15 NOVEMBER 2001

ZURITA AND LINDZEN

3443

The Equilibration of Short Charney Waves: Implications for Potential Vorticity Homogenization in the Extratropical Troposphere PABLO ZURITA

AND

RICHARD S. LINDZEN

Program in Atmospheres, Oceans and Climate, Massachusetts Institute of Technology, Cambridge, Massachusetts (Manuscript received 19 December 2000, in final form 30 April 2001) ABSTRACT In this paper the authors discuss the vertical distribution of the interior potential vorticity gradients in the extratropical troposphere in an idealized setting. This structure is characterized by large positive gradients at the top of the boundary layer and a narrow homogenized region in the interior. To understand this partial homogenization, the authors have looked at how Charney waves of different scales equilibrate. It is shown that when the interior gradients of potential vorticity are small compared to the delta function at the ground, the modes can equilibrate without eliminating the lower-level shear. As is the case in the Eady model, these neutral states can exist even though they violate the Charney–Stern condition for stability. The equilibration process also produces a concentration of the potential vorticity gradients at heights of the order of the Rossby depth, which can be regarded as a boundary to the interior mixing of potential vorticity. The extent of interior potential vorticity mixing required for equilibration depends on the ratio between the boundary and interior potential vorticity gradients. It is implicit in the analysis of previous authors that this ratio is close to 1 in midlatitudes, which implies that an Eady model with full homogenization of the interior potential vorticity gradients is not an appropriate paradigm for the midlatitude troposphere. Nevertheless, it is argued that localized interior potential vorticity mixing could be sufficient for neutralizing the shortest waves.

1. Introduction It has long been recognized that baroclinic eddies play a role in regulating the tropospheric temperature gradients by transporting heat poleward in the midlatitudes. In linear theory, the potential vorticity fluxes (equivalent to heat fluxes at the boundaries) induced by the eddies are downgradient and therefore tend to reduce the gradients of potential vorticity (PV) of the mean state. For this reason, the extratropical troposphere is often regarded as a product of the competing effects of the destabilizing radiative forcing and the stabilizing eddy tendencies. Baroclinic adjustment theories assume that the eddies act on faster timescales than the radiative forcing and envision in consequence a close to neutral tropospheric state, a situation analogous to convective adjustment. However, because in the baroclinic problem the time scale separation is not as clear, this adjustment should be regarded as an idealized limit, as discussed by Barry et al. (2000). The manner and extent to which baroclinic adjustment is achieved by the eddies is still an open issue. The equilibration of the baroclinic eddies implies a Corresponding author address: Dr. Pablo Zurita, Program in Atmospheres, Oceans and Climate, 54-1611, Massachusetts Institute of Technology, Cambridge, MA 02139. E-mail: [email protected]

q 2001 American Meteorological Society

mixing of the potential vorticity gradients in the interior as well as the temperature gradients at the ground, and the analysis of Stone and Nemet (1996) suggests that this is close to observed. Yet the lowertroposphere temperature gradient is significantly different from zero. Although this could be attributed to the strong vertical mixing and damping of thermal anomalies over the boundary layer (e.g., Gutowski et al. 1989; Swanson and Pierrehumbert 1997), it is remarkable that, if we compare the mean isentropic slopes calculated assuming zero interior PV gradient to observations, the agreement is better when the boundary temperature gradient is neglected than when the observed value is used as boundary condition (Stone and Nemet 1996). Given that the quasigeostrophic PV gradient is essentially the vertical derivative of the isentropic slope (see section 2c), the better agreement found when the negative delta-function PV gradient at the ground is neglected suggests that the positive interior PV gradient might also be nonnegligible, but rather have a vertical integral of the same order as the integrated delta function at the ground. In fact, the Charney–Stern criterion (Charney and Stern 1962) is only a necessary condition for instability. Noting that, Lindzen (1993) pointed out that short waves are neutral in the Eady problem regardless of the shear. He also noted that the meridional confinement by the jet puts a constraint on the vertical

3444

JOURNAL OF THE ATMOSPHERIC SCIENCES

penetration of the modes, and argued that baroclinic eddies could equilibrate by mixing the interior PV gradient if the tropopause is high enough, even without eliminating the lower-troposphere temperature gradient. This is also consistent with the conception of the tropopause as a boundary to eddy mixing in the midlatitudes (Held 1982). Lindzen (1993) argued that the interior PV gradient is small enough to be indistinguishable from observational error. However, a more detailed analysis (KirkDavidoff and Lindzen 2000; Solomon and Stone 2001) shows that the interior gradient of potential vorticity has a well-defined structure, with large positive values at the top of the boundary layer and a relatively narrow homogenized region at about 700 hPa. We will show below that, in fact, the interior PV gradient is in some sense comparable to the boundary PV gradient. This is consistent with the fact that observed atmospheric shears tend to be close to the critical shear of the twolayer model (Stone 1978). When extrapolated to a continuous atmosphere of the same depth, this shear is such that the integrated PV gradient over its lower half exactly balances the delta function at the lower boundary. Another remarkable consequence of this scaling (Held 1982, 1999) is the fact that the mean isentrope leaving the tropical boundary layer reaches heights characteristic of the tropopause in polar regions. In this paper we look at the degree of interior PV homogenization and boundary temperature mixing required by the equilibration of Charney waves of different scales. For this purpose, we discuss the vertical structure of the potential vorticity flux for the Charney problem. We show that the PV flux peaks at the steering level of the waves and can be very localized when their vertical scale H is smaller than the Held scale h (Held 1978; see section 2b for definition), or, equivalently, when the interior PV gradient seen by the waves is smaller than the boundary PV gradient. Note that because the thickness of the interior region with large PV flux scales as the imaginary phase speed (Lindzen et al. 1980) these waves typically grow slowly. We will show that these short waves can equilibrate by mixing potential vorticity in a neighborhood of the steering level alone, while still allowing a nonzero temperature gradient at the ground and a significant interior PV gradient. The structure of this paper is as follows. We review in section 2 the basic formulation of the linear instability problem, reinterpret Held’s (1978) scaling in terms of the interior PV gradients, and discuss the degree of PV homogenization found in the observed extratropical troposphere. In section 3 we demonstrate the existence of neutral waves when the interior PV gradient vanishes at the steering level alone [thus violating Bretherton’s (1966) condition]. This is done analytically for an idealized distribution of the interior PV gradient and numerically to test the extent of PV homogenization required for waves of different scales. In section 4, we

VOLUME 58

study the quasilinear equilibration of these Charney waves using a quasigeostrophic model. Finally, we close with a brief summary of our results and a discussion of their implications in section 5. 2. Formulation a. Linear theory The conservation of quasigeostrophic potential vorticity for a Boussinesq fluid, linearized around a zonal basic state with velocity U 5 Lz and stratification N 2 can be written as ]q ]q 1 U 1 y q y 5 0, ]t ]x

(1)

being q 5 ] 2 c/]x 2 1 ] 2 c/]y 2 1 e(] 2 c/]z 2 ) the perturbation potential vorticity, c the geostrophic streamfunction, y 5 ]c/]x the meridional velocity, e 5 f 20/N 2 and q y the interior gradient of potential vorticity in the basic state, defined as q y 5 b 2 f 02

1

2

] 1 ]U . ]z N 2 ]z

(2)

Assuming plane wave solutions c (x, y, z, t) 5 f (z)e i(kxx1kyy2kxct) ,

fzz 1

1

2

qy 1 2 k 2 f 5 0, e Lz 2 c

(3)

with k 2 5 k x2 1 k y2 . Subject to the appropriate boundary conditions, this yields an eigenvalue problem for the phase speed c. The associated eigenvector is f (z), the vertical structure of the modes. When the phase speed c has a positive (negative) imaginary part c i , which also requires a complex f (i.e., a phase tilt with height) the modes are growing (decaying). Bretherton (1966) showed that the domain integral of the perturbation PV flux vanishes:

E

`

y q dz 5 0,

(4)

0

provided that the heat flux at the boundaries is generalized as a delta-function PV flux. He further showed that this PV flux is always downgradient for a growing mode:

yq 5 2

k x ci 2 |h | q y , 2

(5)

where h is the meridional displacement of the basicstate PV contours:

h52

q f 5 . qy L(z 2 z c ) 2 ici

(6)

At z 5 z c the meridional displacement | h | → ` for neutral modes (c i → 0), and residue calculus shows that this implies a nonzero PV flux at the steering level if

15 NOVEMBER 2001

q y (z c ) ± 0. However, Eq. (5) shows that the PV flux must be zero away from the steering level for a neutral mode. Because the domain integrated PV flux must be zero, Bretherton (1966) concluded that neutral modes are not possible whenever there is a PV gradient at the steering level, as is the case in the Charney problem. b. The vertical scale of Charney waves The baroclinic instability problem can be described in terms of the interaction between PV perturbations at different vertical levels (Hoskins et al. 1985). Because the wave activity of a growing mode integrates to zero, instability further requires that the interacting perturbations act on regions with PV gradients of opposite sign. In most classical baroclinic instability problems the only region of negative wave activity is the ground. Hence, the baroclinic modes can be described in terms of the interaction between the lower-boundary edge wave1 and a distributed PV perturbation in the interior and/or at the tropopause. Consider a temperature perturbation with length scale L 5 2p/k at the lower boundary of the Charney problem. Away from the steering level, the PV flux decays with a vertical scale H 5 ½H R 5 ½L(e)1/2 , where the Rossby penetration depth H R gives the height of influence of the lower perturbation. The ½ factor arises from the fact that y q is a quadratic quantity in f, and it is assumed that the structure of the mode is equivalent barotropic (i.e., does not tilt) away from the range of influence of the lower-level perturbation. We introduce the Held scale h (Held 1978): 2 h5

3445

ZURITA AND LINDZEN

E

01

q y dz

0

b

5

Le , b

(7)

where b is the mean value of the interior PV gradient. As discussed by Held (1978), the Rossby depth of the most unstable normal mode in the Boussinesq–Charney problem scales as h. We shall define short Charney waves2 as waves with penetration depths H smaller than, or of the order of, h. Note that h is the height over which the vertically 1 For the purpose of this paper, the edge wave is the geopotential perturbation resulting from the inversion of a temperature perturbation at the boundary. Note that, strictly speaking, the edge wave only exists as an independent solution in a seminfinite fluid with no interior PV gradients (otherwise the edge wave velocity field would create PV anomalies at other levels, which would also contribute to the solution). 2 The notation may induce to confusion and the reader should be careful to understand the subtle difference between our approach and that of Held (1978). The point of that paper was to estimate the vertical scale of the most unstable mode depending on the ratio between the scale height H and h. In our case, however, H is the vertical scale of a mode, which may be set up externally (for instance through the meridional scale), and may not correspond to the most unstable one. In particular, modes with H K h will always be weakly unstable.

integrated PV gradient in the interior balances the negative contribution by the delta function at the ground. In other words, for waves with H , h the vertically integrated potential vorticity gradient in the interior is smaller than the delta function at the ground, while the reverse is true with H . h. Furthermore, for a fluid of depth H R the two-layer shortwave cutoff is H 5 h (Lindzen 1990). Now, if we use Eqs. (4), (5), and the definition of h:

hci |h (0)| 2 5 ci

E

H

|h (z)| 2 dz.

(8)

01

Equation (6) shows that the PV flux is maximized at the steering level z c , where it can be very large as c i → 0, but is bounded away from that level (recall that f is a well-behaved function). Hence, for waves with H K h Eq. (8) can only be satisfied when | h(z c ) | k | h(0) | and the PV flux peaks strongly at the steering level. This suggests that, if such a mode were to equilibrate quasi-linearly according to the vertical profile shown, it would tend to reduce the PV gradient of the basic state primarily in the vicinity of the steering level. This narrow mixing might by itself be able to stabilize the waves because, when H , h, Eq. (8) can never be satisfied once that the PV gradient is removed at the steering level and contributions to the integral in the right-hand side are limited to regions where h(z)/h(0) ; O(1). This is illustrated in Fig. 1. The left panel shows a short Charney wave with H/h 5 0.45 and the right panel the most unstable mode, which has H/h 5 3.9. The dimensional wavenumber and Rossby depth are the same in both cases; only the interior PV gradient, and hence the Held scale h, change. As can be seen, the interior PV flux peaks in both cases at the steering level and decays exponentially in comparable distances (heights are nondimensionalized with the half Rossby depth H). The main difference between them is that the PV fluxes extend over a broader region, and the steering level is lower relative to the Rossby depth, for the case of the most unstable mode. More generally, Fig. 2 shows the dispersion relation for the Charney problem, which in the Boussinesq limit only depends on the dimensionless parameter H/h. (This figure was constructed from a series of runs in which the dimensional wavelength was kept constant and only the interior PV gradient was changed). Note that the most unstable mode has a steering level at a height of just 0.1H. In the following we will examine if this scaling analysis holds, both for waves with H , h and for waves with H 5 O(h). Note that when H 5 O(h) the PV flux still peaks at the steering level but has a broader structure. Hence, the scaling analysis suggests that linear neutrality would require in this case a deeper region with zero PV gradient. Nevertheless, it is plausible that these waves might still equilibrate without eliminating the lower-boundary temperature gradient.

3446

JOURNAL OF THE ATMOSPHERIC SCIENCES

VOLUME 58

FIG. 2. Dispersion relation for the Charney–Boussinesq problem as a function of H/h. The phase speed is nondimensionalized with LH, where H is the half Rossby depth. This curve was constructed changing the interior PV gradient (which enters the parameter h) alone.

theorem, it is the mass-averaged PV gradient that is relevant. Following Stone and Nemet (1996), we will ignore the curvature of the jet and rewrite the interior PV gradient as

For typical midlatitude values, such as an interior PV gradient b 5 1.65 3 10 211 m 21 s 21 , a jet with linear shear and maximum wind speed U 5 50 m s 21 at a height of 10 km, and an inertial ratio e 5 10 24 , Eq. (7) gives a Held scale h 5 30 km. Note that the Eady shortwave cutoff has H/h 5 0.38 for this choice of parameters and therefore a wave short enough to equilibrate according to Lindzen’s (1993) mechanism would also be a short Charney wave.

[

d (he 2z/HS ) dz

5 2be z/HS

d [(H 1 h)e 2z/HS ], dz S

q y (z) 5 b 1 2 e z/HS

FIG. 1. Vertical structure for a short Charney wave (H/h 5 0.45) and the most unstable Charney mode (H/h 5 3.9): streamfunction perturbation (dashed), PV perturbation (thin solid), and PV flux (thick solid). The steering level is the dash–dotted line and only interior values are shown. Vertical scales are nondimensionalized with H.

(9)

where h(z) 5 L(z)e(z)/b, H S is the scale height, and we define g 5 1 2 e 2H/HS . The ratio between the interior and boundary PV gradients is then given by

2

E

H

e 2z/HS q y dz

01

E

01

5

2[(HS 1 h)e 2z/HS ] H01 h(0)

5

g [HS 1 h(H )] 2 Dh , h(0)

q y dz

0

c. How well is the interior PV homogenized? We showed in the previous section that the two-layer shortwave cutoff is H 5 h. Hence, as mentioned in the introduction, the fact that observed shears are close to the two-level critical shear (Stone 1978) suggests that H ; h for tropospheric modes. This in turn implies that the interior PV gradient is as poorly mixed as the boundary temperature gradient and an equilibration model of interior PV homogenization (Lindzen 1993) is not an appropriate paradigm. It is easy to get a more precise result for a nonBoussinesq atmosphere with height-dependent shear and stratification. In that case, from Bretherton’s (1966)

]

(10)

where Dh 5 h(H) 2 h(0). In terms of the isentropic slope in scaled coordinates (S 5 h/H S ):

2

E

H

e 2z/HS q y dz

01

E

5

01

q y dz

g [1 1 S(H )] 2 [S(H ) 2 S(0)] . S(0)

(11)

0

We can estimate these values from Fig. 3, taken from Stone and Nemet (1996). We will choose H to be the distance from the top of the boundary layer (defined for

15 NOVEMBER 2001

ZURITA AND LINDZEN

3447

FIG. 3. Comparison of the vertical structure of observed h at different latitudes with the adjusted profile resulting from full PV homogenization, both at the boundary and in the interior. Taken from Stone and Nemet (1996).

our purposes by the minimum h) to the height at which h is maximized. If q y were zero in the interior, as in the study of Sun and Lindzen (1994), h would describe a line roughly parallel to the adjustment curve, but with h(0) . 0. This is approximately observed up to the h maximum; thereafter q y increases significantly. We can interpret this as a boundary to the vertical penetration of the waves since above that height the EP fluxes tend to be dominated by their horizontal component, as discussed by Edmon et al. (1980). The value of H thus obtained should be regarded as an empirical measure of the penetration of the eddy fluxes, including mixing barrier effects at the jet, rather than as a proper Rossby depth (see section 4). This gives

enough to compensate the beta effect O(g ) and nonBoussinesq corrections O[g h(H)/H S ].

h(0) ø 7 km

a. Analytical model

H ø 6 km,

h(H ) ø 12 km

HS ø 7.5 km (12)

yielding a ratio of interior to boundary PV gradient of 0.82. This number should not be taken too seriously because our estimates of H, h are, to an important extent, ambiguous. Nevertheless, this suggests that the positive interior PV gradient above the boundary layer is comparable with the negative PV gradient at the ground. Note that this net interior PV gradient arises because the rate of change of the isentropic slopes O(Dh/H S ), which is the only negative contribution, is not large

3. Neutral states In this section we prove that short Charney waves become neutral when the PV gradient of the basic state vanishes in a neighborhood of the steering level, as suggested by the scaling argument of the previous section. This is shown analytically for a simple idealized distribution of the interior PV gradient and numerically for more general cases.

The Charney and Green models are known to be unstable for all wavenumbers and shears. This is consistent with the discussion of Bretherton (1966) because in both cases there is a nonzero PV gradient at the steering level. We will consider here a variation of these problems in which the interior PV gradient is not constant but has a vertical structure with a single zero at the steering level. We will show that in the presence of small interior PV gradients elsewhere this is sufficient to stabilize the waves.

3448

JOURNAL OF THE ATMOSPHERIC SCIENCES

R5

VOLUME 58

12

aLD 3 D 5O , e h

(14)

which gives a measure of how important is the integrated PV gradient over the depth D O(aL 2 D 3 ) compared to the delta function at the lower boundary (viz. 7). After nondimensionalizing and applying the change of variable j 5 R1/3 (z˜ t 2 z˜), where z˜ t 5 c˜ 1 k˜ 2 /R is the turning point of the vertical structure equation, this equation becomes

f jj 2 jf 5 0,

(15)

whose solution can be formally written in terms of the Airy functions of the first and second kind A i and B i (Abramowitz and Stegun 1965):

f 5 C1 A i (j) 1 C 2 B i (j).

FIG. 4. Eigenmodes of the problem defined by Eq. (15) for k˜ 5 p, R 5 3 (nondimensionalizing with D 5 H ).

Specifically, we assume that the basic-state interior PV gradient has the form3 q y (z) 5 a(Lz 2 c) 2 ,

(13)

where c is the phase speed of the waves (the eigenvalue of the problem). This idealized distribution can be regarded as the first term in a Taylor expansion satisfying the basic requirements that the PV gradient is positive definite and vanishes only at the steering level. Note that this definition is meaningful only when c is real. However, this is not a limitation because the purpose of this section is to prove the existence of neutral solutions. We will nondimensionalize Eq. (3) as follows: z 5 Dz˜, k 5 (e)1/2 k˜ /D, and c 5 (LD)c˜, with the tilde variables being dimensionless. D is a vertical fluid depth, as of yet unspecified. We will further define the dimensionless parameter: 3 Note that this ad hoc PV gradient is inconsistent with the basic state L, N 2 , assumed constant. However, Lindzen (1993) showed that tiny variations in U and N 2 can produce changes in the PV gradient of order b. In that sense our analysis should be regarded as the leading-order term of a perturbation expansion about some slowly varying functions U(z), N 2 (z) in balance with q y (z).

(16)

For real j both A i and B i are real, which implies that the modes are vertically trapped. Figure 4 shows these two modes for R 5 3, choosing D 5 H 5 p (e)1/2 /k as the height scale (this is equivalent to fixing k˜ 5 p). As can be seen, one of the modes is trapped at the lower boundary, while the other peaks at a height of order j 5 21, z˜ 5 z˜ t 1 R 21/3 , and decays very slowly aloft. Note that both eigenfunctions go to zero as z˜ → `, j → 2`. Hence, they are independently admissible solutions to the modified Charney problem and each of them (as well as any of their combinations) gives a branch of the dispersion relation c˜(k˜ ). In particular, for any choice of C1 , C 2 a solution c˜(k˜ ; C1 , C 2 ) can be found imposing a rigid lid boundary condition at z˜ 5 0. Equivalently, we can always find a pair C1 , C 2 matching any value of c˜. It is possible to show that the eigenvalue c˜ is real and the solution neutral as long as the vertical structure consists of just one eigenfunction A i or B i . However, this is not necessarily the case when a combination of both modes is used, as their interaction can lead to mutual reinforcement much in the manner of the Eady problem. In the case of the modified Green model the situation is a bit more complicated because of the presence of an upper rigid lid at some vertical height D, which supports a second edge wave at that level. As a result, deep modes are expected to be unstable regardless of the interior PV gradient because the two edge waves can still interact in the manner of the Eady problem. For this reason, we will concentrate on modes shorter than the Eady shortwave cutoff, for which the lower boundary edge wave can only interact with the interior PV perturbation. The eigenvalue problem is formulated imposing rigid lid boundary conditions at z 5 0, D:

15 NOVEMBER 2001

ZURITA AND LINDZEN

3449

FIG. 5. Real phase speed for the neutral solutions of the analytical model discussed in the text: R 5 0.5 (solid), R 5 1 (solid), R 5 5 (dashed), and R 5 10 (dot–dashed). The lines are interrupted where the phase speed becomes imaginary or convergence was not reached in 100 iterations.

fj 2 R 21/3 fj 1 R 21/3

f 5 0 at j 5 j 0 5 R 1/3 z˜ t c˜

f 50 1 2 c˜

at j 5 j D 5 R 1/3 (z˜ t 2 1).

(17)

After some algebra, this leads to the dispersion relation4 Ac˜ 2 1 Bc˜ 1 C 5 0,

(18)

where A, B, and C are given by A 5 B9( i j 0 )A9( i j D ) 2 A9( i j 0 )B9( i jD) 21/3 C 5 2R 21/3 A i (j 0 )[B9( Bi (j D )] i jD) 1 R 21/3 1 R 21/3 Bi (j 0 )[A9( A i (j D )] i jD) 1 R

B 5 2A 2 C 1 R 21/3 [A9( i j 0 )Bi (j D ) 2 B9( i j 0 )A i (j D )] 2 R 22 /3 [A i (j 0 )Bi (j D ) 2 Bi (j 0 )A i (j D )].

(19)

Note that although Eq. (18) looks like a quadratic equation there is some additional dependence on the eigenvalue through the coefficients. As a result, the dispersion relation must be solved iteratively starting with some initial guess c˜ 0 . For small R this dependence is weak and the iterative procedure converges rapidly. Figure 5 shows the real phase speed of the neutral solutions to (18) for different values of R. The curves 4 Strictly speaking, the solution should not be called a dispersion relation because the basic state is different for each k˜ .

are interrupted where no convergence was found or when the converged value was imaginary. Note that this typically occurred for those wavenumbers for which the Eady problem itself was unstable, though the shortwave cutoff did show some sensitivity to the magnitude of the interior PV gradient. Note that for small R the solutions agree very well with the Eady dispersion relation, yielding a shortwave cutoff k˜ ø 2.4. Finally, as in the Eady problem, the dispersion relation consists of two branches and there is an upper-level edge wave, as suggested by Rivest et al. (1992). All this suggests that the zero PV gradient at the steering level removes the singularity of the Green problem and yields a solution closer to Eady’s when the interior PV gradient is sufficiently small. In conclusion, at least for this idealized distribution of q y , neutral solutions can be found in the presence of positive interior PV gradients as long as Bretherton’s (1966) condition is violated and the PV gradient at the steering level is zero. b. Numerical results For moderate values of the parameter R, a vertical structure as in (13) gives very small values of the PV gradient across a broad neighborhood of the steering level, even though q y only vanishes exactly at that point. Hence, it would be desirable to examine if neutral solutions can still exist when the region with small PV

3450

JOURNAL OF THE ATMOSPHERIC SCIENCES

VOLUME 58

FIG. 6. Vertical structure of the unstable Green wave at the Eady shortwave cutoff when b 5 1.6 3 10 211 m 21 s 21 . This mode has H/h 5 0.38 and is also short in a Charney sense. The mode becomes neutral when the PV gradient is brought to zero across the shaded region alone.

gradients is narrower. Furthermore, the existence of neutral solutions, as proved above, does not rule out additional unstable solutions since Eq. (18) is transcendental. We have looked at these issues using a linear, timedependent, 1D model, in which the most unstable solutions, when they exist, are found integrating forward in time. For simplicity, a rigid lid is used as the model top boundary condition, though we restrict the analysis to waves shorter than the Eady shortwave cutoff. The methodology is the following. We first run the model with constant interior q y and let the most unstable mode emerge. Then, we solve a new linear problem with modified interior PV gradient; the PV gradient is typically brought to zero across a finite neighborhood of the steering level and left unmodified elsewhere. The depth of the region with zero PV gradients is changed until neutrality is achieved. Figure 6 shows a mode right at the Eady shortwave cutoff, which is made unstable by the addition of interior b. Nevertheless, note that because of the smallness of the interior b the inversion is still dominated by the boundary perturbations as in the Eady problem (including the upper boundary edge wave). This mode has H/h 5 0.38 and is therefore also short in a Charney sense. As shown in Fig. 2 this implies a small growth rate and consequently a PV flux peaking strongly at the steering level (Lindzen et al. 1980). When the PV gradient is brought to zero in the shaded region the wave becomes neutral. The structure of the mode and steering level change very little in the modified state. Hence, it seems a reasonable assumption that such a mode could stabilize quasi-linearly by mixing the basic-state PV at the steering level, rather than by eliminating the shear at the ground or homogenizing the full interior PV.

FIG. 7. Vertical structure and steering level for Charney modes with H/h 5 1.6. (left) Mode with uniform interior PV gradient. (right) Mode when the PV gradient is brought to zero across a broad region surrounding the original steering level. Note that the PV flux vanishes in the region of zero PV gradient, which includes the modified steering level, and yet the mode is still unstable. However, when the PV gradient is brought to zero across the region shaded in the left panel the mode becomes neutral.

However, this is no longer the case when the value of H/h is not so small. This is illustrated in Fig. 7 for a mode with H/h 5 1.6. The left panel, corresponding to the run with uniform q y , shows that the PV flux has much broader structure than in the previous case and extends to the ground. When the PV gradient is smoothed out across a region of order 0.25H surrounding the steering level the mode is still unstable, and has the structure shown in the right panel. This is partly due to the fact that the steering level rises significantly when the positive interior PV gradient is reduced, which qualitatively has the effect of weakening westward propagation relative to the case with full PV gradient. Indeed, note that when the basic-state PV gradient is brought to zero across the narrower region shown shaded in the left panel, centered on the new steering level rather than the original one, the mode becomes neutral. The differentiation between these modes raises one of the main points of this paper, namely that in a continuous atmosphere with fine vertical structure different waves may play different roles in the equilibration. For

15 NOVEMBER 2001

3451

ZURITA AND LINDZEN

However, the situation is more complicated because in reality the lower boundary temperature gradient, which was kept fixed in these linear simulations, should also be mixed, which would weaken eastward propagation relative to the unhomogenized case. In fact, when H , h the modes are essentially edge waves in the sense that the PV inversion is dominated by the boundary perturbation. Hence, the dispersive properties of the modes should be more sensitive to the boundary temperature gradient than to the interior PV gradient. This can be seen in the shortwave asymptotic expansion of Branscome (1983) for the Charney model, which can be rewritten in our notation: cr 5

1

2

LH 1.34 H 12 . p 2p h

(20)

The first term, which shows a 1/k dependence, is simply the phase speed of the edge wave associated to the shear L. The second term, proportional to b/k 2 , always gives westward propagation and can thus be interpreted as a correction to the phase speed when there is a nonzero interior PV gradient. It arises from this expression that for H/h 5 O(1) the first term is 5 times larger, which implies that for short Charney waves the phase speed is primarily a function of the surface temperature gradient. FIG. 8. (top) Growth rate and (bottom) phase speed, nondimensionalized with LH, as a function of the thickness of the region of homogenized PV gradient surrounding the steering level. The different curves correspond to different values of H/h in the unhomogenized profile.

instance, short waves might equilibrate by smoothing out the PV gradient at the steering level while longer waves could also concentrate the barotropic shear of the jet. This scenario of a quasilinear equilibration in which the different modes modify in different ways the mean state raises the interesting question of how sensitive each mode is to the modifications of the basic state induced by the others. Figure 8 shows the dependence of the dispersive properties of waves of different scales on the degree of homogenization. The upper panel shows the growth rate as a function of the thickness of the window around the (modified) steering level where the PV gradient has been brought to zero. The different curves correspond to different values of H/h in the initial profile with uniform q y . As can be seen, even for the Charney problem most unstable mode it suffices to eliminate the PV gradients over a region of thickness 0.25H. The lower panel shows the phase speed, which in all cases increases as the positive interior PV gradient is reduced. The fact that the steering level is so sensitive to the reduction of the basic-state interior PV gradient suggests that the equilibration process would require a fair amount of interior mixing as the steering level rises.

4. Quasi-linear equilibration The linear stability analysis of the previous section suggests that short waves may be able to equilibrate quasi-linearly eliminating the PV gradient at the steering level, whereas for longer waves a broader homogenized region would be necessary. We have tested these ideas in a nonlinear 3D model with 50 vertical levels. As a difference with the study of Gutowski et al. (1989), we use a quasigeostrophic model aiming to concentrate on the PV structure of the equilibrated state. In order to study the quasilinear equilibration of a wave of chosen scale, the channel length is truncated to comprise a single wavelength (4000 km). The meridional structure is also confined by a radiative equilibrium Gaussian jet of the same scale. This yields k x 5 2p/4000, k y ø p/4000, k ø (Ï5/4000)p, and a half Rossby depth H ø 17.9 km. This scale is kept fixed in all the runs to be shown but the ratio H/h is still changed due to the varying interior PV gradient. A rigid lid is used as the top boundary condition, but at enough distance (30 km) to prevent the interaction between both edge waves. The radiative equilibrium jet is chosen to have constant vertical shear, yielding a velocity of 33 m s 21 at a height of 10 km (or 100 m s 21 at the top). The fluid is Boussinesq and the inertial ratio is constant e 5 10 24 . The horizontal resolution is 160 km 3 160 km and the 50 equally spaced levels yield a vertical resolution of 600 m. Many of the grid points are outside the domain of interest as the dimensions are

3452

JOURNAL OF THE ATMOSPHERIC SCIENCES

made large to prevent interactions with the walls, both meridionally and at the top. This ensures that the boundaries to the mixing domain are internally set up by the eddies. The flow is relaxed to its radiative equilibrium configuration with a timescale of 15 days and hyperdiffusion is used to get rid of the small-scale eddies. In some cases Rayleigh damping is also included at the lowest resolved level. Note that because the channel length is truncated so that wavenumber 1 is the most unstable one, the dynamics of the equilibration is quasilinear, in the sense that it is this wave that primarily modifies the basic state. This allows us to study how a given mode equilibrates in isolation. A similar idea was used by Cehelsky and Tung (1991) in the two-layer model. Figure 9 shows the equilibrated PV gradients in a run with b 5 1.0 3 10 211 m 21 s 21 , which gives a ratio of H/h 5 0.54 (not including the curvature contributions to the PV gradient). The ‘‘equilibrated’’ states in this and subsequent figures were calculated by time averaging between days 500 and 800 of the simulation. By that time, the wave has completed most of the modifications of the basic state. Despite the small value of H/ h, the interior PV homogenization seems to proceed efficiently everywhere below the steering level instead of just at that height, though not at the ground, where the temperature gradient fails to be homogenized. The extent of the homogenized region is somewhat surprising in view of the linear analysis of the previous section, which suggested that localized PV mixing could be enough for equilibration. Similar results were found for shorter waves (i.e., with smaller b), although in those cases the neglected curvature PV gradient is actually larger than b. In fact, the snapshot in Fig. 10, taken as the wave starts to modify the basic state at time t 5 135, shows that the interior PV gradient is initially homogenized only at the steering level. Figure 11 shows that, as speculated in the previous section, the steering level drops rather than rises during the equilibration process as the temperature gradient at the ground is mixed. Although the phase speed increases slightly (see lower panel), this increase is much smaller than the increase in the basicstate zonal wind, which results in the observed drop of the steering level. It is remarkable that the weakening in the eastward propagation as the boundary temperature is mixed is compensated by the westerly advection by the enhanced winds, yielding relatively small and positive changes in the phase speed. At all times there is a good correlation between the position of the steering level and the homogenized region. Figure 12 shows the equilibrated state for the same parameters when Rayleigh damping with a timescale of 5 days is included. As can be seen, in this case the homogenization of the basic state is more localized, although PV is still fairly well mixed below the steering level. The equilibrium state is very similar to the snapshot shown in Fig. 10 and the surface wind does not

VOLUME 58

rise beyond 5 m s 21 . It is interesting to note that the phase speed remains again nearly constant during the equilibration, as both the temperature mixing and buildup of the zonal wind are weaker than in the previous case (see Fig. 13). The main difference with the inviscid case is that, because now the basic-state winds also change little, the steering level drops much less, consistent with the observed localization of the homogenized region. The relation between PV homogenization and steering level was found to be very robust for all cases examined. For instance, Figs. 14 and 15 correspond to the inviscid case with b 5 1.3 3 10 211 m 21 s 21 (H/h 5 0.7). As in the previous case, the increase in the phase speed is much smaller than the Doppler shift implied by the enhancement of the zonal wind, due to the fact that eastward propagation weakens with the surface temperature mixing. Consequently, the steering level drops again. However, because now the steering level drops to the ground, the PV gradient is nearly homogenized not only in the interior but also at the surface. When H/h is further increased to 0.86, a similar process is initially observed. However, in this case the interior PV gradient is restored at later times, once the surface temperature gradient is wiped out. There is no interior steering level in this case, and hence no need for interior homogenization, because the single eastward propagating term disappears with the surface temperature gradient. Nevertheless, the presence of weak surface friction suffices to contain the steering level in both cases, thus producing partial PV homogenization and a nonzero surface temperature gradient at equilibration (not shown). This is the case even for longer waves, as demonstrated in Fig. 16 for b 5 3.0 3 10 211 m 21 s 21 (H/h 5 1.61) and damping timescale t 5 1 day. We have also examined the equilibration of the most unstable mode of the Charney problem (not shown). Even then, the equilibrated state was found to be consistent with partial PV homogenization in the presence of strong enough friction. However, an additional complication in this case was the fact that the second harmonic of this mode was also significantly unstable, and also equilibrated at finite amplitude. We found that the primary harmonic, which reached finite amplitude first, was responsible for the initial modification of the mean state, particularly for the barotropic acceleration of the zonal flow and interior PV homogenization characteristic of previous cases. Also as before, the phase speed of the mode again remained relatively stable at this stage, and its steering level dropped and disappeared because of the enhanced surface wind. However, an interesting new feature was the strong increase observed in the phase speed of the second harmonic following the enhancement of the surface flow, instead of remaining constant like the phase speed of the first harmonic. As a result, only this second harmonic kept an interior steering level in the equilibrated state. We shall postpone a more detailed description of this example

15 NOVEMBER 2001

ZURITA AND LINDZEN

FIG. 9. Time-mean equilibrium state for the run with b 5 1.0 3 10 211 m 21 s 21 (H/h 5 0.54). (top left) Absolute value of the zonal mean potential vorticity gradient, normalized by b (contour unit 0.15b). (top right) Profile at the center of the channel at equilibration (solid) and radiative equilibrium (dashed). (bottom left) Absolute value of U 2 c (contour unit 2 m s21). (bottom right) Zonal velocity U at the center of the channel at equilibration (solid) and radiative equilibrium (dashed). The horizontal line highlights the steering level. Regions with values of | q y | and | U 2 c | smaller than the corresponding contour unit are nonshaded in the contour plots. Vertical distances are nondimensionalized with the half Rossby depth H and the lowest-level PV gradient includes the delta function at the ground smoothed at model resolution.

3453

3454

JOURNAL OF THE ATMOSPHERIC SCIENCES

VOLUME 58

FIG. 10. As in Fig. 9 but for a snapshot at time t 5 135, when the wave is just starting to modify the basic state.

and a more general analysis of cases with many waves until a subsequent study. In all cases examined, the homogenization of potential vorticity in the interior is accompanied by the buildup of large PV gradients at the edges of the mixing domain. In the vertical, there is a concentration of the

PV gradient at heights of order 0.5H 5 9 km, roughly consistent with observed tropopause heights. This is not altogether surprising because the static stability is also close to the observed one; our results just prove the relevance of the ‘‘dynamical constraint’’ for the tropopause height proposed by Held (1982). It is also in-

15 NOVEMBER 2001

ZURITA AND LINDZEN

3455

FIG. 11. Time series at the center of the channel for the run with b 5 1.0 3 10 211 m 21 s 21 (H/h 5 0.54). (top) | q y | (contour unit 0.15b). (middle) | U 2 c | (contour unit 2 m s21). (bottom) Surface wind (lower curve) and phase speed of the wave (upper curve). The thickness of the shaded region is related to the height of the steering level.

teresting to note that these large PV gradients are a result of hardly observable changes in the vertical curvature of the wind (indeed, the static stability is fixed in our model). This highlights the relevance of the underlying PV dynamics.

In the horizontal, there is also a concentration of the PV gradients at the margins of the jet. Such concentration seems unrealistic for the real atmosphere but our model does not incorporate any tropical dynamics. Moreover, in the atmosphere the adiabatic mixing of

3456

JOURNAL OF THE ATMOSPHERIC SCIENCES

potential vorticity is not horizontal but along isentropic surfaces, and the mean isentropic slope is such that all the extratropical isentropes intersect either the ground or the tropopause. Thus, in isentropic coordinates the moving earth surface represents an additional barrier to PV mixing. This, together with the steepening of the isentropes by vertical mixing at the atmospheric boundary layer, might explain the large PV gradients observed in the lower troposphere, which our model fails to reproduce. 5. Discussion We have discussed in this paper the observed vertical distribution of the extratropical PV gradients with the help of a very idealized model. It is often argued in the literature (e.g., Swanson and Pierrehumbert 1997; Stone and Nemet 1996) that baroclinic adjustment is more efficient in homogenizing the interior PV gradients than in eliminating the lower-level shear, a fact that is commonly attributed to dissipative effects over the boundary layer. However, we have shown that the interior PV gradients are in some sense comparable to the boundary PV gradients. The failure of the eddies to homogenize the interior PV gradient comes as no surprise when one realizes that the positive (negative) PV gradients are related via the appropriate diffusivities to the equatorward (poleward) eddy PV fluxes, and hence to the poleward (equatorward) mass fluxes (Held and Schneider 1999), which should balance in steady state. As Held and Schneider (1999) have shown using isentropic coordinates, the areas of poleward and equatorward mass flux are comparable and the apparent largeness of the negative PV gradients/equatorward eddy fluxes should be regarded as an artifact of quasigeostrophy, which concentrates the return flux on the lower boundary. Thus, the fact that the interior PV homogenization is incomplete should not be regarded per se as an additional argument against baroclinic adjustment but as a manifestation of this balance. In a steady state, the ratio between the PV gradients should just equal the inverse ratio between the respective average diffusivities (Held 1999), which in the observed atmosphere appears to be of order 1. Hence, to the extent that dissipative effects prevent the eddies from eliminating the surface temperature gradients they prevent as well the homogenization of the interior PV gradient. Similarly, assuming the observed PV diffusivities to be robust, if some other mechanism prevented interior PV homogenization we would expect this mechanism to also set up a limit to the reduction of the surface temperature gradient. The integrated interior PV gradient in the troposphere consists of 3 contributions: the beta effect, O(g ), the non-Boussinesq correction, O[g S(H)], and the rate of change of the isentropic slope, O(DS) (the only negative term). All these terms are comparable and so is the integrated negative PV gradient, O(S 0 ). As a result, the

VOLUME 58

net interior PV gradient is similar to what one would get from just the beta effect and H ; h. This result is in fact apparent in the study of Sun and Lindzen (1994) (cf. their Figs. 3 and 14), although it was not emphasized by the authors. In that paper the mean isentropic slope is better captured when using a potential vorticity gradient of b along the isentropes than when full homogenization is assumed. Given that there is a limit to how well PV can be mixed, a natural question that then arises is what determines the vertical structure of the interior PV gradients. We hypothesized in this paper that the observed homogenization of PV around 700 hPa could be a result of the fact that the mixing of potential vorticity is most efficient close to the critical level. Furthermore, although baroclinic neutrality is not a necessary condition for adjustment in the presence of friction, we have shown that neutrality can actually be achieved without full homogenization in the inviscid problem for waves with H K h. The reason is that the linear diffusivity (basically the Lagrangian displacement h) away from the steering level is bounded. Hence, when the interior PV gradient is much smaller than the boundary contribution unstable normal modes can only be sustained in the presence of a nonzero PV gradient at that level, which drives a very large flux there. The homogenization of a narrow region around the steering level is thus sufficient for neutrality in those cases. We have argued that for waves with H ; h there should still exist a well-mixed region surrounding the steering level, whose dimension would depend on the ratio between the interior and boundary PV gradients. While linear runs suggest that basic states with partial homogenization could be neutral, mechanistically, an important objection against equilibration by localized mixing is the sensitivity of the steering level to the degree of homogenization. For waves of the scales considered the phase speed is more sensitive to the surface temperature gradient than to the interior PV gradient. Hence, in the absence of other effects the phase speed would decrease as the gradients are mixed. In fact, it is observed in the full 3D problem that the westerly acceleration of the zonal wind by the wave induced fluxes tends to compensate the weakening of eastward propagation, in such a way that the phase speed remains nearly constant or increases slightly, depending on the magnitude of the interior PV gradients. In any event, because the increase in the phase speed is still much smaller than the acceleration of the zonal wind, the steering level always drops. This tends to produce a well-homogenized region below the steering level, which only extends to the ground when H/h is large enough. An important factor controlling the drop of the steering level and the thickness of the homogenized region is the strength of surface friction, which basically constrains the acceleration of the zonal wind. Localized PV mixing is also observed in our model for waves with H/h 5 O(1) in the presence of strong enough friction.

15 NOVEMBER 2001

ZURITA AND LINDZEN

3457

FIG. 12. As in Fig. 9 but for the run with b 5 1.0 3 10 211 m 21 s 21 (H/h 5 0.54) and damping timescale 5 days.

The analysis herein has been quasilinear because we have only studied how modes of different, chosen scales would modify the basic state in isolation. A similar idea was used in the two-layer study of Cehelsky and Tung (1991). These authors also found that the longer waves became the main heat carriers in the full nonlinear case as the more unstable modes saturated (see also Welch

and Tung 1998). We have not performed a full nonlinear analysis, though it is possible that something similar might occur in our model in the presence of many modes, provided that they are all able to equilibrate by localized interior PV mixing. In particular, the quasilinear results presented here suggest that the shorter waves need to transport less heat to equilibrate, due to

3458

JOURNAL OF THE ATMOSPHERIC SCIENCES

VOLUME 58

FIG. 13. As in Fig. 11 but for the run with b 5 1.0 3 10 211 m 21 s 21 (H/h 5 0.54) and damping timescale 5 days.

their shallower PV flux. As a result, the longer waves might be able to grow to larger amplitudes in the nonlinear regime. However, other considerations might also be important, as illustrated by the example with two unstable harmonics discussed in the previous section. The main difficulty is, of course, the fact that the

vertical penetration of the waves is a product of the equilibration itself. Not only did we consider waves of vertical scales shorter than the most unstable mode, but we also constrained strongly those scales through the forced radiative equilibrium jet and fixed static stability. In reality, the vertical scale H should also be sensitive

15 NOVEMBER 2001

ZURITA AND LINDZEN

3459

FIG. 14. As in Fig. 9 but for the run with b 5 1.3 3 10 211 m 21 s 21 (H/h 5 0.7) and no damping.

to the wave–mean flow interactions of the full spectrum of modes. It is plausible that the long waves might be able to equilibrate by concentrating the jet and/or raising the tropopause (i.e., by reducing H), at the same time than they partially homogenize the interior PV; indeed, simple offhand calculations suggest that the Eady short-

wave cutoff might also be short in a Charney sense. This argument reconciles the geometric equilibration scenario of Lindzen (1993) with the observed partial homogenization of the PV gradients. While the equilibration process tends to produce features characteristic of the extratropical troposphere, such

3460

JOURNAL OF THE ATMOSPHERIC SCIENCES

VOLUME 58

FIG. 15. As in Fig. 11 but for the run with b 5 1.3 3 10 211 m 21 s 21 (H/h 5 0.7) and no damping.

as the concentration of the potential vorticity gradients at the edges of the mixing domain, it fails to reproduce the observed large PV gradients at the top of the atmospheric boundary layer. In view of the ongoing discussion, this may be very important for the dynamics. We suggest that this is due to the many deficiencies of

the model, particularly the quasigeostrophic framework and the absence of boundary layer dynamics. Models with similar degree of idealization but incorporating these processes (Solomon and Stone 2001) seem to be more successful in reproducing the observed extratropical potential vorticity structure.

15 NOVEMBER 2001

ZURITA AND LINDZEN

3461

FIG. 16. As in Figs. 9, 11 but for the run with b 5 3.0 3 10 211 m 21 s 21 (H/h 5 1.61) and damping timescale 1 day.

Another important inadequacy is the small wave amplitude at upper levels that the model produces compared to observations. This is probably due to the lack of a model tropopause, which would support large wave amplitudes at those levels. Although the equilibration of the Charney waves in our model produces concentration of the PV gradients at upper levels, it is likely

that in the atmosphere the tropopause is also forced by other means. In that case, the question remains of how Eady waves would equilibrate. As our work is only concerned with the equilibration of Charney waves we do not claim to have given a full answer to the equilibration problem. A question arising from this study is what determines

3462

JOURNAL OF THE ATMOSPHERIC SCIENCES

the value of H/h or, equivalently, the isentropic structure of the extratropical troposphere. It would be interesting to see how this ratio changes as the waves equilibrate in a more realistic setting and how this constrains the isentropic structure. In particular, the ratio H/h 5 1 might represent some equilibration limit. However, a more complex model is required to serve this purpose, as the vertical fluxes and diabatic effects are likely to play a role in the equilibration process. Acknowledgments. This work has benefited from useful discussion with Nili Harnik and from the suggestions of two anonymous reviewers. Part of this work was done while the first author was a doctoral student of Edmund Chang at MIT. P.Z. acknowledges the financial support by a doctoral fellowship from Program PG-94 by the Secretaria de Estado, Universidades e Investigacion of Spain, and partial support by NSF Grant ATM-9731393. R.S.L. was supported by NSF Grant ATM-9421195 and DOE Grant FG02-93ER61673. REFERENCES Abramowitz, M., and I. A. Stegun, 1965: Handbook of Mathematical Functions. Dover, 1046 pp. Barry, L., G. C. Craig, and J. Thuburn, 2000: A GCM investigation into the nature of baroclinic adjustment. J. Atmos. Sci., 57, 1141– 1155. Branscome, L. E., 1983: The Charney baroclinic stability problem: Approximate solutions and modal structures. J. Atmos. Sci., 40, 1393–1409. Bretherton, F. P., 1966: Critical layer instability in baroclinic flows. Quart. J. Roy. Meteor. Soc., 92, 325–334. Cehelsky, P., and K. Tung, 1991: Nonlinear baroclinic adjustment. J. Atmos. Sci., 48, 1930–1947. Charney, J. G., and M. E. Stern, 1962: On the stability of internal baroclinic jets in a rotating atmosphere. J. Atmos. Sci., 19, 159– 172. Edmon, H. J., Jr., B. J. Hoskins, and M. E. McIntyre, 1980: Eliassen-

VOLUME 58

Palm cross sections for the troposphere. J. Atmos. Sci., 37, 2600– 2616. Gutowski, W. J., Jr., L. E. Branscome, and D. A. Stewart, 1989: Mean flow adjustment during life cycles of baroclinic waves. J. Atmos. Sci., 46, 1724–1737. Held, I. M., 1978: The vertical scale of an unstable baroclinic wave and its importance for eddy heat flux parameterizations. J. Atmos. Sci., 35, 572–576. ——, 1982: On the height of the tropopause and the static stability of the troposphere. J. Atmos. Sci., 39, 412–417. ——, 1999: The macroturbulence of the troposphere. Tellus, 51AB, 59–70. ——, and T. Schneider, 1999: The surface branch of the zonally averaged mass transport circulation in the troposphere. J. Atmos. Sci., 56, 1688–1697. Hoskins, B. J., M. E. McIntyre, and A. W. Robertson, 1985: On the use and significance of isentropic potential vorticity maps. Quart. J. Roy. Meteor. Soc., 111, 877–946. Kirk-Davidoff, D. B., and R. S. Lindzen, 2000: An energy balance model based on potential vorticity homogenization. J. Climate, 13, 431–448. Lindzen, R. S., 1990: Dynamics in Atmospheric Physics. Cambridge University Press, 310 pp. ——, 1993: Baroclinic neutrality and the tropopause. J. Atmos. Sci., 50, 1148–1151. ——, B. F. Farrell, and K. K. Tung, 1980: The concept of wave overreflection and its application to baroclinic instability. J. Atmos. Sci., 37, 44–63. Rivest, C., C. A. Davis, and B. F. Farrell, 1992: Upper-tropospheric synoptic-scale waves. Part I: Maintenance as Eady normal modes. J. Atmos. Sci., 49, 2108–2119. Solomon, A. B., and P. H. Stone, 2001: Equilibration in an eddy resolving model with simplified physics. J. Atmos. Sci., 58, 561– 574. Stone, P. H., 1978: Baroclinic adjustment. J. Atmos. Sci., 35, 561– 571. ——, and B. Nemet, 1996: Baroclinic adjustment: A comparison between theory, observations, and models. J. Atmos. Sci., 53, 1663–1674. Sun, D. Z., and R. S. Lindzen, 1994: A PV view of the zonal mean distribution of temperature and wind in the extratropical troposphere. J. Atmos. Sci., 51, 757–772. Swanson, K., and R. T. Pierrehumbert, 1997: Lower-tropospheric heat transport in the Pacific storm track. J. Atmos. Sci., 54, 1533– 1543. Welch, W., and K. Tung, 1998: Nonlinear baroclinic adjustment and wave selection in a simple case. J. Atmos. Sci., 55, 1285–1302.

The Equilibration of Short Charney Waves: Implications ...

Nov 15, 2001 - (Manuscript received 19 December 2000, in final form 30 April 2001) ..... zen 1990). Now, if we use Eqs. (4), (5), and the definition of h: H. 2. 2 hc |(0 .... Note that this definition is meaningful only .... tral solutions, as proved above, does not rule out ad- ..... stage, and its steering level dropped and disappeared.

2MB Sizes 0 Downloads 191 Views

Recommend Documents

Potential Momentum and the equilibration of short ...
where αM is a frictional time scale. .... and the net eddy dissipation Deddy includes frictional, thermal and small scale ...... In the new framework, the thermal ho-.

On the equilibration of a symmetrically unstable front ...
Cambridge, MA, 02139, USA. Email: [email protected] ... vertical fluxes induced by the. K-H instability were unable to change the bulk properties of the system.

On the equilibration of a symmetrically unstable front ...
density gradient in the x-direction is in 'thermal wind' balance with a velocity ..... The transfer of energy from the unstable symmetric modes to smaller scales via.

Baroclinic Equilibration and the Maintenance of the ...
Jul 1, 2004 - Sketch illustrating the meaning of the different vertical scales for a short ..... by symmetry, this occurs at the expense of negative PV gradients in.

Universality in the equilibration of quantum systems ...
Mar 11, 2010 - 2Department of Physics and Astronomy and Center for Quantum Information Science & Technology, University of Southern California,.

Universality in the equilibration of quantum ... - APS Link Manager
Mar 11, 2010 - 2Department of Physics and Astronomy and Center for Quantum Information Science & Technology, University of Southern California,.

Waves
How do they compare? Virtual Int 2 Physics .... reflection is used in fibre optics which are used in: medicine ; cable television ; internet ; telephone access.

The Magnitude of the Task Ahead: Macro Implications ...
May 18, 2016 - and inference in macro panel data which are at the heart of this paper. ..... For our empirical analysis we employ aggregate sectoral data for ...

Universality in the equilibration of quantum systems after a small quench
Mar 11, 2010 - Imagine preparing a closed quantum system in a given initial state ρ0 and letting it ... 〈n|V |0〉. En − E0. |n〉 + O(δλ2). (note the plus sign in V ). .... Data were obtained by Lanczos diagonalization of Eq. (8), keeping as

of Wind-Generated Gravity Waves
function ](w,,,'y) has some definition domain deter- mined by the condition y> -—§ and using the approxi- mate relation (2.4). As a result, the non-dimensional.

of Wind-Generated Gravity Waves
to one another by a corresponding system of dynamic equations for ..... Disaster Pre. Res. Inst. Kyoto U ... Systems of wind-generated wave fields. Dc-kl. Akad.