limits on and 45 if we assume i) that deviations of zonal and meridional winds from the Re not greater than ii) that the ratio of vertical eddy to horizontal eddy velocity R is no larger than iii) , means are no larger than the mean zonal wind itself, i.e. that the scale of variations of the eddy fluxes is no smaller than r in the horizontal and H in the vertical. These assumptions are quite reasonable throughout the mesosphere, but may be violated by both internal gravity waves and tides above about 80 kilometers (Hines, 1963). r1_N oT 0 (R < When they are valid, 2 ( OV07 (V Ra o (3.28) and (3.29) If the Rossby number is of order 1/10, may be a very important term is dominated by in equation (3.21), but Furthermore, .5 72 in (3.22). if the inertial terms in (3.22) were to be comparable with the Coriolis term, the dimensional meridional velocity would have to be about 60 m/s (for R0 = 1/10). tainly much less than this, Since the meridional velocity is cer- the only term which can balance the Coriolis -19- term is the term involving . In other words, the mean zonal flow must be in approximate geostrophic balance except perhaps very close to the equator, provided the above assumptions are satisfied and the Rossby number is 1/10 or less. Equation (3.28) gives what is probably a high upper limit to the ageostrophic force in this case, since the limit corresponds to perfect correlation between eddy velocity components. We will assume that the zonal motion is in exact geostrophic balance, and take as the complete expression for this Y(3.30) yZ) Combining this with equation (3.23) gives the thermal wind equation, + _Y_ L (-YZ) I a -2 D _Y d H Y a (3.31) ~ from which it follows that H 7(3.32) We now examine the validity of linearizing equations (3.25) Since the motion and perturbation temperature fields are produced by differential heating, proaches zero. they will approach zero as This suggests the introduction of a parameter, % ap- , and (3.31). (3.21), such that I -20- .< (3.33) and expansion of the dependent variables in power series in E . For the order of magnitude analysis we assume that -e 0- (3.3 4) (3.3:)) and and T . are characteristic dissipation times for ( where Then to first order in Tn 6 *O C4T (3.36) CY0 (3.37) d n, --- Y - W- = 0 Dc~ -+ -A (3.38) T + R, r (3.39) Second order terms give (3.40) -21- L & H 7 Y ay9 g (3.41) -r/ = - D C) - (3.42) 1T, +.~~~~=n, ( + (3.43) Using the fact that all derivatives have order of magnitude unity and assuming that neither (i divided by ) wr ( -4) are greater than one year , we can derive orders of magnitude for the solutions 2 qT of these equations. We find (3.44) H -I .g i: . H H (3.45) The most important term on the left-side of the second order set is IJ 6 Y ru nTYfn I rj (3.46) H A* -22- I* H It follows that 7.- (3.47) 27. (3.48) Since E + T). + ' - -I (3.49) - (3.50) ' n, 71 + - n G + 7l i 7n then f 41 Ro i D -+I SH - ,Y) (3.51) +- * -- The applicability of the expansion in 6 RS 4 - (3.52) and the detailed analysis of the linearized equations clearly depends on the size of the second term in the square brackets; more specifically, -23- it depends on whether 4* D - ___ The parameters or radiative heating field. D field; ' R5 The time scales eddy source terms , (3.53) can all be derived from the is the frequency of oscillation of this the vertical scale and statically stable, e ) , and , in detail in the next chapter. bution. IfJ D 4; C the amplitude will be discussed If the radiative equilibrium state is also depends on the radiative heating distriand depend largely on the but if the effect of the eddy source terms is prima- rily dissipative, these time scales can be derived from the observed phase relation between zonal wind and radiative heating. When these parameters are evaluated in this way, it will be shown that the ratio of the left to right-hand sides of equation (3.53) is not greater than about 1/3. This is not usually regarded as sufficiently small to justify linearization, but it does suggest convergence of the series, and that the solution of the linearized equations should give at least a rough approximation to the true solution. The linearized * equations will thus form the major focus of the remainder of the thesis. But it will be shown in the next chapter that damping by radiative and photochemical effects may also be important. -24- C. Energy Equations For completeness, the energy equations corresponding to a rotating symmetric vortex in geostrophic and hydrostatic balance will be given. These can be written down for both the non-linear and the linearized equations. For the non-linear case the energy equations have been derived by Eliassen (1950), but for a different coordinate system. We shall assume no energy flux through upper and lower bounding pressure surfaces of the mesosphere, and make use of the dimensional equations. In the non-linear case these are (3.12), (3.14), (3.15), and (3.16) together with U AT r,Cos~)L.L~ + =0 (3.54) -r/HA -Tr/i f Multiplying (3.12) and e (3.54) by LL , e and L~ respec- tively, adding, making use of the continuity equation (3.15) and integrating over the entire mesosphere gives the kinetic energy equation - e - dV - JC/HA ^ dV v . is the volume element in ( dV = e COS C> d / cdTrW -25- (3.55) 0 ~ f X . where FU , Tr ) - space: (3.56) After some further manipulation involving the continuity equation and the hydrostatic equation (3.14), c)J (3.55) takes the form .Tr/H ^ 2- -e/H ^ 2 fe- Tr/)4 V H(3.57) A FU.dV The potential-internal energy equation is obtained by multiplying (3.16) by CP and integrating, H V .CP e d then fHRT/ fcP e f~C dv *f e MIN/ H -Tr/)4 1 A A V (3.58) 'dV We note only two unusual features of these equations - the quantity plays the role of density, and the kinetic energy does not A include any contribution from the Lr A or W components. The latter feature of the geostrophic approximation has been pointed out by Eliassen (1950). heating. Omitted from equation (3.58) is a term arising from frictional Although ordinarily negligible in the lower atmosphere, term might be significant in the upper mesosphere (Hines, 1963). The linearized dimensional geostrophic equations are ir Lr F S Q(3.59) -26- this A JLL Sir) LP = C) 0 (3.60) R P H c Tr -Tr I 4 (3.61) I1 rCPscj _ -H A --- P 8f 1 ) = (3.62) A R7 C, H C CP (3.63) where -( S(7T) + P, H. o -o -T )r in a statically stable atmosphere. ;> 0 (3.64) The kinetic energy equation is obtained by a procedure exactly analogous to the non-linear case. The equation is TT/H ^ C et 2 L f v=- A -Tr/ r r v- f _7/Hdv e ua (3.65) On the other hand the potential-internal energy equation is obtained by -27- j multiplying C TP (3.63) by and integrating to give 70S - I T/HCTP -WH (3.66) - f&e-,r/H pC S ,, T T/H e 2Ta volume in ( is the available potential energy per unit - Clearly TO -- X , S p , ~1T ) space (Lorenz, 1955). -28- 4. A. RADIATIVE EQUILIBRIUM AND HEATING General Remarks The preceeding analysis has raised the following two questions which require an investigation of radiative heating in the mesosphere: How should the basic state, T(a), be defined, * i) and is such a state statically stable? ii) What is the distribution of , and does its magnitude satisfy the condition for linearization (3.53)? To these should be added a third question: iii) How do the motions influence the net heating rate through redistributing the temperature? To make these questions more concrete, a somewhat simplified model will be considered. Murgatroyd and Goody (1958) have pointed out that the principle radiative heat balance components between 30 and 90 kilometers are absorption of solar energy by ozone, s , and emission of infrared radiation in the 15 micron bands of carbon dioxide, If r- is the net heating per unit mass, 4 7-)] 4- * [ Unless otherwise specified, all quantities in this chapter are dimensional and Z denotes the geopotential height. - 29-- (4.1) where over symbolizes the linear operation of weighted integration Z , and Bv(T) is the Planck function corresponding to 15 . T- microns and the temperature The heating 7, is a function of temperature as well as position, so that the equilibrium temperature 7-e is defined by the implicit relation 4L B,(r )] ' (?e ) + Now we suppose that the temperature <f T = (4.2) 0 T differs from 7- by an amount which is small in the sense that s (Tr ) +- s (T (T~ (4.3) and B,()=B, + are valid approximations. (4.4) or)T JT; Under these conditions d(1pW) -][+js ) 9 (4.5) 7- T=T7C,. 'n if T = T -+ -r , dT - -I) Te- 7- -30- so that + In particular, F-7-= (T; ~Te) +T p T - (4.6) Clearly a good way of defining 7 e . -F is as the horizontal average of With this definition, the first curly-bracketed term in (4.6) depends only on 7e which in turn depends only on physical properties of the atmospheric gases, the solar spectrum, the earth-sun geometry, and temperatures outside of the region of interest. This term can therefore be called the 'texternal heating" and identified with the forcing function 00e in the analysis of Chapter 3. term depends linearly on Tp The second bracketed J it should therefore by incorporated into the homogeneous part of the dynamic equations and may be called the "internal heating.? This formulation is convenient since it shows how the truly external librium temperature and a statical equilibrium temperature 1 ( ) forcing function is related to the difference between the radiative equi- and it displays the motion-produced radiative heating as a separate term. Its validity depends on both ( 7 - T ) and Tph being small enough that (4.3) and (4.4) are satisfied; it does not depend on the assumption of only a single infrared band but does require that concentrations of radiatively active gases are unchanged by the motions. Examination of the Planck function shows that (4.4) is a good approximation for 1 T 50 0 K when T, is between 200 and 3000K. The valid range of (4.3) will be considered in the next section. Fortunately the internal heating can be easily incorporated into the dynamical model. imation to Zf, Murgatroyd and Goody have shown that a fair approxis -31- j I +4 H- + 90 km - 50 km X 80 km ' 45 km ri 65 km V 40 km o 55 km x dT 1.010 - .02214 BL dt 0 79.2 (omitting 90 km da ta) r x Q x V C2 ty -4 + 0, 0A . IN, .0 -8 -12 1- + + 0 -16 1- erg/cm-sec ) (C0 2 I 100 Fig. 1. I 200 I I I 300 400 500 Cooling rate due to CO2 versus Planck function at 15 microns. the parameters 6 and -4 being functions of 0 only; this is because radiation to space is more important than exchange of radiation between layers in the mesosphere, and is only valid for temperature profiles of approximately similar shape. results for all heights. Figure 1 is a replot of their Linear regression of all points except those at 90 kilometers gives the relation 1.010 - .02214 BW * = with a correlation coefficient of 79.2 . Combining this with the follow- ing linear approximation to the Planck function for the temperature range = 2.055 (-r - 145.8) , 200 to 3000 K gives a "best" linear relation between the heating rate and temperature: o -* *(4.8) , = 5 7.68xlO 5 , and -7 = 5.26x10 . with With these ideas about the relationship between external heating, internal heating and the radiative equilibrium temperature as a background, we shall proceed to a detailed examination of radiative equilibrium and * heating rates in the mesosphere. It should be noted that this high correlation arises to same extent from clustering with respect to temperature of several calculations made at . would be more A considerably larger value of the same height. appropriate for altitudes above 55 kilometers. -33- Radiative equilibrium and heating rate studies have been carried out previously by a number of authors for stratospheric levels. works such as those of Karandikar Early (1946) and Gowan (1947) suffered from inadequate knowledge of the solar spectrum but displayed the major qualitative features of the stratospheric temperature distribution and heating. Brooks (1958) has examined in the lower stratosphere, CO cooling and ozone heating but the most complete study of radiative heating components for the region below 55 kilometers is that of Ohring (1958). He found a net deficit in seasonal and latitudinal averages of heating throughout the lower part of this region. Recently Manabe and Mhler (1961) have presented computations of heating rates as well as equilibrium temperatures for the lower stratosphere. agreement with those of Ohring. Plass Their findings are in general (1956) has computed cooling rates due to the 9.6 micron band of ozone from the ground to 65 kilometers. The only attempt to calculate net heating rates due to all the important components throughout the mesosphere is the work of Murgatroyd and Goody. In addition to ozone heating and COm cooling, they incor- porated cooling by the 9.6 micron band of ozone by making use of Plass' results. All of these studies have one feature in common - a distribution of ozone based either on the assumption of photochemical equilibrium or on observations, or a combination of the two was used. Conversely, attempts to calculate ozone concentrations theoretically have relied on observed temperatures and air densities -34- (Craig, 1950; Dittsch, 1961; Johnson et al., 1951; Paetzold, 1961). The two problems are not inde- pendent, however; the dependence of heating rates and hence of equilibrium temperatures on ozone concentrations is obvious, but it is also known that the ozone concentration depends on the temperature (Craig and Ohring, 1958). The photochemical and radiative aspects of the problem will therefore be considered simultaneously. B. Photochemical Considerations Only reactions involving oxygen allotropes and non-reacting third bodies will be considered. hydrogen, hydroxyl, The neglect of reactions involving atomic and various other compounds of hydrogen and oxygen may be serious in the upper mesosphere, but the occurrence of OH nightglow should be a good indicator of these reactions. There is evidence (Wallace, 1962) that nearly all the OH night-glow originates above 75 kilometers. If this is.true, it should be safe to assume that these reactions do not exert a controlling influence on ozone concentrations below that level. With this restriction, the well-known reactions determining the ozone concentration are given in Table 1. In these expressions JJ is the dissociation rate of molecular oxygen per molecule per second when c, ( 1j) is the absorption cross section of molecular oxygen in ej( cm 2/molecule and dissociation; for ozone, 100 J3 i) is the quantum yield of primary photo, and e3 are the corresponding quantities is the photon flux outside the atmosphere per cm2 and -35- 'A Reaction 0, + v Rate 0 J2 4=ex o2 Iov exP(-x-cX3 X3) J' V >'H 300 A o0÷+ X= .1 O+ M 0 + 03 -+ 204 13 ~ ( Table 1. -- o' 3 x3 ) , =M.0.7 X 10m/ -+Oz +M 0 +02+ M e3 +XI 1,11 e=3 (exP (~o;x v > 8GXO A 0 0, -+ 30 .5- 33 c T's exp (-30Z5/--r) -x ?o c& M/s ) 03 + Reactions determining ozone and atomic oxygen concentrations. d second per wave number, L) is the wave number (cm 1) and X )< and are respectively the total numbers of oxygen and ozone molecules between the point in question and the sun and depend on the zenith angle as well as height. Details of the constants used will be found in appendix 1; about the values of the reaction coefficients and - i , it should be particularly noted that there is considerable uncertainty In the following analysis it will be assumed that photochemical O. changes in concentration are completely negligible, so that always remains perfectly mixed with neutral molecules in the ratio 21:79. 0 (nitrogen and argon) It will also be assumed that all molecules are equally effective as third bodies. For our purposes, the most convenient way of writing the rate equations which result from the reactions in Table 1 is the following d where of 0, 7), , , 0,,, J3 n. , O 13 +n and n f,,, n (4.10) are respectively the concentrations and third bodies. Equation (4.9) gives the rate of change of the total number of odd oxygen atoms, i.e.: the total number of atoms in the form of ozone or available to form ozone, while equation (4.10) may be thought of as expressing the variation of distribution of -37- odd oxygen atoms between ozone and atomic oxygen. .J useful because, in the daytime, This point of view is is so large that equilibrium in equation (4.10) is always achieved in a short time, i.e. from a few minutes up to about an hour depending on height and zenith angle. >> - -11 Since at all heights, this equilibrium can be expressed , by the relation 42 T)3 -n?n , . J (4.11) 3 to a high degree of accuracy. In contrast the equilibrium of total odd oxygen concentration may be achieved in a much longer time. To show this we assume that (4.11) is exactly satisfied in the daytime, then (4.9) becomes (- '.. + Z where Equilibrium will be achieved when -n, T Letting -38- ie -2 43(4.13) (4.12) / I ) // Craig 75 x zenith angle 67* o zenith angle 0* Wallace - -- J- 1 65 / * / 55 tim (das)/ - charctersti 45 zenith angle 0* 4 -4 35 25 10 I Fig. 2. 102 I 10 1 I 1 10 1 10 2 I Characteristic times for equilibrium in the ozone problem. 10 3 10 5 10. --I and 3 and assuming J2 - L - , (4.14) is independent of time leads to the solution 4:J(4.15) is an initial value of where . Clearly A is the appropriate time constant for the odd oxygen atoms, and because changes in the odd oxygen concentration are much slower than the adjustment between atomic oxygen and ozone, A is the crucial time constant for restoration to Wallace equilibrium following any disturbance in concentrations. (1962) has previously derived a solution similar to (4.15) for conditions prevailing above 65 kilometers, while Craig (1950) has obtained an approx- imate solution resembling (4.15) which is valid below 45 kilometers. Equations (4.14) and (4.15) should be valid throughout the stratosphere and mesosphere provided the reactions in Table 1 are the only important ones. -I -j Figure 2 shows the two time scales X and J3 corresponding to a zenith angle of 600 and standard atmosphere temperatures. shown are time constants due to Craig and due to Wallace. -40- Also The discrepancy between the present result and that of Wallace is significant. It is believed to be due to the fact that Wallace's calculation was based on the work of Bates and Nicolet temperature of 45000K near the Schumann-Runge bands. (1950) who assumed a solar blackbody 1) = 50000 cm 1 and assumed eL= 0 for Recent evidence indicates a solar black-body temperature near 5000 K at 50000 cm 1 (Detwiler et al, 1961) and significant predissociation in the Schumann-Runge bands (Wilkinson and Mulliken, 1957). The reader is referred to Appendix 1 for details on the data used in this calculation. A particularly important disturbance of ozone and atomic oxygen concentrations occurs daily, when the sun sets. At this time atomic oxygen disappears by recombination - very rapidly below 70 kilometers, and very slowly above 80 kilometers. Some of this atomic oxygen forms ozone via the three body reaction involving two oxygen atoms or via the two-body reaction. The latter process is greatly enhanced at upper levels at night due to the increase in ozone, and may produce very large decreases in the total concentration of odd oxygen molecules. The degree to which these losses can be restored during the day depends on A . Below about 60 kilometers ozone concentrations can be expected to approach equilibrium values rather early in the day, but a few kilometers higher ozone may remain significantly below its equilibrium concentration throughout the day. For this reason, it may be necessary to seek the equilibrium temperature as the solution of a time dependent problem in which these nighttime variations are specifically taken into account. -41- We are now in a position to discuss the dependence of temperature. on the The solar energy absorbed in any thin layer may be affected both through weak temperature dependence of the absorption coefficient, and changes in the optical depth of ozone, but the most important effect of temperature is in changing the in situ ozone concentration through . the temperature dependence of / AL43 r'77 4ro Below about 65 kilometers so that the equilibrium ozone concentration is, .J3 from (4.11) and (4.13), ~~- 71 2_ (4.16) T Neglecting the effect of temperature on S', n3, Above 65 kilometers hence of 7 dependence of , ( = 43n, J3 , we then have e. (4.17) becomes rapidly independent of -3 and diurnal variations may complicate the picture, but the on temperature at the high levels will certainly S not be greater than that indicated by It follows that 151.5 (4.18) - s (c~ = cFTTe (4.17). T and the next largest term in the expansion of -42- ) 713e , (-Q-+ c) is PPP FT -fT 2. d Near the mesopeak, where these terms are largest, (4.19) CP per second and 15/2.. T .e_ 02 c- ( may then be as high as 2x10 per second, which may be several times as important a damping effect as infrared radiation. Equation (4.19) imposes another type of linearity condition on the problem, independent of (3.53). The validity of the separation of the heating function into an inhomogeneous part plus a linear term depends on the size of the factor multiplying in (4.19). Using typical mesopeak values, we obtain 01 ( - - ( Y2- so that the error in linearizing the heating function will be of the order of 50% for J XT P 500K. The important stabilizing effect of previously by Craig and Ohring (1958). 43 has been emphasized The present discussion is intended to clarify the role of this effect in the present dynamical framework, -43- but is subject to the following caution: the degree of temperature dependence of have used a k3 43 is not well established. Although most workers close to that given by Benson and Axworthy (1957), there is some evidence (Leighton et al, 1959) that this may not be the value applicable to kinetic processes. temperature dependence, however, Even with a much weaker it appears that this particular damping effect would be an extremely important one. C. The Equilibrium Computations The procedure used in these computations was to assume a temperature distribution, then use the approximate equilibrium equations (4.11) and (4.13) with the reaction rates given in Table 1 to deduce the concentrations of atomic oxygen and ozone at each level starting from the highest level (85 kilometers) and working down. Concentrations and amounts of energy absorbed were computed in this way for various zenith angles corresponding to different times of day for each latitude. total energy absorbed during the day, over the day using Simpson's Rule. the 9.6 micron band of ozone, 13.s The , was obtained by integrating To this was added the heating due t6 3 . The latter quantity was computed by application of an empirical linear relation, derived from Plass' work, between in situ heating rate and the 9.6 micron band Planck function at the assumed temperature; this method is the same as that used by Murgatroyd and Goody (see Appendix 1). -44- Cooling due to carbon dioxide was calculated using Curtis' matrix method and Goody, 1958). A new temperature was then computed by solving for in the matrix equation B2wj R. where the L (Curtis, 1956; Murgatroyd and Bv RP. .J - frSi - (4.20) 3L are the Curtis matrix coefficients, correspond to individual heights. and the indices After inverting the resulting set of Planck functions to find the new temperature at each level, the process was repeated, and the entire procedure continued until subsequent temperatures converged to within 10K at all heights. Apparently because of the very strong restoring effect of 3 this procedure led to an oscillation in succeeding temperatures unless some damping was introduced. This was done by choosing at each step a new temperature which was between the computed new temperature and the old temperature but slightly closer to the new temperature. Calculations using this procedure were carried out at latitudes 00 0 o 30 , 60 , o and 90 for December 21st, and 30 , 60 0 o and 90 for June 21st. In addition to the assumptions already mentioned, it was assumed that the temperatures below 22 kilometers and above 85 kilometers as well as the density at 22 kilometers corresponded to the 1959 ARDC Model Atmos1959). It was also assumed that the total number * phere (Minzner, et al, The author is indebted to Dr. of the Curtis Matrices. L.D. Kaplan for kindly providing a copy -45- -J 80 210 20 70 00 1801(6 24023 250 260 2900 b1- 90 70290 300J L'' 90 Fig. 3. 0 (summer) 50 Equilibrium temperatures, JI latitude Iv *K, computed by matrix method. 30 DU / U (winter) 7v 8.o 8.0 7.5 7.5 7.o .7.0 -80 9.5 -70 10.0 -60 10.5 4-i -2J tiD -50 11.0.. 11.5 -40 12.0 i 12.5 -30 13.0 I 20 90 I 70 50 (summer) Fig. 4. 30 10 latitude 10 30 (winter) Log (base 10) ozone concentration, molecules per cc, computed by matrix method. 50 70 of oxygen molecules above 85 kilometers was equal to the number density multiplied by the scale height at 85 kilometers; the total number of ozone molecules above 85 kilometers was assumed to be equal to the ozone number density multiplied by one half of the density scale height, but 85 kilometers is essentially at zero optical depth for ozone in any case. Results for the equilibrium temperature and midday ozone concentrations are given in Figures 3 and 4. D. The Time Dependent Problem Since diurnal variations may have an important effect in the mesosphere, the marching problem was also attacked in an attempt to obtain mean daily temperatures which would be obtained from arbitrary initial conditions. asymptotically For the reason discussed in Section B, it is possible that this type of calculation could yield significantly lower temperatures in the upper mesosphere. In the notation of the previous section, the marching problem is to solve numerically df where C C?=i1 (4.21) is the heat of recombination at the iih level; it can be either positive or negative depending on whether more molecules are dissociating or recombining. The solar energy absorbed, -48- , is a function of equations, and the latter is determined from the photochemical rate -n3 (4.10) together with dtt - , - 3z 3 z.r T,+nJ4. A problem arises in the numerical integration of 3 (4.22) (4.10) and (4.22) since some of the reactions are slow or nonexistent at certain altitudes or times of day but are extremely rapid at other altitudes or times of day. At some altitudes the first order reactions are dominant, while at other heights the second order (non-linear) reactions are comparable to the first order reactions. To solve these equations numerically, the procedure of linearizing the equations about initial values at each time step has been adopted. The exact solution to the linearized equations is then computed, and the solution for the next time step is obtained by linearizing about the new initial values. The time step chosen should then be small enough that changes in and -n, are small within 71, each time step at all levels where the second order reactions are not dominated by first order reactions. We write the equations pertaining to daytime variations in the linearized forms / -49- 2. (4.24) o where the subscript dominates -03 refers to initial values. all other bracketed terms in these equations at all levels even for low zenith angle. The large time increment of one hour has therefore been used during the daytime, except during the periods within an hour of sunrise and sunset when ten minutes was used. compared with the smallest expected value of ( Ten minutes is short - responds to the most troublesome non-linear term. 7,o )1 which cor- The term arising reaction is altogether negligible below 75 kilometers, from the and is only comparable to OT)J above about 83 kilometers. At these heights all recombination reactions have a time scale considerably greater than ten minutes. At night, ten minute intervals were used during the first and last few hours, but some preliminary experiments indicated that a much larger time interval would be adequate during most of the night, i.e. after rapid changes taking place during the first two hours or so have neared completion. Consequently one hour intervals were used during the remainder of the night. the linearized equations are, for daytime: A nJ. ~ e)f 2~~ ~ -k n 300-b -)O(.5 A2 2 -50- (o Solutions to [(Az~ P Z+ 7., cL-n,. 'A ) 3 ( \-1 'I - _-n.l X_ (AAz At (4.26) L where ( CL - 0" ) , -f-- 4 2-. CL - ] 112- (4.27) (6~7~-3 4- 4 ~C * -X (4.28) and (4.29) 303+ (4.30) -n30 -J3 In all the experiments, and expressions for was large enough that the radical in the N was always real, even when the solar 0 zenith angle was greater than 90 T) = ) w e6 . The nighttime linear solutions are (4.31) 0 -51- CL U(4.32) where (4.33) and (4.34) 0-V Twilight radiation (direct solar radiation for zenith angles greater than 90 ) was considered using the method of Pressman (1954). The assump0 tions regarding temperatures below 22 kilometers and air masses of and aboye 85 kilometers were the same as those used in the equilibrium 03 The energy gained through dissociation in the time interval calculations. 3t was calculated from the formula ( =- 34.35) where 2 - is Planck's constant, C the velocity of light, V/ and are the dissociation frequencies of oxygen and ozone respectively, and Jn, , and cfT1. are the changes in r, and -n3 during c5t Integrations were carried out for December 21st at latitudes 00 300and 600 and for 450 on June 21 using the equilibrium results as initial conditions. Integrations were carried out for 10 days at 450, -52- 12 days 0 , and 18 days at 30 and 60 . at In each of these cases the 24 hour temperature changes on the last day did not exceed 0.25 K at any height below 82 kilometers. Above 82 kilometers 24 hour temperature changes up to 0.8 K occurred on the last day indicating assymptotic limiting temperatures a few degrees lower than were obtained at these highest levels. Figure 5 shows the results of an experiment to test the adequacy of the time intervals used. It shows the variations of ozone concentra- tion during a 24 hour period at various heights on the equator using day, night, and twilight time increments half as long as those used in the other experiments. The open circles are results for the longer time increments and identical initial conditions. The agreement is very good except for some slight discrepancy at 49 kilometers. Even at that level the daytime concentrations are in very good agreement. Some of the daily temperature values for the equator are shown in Figure 6. Higher temperatures occurred at the highest levels on the first two days as a result of the release of recombination energy. There- after the temperature fell in a more or less exponentially decaying fashion, apparently approaching an asymptotic value near that reached on the last day. check, This behavior was characteristic of all latitudes. As a an experiment was run for ten days at the equator, starting with the tropical standard atmosphere of Cole and Kantor (1963) as initial temperature distribution. The resulting temperatures, day is also shown in Figure 6, of which the 10th appeared to approach the same asymptotic values. -53- S85 kms. 85kms $ 79 -9 5 1 73 67 07 55 55 o0 55 0mo 0 6 D 61: 49 -73--- 1 2 3 4 5 6 7 Fig. 5. 8 9 10 11 12 Hours Diurnal Variation of 03 , Lot. 0 , Dec. 21 . 13 14 After Noon Initial 15 6 Conditions t Equilibrium 0~ 17 I8 I9 20 21 22 23 24 day 12 day 0 90 I 80 tropica standar atmosphe re (day 10 / day 1 d tropical standard atmosphere (day 0) 70 1- 60 k 50 ] .-W 10 40 k 30 day I day 12 20 1- Sday 0 temperature, 180 Fig. 6. 200 I . -, *K I 220 240 260 Time integrations of temperature at the equator. -55- I 280 I 300 Some details of the ozone distribution are shown in Figure 7. The depleting effect of the nighttime losses shows up clearly above 65 kilometers. The sunrise concentrations are down by a factor of about 6 from the equilibrium values, while the sunset concentrations are down by as much as a factor of 3. As Figure 2 would suggest, the final ozone con- centrations are established in four days or less at all altitudes below 80 kilometers. The slowness of the approach to the asymptotic steady state temperatures is due to the slow radiative processes rather than the photochemistry. Figure 8 shows the noon temperatures obtained by the marching process on the last day. angle at the poles, Since there are no diurnal variations in zenith the equilibrium values were used at these points. The differences between Figures 8 and 3 are small except above 65 kilometers where the marching problem temperatures are somewhat lower. Of more significance is the change in horizontal temperature gradient above the mesopeak in the summer. The gradient is nearly uniform in the equi- librium computation, but shows a definite sharpening in high latitudes in the marching problem. Slightly higher temperatures at very low levels were obtained in the marching problem. This is a consequence of the calculation of heating rates in the equilibrium problem using the equilibrium concentrations corresponding to the zenith angle at each time of day. This procedure evidently underestimates the energy actually absorbed near sunrise and sunset. -56- I . 80- x -. before sunrise 1/2 hr before sunset - -i/2hr after sunrise . /2 hr before sunset day 12 day II 11/2 hr after sunrise day 4 I hr xxx oo x 0. xx -- - x 70- x x Equilibrium day 1I day 4 Concentration (initial values) A A Direct Observation by Johnson et al '60- 50~ - 40 30- 10 10 Log Fig. 7. 1012 10 l 10 base 10 ozone concentration ( Particles per cc Variations of Ozone Concentration on the Equator 103 r-,- - ' f 80 180 160 10 200 25 260 270 280 60 300 290 4 90 50 70 (summer) Fig. 8. I I 30 10 I 10 latitude Equilibrium temperatures, *K, computed by marching method. 30 I 50 70 (vinter) 90 7.o 7.5 8.0 -80 - 8.5 9.0 -70 9.5 10.0 -60 10.5 I' 4J -50 11.0 11.5 -40 12.0 12.5 -30 13.0 I 20 90 I 70 50 (sumer) Fig. 9. 7 30 10 10 latitude 30 50 (winter) Log (base 10) ozone concentration, molecules per cc, computed by marching method. 70 0.5 80 1.0 1.5 2.o 2.5 70 3.0 -60 4-) I 'H 43.o 50 C,) 4..5 30. 4.o - 40 1.5 - 30 1.0 0.5 90 I 70 I 50 I 30 (summer) Fig. 10. Diurnal temperature range, *K. __ 10 10 latitude 30 iI 50 70 (winter) 90 I The noon ozone concentration distribution corresponding to Figure 8 is shown in Figure 9. Again the only important differences with the equi- librium results are the decrease at the highest levels, and the sharpened horizontal gradient toward the summer pole in the upper mesosphere. The diurnal temperature range as a function of latitude and height is shown in Figure 10. This is the range occurring on the last day of the time integrations at each latitude. Above 80 kilometers a correction was made for the overall 24 hour temperature trend. As one might expect, the diurnal range is approximately equal to the absorption of solar energy multiplied by the smaller of the illuminated or dark fractions of the 24 hour period. E. The External Heating The temperatures of Figure 8 have been averaged with respect to area on the earth's surface, and the resulting temperature, has been identified with the basic statical equilibrium state, ( 7 ), which is shown in Figure 11 along with the 1962 U.S. standard atmosphere for comparison. Since the external heating is the heating which the atmosphere would undergo when in statical equilibrium, marching program with each latitude. 7 it has been computed using the as the initial temperature distribution at The marching equilibrium distributions of atomic oxygen and ozone were also used as initial conditions. out at latitudes 90 , 45 and 0 Integrations were carried for June 21st, and 30 , 600, and 90 -60- for 901h T.(Z) 80 1% 1962 U.S. standarc d atmosphere 70 60 f 50 h 4.) 401- 30 20 aI 180 Fig. 11. aA 200 220 Il_ 240 temperature, *K 260 280 The basic statical equilibrium temperature distribution. -61- 300 80 -70 J + + +4 +3 +2 +1 0 -- 5 2 / +8 - 60 0 40 -30 I 90 Fig. 12. i 70 50 30 10 I 10 M 70 I 30 (summer) latitude The "external heating" of an atmosphere whose temperature is T0 (z), 'K/day. 50 (winter) 90 December 21st. The 24 hour temperature change computed in this way gives an approximation to the external heating rate in 0K per day, and is shown in Figure 12. Although calculated quite differently, the results are similar to the net heating rates derived by Murgatroyd and Goody. The largest difference is the relative weakness and lower altitude of the arctic winter heat sink in the present computation. This can be attributed primarily to the warmer temperatures actually occurring above 50 kilometers over the winter pole than occur in the dynamically stable basic state defined here. F. Discussion An indication of the accuracy of these results is given by the comparison of the computed average radiative equilibrium temperature with the U.S. standard atmosphere in Figure 11. no reason why the two should be identical, There is, of course, since there might well be a net import or export of energy across the upper or lower boundaries; nevertheless the consistently higher computed temperatures can probably be attributed largely to errors in computation. Several important possible error sources are the following: i) Inadequacy of the infrared radiative transfer model, particularly the neglect of water vapor emission and very crude treatment of ozone emission. -63- Part of the large discrepancy above 75 kilometers may be due to failure to adequately account for vibrational relaxation of COs which begins at about 75 kilometers. This effect has been discussed by Curtis and Goody (1956) and was taken into account in an approximate way by Murgatroyd and Goody. In the present compu- tation the Curtis matrix coefficients for computing cooling at levels above 75 kilometers were corrected to give cooling rates in agreement with those of Murgatroyd and Goody at those levels, but the approximation may be too crude. ii) Neglect of additional reactions, with hydrogen compounds. particularly those These are probably unimportant below 75 kilometers, but would certainly give a net decrease in ozone concentration at any levels where they are operative. This would alter the temperature in the right direction. iii) Inaccuracies in the solar spectrum, ficients, or reaction rates. absorption coef- Although there is some uncertainty in the dissociation efficiencies of the Schumann-Runge bands as discussed in Appendix 1, by far the largest uncertainty here is in the reaction rates. Reference has been made (see page 44) to evidence for a higher value and weaker temperature -64- dependence of 13 than were used here. Although no temperature dependence was determined, Phillips and Schiff (1962) also report a larger value for at room temperatures. Larger smaller ozone concentrations -, 43 would lead to (below 60 kilometers the ozone concentration decreases approximately as the square-root of temperatures. 3 ) and would cause lower This possibility is particularly appealing since the direct observations of mesospheric ozone concentration by Johnson et al (1951) show lower ozone concentrations than those deduced here at all levels. A particularly good indicator is the region 40 to 60 kilometers, where because of the very short characteristic equilibrium time, day- time ozone concentrations should be very close to their equilibrium values. In addition, below 35 kilometers the results suffer from errors due to the assumption of latitude independent temperatures below 22 kilometers, and from a breakdown of the Curtis cooling scheme below about 25 kilometers. The extremely long characteristic photochemical equilibrium times below 35 kilometers casts doubt on the validity of any scheme such as this much below that level. The equilibrium temperatures over the winter pole should be regarded as only roughly indicative, since the Curtis cooling scheme has been applied here under the assumption that density -65- variations from the standard atmosphere are not too large. Furthermore the temperatures above and below the mesospheric region control the arctic night mesosphere temperatures, and the mechanism determining the temperature in these bounding regions has not been considered. nately, because of the small area involved, Fortu- the arctic winter temperatures 7-. have little effect on The effect of one additional error source, a type of truncation error, can be estimated. Because of the large concentrations of ozone present above 60 kilometers just at sunrise, there will be a spike in the solar energy absorption until the excess ozone is dissociated. The differencing scheme used tends to overestimate this effect because of truncation. We can, however, obtain an estimate of the integrated energy absorption in the spike, and compare this with the computed spike energy of the marching program. n V From equation (4.10), we have at sunrise o e- (4.36) a approximately valid down to about 60 kilometers. absorption per unit mass in the spike, dt where U Idt The rate of energy ,E will then be c J~t(4.37) t cJ.ne is a frequency near the center of the Hartley band, 40000 cm~, and n 3, is the predawn ozone concentration. the total energy in the spike, & E -66- , is say From this, independent of J3 = .35x101 1 .. 30 ergs/gm. (4.38) 71;t . 73 Table 2 gives a comparison between spike size estimated from this formula and computed by the marching program on the llth day at the equator. XE from equation marching (4.38) program 2130 (molecules/cm ) Hei ght (km S) 85 76 67 SE from (10-6 ergs/gm) .503x104 15 .204x10 .756 Table 2. .655x1010 11 .552x10 .11Ox10 12 Error in comp uted heating rate ( K/d ay) 4.6 6.4 +.l 8 9.4 12.3 +.2 9 5.3 8.1 +.2 8 Errors due to overestimating the absorption at sunrise. Below 67 kilometers, the error decreases due to the decrease in the excess From these data one may estimate of ozone concentration over equilibrium. temperature errors in the marching computation to be as much as +5 K near 70 kilometers arising from this cause. In future computations, the details of the sunrise variations should be treated more accurately. In spite of these error sources, the results appear to be reasonably reliable, as Figure 11 suggests, at least in so far as the major features are concerned. In particular, they are probably adequate to answer the two questions posed at the beginning of this chapter. we have defined the basic state, 0 (a rather large static stability everywhere. -67- In the first place, ), and it turns out to exhibit In fact, the maximum lapse rate of 40 per kilometer agrees very closely with that of the standard atmosphere. It is very unlikely that correction of the various error sources mentioned above could reverse this conclusion. In the second place, we are now able to estimate the value of the parameter amplitude of Cp e is about 8 K per day. terms of the non-dimensional heating I11 I = E j C . The maximum Expressing this in from equation (3.20), we obtain .15- Figure 12 also indicates a characteristic vertical scale for the external heating of about 25 kilometers, so that Hvr 1/3. It will be shown in the next chapter that the evidence provided by the phase relation between zonal wind and heating suggests a value of the Prandtl number is of order unity. WIru 0.2, provided that Substituting these values into equation (3.53), we find that the ratio of the second order terms to the first order terms in the dynamical equations is about 1/3. -68- 5. A. SOME LINEAR SOLUTIONS The Equations The basic equations are (3.36) through (3.39), but we modify these by setting D in the analysis. We also reintroduce H to avoid the confusion of two height scales the subscript ordering the E expansion. 'n -n + and suppress Then (5.1) (I )- y2 d/ Y 82 --- Gd and - Ur (5.2) = 0 (5.3) T RS (5.4) It is convenient to introduce a stream-function by means of the identities and to make the substitution -69- We are concerned with periodic response to periodic forcing so e I = 1 , For the annual component, has = the constant component 0 , while any semi-annual component would have (annual mean) Y = 2 . that the time dependence may be represented by the factor With these additional assumptions and definitions, equations (5.1), (5.3) and (5.4) reduce to C) e P r+Rs (5.5) ' - with (5.2) remaining unchanged. (5.6) The dependent variables are now to be interpreted as the complex amplitudes of the harmonic time factor. fix the phase of these quantities, we specify that To is real. -70- j In order to obtain solutions to these equations, it will be assumed that and 0 are related to the mean zonal flow and temperature fields in one of the following ways: 0 = ) ' friction coefficient, while 4. ~,~- is a Rayleigh , i) may be interpreted either as a radiative damping factor, or as Newtonian conductivity, or a combination of the two. e = a0-- and are 4 r 4. and they will be assumed to be constants, Chapter 4, + 4-r) even though, as shown in is a function of both and . 40 - and In terms of the dimensional constants j2 T ii) A/ =-/ nondimensional eddy viscosity coefficient, and eddy heat conductivity. means of - -k * L a nondimensional These are related to the dimensional eddy viscosity and eddy conductivity coefficients, L) HW , is a )< -71- ) and -K< , by Solutions will be obtained both for constant Pi B. and K A/ < and , and for increasing exponentially with height. Boundary Conditions on the Stream Function When any of the forms above are substituted into equations (5.2), (5.5) and (5.6), an elliptic equation in a single dependent variable can be derived. This equation is of second order in the and of either second or fourth order in the on whether Z and a' _ / derivatives, derivatives depending are represented by Rayleigh friction-Newtonian conductivity, or by eddy viscosity and eddy conductivity terms. When the Rayleigh Friction-Newtonian conductivity assumption is made, the solution is completely determined by specifying V' on all boundaries. When eddy viscosity and eddy conductivity are included one additional condition on 2( boundaries . These additional conditions on (or 7 ) is required on each of two horizontal )( will be discussed in Section E of this chapter. From a physical point of view, the boundary conditions on are that there be no mass flux across any boundaries. pressure surface, Z =C, The constant is not the lower boundary, but we assume that it approximates the lower boundary sufficiently well that we can 7=o . This approximation can be * assume no mass flux across The upper horizontal boundary may be at infinity. -72- j justified by the smallness of the parameter . te With this approximation, the lower boundary condition becomes at =O. . (5.7) The side "walls" in the problem are the north and south poles and we = 0 V must certainly require that The along these boundaries. however. actual lateral boundary conditions are somewhat more stringent, From the dimensional continuity equation (3.15), we see that the horizontal divergence of A A ir must remain finite at the poles if ur Now let to remain finite there. Lr where as r (5.8) is a constant and --.- . - -- is (CO V is a quantity which remains finite Then the horizontal divergence of S - COSp (r+ ) (Cosy S&p'V Cos is L- (5.9) This quantity goes to a finite value different from zero only if We can restate equation (5.8) with r= 1 in terms of . -yfinite as -73- >'-* 4 1 r 1 in the form (5.10) It should be noted that as an alternative boundary condition we may require that the vertical component of relative vorticity of the zonal flow must remain finite at the poles. LuCOS f) , I argu- quantity is it follows from an this d - Since, in dimensional terms, 'A - ment identical to the one above that -Y (5.11) finite -- Xy /~ as is required. If the equations are separable (5.11) is automatically satisfied when (5.10) is satisfied. The condition of no mass flux at the upper boundary is simply that 0 the condition is In the case of an atmosphere extending to infinity, somewhat different. infinity, (5.12) on a rigid horizontal boundary. In this case e ur must vanish as -z approaches so that e 0 is a sufficient physical requirement. as - ~+ 00 (5.13) This will turn out to be suffi- cient to specify the problem mathematically as well. We note that all boundary conditions, for for 1P 2X as well as , apply to both the real and imaginary parts of these quantities. C. Solutions in the Case of Rayleigh Viscosity and Conductivity In the following development, R, will be assumed to be constant; an example illustrating the effect of changes in the solutions will be given at the end of this section. the appropriate expressions for ' and Q 1R, on Substituting into equations (5.5) and (5.6), and making use of (5.2), we obtain s +-4 X (5.14) ) (1Y 2 y) A+ + (+- (5.15) ?aYY If the Prandtl Number, P these reduce to -75- , is equal to unity, )-'*X YYZ RS Y2- _ YxL J +~ li (5.17) x ( 1-Y 2. Y3- j d 2.*_ 62_ 81/ It follows that in this case 47 tional to or is independent of On the other hand, . is in phase with T (or 4, ) and is inversely propor- 'X by the phase angle and lags behind 4 4 K There are two limiting cases: .4 i) , corresponding either to a steady - In this case, zonal state or to very large friction. as well as meridional components are in phase with the heating. 1..j ii) - 00 , the case of vanishing friction and conductivity. The zonal component lags in phase by one quarter period, while the meridional component remains in phase with the heating. The agreement in phase between value of -4 The phase of W and regardless of the is a consequence of hydrostatic and geostrophic balance. '2 then depends on the time integrated meridional circu- lation modified by dissipation so that when dissipation vanishes, -76- X is one quarter period behind the heating. P Now suppose I I We write . (5.19) -X =)-r + Z XL and, separating the real and imaginary parts of (5.14), obtain z R 5 ? + I * ) + (5.20) = 5- 11 \ -YJ -! L4 2. L) Y / ( -P)(I 'An + ( R~(W) 6z7- + (1+ (5.21) = j ({\- ) + fI I)I ') E 4?n~~~ y./,~ ~~~ , I(j-p q Equation (5.21) indicates that . Separating the real and imaginary parts of (5.15), one obtains y+ (\ (I- (5.22) - M -- 7- L--Y YZ 62.x Z_ c3Y L 4- 2 X, az -77- I d~ dX, ~P Equation (5.23) suggests, d~(5.23) P even in the case when is not unity, that -4 Since -r results from both radiative damping and damping by some kind of eddy interaction, an upper limit to the ratio .2/.ir A is due to radiative damping alone. can be given by assuming that Using the results of Chapter 4, I. may then range from about 2.5 up to about 10, depending on the size of the dT term. T=T. For the annually varying component of the motion this means that - o.4f quantity is related to OC =+Tn O. , perhaps much less than this. , the phase of X , Since this by the equation (X/'Xr) it is possible to check this conclusion by examining the actual phase lag of the mesospheric zonal motion. Figure 13, taken from Newell (1963) is a plot of rocket network zonal wind observations over a two year period. -78- The points at 60, 48, I IC I 0 0 CM IC 00 0% I 0 F 0 0 % 0 *o. ..*4., q *o .: 0 0 0 \0 IC .0.0 5 0 N 0- 00 J 0 0 0 0 4J o -5 A to 0 -5" 1960 Fig. 13. 1961 Rocket network zonal wind observations, taken from Newell (1963). 1962 and 36 kilometers were fitted with sine curves in an attempt to determine the phase lag relative to the equinoxes; the results are given in Table 3. Phase Lag Relative to 9,- Amplitude of Annual Oscillation Annual Average (m/s) Height (kms.) (m/s) (days) 60 +17+ 5 +71 9+ 3 .15+ .05 48 +22+ 5 +66 17+ 3 .29+ .05 36 +13+ 5 +40 25+ 6 .43+ .10 Table 3. Zonal wind data deduced from rocket network observations. Evidently the phase lag does lie within this upper limit; it appears to range about .25 in the mid-mesosphere corresponding to a characteristic time of 15 days, and decreases with height. The small observed value of . value of 'X'( ~j (i--I) , and corresponding deduced suggests an important simplification to the general time dqpendent problem. involving oC In equation (5.22), we may neglect the terms with an error of order is small, O( . If we also assume we may neglect the terms involving in equation (5.20) with an error of order [ . 7;. This implies a characteristic decay time for momentum of considerably less than 60 days, and seems quite reasonable. Both of these approximations -80- j will be adopted in the analysis involving eddy viscosity terms, as well as in the present case. In the present problem, the result is a simpli- fication from the solution of a fourth order inhomogeneous equation, to the solution of two inhomogeneous second order equations with identical homogeneous parts. When eddy viscosity and eddy conductivity terms are the simplification is from an eighth order equation to two included, fourth order equations with identical homogeneous parts. When terms of order neglected, ( ) have been and equation (5.20) reduces to RS(~ 6 9 _CP2 Y(L (5.24) We assume that this equation is separable in the form LP ,rm(Y) - '1,,7- ZI (5.25) Tn= o -rn =o > (5.26) satisfies Then if Lin Y (- y) dY and (5.27) can be written . 2y Fe -F(Y) at (5.28) -81- I the boundary conditions (5.10) will be satisfied, and the equation for can be written d~, p~-~~L~nP'm provided also that the set of ~(5.29) functions corresponding to different 2- Li values of the separation constant are mutually orthogonal. Ti We first examine the equation for , but instead of (5.27) consider YL where (5.30) Y n is an arbitrarily small positive parameter. Then if 7 ?rn is also required to satisfy a condition analogous to (5.28), we have a Y . - I Inparticular, Sturm-Liouville problem in the range - I - the set of functions are both mutually orthogonal and complete. The solutions of (5.27) with (5.28) are also mutually orthogonal; they cannot be complete however. throughout the interval -/ All solutions to (5.27) which are finite K Y ! + / are special solutions with vanishing second derivatives at the origin. The parameter X arises in a natural physical way if the effect of viscosity is retained in the meridional equation of motion. This effect might be expected to be felt in the immediate vicinity of the equator where geostrophic control is relatively weak, -82- but it will be shown that if is small enough it has little effect on for the lowest values of Lm . and The lowest separation con- stants and solutions therefore reduce to those of (5.27). To solve (5.30), substitute so that ( 4T': dJY ,Lz +l Wjq y+&)~7n (5.32) This closely resembles the spheroidal wave equation in form and has the same singularities (see, for example, Stratton et al, 1957). A convenient way of solving the latter equation is by substitution of a series of Gegenbaur polynomials, for equation (5.32). and the similarity suggests the same approach Let CO F where T_ n is the (5.33) Gegenbaur polynomial of first order And is related to Associated Legendre polynomials by T y) (Morse and Feshbach, Pn_+ 1 (5.34) 1953, P. 782). Using the recurrence formulas for the Gegenbaur polymials we obtain the recursion relation -83- An +B-' + C., C 0 ;CT (5.35) where An= ___~ (z-rl+i)(.n --I) L(Zn sj(n÷2)n +=('n -. C + )(r+1)(n+3) +n -- (5.36) + B n (N+3) + 2 Ln C The set of equations (z-n+5)(2.n+7) (5.35) can be solved for the separation constant _C7f and the infinite solution vector by rewriting them in continued fraction form cx~' Lin -n --+ oo F,' (5.37) /= be finite at -1 means that (5.38) 0,114 4. ) The requirement that ~a7~ An_+ O~a. so that the continued fraction must converge, and as -84- (Ur rLI - A, -"- (5.39) The recursion formula yields a series which is either odd or even, and since AO= At we = 0 (5.40) have OL0 B A. A7 ~ B* - Cz A 4 A3 B- A, ' (5.41) (5.42) B- C 3 A3r which can be solved by successive approximation for the even or odd separation constants. This method is analogous to that used in tidal problems (Lamb, 1932, P. If rapidity. L? 330 to 355). is not too large, the method converges with reasonable It follows that very small values of S cannot much affect the solutions or separation constants in this case. separation constants, For the lowest we may then write ,. 2.. L. (5.43) and use (5.41) and (5.42) in the calculation with & equal to zero. The lowest two separation constants and the corresponding solutions, except for an arbitrary multiplicative factor are given in Table 4. -85- M 0 1 8.1282 Table 4. 12.5439 T7 = 0 1.000000 1.000000 T 2 .134336 .156724 T7 4 .008010 .011125 -n 6 .000263 .000457 _n 8 .000005 .000012 = Solutions and separation constants for the Y equation. Turning our attention to the equation for the - dependence (5.29), we obtain the general solution satisfying the boundary conditions on at = 0 and =co (equations (5.7) and (5.13)) in the form ' + P EM + 1 - where +(I 4RPL) (5.46) and l = RP &(5.47) 3 +-If -86- 9 (5.45) Solutions for all other dependent variables can easily be worked out 'P from the solution for and 'Y, L It should be noted that . n Xr In = 0 (Y) , , '(r for example X M(Z) while the real and imaginary parts of , Yr all separate in the same way as W, (5.48) T separate in the same way as for instance -~ =0 " (5.49) )C Several numerical examples have been worked out. Each of these utilized the heating function , S 00~~~~2~.~ ~= with Z,= 4, ~= ~(5.50) Sao 9, and the amplitude factor chosen to correspond to a maximum heating amplitude at the poles of 80K per day. used in the first example were: fle = 8 .71x10 0 = 1.995x10 H = 8x10 3 m = 3 .1695x10- s-2 -87- m2 2 s 1 Other parameters 180j 0 -70 v -60 +7 -50 +6 +5 +4 +3 +2 +1 0 -1 -2 -3 -4 -5 -6 -7 *- -40 0 30 I 90 I 70 50 (summer) Fig. 14. I 30 I 10 I 10 I 30 latitude The idealized "external heating" used in the dynamical computations. I 70 50 (winter) 90 I N 0 f I ( I 70 I --60 +0.3 +0.1 +0.2 I 1 -0.05 -0.1 -0.2 -0.3 I +0.0 -50 / xx / --40 -7, -30 I 9U Fig. 15. I (~~ 7. ,'I 7U DU IF I 30 10 10 30 50 70 (summer) latitude (winter) Vertical velocity (cm/s) model I, Rayleigh friction - Newtonian conductivity. 90 f~ ') 0. 0.2 v. u-20.5 0.4 0.2 6 80 70 60 -40 -30 I U 90 I I 50 7U (summer) Fig. 16. I 30 a 10 10 30 latitude Meridional velocity (m/s), model I, Rayleigh friction-Newtonian conductivity. 50 70 (winter) 90 -60 -60 -40 -20 0 +20 +40 +60 +L 40 +60 + -2'0 - 40 IQ 80 -70 8 .+80 E w -60 -50 -40 ) C -30 I 90 Fig. 17. I 70 50 (summer) I 30 10 10 30 latitude Zonal velocity (m/s), model I, Rayleigh friction-Newtonian conductivity. I I I 50 70 90 (winter) P = I 10 6 s This example will be referred to as Model I. The field of C. fe derived from equation (5.50) is shown in Figure 14, while the amplitudes A eA of the annual oscillation of 15, 16, and 17. P Since U and are shown in Figures A A. Ur phase of ,the I Gr ,and is A A A identical to that of U , Ur lags behind (L ,while Ore by 11.5 days. In the second example, Model II, all parameters are the same, except that P = 0.3, n* 4m= 0.3x10 19, and 20*. s -l Ur , V The resulting real . A A A parts of the amplitudes of -6 and U are shown in Figures 18, The meridional components of the circulation are decreased relative to the case with P i ,while the increase in the zonal component is small, even though the momentum dissipation parameter has decreased by a factor of 3.3, while the thermal dissipation has remained unchanged. Figure 21 shows the phase relationships in days between the meridional and zonal components and the heating. It is to be noted that the meridional components lead the heating by very nearly (i-'P) A or 29 days, while, although the phase lag of L has increased slightly, it is still quite close to ; if TThe figures are drawn in terms of the pressure coordinate, in terms of geopotential height, the distortion due interpreted TT -surfaces, which in these examples amounts to height to slopes of as 8 kilometers, should be taken into account. much as of changes -92- p - -~ ~------= -- 1 I (80 / I I N 70 / - 4, +.15 -60 +.10 +.05 -. 05 -. 10 .15 4-5 / -50LA \1 -40 -30 I I I I I I 30 10 10 30 i I 90 Fig. 18. 70 50 (summer) latitude Vertical velocity (cm/s), model II, Rayleigh friction - Newtonian conductivity. 70 50 (winter) 90 I 0.10 0.20 0. 0 0. 0 80 .30 30 - 70 - 60 - 50 - 40 0.0 0.0 - 30 4- I 90 Fig. 19. i 70 I 50 (summer) a 30 -- L- I A I 10 10 30 50 latitude -I 70 (winter) Meridional circulation (m/s), model II, Rayleigh friction - Newtonian conductivity. I 90 80 -70 K w E 100 I 00 80 80 60 60 40 40 20 20 -60 cm -50 F40 0 0 -30 9V 50 7U (summer) Fig. 20. 30 I 10 10 30 latitude Zonal velocity (m/s), model II, Rayleigh friction - Newtonian conductivity. 70 50 (winter) 90 80I- 70Phase lead of or Phase lead of 0~ AA 60F- 501I 401bO 30h I I I +30 +20 Fig. 21. Phase lead, dgys +10 Phase relationships between heating, P = 0.3. I 0 A A -10 A u., T , kr -96- A Ur I -20 and the external These results, as well as the differences in the amplitudes P= between this case and the case with i , are a consequence of the dominance of the dissipation term over the vertical velocity term in It is this term which controls the amplitude the thermodynamic equation. of the zonal motion, while the control on the meridional motion by the thermodynamic equation is relatively weak. The major control on the We can conclude meridional motion is through the momentum equation. that in this type of system, phase lag information, Table 3, such as that in can be interpreted in terms of a characteristic time for thermal damping rather than momentum damping. The differences between the two examples also illustrate the role RS of the static stability parameter portional to P Rs P i , and otherwise depends only on Model II by Rs multiplied by 0.3. % RSP Lf P real parts of the solution for the case the old is pro- J The solution for otherwise it depends only on the combination while the solution for , real solutions of . , is proportional to . Hence, if we multiply the these can be interpreted as the P = 1 and a new R, equal to With this interpretation, we see that a rather large decrease in static stability has only a small effect on the solutions; amplitudes of both zonal and meridional components are increased slightly, the height of the maximum rises, and the rate of decay of the motions above the heat sources is diminished. -97- D. Solutions for Constant and Exponentially Increasing Viscosity and Conductivity. Let K = K N = N. e NO where Ko , P and are non-negative constants. This is a generalization which includes both the cases of constant eddy viscosity and conductivity ( P = 0), ( f= 1). The equations to be solved are then (~ Y~)X together with (5.2). , , -- 0(5.51) a (<0 iT and molecular viscosity and conductivity (5.52) If one again assumes that the terms containing can be neglected in the equations for the real and parts, solutions can be obtained with the aid of the substitutions Xe- I Z (5.53) One obtains Z (5.54) Variables can be separated in the same way as for equation (5.24), and furthermore, since this equation is of second order in the vertical, -98- we need specify no new boundary conditions. mass flux vanish at 2 0 = and Again we require that the . The solution for satisfying these conditions is ft fCO oo [ ?~ (5. 55) eij 't~ie~K with L'~ '- /U f34~ When = (-3)÷41~L~j(5.56) +.,(-) PL 2, (5.57) 0, equations (5.55) through (5.57) reduce to the solutions for Rayleigh friction and Newtonian conductivity. obtained from the solution for the terms involving ',. 'Xi q) can be by integrating (5.51) twice with neglected, provided boundary conditions on have been specified. Although similar in form, different from that when the solution when 0=; when -99- L ')(, 3 = =1, we have 1 is very fR., , - 3 + (5.58) R - (5.59) Consider the motion which would occur above some pressure level above all heat sources. The solution for Y!r, Z. which gives vanishing mass flux at infinity (see equation (5.13)) is O e(5.60) Since in the actual atmosphere for global scale motions R.PL <1 this solution gives increasing meridional components with height, even though the mass flux vanishes at infinity. geostrophic assumption when Z Such a flow must violate the becomes large enough that both dissi- pative and inertial terms become large in the meridional equation of motion. Nevertheless this type of solution may be applicable over a height range of several scale-heights in the lower thermosphere where molecular viscosity and conductivity first become important. The separated momentum equation is a7 z/ (5.61) -100- from which we obtain X~ pEL(RsPL)Y-z) ~,zoex X, as the formal solution for (5.62) which vanishes at infinity. if we are to have vanishing of the zonal velocity at infinity, im and must be of opposite sign. Thus, Xr, This is in contrast to the behavior of these quantities in the case of Rayleigh friction and Newtonian conductivity. In the latter case, in a region above all heat sources, Xrr Since E. OA-i) < .(5.63) o , and Boundary Conditions On 'X W are of the same sign. and an Example With Eddy Viscosity and Eddy Conductivity has been obtained, When the solution for X, can be determined from equation (5.61) provided that we specify two additional boundary conditions on ' , one at the lower boundary and one at the upper boundary. In a region bounded above and below by rigid horizontal surfaces, the necessary conditions are the no-slip conditions: ) = on both rigid horizontal boundaries. -101- (5.64) When the atmosphere extends from the earth's surface to infinity, lower boundary condition is still the (5.64) and a suitable upper boundary condition is that momentum flux vanishes as 3- approaches infinity, or 0 as ~ Condition (5.64) does not apply at ;F= 0 -4 , W (5.65) since this is a constant pressure surface rather than a rigid horizontal surface. The actual phenomena taking place near the lower boundary are extremely complex, and cannot be considered in detail here. ever, We make the assumption, how- that pressure fluctuations vanish as Z . -+0 consistent with the assumption of vanishing mass flux at This is '= 0 , and corresponds to the physical condition of no energy flux across that pressure surface. It is also consistent with the finding of Charney and Drazin (1960) that the energy in tropospheric baroclinic waves is normally trapped below the mesosphere . With this assumption, the lower boundary condition can be written 'X= C = at . Solutions satisfying the upper boundary condition have 'X approaching a constant as 2 goes to infinity. (5.66) (5.65) will In general, Assuming that pressure perturbations vanish at the ground is not necessarily the best lower boundary condition. Solutions obtained by assuming that vertical energy flux and pressure perturbations vanish in the lower stratosphere will be considered in Section C of Chapter 6. -102- this constant will be different from zero, in particular with constant eddy viscosity and conductivity coefficients, different from zero. this constant will be This is an unsatisfactory state of affairs; since we know that dissipative effects will ultimately increase exponentially at high levels, while differential heating does not, we would expect oscillatory components of the zonal velocity to vanish at infinity. This behavior of 2X can be obtained by joining the mesospheric solution for constant coefficients of eddy viscosity and conductivity to a solution for the thermosphere where dissipation is by exponentially increasing molecular viscosity and conductivity. We know that the latter solution becomes invalid at great heights because of the breakdown of the geostrophic approximation, and the solutions in the joining region will probably not adequately represent conditions there, but if the resulting solutions do not have much effect on the meridional components at levels below the joining region while giving the correct behavior of the zonal component above the mesospheric heat sources, this should be a valid procedure. We assume that the solutions are to be joined at some height above all mesospheric heat sources. continuity of mass, or Y? continuous; because the model is a viscous one, we must have continuity of If .Xm, At the interface, we must have ]X and Tm. or dXm as well. and the mass flux are to vanish at infinity, we have from equation (5.62), that (5.67) NIiRs Paln -103- furthermore, dm < '" , From the continuity of and cixM ~& ' in the upper layer. (5.68) equations (5.67) and (5.68) must also be satisfied by the lower layer solution at the interface level, Zo . To satisfy both of these con- ditions we replace the previously obtained "well-behaved" solution for Tn by a new stream function, __ 1 T = where %, is the to be determined. "boundary" it well-behaved" solution, and Because solution, eA3 - -, is quite large, (5.67), (5.68), - RPk -ZoA in particular ZO , using equations and assuming equality of the dissipation coefficients at the interface, ~ ~ Izo)- or of the lower boundary condition. (5.61) twice from 0 to and (5.69), is a constant the second term, is important only near the interface, does not affect the satisfaction Integrating equation Cb R,PC,,.- we obtain: (t 3-') ( I--)C6 (5.70) ( z) 4RPL, (/U7I) + C6 (5.71) -104- I P where It (Ei) and A for the unknowns C, and are known, we can solve this system - I, (2) Since This method has been applied, assuming A and the resulting (AT Cr , tAT and , tr~ and E iLL solutions are shown in Figure 22. corresponding to / on both = 108 kilometers, together with the amplitudes of the "well-behaved" -4 have been replaced by LA. 4 dependence of the real parts of the amplitudes of = k and E All parameters AM 3 the dissipation coefficients = 1200 m2/s . = Cr Although at distances greater than three or 4 is continuous across Zo , tinuity, which leads to a discontinuity in been assumed that any g / and = 92.8 Figure 23 shows the dependence It is clear that the "boundary" solution has little effect and in this in the first example, except that example are the same as those of n viscosity does not act on on W-r four scale heights below E, its gradient suffers a disconIf Uf at , Eo . Since it has this does not violate features of the model, but is nevertheless an unrealistic situation. The actual transition to molecular dissipation is not sharp, so that the -105- I well-behaved~ solutions I I -100 1 I' I I - I I I 80 I 1 - 60 I-J - 40 20 0 vertical velocity scm/s) meridional -2 Fig. 22. -1 0 +0.5 -1 elocity (m/s) 0 +1 zona; velojity (9/s) +2 0 20 I 40 i I 60 80 Comparison of "well-behaved" solution, and solution matched to molecular dissipation layer. radiative coupling. No 80 n -20 70 -40 -60 -80 E -80 -60 -40 I-I -20 20 0 40 60 80 W 80 60 40 20 I -60 -5( I/ U) -4( I 2( I 90 I 70 I 50 (summer) Fig. 23. I I 30 10 latitude I 10 Zonal wind (m/s), model I with eddy viscosity and conductivity, 30 I (winter)50 IV 70 I 90 solution in the vicinity of heights below is not reliable, but since a few scale 7.o -Z. the actual solution for the meridional components does not depend strongly on the assumptions made about the transition, the solutions in the mesosphere may be valid. The equations for RSpbY) L'; and i can be written V (5.72) and RSP 7 e - 2 Z -- 5.3 (5.73) K which is similar to the phase relation for Newtonian conductivity case. X On the other hand, in the Rayleigh frictionthe phase of I is given approximately by I -('- (5.75) -108- 90 80 1- 70 - 60 50 40 k 4-i 30 H 20 I- phase lag I 0 Fig. 24. 20 10 Phase lag of A U (days) I 30 40 model I with eddy viscosity and eddy conductivity, P= 1. -109- In terms of the characteristic vertical scale, _ z H D P) H..- D we have , so that we have the order m magnitude phase relation for In the present example, we have taken from equation Wi M o can be calculated without additional approximation ' and the phase of P = 1, so that , (5.76) (5.74). The resulting phase lag of ( (or Ui ) is shown in Figure 24. F. Eddy Viscosity, Eddy Conductivity and Radiative Coupling In this case the momentum and thermodynamic equations are (- 1 )X -I- -- _ ~= (5.77) and s(L 4K- and the thermal wind equation T and (5.2) remains unchanged. (5.78) Again neglecting in the equations for the real parts, we obtain -110- +GP RSP( c" 9r __Y Z 2- az (5.79) - GP I, Y =P(9 where No (5.80) Again variables separate as in equations d' d2 -'m 2 dependence of the stream function is obtained as _ - (' P+-R ?M)d ' ez d -a = (5.81) xj ) , we obtain For the imaginary part of a3 d)W 3 +G + equation for the (5.25) and (5.26), and the - 4L 4 ~ = YP 7P(-~ & X,?Z (5.82) with the corresponding separated equation for the vertical dependence d3 r 10P' -G+ (5.83) -111- L ., where the functions correspond to the real part of 'x The general solution of equation (5.81) can be written A Lc and IP3 -- Q,-if~e"dj e"P, C where /M are the roots of the cubic equation % 3 ~s P, (5.85) &4-& C and the constants , C, and + C3 are to be determined from the boundary conditions. By considering the function three roots of (,), we will show that the (5.85) are real for non-negative Ps? and rP, and they can be ordered as follows: < 0 1 -112- (5.86) and , - 00 -F , also goes to plus infinity and minus infinity respectively. A = + 1 4 , (.) to the right of = + - RsPLm . When ,so , that one real root lies -F , 0 so that there are two other real roots, one between = and the other to the left of = 0 case there is a root at roots of . o When I) /= GP = 0 , unless = 0 CP U=+I and , --3 , # In the limits, , in which But in this case the other two (5.85) reduce to (5.46) and (5.47) so that the ordering (5.86) still holds. Figure 25 illustrates the dependence of the roots of R(PL (5.85) on 6.P for three values of The three constants are determined by the boundary conditions: vanishing mass flux at flux as in r-O and R- = 00 , and vanishing momentum If mass flux is to vanish as . D , we must have C3 = 3 CI 4 ) f S'e (44- (5.87) and vanishing momentum flux requires that co e JS, o C2 ( =-f /3) -- d (5.88) as well. The vanishing of mass flux at C, =- ( Z =0 C ) -113- then requires that (5.89) ~jJ (1.0, (0.5 t 2) (.0, p 3 ) (0.5 P) /43) 0 (0.I,(0) I-J (0- , I-J /2) (0.1, -2.0 Fig. 25. -1.0 Roots of the equation 0 +1.0 of Cin+Gthss v3 4P in parentheses. , values 3) +2.0 +3.0 These constants together with the condition =:- 0 at completely specify the solution. X when L A convenient way of obtaining the has been determined is by direct inte- gration of the momentum equation. are , (.=?o (5.90) The two arbitrary constants arising , which is given by (5.90) and which can be calculated from the thermodynamic equation. At F = , solution for = 0 . the latter equation takes the form GIPT d ,T or e~o dz If we denote f 1%! A -t +AM~ ,+U}-(pVz C; ,+,(p // A = d7 by d4 =O. A ,we dz e=0 find that (5.91) - r The solution determined in this way will be called the "well-behaved" solution. An example showing the vertical variation of the real parts of the maximum amplitudes of illustrated in Figure 26. to L)= A A A Ur U and In this example 300 m2 /s, and -115- GP = Li N for such a solution is = a = 23.5 corresponding .21333 corresponding to 10-6 I 100 I I / I / I / / / / 801L / / I 60 I-. 7 Q) I' N 'I / 40 L 7 / / ,1 20r / I 0 vertical ve ocity (cm/sl meridiona velocity (qms 1 -0.2 Fig. 26. 0 +0.2 -20 0 zonql velqcity (m/s) veoiy(ms meiioa +20 0 20 40 60 80 100 Well-behaved solutions (dashed), and solutions matched to eddy dissipation layer (solid), radiative coupling included. 120 140 All other parameters and the heating are the same as in per second. the Rayleigh friction-Newtonian conductivity examples. X vanishing of does not permit W, Again the "well-behaved" solution for Such a solution may again be ob- at high levels. tained by joining the solution to an upper layer with molecular viscosity The upper boundary conditions on the lower layer and conductivity. (5.67) and (5.68), and we fit these by writing solution are T fe _~ eJQ 4 C e/~( is the "well-behaved" solution. where (5.92) The term involving Cb will be negligible at low levels, so that the lower boundary condition A (/-)jJC'. =1-f ,( ,-i)C, +/ -QcaAZ +) P,)C C3 +- (5.93) O Introducing (5.92) into (5.67) and (5.68) and utilizing the zonal momentum equation, we obtain - )U ~ ~ ~ (frrI(e/Aa ,)itR)L~ ("L - 5L , R=LJ; _ (Ju- - - Equation (5.91) now generalizes to is satisfied by (5.92). 14rn on fr.3 + fr- =t t -117- R5 (5.94) Q#K-1 PL fR {(IA JI)e/ULo + ( R5PL~m) (Aii PL{ ) and Auz7-0 (ze Cb6 s(/"3-1) RPL~ *I.(- ) -- (5.95) when terms of order Cb I) df (~) and /o3 are neglected. o Since ) ) f e are known, equations solved for the quantities A CO- , (5.93) through (5.95) can be and 0 b which determine the solution. An example of this type of solution, which will be referred to as Model III, has been worked out, using the "well-behaved" solution as 4 R3 Pom a basis and with to be 108 kilometers. AA (Ar again equal to .4296 and taken Z.e The resulting real parts of the amplitudes of A , , and LL are shown in Figures 27, 28, and 29*. * A Note the different isoline spacings used in this example for than in Models I and II. -118- tY and LP 80 70 +.125 +.1( )0 +.075 +.050 +.025 0 -. 025 5-. 0500 .075 -. 100 -. 125 60 - -5(f 0 - 0 40 -30 -. 050 -. 025 I 90 70 50 +.025 I I 30 10 (summer) Fig. 27. Vertical velocity (cm/s), I 10 latitude model III, I 30 50 +.050 I 70 (winter) eddy viscosity and conductivity with radiative coupling. I 90 ( 80 70 [- +0.2 60 |+0.1 0.0 50 4 bb .H fe 10 30 - 40 -0.1 90 70 50 30 (summer) Fig. 28. 10 10 latitude 30 (winter) Meridional velocity (m/s), model III, eddy viscosity and eddy conductivity 50 70 with radiative coupling. 90 I I (If - 80 E 'I' I w 70 1 I I I I I +100 -100 60 -80 b1~ +80 50 -60 +60 40 -40 +40 30 -20 90 I 70 I 50 (summer) Fig. 29. Zonal +20 30 10 I 10 latitude I 30 50 (winter) winds (m/s), model III, eddy viscosity, eddy conductivity and radiative coupling. I 7'0 I 90 / 90 80 I I t- 70 r 60 r 50 N 40 A . 30 1- . N phase of LL TF phase of IW 20 t- L~ -100 ~ -80 - - I -60 -40 I- II I -20 0 +20 I i----------I +40 +60 +100 +80 phase lead (days) Phase relationships between -122- U. T tr , model III. and , Fig. 30. Equations P I , (5,82) and (5.83) indicate that even in the case when there is a phase lag of which is proportional to GP In the example described above the phases of P and X relative to the heating have been obtained by solving (5.83) and the imaginary part of the zonal momentum equation. The results are shown in Figure 30. The solutions for this case exhibit surprising differences from either the Rayleigh-Newtonian or eddy dissipation models. The most important difference is the replacing of the single meridional circulation cell by a much weaker two-cell pattern. The lower cell is closed by a weak return circulation in the lowest levels; in the example chosen, it reaches a maximum amplitude of four centimeters per second at the ground. The zonal circulation is much weaker in this case than in the case where only eddy viscosity and eddy conductivity act. These effects are again a consequence of the dominant control exerted by the dissipation terms in the thermodynamic equation. only eddy viscosity and eddy conductivity act, When the region of large and nearly constant negative shear of the zonal wind below the heat source leads to a large perturbation temperature, but negligible curvature of the temperature. Divergence of the diffusive flux of both momentum and temperature is therefore small in this region. When the simple linear radiative coupling term is included, such large vertical shear would lead to a substantial heat sink in the region below external heat sources, and would drive strong downward motions in this region; upward motions are still required in the external heating region just above. -123- The net result of these effects is greatly reduced perturbation temperature and zonal wind shear below the external heat sources, and weak downward motion in the same region. Although the solutions for this case reduce to those for eddy viscosity and eddy conductivity when 0 P - 0) , as they should, the same qualitative features are obtained for much smaller values of than that used in this example. G. A Solution in the Case of a Bounded Atmosphere The solutions obtained in the previous sections are very incon- venient for taking general ) variations of the heating into account. This is because of the slow convergence of the series and the difficulty involved in calculating each of these functions. A simpler solution can be obtained for the Rayleigh friction-Newtonian conductivity model when vertical velocity is required to vanish on two constant We use equation (5.24), and introduce the substi- pressure surfaces. tutions u 2(t h. equation (5.24) then takes the form -124- The vanishing of Wi- on upper and lower boundaries means that vanishes on these surfaces. a sine series in (V A general solution can then be obtained as Z q,./V~nY) in (-f)i!-(5.97) with Z7m(Y) = where D gi (5.98) is now the distance between the constant pressure surfaces. Y ) has again been introduced to insure satisfaction of The factor (1- the side boundary conditions. Vm must satisfy the equation An dY d (5.99) and (5.100) The homogeneous part of equation (5.99) is again similar to the spheroidal wave equation, except that no solutions to the wave equation exist which satisfy the boundary conditions for the separation constant equal to 2 as in this case. We conclude that equation (5.99) has only inhomogeneous solutions, as we would expect from the elliptic nature of the problem. To solve, we again assume a series of the form -125- Co en Vrn In n(5.101) as well as )(5.102) " -d 7-= Instead of the infinite homogeneous set of equations, (5.35), we have to solve the infinite inhomogeneous set A' where A Y1+2 and C. +n C.3 O =1 (5.103) are defined as in (5.36), but now Bn = [+ Z5.104) ( This system can again be solved by the continued fraction method if it is assumed that all of the c4 are zero for 1 greater than some finite value. An example has been worked out for this case using essentially the same heating function which Murgatroyd and Singleton (1961) used (Model IV). This heating function in turn was based mainly on the work of Murgatroyd and Goody (1958), and Ohring (1958). series and five terms in the Ti Only the first four terms in the series expansions of The coefficients in this series are given in Table 5. -126- 06- in were calculated. i I +4 80 L. + +2. +1 o n -i -z -3 -4-5--6 -7 -9 -10 70l. -11 60 k I-A ['3 -6 ~B .4' +6 50 -3 - 40 +3 30L +. -i 0 10 10 ~30 50 70 90 90 to (summer) -1) JU1UI 3U latitude Fig. 31. Heating used by Murgatroyd and Singleton (1961), 0 JO (winter5 0 K/day 9U 0 80 +1 +9 4 70 -10% .1 00 50 *r4 G) -6 +3 40 +1 301-. 4-2_. 0 a 90 Fig. 32. 70 50 (summer) 30 10 latitude 10 a 30 (winter), Truncated series approximation to the heating of figure 31, OK/day. I 70 I 90 0.0 80 0+ 004 +.*08 +.12 70 -0 .12 60 +.2 ;50 ci) +.08 40 +.04 30 -e2 l 0 90 Fig. Ii 00 33. 70(winter) 50 30 Vertical velocity (cm/s), model IV. 10 II latitude r f 10 t I I I 30 I I i I (summer)50 I I % 70 I I I I 90 O.4 0.2 0.6 1.0 1 .4 1.2. 1.0 0.8 0.6 .0 0.8 0.f a.z 0 0 I 607 00 -50 -40 0.0 -30 N 90 I 70 (winter) Fig. 34. 50I 30 1Ai latitude Meridional velocity (m/s), model IV. ' -*1 / I (summer) .. 70 +20+4/ 1.O 40 vI*. +. 0 -0+ 0 +44o +2.0 o 4 -4o - 0 0 -80 _ 40 -20 80 -130 -- 70 +10 -10 -10/ 40 - 0 0 30 - -2O'~ I I VU 7U Fig. 35. -.. F -10 I'- (winter) D0 Zonal winds (m/s), model IV. 30 I-f 1u latitude 10 30 (summer) 50 70 90 ,rl 1 2 3 4 0 +.0034 -.0041 -.0260 -.0151 1 -. 4886 -. 1512 -. 0417 -. 0979 2 +.1526 +.0787 +.0849 +.1307 3 -.0712 +.0307 -.0424 -.0349 4 -. 1041 -. 0938 -. 0370 -. 0534 17.3622 37.2338 70.3553 116.7206 Coefficients in representation of Murgatroyd - Table 5 - Singleton heating. The resulting heating field is given in Figure 32. Because of the asym- metry in the heating, which is due to difference between the winter and summer heating in one hemisphere, the results should be interpreted on the basis of a steady state approximation to winter and summer conditions within one hemisphere. Parameters used in this calculation were 4/-_( t 8.71x105 D = 60x103 H = 7.5x10 3 m 4 r Vt = .743x10- s-1 A A , 2 -3 -2 ,361x10 s No Resulting fields of 2 = ,and UA are shown in Figures 33, 34 and 35. -132- j 6. A. DISCUSSION OF RESULTS Comparison with Observations Mean cross-sections of mesospheric zonal winds have been published by Murgatroyd (1957), Batten (1961), and Newell from Batten, is typical of these cross-sections. (1963). Figure 36 taken These data can be compared with the results of the preceding chapter, particularly with reference to the following features of the zonal wind distribution: amplitudes of the maxima, heights and latitudes of the maxima, phase lag of the zonal winds relative to the external heating, between summer and winter. and differences In Table 6, such comparisons are made between the cross-sections of Batten and of Newell and the various models which have been examined in Chapter 5. It should be noted that the latitude dependence for Models I, II and III reflects the particular eigenfunction chosen to represent the variation of the heating. Comparison of Figures 12 and 14 suggests that this function is more concentrated toward the poles than the actual heating, consequently the deduced zonal winds are concentrated more toward the poles than they would be if a more realistic representation of the )/ variation had been used. Model IV, which is based on more realistic latitudinal variations of the heating, gives better latitudinal agreement between the deduced zonal winds and the observations. -133- +10 +20 0 -20 -20 +2 0 I' 80 / / / +6 +80 I - 70 4 7 60 - I-1 w /- 50N E 50 bb wr I I I - 40 /0 - 30 I I to U I I 70 (winter) Fig. 36. Observed zonal E I 50 30 10 latitude winds (m/s), after Batten (1961). 10 30 (summer) 50 70 90 Dissipation Model Amplitude of Zonal Wind Summer Winter Parameters Phase of Zonal Wind Height of Latitude Wind Max- of Wind Maximum imum I = 10-6 s- (Rayleigh friction, Newtonian Conduc= 1) tivity, -6 10 V =4 10 (Rayleigh friction, Newtonian Conductivity, = 0.3) -6 0.3x10 = 10-6 III 81 m/s 80 M/s 80 m/s 1200 m2/s * II m/s 81 -l =X 2 1200 m /s Ia (Eddy Viscosity and Conductivity) 5 -1 s -6 s -l -13 days 66 kms 560 -14 to 68 kms 56 -19 days 112 m/s 112 m/s -16 days 72 kms 56 118 m/s 118 m/s -20 to 75 kms 560 75 kms 42 65 kms 300 65 kms 40 -1 r (Radiative Coupling included) 1J = 300 m 2/s *= IV 300 m 2 /s .74x10-6 s -6 s -l -25 days 120 m/s 85 m/s Observations (Batten) 85 m/s 60 m/s (Newell) 70 m/s (Bounded Atmosphere) -15 days .74x10 Observations Table 6. -10 to -20 days Comparison of deduced and observed zonal winds. 0 Another interesting feature of Model IV, which is not based on strictly antisymmetric heating, is the good correspondence of the rela- tive amplitudes of the winter and summer wind maxima with those observed. This suggests that the heating of Murgatroyd and Singleton, which is not anti-symmetric about the equator, may be a better representation to the true driving force than the nearly anti-symmetric heating derived in Chapter 4 (Figure 12). This anti-symmetry of the heating in Figure 12 is rather surprising in view of the very definite deviations from antisymmetry of the equilibrium temperatures. two effects. This is mainly a result of One is the photochemical effect of 4 . Where solar absorption is large, the tendency to return to equilibrium is faster because of the temperature dependence of 13 . Since solar absorption is largest near the summer pole and vanishes at the winter pole, this effect will not be anti-symmetric, and will partially compensate for the departure from anti-symmetry of the equilibrium temperature. The second effect arises from the non-linearities in the dependence of heating on the temperature. For example, - in equation (4.6), d -L is an increasing function of temperature; this will tend to decrease the amount of cooling taking place in an atmosphere which is warmer than equilibrium relative to the warming in an atmosphere which is cooler than equilibrium. To the extent that these non-linear effects are important, Figure 12 does not represent the true driving force. The heights of the maxima in the unbounded atmosphere models are in reasonable agreement with the observations. -136- This height in the models depends mainly on the height of the heating maximum, which was taken as 54 kilometers to agree with Figure 12. In the cases where the thermal dissipation dominates momentum dissipation, the heights of the maxima appear somewhat too high; much better agreement is achieved when P = 1. The most striking feature of Table 6 is the joint agreement in amplitude and phase of the wind maxima with the observations. these two features depend on the single parameter P r , Since at least when is not too greatly different from unity, this result lends support to the validity of the dynamical model. Whether dissipation is given the eddy viscosity-eddy conductivity interpretation or not, models support a thermal damping constant, all of the -4v 10-6 per second. This is also the right order of magnitude for the radiative and photochemical damping constant derived in Chapter 4. Apparently radiative and photochemical damping are very important factors in mesospheric dynamics. No direct observations of zonal mean meridional circulations in the mesosphere have been made, nor can such observations be made with present techniques. However, Oort (1962) has measured meridional circu- lations in the stratosphere up to 30 millibars (about 26 kilometers). His results indicate motions in which the annual average dominates the seasonal oscillation. In middle and low latitudes there is equatorward flow with amplitudes up to 40 centimeters per second, while north of latitude 60 there is poleward flow. In contrast the circulations deduced in Chapter 5 all have flow away from the winter pole and toward the summer -137- PF pole at these levels, with maximum amplitudes ranging from 2 to 10 centimeters per second. There are several possible reasons for this disagreement. In the first place, the circulations observed by Oort appear to be driven by momentum sources associated with large scale horizontal eddies of the troposphere. This is indicated by the qualitative agreement of Oort's circulation with circulations deduced by Dickinson (1962) which are needed to balance the momentum sources provided by these eddies. is also suggested by the general decrease with height of This Oort's circu- lation in the same region where eddy momentum flux is decreasing with height. It is also possible that an equator to pole thermal forcing component in the annual mean, which has not been considered in Chapter 5, makes an important contribution at these levels. Such a heating compo- nent has been demonstrated by Ohring (1958) for the stratospheric region below 21 kilometers. It arises primarily from latitudinal variations in tropopause height and lower stratosphere temperatures, been considered here. which have not Although such a distribution of heat sources would produce an annual mean component of meridional circulation, they would be expected to produce a circulation of the type postulated by Brewer (1949) i.e., of opposite sense to that found by Oort in middle and low latitudes. If a circulation of the type deduced in this study exists, it would certainly be masked in the lower stratosphere by the much larger amplitude circulation found by Oort. -138- The only other attempt to deduce quantitative circulations in the mesosphere itself is the study of Murgatroyd and Singleton. They obtained mesospheric meridional velocities up to 4 meters per second above 50 kilometers. The lower values found in the present study are a consequence of the control exerted by the damping terms, equation. In this model, especially in the thermodynamic these terms determine the amplitudes of meridional as well as zonal components. Since the thermal damping constant is inde- pendently known to be at least of the order of 10-6 per second, this control will be important even if non-radiative damping effects are very weak. B. The Evidence From Radioactive Tracers During the first half of August, 1958, two high altitude nuclear detonations deposited substantial amounts of the tracer Rhodium 102 into the high atmosphere at latitude 17 North. These detonations were unique, in that they were the only ones containing large amounts of Rhodium 102. Most of the material was originally injected at mesospheric levels and above (Kalkstein, 1962), and the subsequent distribution of Rhodium 102 was monitored by aircraft covering a wide range of latitude and height in the stratosphere up to 27 kilometers. These observations, which have been reported on by Kalkstein (1962) and by List and Telegadas (1961), may indicate some features of meridional transport processes above 27 kiloThe essential features of the observed stratospheric distribution * meters. are the following All concentrations of Rhodium 102 reported by Kalkstein and by List and Telegadas were corrected for a 210 day half life. -139- i) No Rhodium 102 attributable to the high level injections was observed until the middle of 1959 when there was evidence of a rise in concentrations in the southern hemisphere, particularly between 100 and 300. After this initial rise, concentrations remained essentially constant through the southern hemisphere summer. ii) Northern hemisphere concentrations remained low until the latter part of 1959 when sharp rises took place, particularly north of 30 0 . No important trends were observed in the northern hemisphere stratosphere between the spring of 1960 and the end of 1961. iii) No samples were taken in the southern hemisphere at middle and high latitudes between the southern winter of 1959 and the southern winter of 1960. latitude sampling resumed in June 1960, When higher there had been a very large increase in Rhodium 102 concentrations to values comparable with those at high latitudes (above 30 ) in the northern hemisphere. Low latitude concentrations also increased during the middle of 1960, iv) List and Telegadas have estimated that only 10 to 15% of the total injection of Rhodium 102 had reached the lower stratosphere by May, -140- 1960. . 0 but did not reach values as high as those south of 30 Friend, Feely, Krey, Spar and Walton (1961) have discussed this evidence, and have compared it with the distribution of Strontium 90. They have also attempted to deduce the origin of the Strontium 90 by measuring the ratio of Cerium 144 to Strontium 90. High altitude injec- tions of Strontium 90 should have a higher Ce 144/Sr90 ratio than high yield low altitude injections. The evidence of the Ce 144/Sr90 distri- bution supports the general features of the Rhodium 102 pattern, except that there is evidence that the initial mid-1959 southern hemisphere rise of Strontium 90 of high altitude origin took place primarily south of 300. There is no unique interpretation for these tracer distributions; the final concentrations will depend on the initial cloud size and distribution, and on the size distribution of particles since relative fall rates will be important, although the lack of any early appearance of Rhodium 102 near the latitude of injection at the sampling altitudes, and the small total fallout from injection until May, 1960 indicates that fall rates cannot be very important, greater than 27 kilometers. except perhaps at heights much Kalkstein has given one possible interpre- tation in terms of upper atmosphere transport processes. Since the initial particle distributions with height and size are not known, no attempt will be made to assess the vertical transport process, we will, however, examine the efficacy of the circulations derived in Chapter 5 in producing the observed latitudinal distributions. The strongest circulation derived is that of Figure 34. -141- If we assume that the particles observed had the benefit of transport by the strongest meridional winds in the circulation, these could not have traveled farther south than latitude 10 South between the early August injection, and the circulation phase shift on September 21 which corresponds to a Prandtl number of unity. During the subsequent northern hemisphere winter, the debris could have been carried northward by the circulation to latitude 45 or 50 0 . If the Prandtl number is actually less than unity the debris would have begun returning north before September 21, and could not have reached as far as 100 South. If the Prandtl number were greater than unity, the phase shift of the meridional circulation would have been later, and the circulation itself would have been somewhat stronger, allowing greater southward transport of debris, but it is difficult to account for a Prandtl number much greater than unity, since this would require prohibitively high eddy viscosity. If only the meridional cir- culation were acting, the debris would not be able to move beyond the limits of 100 South and 50 North in subsequent seasons. As a consequence, the increase in Rhodium 102 observed near 300 South in the southern winter of 1959, and the poleward gradient of Rhodium 102 in the southern hemisphere beginning with the winter of 1960 indicates the presence of lateral mixing processes at levels above the highest sampling altitude (27 kilometers). Although the distribution of tracers could be accounted for entirely by mixing processes, a circulation of the type discussed here may well be an additional important factor in producing the observed distributions. The tracer observations clearly indicate stronger vertical mixing during the winters of both hemispheres. -142- C. Eddy Viscosity and Eddy Conductivity The theoretical model permits the interpretation of the zonal wind data in terms of an eddy conductivity parameter. basis of the zonal wind alone, We are unable, on the to deduce the Prandtl number, but the actual height variations of the zonal wind suggest that this parameter is of order unity. In the following discussion, it will be assumed that eddy viscosity and eddy conductivity coefficients are equal. Table 6 shows that agreement with the zonal wind observations can * be obtained using values of 1200 m2/s. and -K* ranging from 300 m2/s to The higher values may be due in part to the use of unrealistic boundary conditions on zonal wind above and below the mesosphere. Figure 37 is intended to illustrate the effect of varying the boundary conditions. It shows wind profiles for latitude 30 in the winter corresponding to the meridional circulation of Model IV and eddy dissipation parameters =- =300 m2 /s . The boundary conditions used were 'X = 0 at 20 kms and L~ =-Xx for three values of Nb at 82.5 kms . Since the smaller values of are more appropriate, if the upper boundary condition is determined by a molecular dissipation layer, of about two. * Under these boundary conditions, the most appropriate value would be about 600 m 2 /s . of the wind values derived are still too high by a factor -143- 80 =3.0 1- ,X=.750 )g=,3ll 70 1 60 F 50 Wind Profiles at 300, winter assuming that the boundary condition at 82. 5 kms is 40 for three values of 30 I 20 Fig. 37. 40 60 80 100 120 140 I I I 160 180 200 Zonal wind (m/s) for model IV with eddy viscosityand eddy conductivity, showing the effect of various boundary conditions on % . 0 -144- Considering all of the models of Chapter 5, it appears that the best agreement with zonal wind observations is obtained if the eddy viscosity and eddy conductivity are in the range 500 m 2/s to 1000 m2 s. Of course there is no reason to expect the actual eddy viscosity and eddy conductivity to be independent of height. These values should be most representative of the region of maximum wind, i.e., from 60 to 75 kilometers. if any, may be responsible for It is not clear what processes, effects resembling eddy viscosity. It is well-known that large-scale eddies, of the order of thousands of kilometers are quasi-horizontal and act to transport momentum and heat in a highly systematic fashion, against the gradients of momentum and heat often (see, for example, Starr, 1954). It is therefore unlikely that these processes could be successfully parameterized by means of vertical eddy viscosity and conductivity coefficients. at the very smallest scales - of the order of meters - On the other hand, it is probable that eddy processes would approach the idealized case of homogeneous and isotropic turbulence, and conditions may resemble the inertial subrange of Kolmogoroff theory. At slightly larger scales, Bolgiano (1959) has predicted a "buoyancy subrange", forces strongly modify shear turbulence. where buoyancy These small scale turbulent motions are very likely to produce eddy viscosity and eddy conductivity effects. Over the very large intermediate range of scales, practically nothing is known about mesospheric motions, although Mantis -145- (1963) has reported observations indicating a maximum in the horizontal kinetic energy of the upper stratospheric zonal easterlies at a time period of 10 hours. At levels above 80 kilometers, motions which appear to be internal gravity waves are found (Hines, 1961; Witt, 1962). If the gravity wave interpretation is correct, they should also occur at lower mesospheric heights, but with reduced amplitude. amplitude, Although large in these motions would not transport momentum or heat unless nonlinear interactions were important. There is some observational evidence bearing on the intensity of At stratospheric levels, up to 19 kilometers, * small scale turbulence. Kellogg (1956) has measured the dispersion of smoke puffs. He finds that the dispersion cannot be well described by a constant eddy viscosity coefficient, since the puffs appear to be continually acted on by larger and larger eddies as they grow. Root mean square velocities of these eddies were deduced to be of the order of 0.1 m/s. If it is assumed thatthis is representative of the vertical velocities of turbulent eddies, it is possible to derive an upper limit on their vertical scale since the work done by such an eddy against buoyancy forces cannot exceed its kinetic energy. If the logarithmic potential temperature difference between an at its maximum vertical eddy element and its environment is displacement, , the work done per unit mass against gravity during the displacement is approximately The measured dispersions were mainly in the horizontal plane. -146- where V is the characteristic eddy speed. and for an adiabatic process, In an isothermal atmosphere this leads to: V This is not in disagreement with the observed dispersions of the puffs If we now insist on which were generally between 100 and 200 meters. interpreting these turbulent dimensions in terms of an eddy viscosity, we obtain \Ye r0.5~ <T nS Kellogg notes that dispersions and derived eddy velocities increase with height, and the value given above can be taken as a plausible lower limit on an eddy viscosity coefficient for the mesosphere, even though the vertical scale of stratospheric turbulence may be somewhat less than 5 meters. Above 80 kilometers information on turbulence has been derived by Greenhow (1959) and Greenhow and Neufeld (1959) from observations of radio-meteor trails. Applying the Kolmogoroff theory to the observed dispersion, they deduce a turbulence power, The same value of , of 7x10-3 watts per was deduced by Blamont and de Jager * kilogram. , (1962) from the observed dispersion of sodium-vapor trails. They also There is some question that the turbulence responsible for the dispersions of both meteor trails and sodium-vapor clouds may be produced by the meteors or rockets involved. -147- report that the observed turbulent elements have diameters of 0.5 km. If these eddies are in the inertial subrange, we should have, approx- imately V 400 m2s. V3 On the other hand if there were a constant coefficient of eddy viscosity (independent of scale) the dispersions of the meteor and sodium clouds should, after a sufficiently long time, exhibit a linear dependence of mean square radius on the time. There is no evidence for such behavior in the observations of Greenhow, or in those of Blamont and de Jager. This fact can be used to place a lower limit on a scale-independent viscosity coefficient from the relation 4-te where r;- is the dispersion at the time of the final observation, 1. On this basis, both the observations of Greenhow and those of Blamont and de Jager give 700 mIs. There is no evidence of the behavior forecast by Bolgiano (1959) for the bouyancy subrange in either the experiments of Greenhow and Neufeld or in those of Blamont and de Jager. On the basis of a Richardson number argument, the eddy viscosity in the middle-mesosphere should be considerably less than that above -14S- 85 kilometers. The turbulence producing shears probably arise from gravity waves and tidal motions. If this is so, the shear should be only about 1/3 as large at 70 kilometers as at 85 kilometers, while the static stability at 70 kilometers is about half of that at 85 kilometers. These values imply a Richardson number which is twice as large at 70 kilometers as at 85 kilometers and consequently turbulence should be less intense. The inference concerning relatively lower shears at mid-mesospheric heights is borne out by the observations of aufm Kampe et al (1962). To sum up: there is no evidence for a true scale-independent eddy viscosity coefficient, but if one insists on interpreting observations of turbulence in terms of an eddy viscosity, probable lower and upper limits on a mid-mesospheric value are 0.5 < V k 400 m2/s. This range suggests that small-scale turbulence by itself may not be sufficiently intense to produce the required diffusion of the momentum and temperature fields. It is possible, however, that motions at some- what larger scales may produce effects resembling dissipation by eddy viscosity and eddy conductivity. In particular, non-linear interactions between gravity waves and the zonal flow might be such a mechanism. D. Some Implications of the Deduced Circulations There are a number of implications which circulations of the type deduced here would have if they actually exist in the mesosphere. -149- Only a few will be discussed here. Any type of convective transport in a stratified fluid will tend to increase the static stability of the fluid. In the present case it is possible to estimate the static stability increase which would result The vertical divergence of the heat flux from the deduced circulations. can be written where hand, L ] now indicates the latitudinal average. On the other the additional heat loss due to the vertical circulation is approximately Cpe LT* where is the temperature increase arising from a circulation pro- duced convergence of heat flux. [T*J will be determined by the requirement that these quantities balance, or r* Using H = = 5x10 LLA (6.1) PJ per second, r = 0.2 cm/sec, 7 = 30 K and 8 kilometers, we obtain [T] Nu 13 K. To obtain the static stability change, we must divide this quantity by the characteristic scale of the mesospheric circulations. -150- If we take D to be 15 kilometers, to we find that the static stability change corresponds lo /km or a percentage change in of less than D 15%. This result can be compared with a similar calculation for largeA scale horizontal eddies. We can again use equation (6.1), but now correspond to the eddies. TP utr and As an example, we may take values for these quantities corresponding to the polar night jet, or i.JrN I cm/s, T 200 K. Such values will lead to [T n 30 K/km, and to static stability modifications of up to 50% of the observed values. It is clear that eddy convection is a much more effective process than the simple cellular overturning considered here as far as static stability changes are concerned. Because of this the good agreement between the over-all observed mesospheric stability and the mean static stability deduced from radiative and photochemical considerations (see Figure 12) supports the hypothesis of a symmetric meridional circulation as the primary convective element in much of the mesosphere. Kellogg (1961) has suggested that an important heating agent in the polar winter mesosphere is the recombination of atomic oxygen in descending air. He estimates that descending velocities as low as .05 centimeters per second at 95 kilometers would be sufficient to produce a heating rate of 100K per day. Such velocities at 95 kilometers are apparently quite feasible within the framework of the models studied here. -151- It should be noted however that if such a large heat source were present at high mesospheric levels, it would have a considerable effect on the circulation, and would in fact probably serve as an important damping agent. It has been pointed out by Newell (1963) that the anamolously warm temperatures in the polar winter mesosphere need not be explained by photochemical action, but may instead be a consequence of eddy motions driven by instabilities at lower levels and characterized by a countergradient heat flux in the upper mesosphere much like the counter-gradient heat flux which has been observed in the stratosphere (White, 1954). The calculations of Chapter 5 indicate that such a temperature distribution may also result from symmetric convection, the motions at the higher levels being simply driven by differential heating at lower levels. Figure 38, showing the temperatures deduced from Model I, this feature. illustrates Either of these two dynamical explanations of the polar winter warmth has the advantage over the chemical heating hypothesis of also explaining the extremely cold temperatures occurring in the upper summer mesosphere. -152- 180 80 200 220 O)A r% 70 1- 60 H 220 240 260 280 340 320 50 -i 40- 30 I vu Fig. 38. - 7U (summer) Temperature OU JU -I latitude-"' distribution corresponding to model I (OK). I AI I (winter) 50 .3L 70 90 7. CONCLUSIONS AND SUGGESTIONS FOR FURTHER RESEARCH Following are the principle conclusions which the analysis presented in the preceeding chapters suggests: i) Calculation of equilibrium temperatures on the basis of presently available radiative and photochemical data indicates that the radiative-photochemical equilibrium state of the mesosphere is everywhere statically stable, at least when possible reactions involving hydrogen are neglected. ii) The relevant non-dimensional parameter determining the linearity of a thermally driven symmetric circulation is proportional to the product of the Prandtl number, the annual frequency, the vertical scale of heating and motions and the characteristic time for thermal damping. the scale height. It is inversely proportional to On the basis of radiative calcu- lations and the observed phase lag of the mean zonal wind, and assuming a Prandtl number of unity, this parameter, which approximates the ratio of the nonlinear to linear contributions to the solution, has the value 1/3 for the mesosphere. The remaining conclusions are therefore based on the linearized equations. -154- M iii) A thermal damping parameter arising from infrared radiation and from temperature dependence of the photochemical reaction rates can be derived, and it has the order of magnitude 10-6 per second. This is large enough to have an important damping effect on any symmetric thermally driven circulation in the mesosphere. iv) When damping is accomplished entirely by eddy viscosity and eddy conductivity, the deduced mesospheric heat sources produce a single cell meridional circulation from summer pole to winter pole at high levels with a much weaker return circulation below 35 kilometers. The meridional circulation is approximately proportional to the Prandtl number, Prandtl number equal to unity, and for the the maximum horizontal velocity is about 80 cm/s occurring near 65 kilometers and in middle and low latitudes. The phase of the cir- culation relative to the external heating is roughly proportional to one minus the Prandtl number P i.e., the meridional circulation leads the heating when (I-P) >0 and lags behind the heating when (I-P) < 0 v) When linear thermal damping of Newtonian type -155- cor- responding to radiative and photochemical effects is included in a model with constant eddy viscosity, the meridional circulation consists of two cells, one above the other; the strongest branch is from summer pole to winter pole above 40 kilometers, weaker flow from winter pole to summer pole between 8 and 40 kilometers, and much weaker return flow below 8 kilometers. An example of such a circula- tion calculated with the radiative damping constant = 10 6 sec and eddy viscosity and eddy conductivity coefficients V = -K = 300 m2 gives maximum amplitudes of 30 cm/sec, 10 cm/sec and 4 cm/sec for the three branches. The circula- tion amplitudes are roughly proportional to the ratio DX +X where V is a suitable characteristic vertical scale of the heating. The radiative damping constant introduces a phase difference between meridional circulation and heating even in the case when irv) For the zonal circulation, P = 1. the amplitude and phase lag are both roughly proportional to the characteristic -156- time for thermal dissipation i.e., "'/ K- when Newtonian thermal damping is not included and :D. ( N when it is included. + Reasonable agreement with the observed zonal winds in amplitude, phase lag, of the wind maxima and height and latitude can be obtained by assuming that the upper boundary condition on the zonal wind is determined by molecular viscosity and conductivity. vii) The best agreement with zonal wind observations is obtained by assuming These values of t/ are somewhat higher than estimates based on deductions about small-scale turbulence from observations of the dispersion of sodium-vapor clouds and meteor trails above 80 kilometers, and smoke puff diffusion in the stratosphere would suggest. viii) In an atmosphere in which molecular viscosity and conductivity are responsible for diffusion of the zonal momentum and temperature fields, and at altitudes above all heat sources, the meridional components of motion increase exponentially with height while the vertical mass flux decreases exponentially with height. The zonal velocity and -157- the temperature decrease exponentially with height. Such a solution can only be valid in a limited height range because it implies a breakdown of the geostrophic approximation at very high levels. ix) The circulations deduced are not strong enough to account for the observed latitudinal distribution of the artificial tracer Rhodium 102 without additional dispersion by horizontal mixing processes. The circulations would assist any lateral mixing to produce the observed distributions, however. Several avenues of research which are now being pursued may shed additional light on the problems considered here. Attempts to evaluate the ozone reaction constants in the laboratory are going on, and the results that have been obtained in this work suggest that the value of together with its dependence on temperature is the most critical quantity for the understanding of the dynamics of the mesosphere. Another important question to be answered is: how much water vapor is there in the stratosphere and mesosphere? This is important, both because water vapor may make an important contribution to the heat balance through infrared radiation, and because this information is fundamental to the photochemical problem of the reactions involving hydrogen at higher levels. Further refinement of the photochemical-radiative equilibrium and heating rate calculations presented in Chapter 4 should probably include water vapor radiation and photochemistry. -158- Further direct observations of the ozone concentration in the mesosphere, such as that of Johnson, Purcell, and Tousey (1951) would be valuable, especially, since the ozone concentration at lower mesospheric levels is a sensitive indicator of the accuracy of the parameters in the photochemical theory. On the theoretical side, study of the non-linear interactions of tidal and gravity waves with the zonal flow could shed light on what may turn out to be important mechanisms for the transport of momentum and heat in the mesosphere. Measurements of the intensity of small scale turbulence by observing the dispersion of artificially produced clouds in the main part of the mesosphere, analogous to the observations of sodium vapor clouds at higher levels, would be valuable. Finally, it seems clear that no really conclusive statements as to the character of the meridional circulations in the mesosphere can be made until momentum and heat fluxes due to the large scale eddies is known with some degree of confidence. This knowledge will require a good distribution of observations with longitude in order to eliminate the effect of standing waves, and a good distribution with time of day in order to allow separation of tidal effects. When these fluxes and their divergences become known, it should be possible to determine whether the resulting sources are "small " in the sense of the analysis of Chapter 3. If they are small, it should be possible to incorporate their effect into the present theoretical framework to evaluate their -159- contribution to the meridional circulation in much the same way that Kuo (1956) evaluated the effect of empirically determined momentum fluxes on tropospheric mean meridional circulations. -160- Appendix 1. A. Details of the Photochemical & Radiative Calculations Photochemical Data - For the photochemical computations, the follow-ing information is necessary: the spectral distribution of solar energy outside of the earth's atmosphere, the spectral distribution of the absorption coefficients of ozone and molecular oxygen and the dependence of these coefficients on temperature and pressure, the values of the reaction rate coefficients , 1- , and 3 and their dependence on temperature, and the spectral distribution of the quantum yield factors, Dutsch e.. and e3 (1961) has reviewed out knowledge of these quantities, but more recent data is now available in all but the last of these areas. It is convenient to discuss the data according to spectral regions. i) From 13530 cm~ to 20700 cm~ (7392 to 4831 A), Chapuis bands of ozone must be considered. the These are relatively weak, but they are of importance in determining ozone concentrations below 35 kilometers. Ozone absorption coefficients in this spectral region were taken from Vigroux (1953). Absorption in these bands is essentially independent of pressure, and Vigroux's work also indicates that they are quite insensitive to temperature as well. workers In accordance with earlier e.g.: Craig (1950), Diitsch (1961), it has been assumed that all solar absorption by ozone at wave-lengths shorter than 11,000 A leads to direct A-1 e dissociation, so that has been taken to be unity for all spectral intervals considered. The solar spectral intensity for this region was taken from Johnson (1954). ii) From 28750 cm-1 to 40000 cm 1 (3475 to 2500 A) absorption is by the Hartley-Huggins bands of ozone. Absorption coefficients for this region were also Temperature dependence is taken from Vigroux. slightly more important in the long-wave portion of this region than in the Chapuis bands, but this effect has again been neglected. Johnson's paper was again used as the source for the solar spectral intensity data. iii) -l (2500 to 2000 A) ozone From 40,000 to 50,000 cm absorption coefficients were taken from Inn and Tanaka (1953). Again it is safe to neglect pressure and temperature dependence. Beginning at 41400 cm 1 (2720 A), the Herzberg continuum of molecular oxygen produces direct dissociation ( e4= 1). This absorption is pressure dependent, but the recent work of Ditchburn and Young (1962) indicates that at pressures less than about 30 millibars this effect can be safely neglected, so that these coefficients have been assumed independent of pressure in this work. A-2 The absorption coefficients of Ditchburn and Young have been used. Solar spectral intensities were taken from Detwiler, Garrett, Purcell, and Tousey (1961); their values overlap with those of Johnson near 40,000 cm-1. iv) From 50,000 to 60,000 cm~ 1 (2000 to 1667 A) absorp- tion is primarily by the Schumann-Runge bands, although the Schumann-Runge continuum occupies the short wave end of this region. These are very sharp bands and do not lead to direct dissociation. Dissociation may take place, however, as a result of the following processes: direct dissociation to Z the state in the Herzberg continuum which underlies the long wave end of the region, direct dissociation in a continuum resulting from a transition from the ground state to the 3 T, state, and predissociation resulting from perturbation of the upper state of the Schumann-Runge transition 3 ( E ) with the -TT, state. The intensity of the Herzberg continuum has been estimated theoretically by Ditchburn and Young for frequencies greater than 50,000 cm 1, while Wilkinson and Mulliken (1957) have observed the intensity of an underlying continuum, which they assumed was due to the 3 A-3 TT, transition, at 56,200 cm 1 and 55,600 cm 1. The latter authors also observed well-defined broadening attributable to predissociation at the Lr= 12 level. Absorption intensities in the Schumann-Runge bands have been measured by A. Vassy (1941) and by Watanabe, and Zelikoff (1953). rapid; Inn, Intensity variations are very in the present calculation a smoothed absorp- tion spectrum was used based on the observations of Watanabe et al for the short-wave end of the region and on Vassy for the longer wave-lengths. Pressure and temperature dependence has again been neglected. The quantum yield factor, e , . was estimated from the results of Wilkinson and Mulliken, and Ditchburn and Young. The solar spectrum for this region was based on recent data which were kindly provided by Dr. R. Tousey of the Naval Research Laboratory. These data are shown in Figure Al and exhibit lower intensities than those published by Detwiler et al by as much as a factor of two near 1880 A. Ozone absorption for this region was taken from Tanaka, Inn and Watanabe (1953), and was again assumed to be independent of pressure and temperature. Below 1667 A (60,000 cm 1) there is a strong absorption by molec- ular oxygen in the Schumann-Runge continuum, A-4 but this radiation does not 5500 --- -Si -- -- ~~ ~ UI (1) z U)500K1.0 - Al M (1) S i II (1)A 0* ---ll - - Al I Mg 1 (2) -- ) 10.0 SI (2 Ln 4500*K_~ J a- 01 0 U) 1750 1800 1850 1900 1950 2000 WAVELENGTH (A) Fig. Al. Solar spectrum from 1750 to 2100 A. Photgraph provided by Dr. R. Tousey of the Naval Research Laboratory. 2050 2100 penetrate below 85 kilometers, and was not considered. The spectral data actually used in the computations are given in Table Al. Integration over the spectrum was done by simply summing over the spectral intervals. The height increment used in the photochemical computations was 1.5 kilometers. Table A2 gives the data used in computing the time constants illustrated in Figure 2. Experimental data bearing on the reaction rates ,and have recently been reviewed by Harteck and Reeves Kaufman and Kelso (1961). Attempts to measure (1961), and by and by no less than nine different groups have been reported since 1957, but there is a wide divergence of results, especially for only three recent attempts to measure - . On the other hand, have been reported (Benson and Axworthy, 1957; Leighton et al, 1959; Phillips and Schiff, 1962), and only two of these have measured the temperature dependence. There is considerable disagreement between these two (Benson and Axworthy and Leighton et al), For these reasons, the rate constants I are the source of greatest uncertainty in the calculations. A-6 and I sity (photons/cm -sec per wave number multi- plied by 10 13500. .000 258 14500. 15500. 16500. .000 243 .000 223 .000 205 17500. 18500. .000 183 19500. 20500. .000 130 .000 119 29375. 30625. 31875. 33125. .000 022 .000 018 -17 Absorption Coefficients 2 (cm /molecule multiplied by 10 Molecular Oxygen .000 160 Width of Spectral Interval (cm -1 .000 035 .000 130 1000. .000 275 .000 460 1000. .000 .000 .000 .000 440 1000. 290 1000. 175 1000. 085 1000. .000 012 .000 008 5 .025 34375. 35625. 36875. .000 007 1 38125. .000 001 5 .11 .36 .65 .91 39375. 40625. 41875. 43125. .000 000 79 .000 000 63 .000 000 45 .000 000 42 .000 000 03 .000 000 11 .93 .74 .000 000 25 .50 44375. 45625. 46875. 48125. .000 000 38 .000 000 31 .000 000 20 .000 000 11 .000 000 37 .000 000 64 .000 000 86 .000 001 02 .305 Table Al. efficiency for oxygen dissociation Ozone .000 12 .001 1 .005 2 .000 003 4 .000 002 4 Quantum 17 ) 2 ) -1 ) (cm Solar Spectral Inten- ) Mean Wave Number 1.00 .172 .082 .040 Spectral data used in photochemical calculations. 1000. 1000. 1250. 1250. 1250. 1250. 1250. 1250. 1250. 1250. 1.00 1.00 1250. 1250. 1.00 1.00 1250. 1250. 1.00 1.00 1.00 1.00 1250. 1250. 1250. 1250. (Page 1 of 2) I ) (cm Solar Spectral Intensity (photons/cm2-sec per wave number multi) plied by 1017 00 Absorption Coefficients (cm2/molecule multiplied by 10 1) Molecular Oxygen 49375. .000 000 062 .000 001 16 50312.5 .000 000 040 .000 001 50 50937.5 51562.5 .000 000 035 .000 002 50 .000 000 023 .000 005 52187.5 .000 000 017 .000 013 5 52812.5 .000 000 014 .000 071 53437.5 54052.5 .000 000 012 .000 000 011 .000 26 54677.5 55312.5 55937.5 56875. .000 000 012 .000 000 006 2 .000 000 003 2 .001 .002 .005 .016 58125. 59375. .000 000 002 32 .052 .000 000 001 72 .1 .000 000 004 8 Table Al. .000 62 39 65 4 8 Quantum efficiency for oxygen dissociation Ozone .030 .034 .038 .042 1.00 .94 .60 .33 .049 .056 .061 .067 .057 .028 .014 .015 .072 .071 .074 .077 .22 .080 .25 1.00 .080 .080 1.00 1.00 Spectral data used in photochemical calculations. Width of Spectral Interval (cm -1 ) Mean Wave Number (Page 2 of 2) 1250. 625. 625. 625. 625. 625. 625. 625. 625. 625. 625. 1250. 1250. 1250. Height (kilometers) Oxygen Concentration (partigles per cm 85 .301x10 14 . 548x106 .169x10 7 .953x10-2 .183x10- 79 .820x1014 .406x10- 1 6 .722x10- 8 . 953x10- 2 .327x10- 5 73 .219x1015 .927x10- 16 .334x10-8 .952x10- 67 .548x1015 .278x10- 1 5 61 .127x10 1 6 .950x10- (sec -1 ) ) (sec 3 -1 ) (sec -1 ) J 3 (cm /sec) 15 4 55 .274x1016 .213x10O' 49 .576x1016 4 .197x10 14 43 .124x10 7 .122x10~ 14 2 .195x10- 8 .949x10-2 .146x10- 8 .928x10- .122x10- 8 .100x10-8 .640x10~ 9 9 5 .617x10- 5 .146x10~ 4 2 .532x10~4 2 .762x10- .167x10~ 4 .638x10- 2 .184x10- 3 .282x10- 2 .657x10~0 .921x10- 3 .904x10- 5 .277x1017 .475x10-15 .196x10~ 31 .668x10 7 .185x105 .205x10- 1 0 .530x10-3 .889x10- 6 25 .169x10 1 8 . lllxo- 15 .208x10- 12 .416x10- 3 .382x10 7 Table A2 - 37 Calculation of time constants used in figure 2. A-9 I We have used the values -4 -32 = .27x10 6 cm /molecule in agree- ment with Reeves, Mannella and Harteck (1960), and with Kaufman and ,the For Following Benson and Axworthy, we have used 3 = 5 X 10 exp (-3o25-/7) value 1.lx1- Kelso has been adopted. minations of cI/sec . Kelso (1961). cm /sec which was found by Kaufman and This is the lowest of recent experimental deter- reported in the literature, and was chosen primarily because of the finding by Phillips and Schiff that the rate of the reactions Q + OH -+H + H-+-02+M -+ Oz. H02+M is extremely fast, so that higher values of obtained in other ex- periments may have arisen because of contamination by a reaction chain initiated by these reactions. Temperature dependence of should be unimportant on theoretical grounds and (Bates and Nicolet, 1950), and it has been neglected. B. Calculation of C O2 Emission - We have attempted to adhere as closely as possible to the procedure followed by Murgatroyd and Goody. The Curtis matrix method assumes non-overlapping lines, and the cooling at the level resulting from each such non-overlapping band is given by . P. A-10 t th T) is the Lorentz half-width of the lines at the pressure 0C, , where is the number of lines in the band, and is the line strength parameter: O band (cm2 -wave number per gram), IAL K has the value 106. the mass mixing ratio of T If the acceleration of gravity. per day, is the mean line strength in the is in degrees Kelvin Murgatroyd and Goody divided the lines in the 15 micron bands into two groups: "tweak lines", and "strong lines", Y- 933, Tis = 108. 'Y The value of = 33, C = 372, In used was The Curtis matrices have been .064 cm-1 at one atmosphere pressure. computed for CO. and Y = 5, 100, 560, and 950. The matrix for 950 was used , but for the strong lines with a slight compensating correction to T)l a new matrix was interpolated graphically for the weak lines. Graphical interpolation of the elements on the principle diagonal of the latter matrix was guided by the requirement that the radiation to space for each row, i.e.: 5 , should also interpolate smoothly. The resulting weak-line matrix was compared with that of Murgatroyd and Goody by doing several sample calculations using vertical temperature distributions of Murgatroyd (1957). The results were in very good agreement with those of Murgatroyd and Goody suggesting that the cooling rates are not very sensitive to the exact interpolation used. The matrix elements were computed with pressure as vertical coordinate. This coordinate was converted to a height scale with A-11 increments of 3 kilometers between adjacent layers. The error arising from this conversion amounts to a height difference of not more than 4 kilometers, except in the computation of equilibrium temperatures in the polar night zone. As discussed by Curtis and Goody (1956), above 75 kilometers the decrease in the ratio of collision frequency to vibrational emission frequency causes deviations from local thermodynamic equilibrium. This effect was taken into account approximately by Murgatroyd and Goody, and in the present calculation the matrix rows corresponding to levels above 75 kilometers were simply multiplied by constant factors to bring the results into agreement with those of Murgatroyd and Goody. C. Calculation of Ozone Emission - Figure A2 is a plot of the cooling rates calculated by Plass (1956) versus the Planck function at the center of the 9.6 micron band, B . These data suggest the following formula which was used for ozone infrared cooling: CP C + C B0 dT .at where the value of C, and C in ergs/gm-sec and cm 2-wave number/gm are given in Table A3. A-12 ML 11 10 X 20 kms o 30 kms & 40 kms + 50 kms a 60 kms t 9 8 7 6 IN 5 0 4-) 4 4-) 3 At + 2 1 a a 0 0 x x I 100 i 200 I 300 400 B 0 3 (ergs/cm Fig. A2. I 2 - 500 sec - wave - 600 number) 700 Cooling rate due to ozone infrared emission versus the 9.6 micron Planck function. from Pless (1956). 800 900 Data taken I. Height, Kilometers C, 64 0. 0. 61 31.5 -0.69 58 52.0 -1.15 55 65.5 -1.46 52 70.0 -1.55 31-49 70.8 -1.571 28 82.0 -1.571 25 95.0 -1.571 22 110.0 -1.571 Table A3. Constants used in calculating cooling due to ozone infrared emission. A-14 REFERENCES Aufm Kampe, H.J., M.E. Smith, R.M. Brown, 1962: 110 kms. J. Geophys. Res., 67, 4243-4257. Winds between 60 and Bates, D.R., and M. Nicolet, 1950: The photochemistry of atmospheric water vapor. J. Geophys. Res., 55, 301-327. Batten, E.S., 1961: Wind systems in the mesosphere and lower ionosphere. J. Meteor., 18, 283-291. Benson, S.W., and A.E. Axworthy, Thermal decomposition of ozone. 1957: Mechanism of the gas phase. J. Chem. Phys., 26, 1718-1726. Blamont, J.E., and C. de Jager, 1962: Upper atmosphere turbulence determined by means of rockets. J. Geophys. Res., 67, 3113-3119. Bolgiano, R., Jr., 1959: Turbulent spectra in a stably stratified J. Geophys. Res., 64, 2226-2229. atmosphere. Brewer, A.W., 1949: Evidence for a world circulation provided by the measurements of helium and water vapour distribution in the stratosphere. Quart. J. R. Meteor. Soc., 75, 351-363. Brooks, D.L., 1958: The distribution of carbon dioxide cooling in the lower stratosphere. J. Meteor., 15, 210-219. Charney, J.G., 1948: 17, no. 2, 17 pp. On the scale of atmospheric motions. Geof. Publ., Propagation of planetary scale Charney, J.G., and P.G. Drazin, 1960: J. Geophys. Res., disturbances from the lower into the upper atmosphere. 66, 83-109. On the stability of baroclinic jets Charney, J.G., and M.E. Stern, 1962: in a rotating atmosphere. J. Atmos. Sci., 19, 159-172. Cole, A.C., and A.J. Kantor, J. appl. Meteor., 2, 90-98. 1963: Tropical and subtropical atmospheres. Craig, R.A., 1950: The observations and photochemistry of atmospheric ozone and their meteorological significance. Meteor. Monographs, 1, no. 2, American Meteorological Society, Boston. 50 pp. Craig, R.A., and G. Ohring, 1958: The temperature dependence of ozone radiational heating rates in the vicinity of the mesopeak. J. Meteor., 15, 59-62. The computation of radiative heating rates in the Curtis, A.R., 1956: Soc., A236, 156-159. R. atmosphere. Proc. Curtis, A.R., and R.M. Goody, 1956: Thermal radiation in the upper atmosphere. Proc. R. Soc., A236, 193-206. Detwiler, C.R., D.L. Garrett, J.D. Purcell, and R. Tousey, 1961: The solar spectrum in the extreme ultraviolet. Ann. de Geophysique, 17, 9-18. Dickinson, R.E., 1962: Momentum balance of the stratosphere during the I.G.Y. S.M. Thesis, Dept. of Meteor., Mass. Inst. of Technology, 32 pp. Ditchburn, R.W., and P.A. Young, 1962: The absorption of molecular oxygen between 1850 and 2500 A. J. Atmos. and Terres. Phys., 24, 127- 139. Origin and distribution of the polyatomic molecules Dobson, G.M.B., 1956: in the atmosphere. Proc. R. Soc., A236, 187-192. Current problems of the photochemical theory of DU*tsch, H.U., 1961; atmospheric ozone. In Chemical Reactions in the Lower and Upper Atmosphere (proceedings of an international symposium at Stanford Interscience Pub., New York. Research Institute), 167-180. Eliassen, A., 1949: The quasi-static equations of motion with pressure as independent variable. Geof. Publ., 17, no. 3, 44 pp. Eliassen, A., 1950: Slow thermally or frictionally controlled meridional circulation in a circular vortex. Astrophys. Norv., 5, 19-60. Eliassen, A., and E. Kleinschmidt, 1957: Handbuch der Physik, 48, 1-154. Dynamic meteorology. In Friend, J.P., H.W. Peely, P.W. Krey, J. Spar, and A. Walton, 1961: Discussion of HASP results. Final Report, volume III, High Altitude Sampling Program, DASA, Washington, D.C. 307 pp. Gowan, E.H., 1947: Ozonosphere temperatures under radiative equilibrium. Proc. R. Soc., A190, 219-225. Greenhow, J.S., 1959: Eddy diffusion and its effect on meteor trails. J. Geophys. Res., 64, 2208-2209. Greenhow, J.S., and E.L. Neufeld, 1959: Measurements of turbulence in the upper atmosphere. Proc. Phys. Soc. of London, 74, 1-10. Harteck, P., and R.R. Reeves, 1961: Recent investigations of chemical reactions of fundamental importance in the atmosphere. In Chemical Reactions in the Lower and Upper Atmosphere (proceedings of an international symposium at Stanford Research Institute), 219-238. Interscience Pub., New York. Haurwitz, B., 1961: Frictional effects and the meridional circulation in the mesosphere. J. Geophys. Res., 66, 2381-2391. Hines, C.O., 1961: Internal atmospheric gravity waves at ionospheric heights. Canadian J. Physics, 38, 1441-1481. Hines, C.O., 1963: The upper atmosphere in motion. Quart. J. R. Meteor. Soc., 89, 1-42. Holl, M.F., 1961: Thermally driven axially symmetric circulations of a planetary atmosphere. Sci. Rpt. no. 3, Stanford Research Institute, Contract number AF 19(604)-5965. Inn, E.C.Y., and Y. Tanaka, 1953: Absorption coefficients of ozone in the ultraviolet and visible regions. J. Opt. Soc. Amer., 43, 870-873. Johnson, F.S., 1954: The solar constant. J. Meteor., 11, 431-439. Johnson, F.S., J.D. Purcell, and R. Tousey, 1951: Measurements of the vertical distribution of atmospheric ozone from rockets. J. Geophys. Res., 56, 583-594. Kalkstein, M.I., 1962: Science, 137, 645-652. Rhodium-102 high altitude tracer experiment. Kampe de Feriet, J., 1951: in atmospheric turbulence. Averaging processes and Reynolds equations J. Meteor., 8, 358-361. Karandikar, R.V., 1946: Radiation balance of the lower stratosphere, Proc. Ind. Acad. Sci., 23, 70-96. I. Kaufman, F., and J.R. Kelso, 1961: The homogeneous recombination of atomic oxygen. In Chemical Reactions in the Lower and Upper Atmosphere (proceedings of an international symposium at Stanford Research Institute), 255-268. Interscience Pub., New York. Kellogg, W.W., 1956: 13, 241-250. Diffusion of smoke in the stratosphere. J. Meteor., Xellogg, W.W., 1961: Warming of the polar mesosphere and lower ionosphere in winter. J. Meteor., 18, 373-381. Kellogg, W.W., and G.F. Schilling, 1951: A proposed model of the circulation of the upper stratosphere. J. Meteor., 18, 373-381. Kuo, H.L., 1954: Symmetrical disturbances in a thin layer of fluid subject J. Meteor., 11, 399-411. to a horizontal temperature gradient and rotation. Kuo, H.L., 1956: Forced and free meridional circulations in the atmosphere. J. Meteor., 13, 561-568. Lamb, H., 1932: Hydrodynamics. New York. 738 pp. Reprinted 1945 by Dover Publications, Leighton, F., H.B. Uhrbach, J.A. Wojtowicz, and J. Zaslowsky, 1959: The kinetics of the reaction of atomic oxygen with ozone. Abstract of papers presented at A.C.S. Meeting, Atlantic City, N.J., Sept. 1959, p. 465. Libby, W.F., and C.E. Palmer, 1960: Stratospheric mixing from radioactive fallout. J. Geophys. Res., 65, 3307-3317. The pattern of global atmospheric List, R.J., and K. Telegadas, 1961: radioactivity - May 1960. In Health and Safety Laboratory, Fallout Program, quarterly summary report, April, 1961, US AEC, New York, 186-197. Lorenz, E.N., 1953: A proposed explanation of two regimes of flow in a In Fluid Models in rotating symmetrically-heated cylindrical vessel. Geophysics (proceedings of a symposium at Johns Hopkins University), Supt. of Documents, Washington, D.C., 73-80. Lorenz, E.N., 1955: Available potential energy and the maintenance of of the general circulation. Tellus, 2, 157-167. Machta, L., 1959: Transport in the stratosphere and through the tropopause. Adv. in Geophys., 6, International Symposium on Atmospheric Diffusion and Air Pollution, 273-288. Manabe, S., and F. Niller, 1961: On the radiative equilibrium and heat balance of the atmosphere. Mon. Wea. Rev., 89, 503-531. Mantis, H.T., 1963: Note on the structure of the stratospheric easterlies of mid-latitude. J. Appl. Meteor., 2, 427-429. Minzner, R.A., K.S.W. Champion, H.L. Pond, 1959: The ARDC model atmosphere, 1959. Air Force Surveys in Geophsics, no. 115, AFCRC-TR-59267, Geophysics Research Directorate, Air Force Cambridge Research Center, Air Research and Development Command, United States Air Force, Bedford, Mass. Morse, P.M., and H. Feshbach, 1953: Methods of theoretical physics. 997 pp. Vol. I. McGraw-Hill Company, New York. - Murgatroyd, R.J., 1957: Winds and temperatures between 20 km and 100 km a review. Quart. J.R. Meteor. Soc., 83, 417-458. Sources and sinks of radiative Murgatroyd, R.J., and R.M. Goody, 1958: Soc., 84, 225-234. Meteor. J.R. Quart. 90 km. 30 to energy from Murgatroyd, R.J., and F. Singleton, 1961: tions in the stratosphere and mesosphere. Possible meridional circulaQuart. J.R. Meteor. Soc., 87, 137-158. The transport of trace substances in the atmosphere Newell, R.E., 1961: and their implications for the general circulation of the stratosphere. Geofisica Pura e Applicata, Milano, 49, 137-158. preliminary study of quasi-horizontal eddy fluxes Newell, R.E , 1963: from meteorological rocket network data. J. Atmos. Sci., 20, 213-225. Nordberg, W., and W.S. Stroud, 1961: Results of I.G.Y. rocket-grenade experiments to measure temperature and winds above the island of Guam. J. Geophys. Res., 66, 455-464. On the phenomena of motion in the atmosphere. Oberbeck, A., 1888: Sitzberichte der Kniglichen Preussischer Akademie der Wissenschaffen, translated by C. Abbe, and reprinted in Selected Papers on the Theory of Thermal Convection, B. Saltzman, ed. Dover Publications, New York. Ohring, G., 1958: The radiation budget of the stratosphere. J. Meteor., 15, 440-451. Oort, A.H., 1962: Direct measurement of the meridional circulation in the stratosphere during the I.G.Y. Report no. 6, Contract numbers AT (30-1)2241 and AF 19(604)-5223, Planetary Circulations Project, Dept. of Meteor., Mass. Inst. of Technology. Paetzold, H.K., 1961: Photochemistry of the atmospheric ozone layer. In Chemical Reactions in the Lower and Upper Atmosphere (proceedings of an international symposium at Stanford Research Institute), 181-195. Interscience Pub., New York. Phillips, L.F., and H.I. Schiff, 1962: Mass spectrometric studies of atomic reactions, III. J. Chem. Phys., 37, 1233-1238. The influence of the 9.6g ozone band on the Plass, G.N., 1956: atmospheric infra-red cooling rate. Quart. J. R. Meteor. Soc., 82, 30-44. The latitudinal and seasonal variations in the Pressman, J., 1954: absorption of solar radiation by ozone. J. Geophys. Res., 59, 485-498. Reeves, R.R , G. Mannella and P. Harteck, 1960: of oxygen atoms. J. Chem. Phys., 32., 632-633. Rate of recombination Perturbation equations for the time average state Saltzman, B., 1961: of the atmosphere including the effects of transient disturbances. Geofisica Pura e Applicada, Milano, 52, 173-188. Stratton, J.A., P.M. Morse, L.J. Chiu, J.D.C. Little, F. Corbat6, 1957: Spheroidal Wave Functions. John Wiley and Sons, New York. 613 pp. Stroud, W.G., W. Nordberg, W.R. Bandeen, F.L. Bartman and P. Titus, 1960: Rocket-grenade measurements of temperature and winds in the mesosphere over Churchill, Canada. J. Geophys. Res., 65, 2301-2323. Stroud, W.G., W. Nordberg,and J.R. Walsh, 1956: Atmospheric temperatures and winds between 30 and 80 km. J. Geophys. Res., 61, 45-56. Starr, V.P., 1954: Studies of the atmospheric general circulation. Final Rpt., Part I, General Circulation Project, Contract number AF 19-122-153, Dept. of Meteor., Mass. Inst. of Technology, 535 pp. Tanaka, Y., E.C.Y. Inn, and K. Watanabe, 1953: of gases in the extreme ultraviolet, Part IV: Absorption coefficients Ozone. J. Chem. Phys., 21, 1651-1653. Teweles, S., and F.G. Finger, 1962: Synoptic studies based on rocket sonde data. In Space Research III, Amsterdam, North Holland Publishing Company. Vassy, A., 1941: L'absorption de l'atmosphere dans l'ultraviolet. Ann. de Physique, 16, 145-203. Vigroux, E., 1953: Contribution a l'etude expdrimentale de l'absorption de l'ozone. Ann. de Physique, 8, 709-762. Watanabe, K., E.C.Y. Inn, and M. Zelikoff, 1953: Absorption coefficients of oxygen in the extreme ultraviolet. J. Chem. Phys., 21, 1026-1030. Wallace, L., 1962: The OH nightglow emission. J. Atmos. Sci., 19, 1-16. White, R.M , 1954: The counter-gradient flux of sensible heat in the lower stratosphere. Tellus, 6, 177-179. Wilckens, F., 1962: Bemerkungen zur derzeitegen kenntnis der meridionalen zirkulation in der stratosphare und mesosphlre. Meteorologische Rundschau, 15, 23-27. Wilkinson, P.G., and R.S. Mulliken, 1957: Dissociation processes in oxygen above 1750 A. Astrophys. J., 125, 594-600. Witt, G., 1962: Height, structure and displacement of noctilucent clouds. Tellus, 14, 1-18. .=a Biographical Note The author was born July 16, 1933 in Hermosa Beach, California. He attended the University of Southern California from 1950 to 1954 when he received his A.B. in the Division of Physical Sciences and Mathematics. From 1954 to 1958, he served in the Air Force as a weather forecaster in Korea and in Florida, and as a weather reconnaisance observer in the Pacific. His forecaster training was at the University of California at Los Angeles. graduate student at Since 1958, he has been a Massachusetts Institute of Technology, and has occasionally been a consultant to the Rand Corporation. to the former Janet Seitz, and has two children. He is married