Gravity Waves Generated by Sheared Potential Vorticity Anomalies

François Lott Laboratoire de Météorologie Dynamique du CNRS, Ecole Normale Supérieure, Paris, France

Search for other papers by François Lott in
Current site
Google Scholar
PubMed
Close
,
Riwal Plougonven Laboratoire de Météorologie Dynamique du CNRS, Ecole Normale Supérieure, Paris, France

Search for other papers by Riwal Plougonven in
Current site
Google Scholar
PubMed
Close
, and
Jacques Vanneste School of Mathematics, and Maxwell Institute for Mathematical Sciences, University of Edinburgh, Edinburgh, United Kingdom

Search for other papers by Jacques Vanneste in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

The gravity waves (GWs) generated by potential vorticity (PV) anomalies in a rotating stratified shear flow are examined under the assumptions of constant vertical shear, two-dimensionality, and unbounded domain. Near a PV anomaly, the associated perturbation is well modeled by quasigeostrophic theory. This is not the case at large vertical distances, however, and in particular beyond the two inertial layers that appear above and below the anomaly; there, the perturbation consists of vertically propagating gravity waves. This structure is described analytically, using an expansion in the continuous spectrum of the singular modes that results from the presence of critical levels.

Several explicit results are obtained. These include the form of the Eliassen–Palm (EP) flux as a function of the Richardson number N22, where N is the Brunt–Väisälä frequency and Λ the vertical shear. Its nondimensional value is shown to be approximately exp(−πN/Λ)/8 in the far-field GW region, approximately twice that between the two inertial layers. These results, which imply substantial wave–flow interactions in the inertial layers, are valid for Richardson numbers larger than 1 and for a large range of PV distributions. In dimensional form they provide simple relationships between the EP fluxes and the large-scale flow characteristics.

As an illustration, the authors consider a PV disturbance with an amplitude of 1 PVU and a depth of 1 km, and estimate that the associated EP flux ranges between 0.1 and 100 mPa for a Richardson number between 1 and 10. These values of the flux are comparable with those observed in the lower stratosphere, which suggests that the mechanism identified in this paper provides a substantial gravity wave source, one that could be parameterized in GCMs.

Corresponding author address: François Lott, LMD, Ecole Normale Supérieure, 24 rue Lhomond, 75231 Paris, CEDEX 05, France. Email: flott@lmd.ens.fr

Abstract

The gravity waves (GWs) generated by potential vorticity (PV) anomalies in a rotating stratified shear flow are examined under the assumptions of constant vertical shear, two-dimensionality, and unbounded domain. Near a PV anomaly, the associated perturbation is well modeled by quasigeostrophic theory. This is not the case at large vertical distances, however, and in particular beyond the two inertial layers that appear above and below the anomaly; there, the perturbation consists of vertically propagating gravity waves. This structure is described analytically, using an expansion in the continuous spectrum of the singular modes that results from the presence of critical levels.

Several explicit results are obtained. These include the form of the Eliassen–Palm (EP) flux as a function of the Richardson number N22, where N is the Brunt–Väisälä frequency and Λ the vertical shear. Its nondimensional value is shown to be approximately exp(−πN/Λ)/8 in the far-field GW region, approximately twice that between the two inertial layers. These results, which imply substantial wave–flow interactions in the inertial layers, are valid for Richardson numbers larger than 1 and for a large range of PV distributions. In dimensional form they provide simple relationships between the EP fluxes and the large-scale flow characteristics.

As an illustration, the authors consider a PV disturbance with an amplitude of 1 PVU and a depth of 1 km, and estimate that the associated EP flux ranges between 0.1 and 100 mPa for a Richardson number between 1 and 10. These values of the flux are comparable with those observed in the lower stratosphere, which suggests that the mechanism identified in this paper provides a substantial gravity wave source, one that could be parameterized in GCMs.

Corresponding author address: François Lott, LMD, Ecole Normale Supérieure, 24 rue Lhomond, 75231 Paris, CEDEX 05, France. Email: flott@lmd.ens.fr

1. Introduction

It is well established that gravity waves (GWs) have a substantial influence on the large-scale atmospheric circulation, particularly in the middle atmosphere. As a result, sources of atmospheric GWs have received a great deal of attention. Significant tropospheric sources include topography, convective and frontal activities (Bretherton and Smolarkiewicz 1989; Shutts and Gray 1994), wind shear instabilities (Lalas and Einaudi 1976; Rosenthal and Lindzen 1983; Lott et al. 1992), nonmodal growth (Lott 1997; Bakas and Ioannou 2007), and geostrophic adjustment.

When it comes to geostrophic adjustment, one should distinguish classical adjustment, in which an initially unbalanced flow radiates GWs as it returns to near-geostrophic balance (Rossby 1937, Blumen 1972, Fritts and Luo 1992), from spontaneous adjustment, in which a well-balanced flow radiates weakly GWs in the course of its (near balanced) evolution (Ford et al. 2000). The source of GW activity differs between the two types of adjustment. In the first case, GW generation should be attributed to the mechanism responsible for the initial imbalance rather than to the adjustment.1 In the second case, spontaneous adjustment itself is the GW source.

Spontaneous adjustment, on which the present paper focuses, has been the subject of intense research activity. Much of this has been motivated by theoretical issues related to the limitations of balanced models and the nonexistence of exactly invariant slow manifolds (e.g., Vanneste 2008 and references therein). There are, however, practical implications, in particular for the parameterization of nonorographic GWs in general circulation models, since the (pseudo) momentum flux associated with the GWs generated spontaneously may well be significant. To assess this, it is important to quantify the GW activity generated by simple, physically plausible processes. This is the main aim of this paper.

The process that we examine is the generation of GWs that results from the advection of potential vorticity (PV) anomalies by a vertically sheared wind. Since we consider background flow with uniform PV, it is related to the work of Plougonven et al. (2005) on the unbalanced instabilities associated with surface edge waves (see also Molemaker et al. 2005). In our case, however, there is no boundary, and the PV disturbance is imposed within the flow and can have a finite vertical extent. The process that we examine is also related to the work of Vanneste and Yavneh (2004) and Olafsdottir et al. (2008) on GWs generated by PV anomalies in a horizontal shear with the difference that the wind shear is vertical in our case. A common feature of all of these processes, one which is likely generic for spontaneous-generation phenomena (Vanneste 2008), is that the GW activity is exponentially weak in the limit of small Rossby number, or equivalently large Richardson number. Our study, which is not restricted to this limit, confirms this conclusion; it nevertheless suggests that for reasonable values of the parameters, the GW amplitudes generated by PV anomalies in a vertical shear can be significant—comparable, for instance, with those observed in the low stratosphere by constant-level balloons far from mountain ranges (Hertzog et al. 2008).

In the background flow that we consider, with uniform PV, the small-amplitude PV anomalies are advected passively by the shear. The spectral representation of this dynamics involves a continuous spectrum of singular modes whose phase velocities are in the range of the basic-flow velocity, (−∞, ∞) in the unbounded domain that we assume. The vertical structure of the PV associated with the singular modes is simple: each mode is represented by a Dirac distribution centered at the critical level, where the phase velocity equals the basic flow velocity [see Pedlosky (1979) for a discussion of the analogous continuous spectrum in the quasigeostrophic (QG) Eady problem]. The fields decay rapidly above and below the critical level as far as the inertial levels where the Doppler-shifted frequency is equal to the Coriolis parameter (Jones 1967). Beyond these levels, the structure is oscillatory and can be identified as the GW signature of the (Dirac) PV anomaly.

The first purpose of this paper is to obtain the vertical structure of the singular modes with Dirac PV analytically. The second is to deduce, by integration over the continuous spectrum, the GW response to a vertically smooth, localized PV distribution. The third is to show that this GW response can yield substantial Eliassen–Palm (or pseudomomentum) fluxes at large vertical distances from the PV anomaly. Interestingly, we find that about half (exactly so in the limit of large Richardson number) of the pseudomomentum generated by the advection of PV is transported by GWs over arbitrarily large distances; the other half is deposited in an inertial layer, where substantial wave–flow interactions likely take place.

The plan of the paper is as follows: The general formulation of the problem and its transformation to a dimensionless form are given in section 2. Section 3 is devoted to the derivation of the vertical structure of the singular modes associated with a Dirac in PV. In section 4 we rephrase the result of section 3 in dimensional terms and estimate the amplitude of the pseudomomentum fluxes that can be expected from horizontally monochromatic PV anomalies; these compare with those measured in the lower stratosphere during field campaigns. We also consider PV distributions that are localized horizontally and have a finite depth so that the GW response is transient. In section 5, we summarize our results and discuss their significance for (i) the parameterization of GWs in GCMs, (ii) the transient evolution of baroclinic disturbances, and (iii) the treatment of the more general initial value problem. An appendix is devoted to approximate solutions valid in the limit of large Richardson number.

2. General formulation

a. Disturbance equations and potential vorticity

In the absence of mechanical and diabatic forcings, the hydrostatic–Boussinesq equations for the evolution of a two-dimensional disturbance in the uniformly sheared flow u0 = (Λz, 0, 0), where Λ > 0 denotes the shear, read as
i1520-0469-67-1-157-e21a
i1520-0469-67-1-157-e21b
i1520-0469-67-1-157-e21c
i1520-0469-67-1-157-e21d
and
i1520-0469-67-1-157-e21e
Here u′, υ′, and w′ are the three components of the velocity disturbance, p′ is the pressure disturbance, ρr is a constant reference density, θ′ is the potential temperature disturbance, θr is a constant reference potential temperature, g is the gravity constant, f is the Coriolis parameter, and N2 = gθ0z/θr is the square of the constant Brunt–Väisälä frequency, with θ0(y, z) the background potential temperature. Note that Λ appears in (2.1d) because the background shear flow is in thermal wind balance, θ0y = −θr fΛ/g.
Equations (2.1a)(2.1e) imply the conservation equation,
i1520-0469-67-1-157-e22
for the PV perturbation,
i1520-0469-67-1-157-e23
Note here that (2.2) and (2.3) correspond to the linearization of the conservation equation for the hydrostatic and Boussinesq approximation of the Ertel PV, the background PV, Λθ0y + fθ0z, being constant. It follows from (2.2) that the PV at any time t is given explicitly in terms of the initial condition q0(x, z) = q′(x, z, t = 0) by
i1520-0469-67-1-157-e24

b. Normal-mode decomposition

To evaluate the disturbance field associated with the PV anomaly (2.4), we express this solution in Fourier space:
i1520-0469-67-1-157-e25
where 0 is the Fourier transform of q0, satisfying
i1520-0469-67-1-157-e26
We rewrite (2.5) in the form
i1520-0469-67-1-157-e27
where δ(ξ) is the Dirac function of the variable
i1520-0469-67-1-157-e28
Note that (2.7) can be interpreted as the expansion of the perturbation PV in the (singular) normal modes of (2.3); these modes form a continuum, parameterized by the phase speed Λz′. The scaling used in (2.8) places the inertial levels of these modes at ξ = ±1 (Inverarity and Shutts 2000).
We are interested in the response of other fields, which can display GW activity, to the evolving PV (2.5). As a representative of these fields, we mainly focus on the perturbation streamfunction ψ′, related to the perturbation velocity in the (x, z) plane according to
i1520-0469-67-1-157-e29
The expansion of the streamfunction corresponding to the expansion (2.7) of the PV can be written as
i1520-0469-67-1-157-e210
where ψ̂0(k, z′) is the amplitude of the normal mode and Ψ(ξ) its vertical structure. Note that this expansion, which describes the part of ψ′ slaved to the PV, is not complete: an additional continuum of singular modes, representing free sheared GWs, would need to be added to the expansion to solve an arbitrary initial value problem.
The velocities u′, υ′, w′ and the potential temperature θ′ have expansions analogous to (2.10), with ψ̂0 replaced by û0, υ̂0, ŵ0 and θ̂0, and with Ψ replaced by U, V, W. Introducing these expansions into (2.1a)(2.1e) and choosing
i1520-0469-67-1-157-e211
gives
i1520-0469-67-1-157-e212
where we have introduced the Richardson number
i1520-0469-67-1-157-e213
We now introduce (2.11) and (2.12) into the expressions (2.3) and (2.7) for the PV. Choosing the streamfunction amplitude
i1520-0469-67-1-157-e214
then leads to the differential equation for the streamfunction structure Ψ(ξ):
i1520-0469-67-1-157-e215
We solve this equation explicitly in the next section.

3. Evaluation of Ψ(ξ)

To find a solution to (2.15), we first derive its homogeneous solutions for ξ > 0 and impose a radiation condition for ξ ≫ 1 to obtain a solution that represents an upward-propagating GW. We deduce from this a solution valid for ξ < 0, which represents a downward-propagating GW for ξ ≪ −1. The amplitudes of these two solutions are then chosen to satisfy the jump conditions associated with (2.15),
i1520-0469-67-1-157-e31

a. Homogeneous solution for ξ > 0

The change of variable η = ξ2 transforms (2.15) into the canonical form of the hypergeometric equation [Eq. (15.5.1) in Abramowitz and Stegun (1964), hereafter AS]:
i1520-0469-67-1-157-e32a
where
i1520-0469-67-1-157-e32b
Note that a + bc = 0, a relation that is related to the fact that the two inertial levels at ξ = ±1 are logarithmic singularities of (2.15), as described by Jones (1967).
For ξ > 1 we retain one of the two independent solutions of the hypergeometric equations [see AS, (15.5.8)], written as
i1520-0469-67-1-157-e33
where F denotes the hypergeometric function, a′ = a*, and b′ = 1 − a. We retain this solution because its asymptotic form,
i1520-0469-67-1-157-e34
corresponds to a GW propagating upward (Booker and Bretherton 1967). The other solution [given by (15.5.7) in AS] corresponds to a GW propagating downward.

It is important to note here that, without the hydrostatic approximation, the equation for Ψ in (2.15) would present turning points in the very far field. At these turning points, the GWs would be entirely reflected back toward the critical levels. This effect, and the fact that we are only interested here in the emission of the GWs by the PV, has motivated our choice to make the hydrostatic approximation.

For 0 < ξ < 1 the solution to (3.2a) is best written as a linear combination of the two independent solutions:
i1520-0469-67-1-157-e35
where a″ = 1 − a*, b″ = 1 − a, and A and B are two complex constants [(15.5.3–4) in AS]. To connect this solution to (3.3), we use a transformation formula for F [(15.3.10) in AS] and obtain the asymptotic approximations
i1520-0469-67-1-157-e36a
and
i1520-0469-67-1-157-e36b
In these expressions,
i1520-0469-67-1-157-e37
where Γ is the gamma function and ψd is the digamma function (see AS, chapter 6). The other coefficients (α′, β′) and (α″, β″) are defined by the same formulas with (a, b) replaced by (a′, b′) and (a″, b″), respectively.
To continue the solution (3.6a) below the inertial level at ξ = 1, we follow Booker and Bretherton (1967) and introduce a infinitely small linear damping that shifts the real ξ axis into the lower half of the complex plane so that
i1520-0469-67-1-157-e38
Thus, (3.6a) matches (3.6b) provided that
i1520-0469-67-1-157-e39
Solving for A and B gives
i1520-0469-67-1-157-e310
after simplifications using reflection formulas for the gamma and digamma functions [(6.1.17) and (6.3.7) in AS]. This completes the determination of Ψ(u)(ξ).

b. Solution over the entire domain

The solution for ξ < 0 can be deduced from Ψ(u)(ξ) by noting that (2.15) only contains real coefficients and is even. A possible solution is simply
i1520-0469-67-1-157-e311
This satisfies the radiation condition for ξ → −∞ since
i1520-0469-67-1-157-e312
which represents a downward-propagating GW.
The two solutions Ψ(u) and Ψ(d) can be combined to obtain a solution valid in the entire domain that satisfies the jump condition (3.1). This is given by
i1520-0469-67-1-157-e313
The jump condition is readily verified by noting that, if |ξ| < 1,
i1520-0469-67-1-157-e314a
i1520-0469-67-1-157-e314b
The first terms on the right-hand sides of (3.14a) and (3.14b) are identical and ensure the continuity of Ψ in ξ = 0 and do not contribute to the jump in (3.1); the second terms do not affect the value of Ψ in ξ = 0 but combine so that Ψξ/ξ2 jumps by 1 at ξ = 0 as required.
Several conclusions can be derived from the explicit form (3.13) of the streamfunction Ψ(ξ). First, for small |ξ|, Ψ(ξ) approaches the value
i1520-0469-67-1-157-e315
where the symbol is used to denote the asymptotic behavior for large J. This asymptotic estimate, derived here using Stirling’s formula for the gamma function [see AS’s (6.1.37)], is also obtained when the quasigeostrophic approximation of (2.15) is solved (see section 1 of the appendix). Second, we obtain from (3.4) and (3.13) that
i1520-0469-67-1-157-e316
where
i1520-0469-67-1-157-e317
The behavior as ξ → −∞ is similar. Accordingly, the amplitude of the GWs in the far field (|ξ| ≫ 1) is
i1520-0469-67-1-157-e318
The large J approximation in (3.18) is one of the main results of this paper. Obviously it cannot be recovered in the quasigeostrophic approximation, which filters out GWs completely, but it can be recovered by a WKB treatment of (2.15). This is demonstrated in section 2 of the appendix.
A third result derived from (3.13) concerns the EP flux (Eliassen and Palm 1961), or pseudomomentum flux, associated with the solution Ψ(ξ). Multiplying (2.15) by J3/2Ψ* and integrating by parts results in a conservation relation for the nondimensional EP flux
i1520-0469-67-1-157-e319
that is valid away from ξ = 0, ±1. Using the asymptotic approximation (3.16) shows that Fξ = μJ3/2|E|2/2 for |ξ| > 1. The flux Fξ is discontinuous across the inertial level ξ = 1. To evaluate its jump, we use the approximations (3.6a) and (3.6b) valid for ξ → 1± to find that Fξ(1+) − Fξ(1) = −πJ3/2|E|2|α″|2. Finally, the Taylor series expansions of (3.14a) and (3.14b) near ξ = 0 shows that the flux is continuous across ξ = 0: Fξ(0+) − Fξ(0) = 0. Using the explicit expression of α″ and Stirling’s formula, our results for the EP flux are summarized as follows:
i1520-0469-67-1-157-e320
This shows, in particular, that the GWs produced by the PV anomalies deposit almost as much momentum at the inertial levels as they transport in the far field.

c. Results

The four panels in Fig. 1 show the solution (3.13) for Ψ for four values of the Richardson number J. For all values of J, the streamfunction amplitude decays away from ξ = 0 in the region between ξ = ±1, where it has an almost constant phase, indicating a nonpropagating character. The decay for small ξ is well predicted by the quasigeostrophic approximation Ψg (also shown), which behaves essentially like exp(−J|ξ|) (see section 1 of the appendix).

Beyond the inertial levels, that is, for |ξ| > 1, the disturbance is propagating, with the real and imaginary part of Ψ in quadrature. The asymptotic behavior 1/2+ as |ξ| → ∞ is also shown in Fig. 1: it corresponds to a pure GW in our context. The amplitude |E| of the far-field GW, given in (3.18), is compared in Fig. 2 with its large J asymptotic estimate, also given in (3.18). The figure confirms the validity of this estimate and shows that it remains useful for values of J as small as 1. The figure also highlights the rapid decrease of the GW amplitude with J that is encapsulated in the exponential factor exp(−Jπ/2) appearing in the asymptotic estimate. A crude argument is suggestive of this dependence: if the quasigeostrophic approximation is used (well beyond its range of validity) up to the inertial levels ξ = ±1, the amplitude attained there, and hence the GW amplitude, is predicted to be roughly exp(−J). The presence of the factor π/2 can be traced to the breakdown of the quasigeostrophic approximation for ξ = O(1) and the replacement of the decay in exp(−J|ξ|) of the quasigeostrophic solution by a decay in exp(−J|sin−1ξ|). This is demonstrated explicitly by the WKB solution in section 2 of the appendix. Figure 2a also compares the value of Ψ(0) with its large J (quasigeostrophic) estimate, confirming the validity and usefulness of the estimate.

The amplitude of the nondimensional EP fluxes, between and beyond the inertial levels ξ = ±1, is shown in Fig. 2b. The exact results are compared with the large J estimates, which again prove accurate for J as small as 1. The figure, like Fig. 2a, illustrates the strong sensitivity of the GW generation to J.

4. Solutions for various PV distributions

To gauge the possible importance of sheared PV anomalies as a mechanism of GW generation, we now evaluate the GW field produced by a variety of localized anomalies. From now on, we report the results in dimensional form.

We compute the perturbation fields on a grid by evaluating the integral (2.10) numerically for different initial PV distributions q0(x, z) using the analytical solutions derived in section 3. The computational domain −X < x < X and −Z < z < Z is discretized in a regular grid with 2L + 1 points in the horizontal and 2M + 1 in the vertical. Correspondingly, the Fourier transform 0(k, z) is also discretized and given as 0(kl, zm), where kl = πl/X, l = −L, … , L and zm = πm/Z, m = −M, … , M. In the computations presented below, we typically take X = 4000 km, Z = 10 km, L = 2048, and M = 200, which ensures excellent resolutions in both directions of propagation as well as in the spectral space.

Given 0(kl, zm), the perturbation PV at different times is computed as
i1520-0469-67-1-157-e41
which is the discretized version of (2.7), where the values of x and z lie on the grid, Δk = π/(LX) and Δz = Z/M. The perturbation streamfunction then takes the analogous form
i1520-0469-67-1-157-e42
where ψ̂0(kl, zm) = r q̂0(kl, zm)/(rΛ2) [see (2.10) and (2.14)]. Similar expressions can be written down for u′, υ′, w′, and θ′ using (2.11) and (2.12).

a. Monochromatic, infinitely thin PV

The simplest case that we consider is that of uniform PV, with value qr, in a layer of depth σz that varies monochromatically with wavelength 2π/kr. Modeling this layer as infinitely thin corresponds to taking
i1520-0469-67-1-157-e43
The associated streamfunction is then simply
i1520-0469-67-1-157-e44
where ψr = r qr/(krθrΛ2). This is of course nothing other than the solution discussed in section 3 except for the dimensional factor. We focus on the dimensional aspect and examine the dimensional EP flux,
i1520-0469-67-1-157-e45
where the overbar denotes the horizontal average.
Let us estimate an order of magnitude for this EP flux. If we consider a 1-km-thick layer of stratospheric air entering in the troposphere, we can take a PV amplitude of ρr qr = 1 PVU, yielding ρr qr σz = 10−3 K s−1. Assuming that this air enters the troposphere at midlatitudes, we take ρr = 1 kg m−3, N = 0.01 s−1, θr = 300 K, and f = 10−4 s−1. For these parameters the dimensional factor, in Eq. (4.5),
i1520-0469-67-1-157-e46
Thus, the value of the nondimensional flux Fξ needs to be multiplied by 10 Pa to be dimensionally meaningful. This scaling, which is independent of the horizontal wavenumber kr, is used in Fig. 2 for the axis on the right of the panel. For J between 1 and 10, the EP flux is seen to be between 0.1 and 100 mPa. This covers the range of values measured in the lower stratosphere away from mountains during the constant-level-balloon Vorcore campaigns (Hertzog et al. 2008).

b. Horizontally localized, infinitely thin PV

We now consider a PV distribution that is localized in the horizontal but can still be modeled as infinitely thin in the vertical. Choosing a Gaussian profile in the horizontal, we represent the initial PV as
i1520-0469-67-1-157-e47
where σx gives the width of the PV distribution, by taking in (4.1)
i1520-0469-67-1-157-eq1

The vertical velocity field corresponding to this distribution of PV is shown in Fig. 3. The width of the PV anomaly has been taken as σx = 40 km (which corresponds to about 0.5° of longitude when f = 10−4 s−1); the other parameters are as in the preceding subsection. The four panels are obtained for different values of the shear Λ and hence of J, with the value of N kept constant.

For small values of J (J = 2 and J = 5), there are clear differences in the response to the PV anomaly between a region immediately surrounding the PV anomaly and the two far-field regions. The transition between these three regions can be located around the altitudes zI = ±x/Λ of the inertial levels of disturbances with wavelength 1/σx (for J = 2 and 5, zI = ±500 m and zI = ±1 km, respectively). In what follows, we call the transition regions “inertial layers.” Between the inertial layers, the vertical velocity is everywhere positive to the east of the positive PV disturbance (i.e., for x > 0) and negative to the west. This is because the transverse wind υ′ is cyclonic (with υ′ < 0 for x > 0 and υ′ < 0 for x < 0) and the meridional advection of background potential temperature, −fΛυ′ in (2.1d), is balanced by the vertical advection N2w′. Note also that the vertical velocity is almost untilted in the vertical in this region and decreases in amplitude when |z| increases. These indicate that the dynamics near z = 0 is well predicted by the quasigeostrophic theory.

Above and below the inertial layers, the disturbance has a propagating character, with w′ changing sign with altitude at a given horizontal location. In these two regions w′ is also tilted against the shear, indicating upward propagation in the upper region and downward propagation in the lower one. The GW signal is comparable in magnitude with the signal near the PV anomaly for J = 2 (Fig. 3a) but substantially smaller when J = 5 (Fig. 3b). For even larger J (Figs. 3c and 3d) the GW signal becomes very weak—so weak as to be undetectable for J = 25.

As noted, the fact that the PV disturbance is not monochromatic spreads the inertial levels over an inertial layer of finite depth. To illustrate how this affects the interactions with the large scales, Fig. 4 shows the averaged EP flux evaluated as
i1520-0469-67-1-157-e48
from the discrete approximations of u′, w′, υ′ and θ′. The EP flux is almost constant between the inertial layers, falls off by a factor of about 2 across the inertial layers, and is constant beyond. This suggests that substantial wave–mean flow interactions can occur at distances of up to a few kilometers from PV anomalies.
Note that the value of the EP flux in the far field can be estimated analytically. Returning to the continuous formalism, using the Gaussian distribution (4.7) and that Fξ is constant and independent of k for ξ > 1, gives
i1520-0469-67-1-157-e49
as in the monochromatic case (4.5), up to the multiplicative factor π.

c. Horizontally localized, finite-depth PV

The PV anomalies examined in sections 4a and 4b are approximated as infinitely thin layers. This approximation neglects the vertical shearing of the PV and hence the changes that this induces in the horizontal distribution of the PV. If this shearing occurs very rapidly, the results obtained so far are only relevant for a short time, after which the GW emission stops. To assess this, and more generally to demonstrate how our results predict a time-dependent GW generation, we now consider a PV distribution that has a finite depth. For simplicity, we take a PV distribution that is separable in x and z at t = 0. Specifically, we choose
i1520-0469-67-1-157-eq2
which has the same vertical integral as (4.7).

The time-dependent streamfunction corresponding to the PV is computed according to (4.2). Because x and t enter (4.2) only in the combination x − Λzmt, it is possible to reduce considerably the computations involved by summing vertical and horizontal translations of the solution in section 4b. To illustrate briefly how the horizontal translations are done, for fixed zm, the sum over the index l of the wavenumber kl at some time t can be inferred from the corresponding sum for t = 0 provided that t is a integer multiple of Δx/(ΛΔz). We have adjusted our grid size to take advantage of this for the values of t chosen for the results presented.

Figure 5 shows the evolution of the disturbance PV (gray shading) and of the vertical velocity it produces for σz = 1 km, J = 5, and all other parameters as in the previous sections. The solution is shown only for negative values of t. Indeed, the symmetries in our problem are such that, for the shallow PV disturbances considered, the solutions for positive t are almost symmetric to that at negative t. The background velocity shears the PV, which is strongly tilted against the shear for large negative time and with the shear for large positive time (not shown). Accordingly, the width of the PV distribution is deeply altered, decreasing here by a factor of ∼2 from t = −24 h to t = 0. As a result, the vertical velocity signal has a decreasing width and increasing amplitude as t increases toward 0.

Comparing the four panels in Fig. 5 to the (time independent) disturbance produced by the infinitely thin distribution of Fig. 3b indicates that the GW patterns in the far field are comparable at t = −12 h and almost identical at t = 0 h. Accordingly, and because the EP flux is a quadratic quantity, it is only in a time interval of a day or so that we can expect the EP flux to approach the values shown in Fig. 4. This last point is confirmed by Fig. 6, which displays the evolution of Fz evaluated at z = +10 km for the finite-depth PV anomaly. For all values of J, the EP flux peaks at t = 0, with peak values that compare well with the (time independent) values obtained for the infinitely thin distribution [see Fig. 4 and Eq. (4.9)]. The peak in EP flux shown in Fig. 6 broadens as J increases because the PV advection also slows. Nevertheless, the characteristic durations of the burst in EP flux remain of the order of a day in the range of J considered here.

Although this last result is quite sensitive to the spatial extent of the PV distribution (with the characteristic duration of the EP flux bursts decreasing when the depth of the PV disturbance increases and/or when its width decreases), it clearly illustrates that our mechanism of GWs generation is quite robust: the EP flux in the far field predicted by (4.9) at some time are representative of the flux within a few hours of that time. This is important for the parameterization of the GWs produced by PV anomalies in GCMs since these parameterization schemes can realistically be updated every few hours.

5. Summary and applications

a. Summary

In the presence of a uniform vertical shear, and in the absence of boundaries in the vertical, localized PV anomalies produce disturbances associated with two inertial layers. These layers are located above and below the PV anomaly, at a distance zI = σxf /Λ, where σx is the typical width of the PV anomaly. Between these two inertial layers, the form of the disturbance is qualitatively well predicted by the quasigeostrophic theory, and geostrophic balance is a good approximation for the meridional wind in this region. Accordingly, the decay with altitude of the disturbance amplitude is exponential with a decay rate of the order of N/( x). Beyond the inertial layers, the intrinsic frequency of the disturbance is larger than f so that the disturbance propagates vertically in the form of a GW. If the disturbance amplitude is substantial at the inertial level, the GW amplitudes produced by this mechanism can be significant. This condition is satisfied provided zIN/( x) = N/Λ = J is not large. More specifically, we show that the amplitude of the GW is near exp(−Jπ/2)/(2J) for J > 1. Correspondingly, the Eliassen–Palm flux associated with the GW depends also exponentially on J, scaling almost like exp(−πJ)/8 beyond the inertial levels and almost like exp(−πJ)/4 between them.

The robustness of these results has been tested numerically and for the case of a PV disturbance localized horizontally and of finite depth. Horizontal localization spreads the inertial level of the monochromatic case over an inertial layer, and finite depth leads to a time-dependent perturbation. It is nevertheless shown that under these circumstances the GW amplitudes and the EP fluxes in the far field remain well predicted by the monochromatic results, with multiplicative factors that are O(1) [see (4.9)]. The characteristic time over which they evolve is also found to exceed a few hours, suggesting that our formula can usefully predict the EP fluxes produced by PV anomalies (if these anomalies are diagnosed every hour, for instance).

b. Applications

These results can be directly useful for the parameterization of GWs in GCMs that include the middle atmosphere. The dimensional results of section 4 suggest that the EP fluxes in the far field produced by localized PV anomalies compare in amplitude with the EP fluxes measured in the lower stratosphere by almost constant-level balloons during the Vorcore campaign (Hertzog et al. 2008). Although the analytical expressions we give are well suited to this context, notably because the GW parameterization routines typically reevaluate the EP fluxes according to the large-scale flow every hour, they require an estimation of the PV anomalies at subgrid scales.

In most of the parameterizations used currently, the GW EP fluxes from the troposphere toward the middle atmosphere are imposed regardless of the GWs tropospheric sources. There are nevertheless some exceptions, as in Charron and Manzini (2002), where the GW amplitude is larger if fronts are identified. Our formula in (3.20) could well be used in this context since large PV anomalies form during frontogenesis. It could therefore be used to parameterize quantitatively the GW radiated by fronts as well as by other processes that induce localized PV anomalies.

Our results are also relevant to the problem of nonmodal growth of baroclinic disturbances (Farrell 1989). In fact, the inertial levels, which are ignored in balanced approximations, change the boundary conditions at large distances from the critical level from decay conditions into radiation conditions; as a result, the global structure of the solutions in the continuous spectrum is deeply altered. Consequently an inflow PV anomaly of short horizontal scales can have a much larger surface signature than predicted by balanced models. This is because the disturbances produced in this case can have an inertial layer between the PV and the surface. As our solutions are the nongeostrophic counterpart of the building blocks that are used by Bishop and Heifetz (2000) to explain the triggering of storms by upstream PV anomalies or to explain the optimal perturbations evolution by de Vries and Opsteegh (2007), they may be useful to examine these problems in the nongeostrophic case.

Finally, our results could be useful to study the evolution of initial disturbances or of disturbances produced by any external causes (the classical adjustment problem). For this purpose, it should nevertheless be noticed that the modes with Dirac PV are not the only modes associated with the continuous spectrum. In fact, for each value of the phase velocity, there are two further singular modes, with zero PV and singular behavior of the other fields at the inertial levels. These modes are essential to the completeness of the modal representation of the solutions.

Acknowledgments

This work was supported by the Alliance Programme of the French Foreign Affairs Ministry and British Council. J.V. acknowledges the support of a NERC grant. R.P. acknowledges the support of the ANR project FLOWING.

REFERENCES

  • Abramowitz, M., and I. A. Stegun, 1964: Handbook of Mathematical Functions. 9th ed. Dover, 1045 pp.

  • Bakas, N. A., and P. J. Ioannou, 2007: Momentum and energy transport by gravity waves in stochastically driven stratified flows. Part I: Radiation of gravity waves from a shear layer. J. Atmos. Sci., 64 , 15091529.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., and E. Heifetz, 2000: Apparent absolute instability and the continuous spectrum. J. Atmos. Sci., 57 , 35923608.

  • Blumen, W., 1972: Geostrophic adjustment. Rev. Geophys., 10 , 485528.

  • Booker, J. R., and F. P. Bretherton, 1967: The critical layer for internal gravity waves in a shear flow. J. Fluid Mech., 27 , 513539.

    • Search Google Scholar
    • Export Citation
  • Bretherton, C. S., and P. K. Smolarkiewicz, 1989: Gravity waves, compensating subsidence, and detrainment around cumulus clouds. J. Atmos. Sci., 46 , 740759.

    • Search Google Scholar
    • Export Citation
  • Charron, M., and E. Manzini, 2002: Gravity waves from fronts: Parameterization and middle atmosphere response in a general circulation model. J. Atmos. Sci., 59 , 923941.

    • Search Google Scholar
    • Export Citation
  • de Vries, H., and H. Opsteegh, 2007: Resonance in optimal perturbation evolution. Part I: Two-layer Eady model. J. Atmos. Sci., 64 , 673694.

    • Search Google Scholar
    • Export Citation
  • Eliassen, A., and E. Palm, 1961: On the transfer of energy in stationary mountain waves. Geophys. Publ., 22 , 123.

  • Farrell, B. F., 1989: Optimal excitation of baroclinic waves. J. Atmos. Sci., 46 , 11931206.

  • Ford, R., M. E. McIntyre, and W. A. Norton, 2000: Balance and the slow quasimanifold: Some explicit results. J. Atmos. Sci., 57 , 12361254.

    • Search Google Scholar
    • Export Citation
  • Fritts, D. C., and Z. Luo, 1992: Gravity wave excitation by geostrophic adjustment of the jet stream. Part I: Two-dimensional forcing. J. Atmos. Sci., 49 , 681697.

    • Search Google Scholar
    • Export Citation
  • Hertzog, A., G. Boccara, R. A. Vincent, F. Vial, and P. Cocquerez, 2008: Estimation of gravity wave momentum flux and phase speeds from quasi-Lagrangian stratospheric balloon flights. Part II: Results from the Vorcore campaign in Antarctica. J. Atmos. Sci., 65 , 30563070.

    • Search Google Scholar
    • Export Citation
  • Inverarity, G. W., and G. J. Shutts, 2000: A general, linearized vertical structure equation for the vertical velocity: Properties, scalings and special cases. Quart. J. Roy. Meteor. Soc., 126 , 27092724.

    • Search Google Scholar
    • Export Citation
  • Jones, W. L., 1967: Propagation of internal gravity waves in fluids with shear flow and rotation. J. Fluid Mech., 30 , 439448.

  • Lalas, D. P., and F. Einaudi, 1976: On the characteristics of gravity waves generated by atmospheric shear layers. J. Atmos. Sci., 33 , 12481259.

    • Search Google Scholar
    • Export Citation
  • Lott, F., 1997: The transient emission of propagating gravity waves by a stably stratified shear layer. Quart. J. Roy. Meteor. Soc., 123 , 16031619.

    • Search Google Scholar
    • Export Citation
  • Lott, F., 2003: Large-scale flow response to short gravity waves breaking in a rotating shear flow. J. Atmos. Sci., 60 , 16911704.

  • Lott, F., H. Kelder, and H. Teitelbaum, 1992: A transition from Kelvin–Helmholtz instabilities to propagating wave instabilities. Phys. Fluids, 4 , 19901997.

    • Search Google Scholar
    • Export Citation
  • Molemaker, M. J., J. C. McWilliams, and I. Yavneh, 2005: Baroclinic instability and loss of balance. J. Phys. Oceanogr., 35 , 15051517.

    • Search Google Scholar
    • Export Citation
  • Olafsdottir, E. I., A. B. O. Daalhuis, and J. Vanneste, 2008: Inertia–gravity-wave radiation by a sheared vortex. J. Fluid Mech., 596 , 169189.

    • Search Google Scholar
    • Export Citation
  • Pedlosky, J., 1979: Geophysical Fluid Dynamics. Springer, 624 pp.

  • Plougonven, R., D. J. Muraki, and C. Snyder, 2005: A baroclinic instability that couples balanced motions and gravity waves. J. Atmos. Sci., 62 , 15451559.

    • Search Google Scholar
    • Export Citation
  • Rosenthal, A. J., and R. S. Lindzen, 1983: Instabilities in a stratified fluid having one critical level. Part I: Results. J. Atmos. Sci., 40 , 509520.

    • Search Google Scholar
    • Export Citation
  • Rossby, C. G., 1937: On the mutual adjustment of pressure and velocity distributions in certain simple current systems. J. Mar. Res., 21 , 1528.

    • Search Google Scholar
    • Export Citation
  • Scavuzzo, C. M., M. A. Lamfri, H. Teitelbaum, and F. Lott, 1998: A study of the low-frequency inertio-gravity waves observed during the Pyrénées Experiment. J. Geophys. Res., 103 , 17471758.

    • Search Google Scholar
    • Export Citation
  • Shutts, G. J., and M. E. B. Gray, 1994: A numerical modeling study of the geostrophic adjustment process following deep convection. Quart. J. Roy. Meteor. Soc., 120 , 11451178.

    • Search Google Scholar
    • Export Citation
  • Vanneste, J., 2008: Exponential smallness of inertia–gravity wave generation at small Rossby number. J. Atmos. Sci., 65 , 16221637.

  • Vanneste, J., and I. Yavneh, 2004: Exponentially small inertia–gravity waves and the breakdown of quasigeostrophic balance. J. Atmos. Sci., 61 , 211223.

    • Search Google Scholar
    • Export Citation

APPENDIX

Approximate Solutions

Quasigeostrophic approximation

The quasigeostrophic approximation can be recovered by approximating the perturbation PV as
i1520-0469-67-1-157-ea1
Using the polarization relations (2.11) and (2.12), the QG PV equation has a streamfunction solution ψg that can be written as a spectral expansion similar to (2.10), where ψ̂0 is as in (2.14) and Ψ(ξ) is replaced by the structure function Ψg(ξ). This new structure function Ψg(ξ) is a solution of the QG approximation of the structure equation (2.15), namely
i1520-0469-67-1-157-ea2
The solution vanishing for |ξ| → ∞ is given by
i1520-0469-67-1-157-ea3

WKB approximation

In this section, we solve the differential Eq. (2.15) asymptotically using a WKB method. This method has the interest of providing the solution in terms of (mostly) elementary functions: these reveal the structure of the solution more explicitly than the exact solutions do.

To derive an approximate homogeneous solution Ψ(u)(ξ) valid for ξ > 0, we distinguish four different regions in which Ψ(u)(ξ) takes a different asymptotic form: (i) an inner region with ξ ≪ 1; (ii) an outer region with O(1) = ξ < 1; (iii) an inner region for ξ ≈ 1; and (iv) a second outer region for ξ > 1. In the outer regions (ii) and (iv), two independent solutions are found by introducing the WKB ansatz
i1520-0469-67-1-157-eqa1
to find that
i1520-0469-67-1-157-eqa2
for some constant A. Thus, we obtain
i1520-0469-67-1-157-ea4
in region (ii) and
i1520-0469-67-1-157-ea5
in region (iv), where A(ii), B(ii), A(vi), and B(vi) are arbitrary constants. The proper radiation condition as ξ → ∞ is satisfied by (A.5) provided that
i1520-0469-67-1-157-eqa3
The WKB solutions (A.4) and (A.5) break down as the inner regions are approached and (2.15) needs to be rescaled. For region (i), we introduce Z = Jz into (2.15) and obtain at leading order the equation
i1520-0469-67-1-157-eqa4
which is equivalent to the quasigeostrophic approximation (A.2), giving
i1520-0469-67-1-157-ea6
In region (iii), we introduce the scaled variable ζ = J(ξ − 1) and obtain the leading-order equation
i1520-0469-67-1-157-eqa5
The solution can be conveniently written in terms of Hankel functions (see chapter 9 of AS):
i1520-0469-67-1-157-ea7
The response to a Dirac of PV can then be constructed as in (3.11) by combining Ψ(u)(ξ) for ξ > 0 with [Ψ(u)(−ξ)]* for ξ < 0. The jump condition at ξ = 0, applied to (A.6), gives
i1520-0469-67-1-157-ea8
All the constants can then be obtained by matching the various asymptotic results across regions. Matching between regions (i) and (ii) gives
i1520-0469-67-1-157-eqa6
To match between (ii) and (iii), we recall the branch choice (3.8) and note the asymptotic formulas
i1520-0469-67-1-157-ea9
valid for |x| → ∞, and −π < argx < 2π and −2π < argx < π, respectively [see (9.2.3–4) in AS]. Using this in (A.7) and comparing with the expansion of (A.4) for ξ → 1 gives
i1520-0469-67-1-157-eqa7
Similarly, comparing with the expansion of (A.5) for ξ → 1+ gives
i1520-0469-67-1-157-eqa8
Since the radiation condition for large ξ implies that B(iv) = 0, we conclude that
i1520-0469-67-1-157-eqa9
It follows that A(i) = 1/(2J3/2) and, hence, that
i1520-0469-67-1-157-eqa10
Thus, we find that Ψ(0) ∼ 1/(2J3/2) and that the gravity wave amplitude is |A(iv)| ∼ eJπ/2/(2J), consistent with (3.15) and (3.18), respectively. The EP flux FξeJπ/8 for |ξ| > 1 follows immediately, consistent with (3.20). The evaluation of the EP flux for |ξ| < 1 is more delicate because it includes terms that are exponentially small in J and are ignored here.

Fig. 1.
Fig. 1.

Structure function Ψ(ξ) associated with a monochromatic PV distribution equal to δ(ξ) for a Richardson number (a) J = 2, (b) J = 5, (c) J = 10, and (d) J = 25. The thick black curves and thick dotted curves show the real and imaginary parts of Ψ, respectively; the thick gray curves show the quasigeostrophic approximation Ψg; and the gray dots show the real part of the far-field GW approximation 1/2+.

Citation: Journal of the Atmospheric Sciences 67, 1; 10.1175/2009JAS3134.1

Fig. 2.
Fig. 2.

Characteristic amplitudes of the disturbances produced by the Dirac PV anomaly: (a) exact and approximate values for the GW amplitude |E| given by (3.18) (black solid and gray dashed curves, respectively) and exact and approximate values of Ψ(0) given by (3.15) (gray solid curve and black dots, respectively), (b) exact and approximate values of the EP flux F ξ between the inertial levels (black solid curve and gray solid curve) and exact and approximate values of F ξ beyond the inertial levels (black dashed curve and gray dashed curve); see (3.20). The axis on the right of (b) gives the results in the dimensional form corresponding to the parameter choice (4.6).

Citation: Journal of the Atmospheric Sciences 67, 1; 10.1175/2009JAS3134.1

Fig. 3.
Fig. 3.

Vertical velocity induced by a horizontally localized but infinitely thin PV anomaly [see the PV horizontal profile in (4.7)] for (a) J = 2, (b) J = 5, (c) J = 10, and (d) J = 25, with dashed curves indicating negative values. The contour interval, indicated on each panel, varies.

Citation: Journal of the Atmospheric Sciences 67, 1; 10.1175/2009JAS3134.1

Fig. 4.
Fig. 4.

Vertical profiles of Eliassen–Palm flux for the four solutions in Fig. 3. Note the rescaling of the flux for the different values of J.

Citation: Journal of the Atmospheric Sciences 67, 1; 10.1175/2009JAS3134.1

Fig. 5.
Fig. 5.

Evolution of the vertical velocity field associated with the evolution of a PV disturbance of finite depth σz = 1 km, finite width σx = 40 km, and maximum value of ρr qr = 1 PVU, and when the Richardson number J = 5. PV values above 0.1 PVU are shaded; contours for the vertical velocity as in Fig. 3b.

Citation: Journal of the Atmospheric Sciences 67, 1; 10.1175/2009JAS3134.1

Fig. 6.
Fig. 6.

Temporal evolution of the far-field EP flux for finite-depth, finite-width PV anomaly and for different values of the Richardson number J. Note that the EP flux has been rescaled as in Fig. 4.

Citation: Journal of the Atmospheric Sciences 67, 1; 10.1175/2009JAS3134.1

1

It is in this context that Scavuzzo et al. (1998) and Lott (2003) attributed the presence of inertia–gravity waves near mountains to a large-scale adjustment to breaking small-scale gravity waves.

Save
  • Abramowitz, M., and I. A. Stegun, 1964: Handbook of Mathematical Functions. 9th ed. Dover, 1045 pp.

  • Bakas, N. A., and P. J. Ioannou, 2007: Momentum and energy transport by gravity waves in stochastically driven stratified flows. Part I: Radiation of gravity waves from a shear layer. J. Atmos. Sci., 64 , 15091529.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., and E. Heifetz, 2000: Apparent absolute instability and the continuous spectrum. J. Atmos. Sci., 57 , 35923608.

  • Blumen, W., 1972: Geostrophic adjustment. Rev. Geophys., 10 , 485528.

  • Booker, J. R., and F. P. Bretherton, 1967: The critical layer for internal gravity waves in a shear flow. J. Fluid Mech., 27 , 513539.

    • Search Google Scholar
    • Export Citation
  • Bretherton, C. S., and P. K. Smolarkiewicz, 1989: Gravity waves, compensating subsidence, and detrainment around cumulus clouds. J. Atmos. Sci., 46 , 740759.

    • Search Google Scholar
    • Export Citation
  • Charron, M., and E. Manzini, 2002: Gravity waves from fronts: Parameterization and middle atmosphere response in a general circulation model. J. Atmos. Sci., 59 , 923941.

    • Search Google Scholar
    • Export Citation
  • de Vries, H., and H. Opsteegh, 2007: Resonance in optimal perturbation evolution. Part I: Two-layer Eady model. J. Atmos. Sci., 64 , 673694.

    • Search Google Scholar
    • Export Citation
  • Eliassen, A., and E. Palm, 1961: On the transfer of energy in stationary mountain waves. Geophys. Publ., 22 , 123.

  • Farrell, B. F., 1989: Optimal excitation of baroclinic waves. J. Atmos. Sci., 46 , 11931206.

  • Ford, R., M. E. McIntyre, and W. A. Norton, 2000: Balance and the slow quasimanifold: Some explicit results. J. Atmos. Sci., 57 , 12361254.

    • Search Google Scholar
    • Export Citation
  • Fritts, D. C., and Z. Luo, 1992: Gravity wave excitation by geostrophic adjustment of the jet stream. Part I: Two-dimensional forcing. J. Atmos. Sci., 49 , 681697.

    • Search Google Scholar
    • Export Citation
  • Hertzog, A., G. Boccara, R. A. Vincent, F. Vial, and P. Cocquerez, 2008: Estimation of gravity wave momentum flux and phase speeds from quasi-Lagrangian stratospheric balloon flights. Part II: Results from the Vorcore campaign in Antarctica. J. Atmos. Sci., 65 , 30563070.

    • Search Google Scholar
    • Export Citation
  • Inverarity, G. W., and G. J. Shutts, 2000: A general, linearized vertical structure equation for the vertical velocity: Properties, scalings and special cases. Quart. J. Roy. Meteor. Soc., 126 , 27092724.

    • Search Google Scholar
    • Export Citation
  • Jones, W. L., 1967: Propagation of internal gravity waves in fluids with shear flow and rotation. J. Fluid Mech., 30 , 439448.

  • Lalas, D. P., and F. Einaudi, 1976: On the characteristics of gravity waves generated by atmospheric shear layers. J. Atmos. Sci., 33 , 12481259.

    • Search Google Scholar
    • Export Citation
  • Lott, F., 1997: The transient emission of propagating gravity waves by a stably stratified shear layer. Quart. J. Roy. Meteor. Soc., 123 , 16031619.

    • Search Google Scholar
    • Export Citation
  • Lott, F., 2003: Large-scale flow response to short gravity waves breaking in a rotating shear flow. J. Atmos. Sci., 60 , 16911704.

  • Lott, F., H. Kelder, and H. Teitelbaum, 1992: A transition from Kelvin–Helmholtz instabilities to propagating wave instabilities. Phys. Fluids, 4 , 19901997.

    • Search Google Scholar
    • Export Citation
  • Molemaker, M. J., J. C. McWilliams, and I. Yavneh, 2005: Baroclinic instability and loss of balance. J. Phys. Oceanogr., 35 , 15051517.

    • Search Google Scholar
    • Export Citation
  • Olafsdottir, E. I., A. B. O. Daalhuis, and J. Vanneste, 2008: Inertia–gravity-wave radiation by a sheared vortex. J. Fluid Mech., 596 , 169189.

    • Search Google Scholar
    • Export Citation
  • Pedlosky, J., 1979: Geophysical Fluid Dynamics. Springer, 624 pp.

  • Plougonven, R., D. J. Muraki, and C. Snyder, 2005: A baroclinic instability that couples balanced motions and gravity waves. J. Atmos. Sci., 62 , 15451559.

    • Search Google Scholar
    • Export Citation
  • Rosenthal, A. J., and R. S. Lindzen, 1983: Instabilities in a stratified fluid having one critical level. Part I: Results. J. Atmos. Sci., 40 , 509520.

    • Search Google Scholar
    • Export Citation
  • Rossby, C. G., 1937: On the mutual adjustment of pressure and velocity distributions in certain simple current systems. J. Mar. Res., 21 , 1528.

    • Search Google Scholar
    • Export Citation
  • Scavuzzo, C. M., M. A. Lamfri, H. Teitelbaum, and F. Lott, 1998: A study of the low-frequency inertio-gravity waves observed during the Pyrénées Experiment. J. Geophys. Res., 103 , 17471758.

    • Search Google Scholar
    • Export Citation
  • Shutts, G. J., and M. E. B. Gray, 1994: A numerical modeling study of the geostrophic adjustment process following deep convection. Quart. J. Roy. Meteor. Soc., 120 , 11451178.

    • Search Google Scholar
    • Export Citation
  • Vanneste, J., 2008: Exponential smallness of inertia–gravity wave generation at small Rossby number. J. Atmos. Sci., 65 , 16221637.

  • Vanneste, J., and I. Yavneh, 2004: Exponentially small inertia–gravity waves and the breakdown of quasigeostrophic balance. J. Atmos. Sci., 61 , 211223.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    Structure function Ψ(ξ) associated with a monochromatic PV distribution equal to δ(ξ) for a Richardson number (a) J = 2, (b) J = 5, (c) J = 10, and (d) J = 25. The thick black curves and thick dotted curves show the real and imaginary parts of Ψ, respectively; the thick gray curves show the quasigeostrophic approximation Ψg; and the gray dots show the real part of the far-field GW approximation 1/2+.

  • Fig. 2.

    Characteristic amplitudes of the disturbances produced by the Dirac PV anomaly: (a) exact and approximate values for the GW amplitude |E| given by (3.18) (black solid and gray dashed curves, respectively) and exact and approximate values of Ψ(0) given by (3.15) (gray solid curve and black dots, respectively), (b) exact and approximate values of the EP flux F ξ between the inertial levels (black solid curve and gray solid curve) and exact and approximate values of F ξ beyond the inertial levels (black dashed curve and gray dashed curve); see (3.20). The axis on the right of (b) gives the results in the dimensional form corresponding to the parameter choice (4.6).

  • Fig. 3.

    Vertical velocity induced by a horizontally localized but infinitely thin PV anomaly [see the PV horizontal profile in (4.7)] for (a) J = 2, (b) J = 5, (c) J = 10, and (d) J = 25, with dashed curves indicating negative values. The contour interval, indicated on each panel, varies.

  • Fig. 4.

    Vertical profiles of Eliassen–Palm flux for the four solutions in Fig. 3. Note the rescaling of the flux for the different values of J.

  • Fig. 5.

    Evolution of the vertical velocity field associated with the evolution of a PV disturbance of finite depth σz = 1 km, finite width σx = 40 km, and maximum value of ρr qr = 1 PVU, and when the Richardson number J = 5. PV values above 0.1 PVU are shaded; contours for the vertical velocity as in Fig. 3b.

  • Fig. 6.

    Temporal evolution of the far-field EP flux for finite-depth, finite-width PV anomaly and for different values of the Richardson number J. Note that the EP flux has been rescaled as in Fig. 4.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 309 74 3
PDF Downloads 233 61 5