1. Introduction
The nature of topographically induced perturbations in stratified flows has been a widely studied subject due to its applicability to airflows over complex terrain and the associated mesoscale phenomena. The variety of the topographically induced perturbations reflects the sensitivity of the system to the nature of forcing as well as to the properties of the medium through which they propagate. In the atmosphere, where (in general) both wind and stability vary with height, the impact of this variation on the vertical structure of internal gravity waves can be profound. This is especially evident if the mean flow contains critical levels for mountain-generated internal gravity waves.
A critical level is one at which the intrinsic frequency ω′ = U·κ − ω of a plane wave with horizontal wavenumber vector κ andfrequency ω becomes zero. In other words, it is a level where wave phase speed becomes equal to the component of the mean velocity parallel to the horizontal wavenumber vector. For mountain waves resulting from a steady forcing by the flow over topography, ω = 0, and the critical levels arise wherever U·κ = 0. The name critical derives from the singular behavior of the linearized inviscid equations at such levels, recognized originally in the context of hydrodynamic stability of parallel shear flows [see Drazin and Reid (1971) for a review]. The seminal contributions by Bretherton (1966) and Booker and Bretherton (1967) were first to describe the effects of critical levels on the propagation of internal gravity waves, analyzing both the transient and steady-state aspects of the linearized inviscid 2D problem. Using the WKB approximation (assuming the mean-flow Ri is large), Bretherton (1966) showed that as a wave packet approaches a critical level, the vertical component of the group velocity and the vertical wavelength tend to zero. The wave packet cannot reach the critical level in finite time, and its energy becomes absorbed by the mean current without any reflection. Booker and Bretherton (1967) investigated the interaction of monochromatic internal waves generated by a flow over a wavy boundary and a critical level with Ri > ¼. Their results showed that the vertical wavenumber and horizontal velocity become infinite at the critical level. Aloft, the amplitude of the waves as well as energy and momentum fluxes are strongly attenuated. The effective absorption of waves (i.e., the deposition of the wave momentum and energy into the mean flow) occurs below the critical level.
The singular behavior of the linear, inviscid, time-independent equations indicates that the transience, momentum, and heat diffusion, as well as nonlinear steepening and amplification of waves, are important in a vicinity of the critical level (i.e., within a critical layer). Different balances between these effects can lead to a variety of flow evolution scenarios, from linear viscous wave absorption to strongly nonlinear “cat’s eye” circulations within critical layers, as demonstrated in a number of studies (theoretical, numerical, and laboratory) that have examined the interaction of small- and large-amplitude waves with a critical level. For reviews see Kelly (1977), Maslowe (1986), and chapter 4.11 in Baines (1995). For Ri > ¼ at the critical level, two aspects of the gravity wave–critical level interaction are well established. First, in agreement with the linear prediction, there is no significant transmission of wave energy through the critical level. Second, for sufficiently large-amplitude incoming waves, nonlinearities dominate viscosity and diffusion within the critical layer, resulting in wave steepening and overturning. Thus, the critical level is a preferred location for internal wavebreaking (Winters and D’Asaro 1989; Winters and Riley 1992; Fritts et al. 1994). The resultant convective and shear instabilities lead to the formation of a homogeneous mixed layer below the critical level that acts as a perfect reflector to all incoming waves.
Such a scenario of the local flow evolution has been identified by Clark and Peltier in their numerical simulations of 2D finite-amplitude mountain waves (Clark and Peltier 1977). They have attributed the formation of the downslope windstorms (cf. Lilly and Zipser 1977) to the wave–critical level interaction. Here, thecritical level and the associated critical layer are wave-induced local features corresponding to a quasi-stagnant homogeneous region of finite thickness and horizontal extent that forms as a result of mixing in the overturning mountain waves above the lee slopes. The reflective behavior of the critical level decouples the flow aloft from that at lower levels, whereupon the low-level stratified flow yields a sub- to supercritical transition over the obstacle for certain critical-level heights (Peltier and Clark 1979; Smith 1985). A similar flow evolution has been simulated by substituting the local, wave-induced critical level with a global critical level in the ambient flow (Bacmeister and Pierrehumbert 1988; Durran and Klemp 1987). By specifying an ambient critical level, the height of the wave reflecting level could be controlled, which has led to the numerical solutions consistent with the nonlinear analytic model of Smith (1985).
Although nonlinearity seems to be an inherent part of the 2D gravity-wave–critical-level interaction, the recent study by Wurtele et al. (1996) documents that this does not necessarily extrapolate to all types of flows with critical levels. Their investigation addresses the propagation of small-amplitude 2D gravity-inertia waves below corresponding critical levels (Rossby singular levels) and provides an example of linear critical-level behavior. For a monochromatic gravity-inertia wave they found that the wave–critical-level interaction leads to a nonlinear reflection similar to the case of a gravity wave. However, when a prescribed forcing excites a continuous wave spectrum, with the Rossby critical levels different for every wave component, the problem is essentially linear as the critical levels produce no singular effects even in the absence of viscous dissipation.
While the understanding of the gravity-wave–critical-level interaction in 2D orographic flows is still incomplete, recently, interests have shifted toward 3D flows (Nappo and Chimonas 1992; Shutts 1995; Broad 1995). In general, 3D gravity waves encounter their critical levels at different altitudes (similar to the 2D gravity-inertia waves), where the projection of the mean velocity U(z) on a given wavenumber vector κ = (k, l) vanishes. Using linear analysis, Shutts (1995) and Broad (1995) have examined the effects of the directional wind shear (and the associated continuous distribution of critical levels) on the vertical flux of horizontal momentum. In particular, they showed that the stress force due to wave absorption at any critical level is normal to the local wind direction—a finding important for parameterizing the gravity wave drag in large-scale models. A single critical level in a unidirectional mean flow has been addressed by Nappo and Chimonas (1992). Employing linear theory, they investigated the effects of wave absorption below the critical level within a stably stratified boundary layer. The surface wave drag and the stress due to the critical-level absorption were found to be of the same order of magnitude as the drag conventionally associated with the surface friction.
In this study, we address unidirectional steady flows with the critical level located where U = 0 and describe the 3D steady-state wave pattern forced by an isolated axisymmetric obstacle. We combine the approaches of Booker and Bretherton (1967) and Smith (1980) to provide linear hydrostatic solutions for a class of constant-shear flows past a bell-shaped mountain. In thezero-shear limit, our solutions reduce to those of Smith (1980) for the uniform-wind case. Similar to the equivalent 2D problem (Smith 1986), we find that the wave drag (as well as the solutions themselves) depends on the mean shear. However, in the limit of Ri d⃗ ¼, this dependence is essentially different from that in 2D, due to an inherently three-dimensional effect of lateral wave dispersion. Since linear inviscid solutions are singular at the critical level we use a numerical model to assess the impact of dissipation and local nonlinearities on the solutions. We determine the bounds, in the parameter space spanned by nondimensional mountain height
Numerical simulations also indicate that the wave breaking below a single critical level in 3D orographic flows can be induced for relatively small mountains, a behavior already familiar from the 2D orographic flows. Thus, by introducing a critical level in an ambient flow, the wave breaking aloft can be examined separately from other nonlinear aspects of the 3D flows. This provides an opportunity to clarify the role of wave breaking in fostering the transition from a linear flow regime dominated by mountain waves to a nonlinear regime characterized by splitting of the low-level upwind flow and the formation of lee eddies (Hunt and Snyder 1980; Smolarkiewicz and Rotunno 1989, 1990; Smith 1989; Crook et al. 1990; Miranda and James 1992; Smith and Grønås 1993; Schär and Durran 1997 and references therein). Such finite-amplitude effects in 3D flows with critical levels will be addressed in future studies.
The paper is organized as follows. In section 2, we present the linear equations and discuss their solutions. In section 3, we present the numerical model and compare the results of selected simulations with the linear theory predictions. Section 4 concludes the paper.
2. Linear model
a. Governing equations
In the next section we discuss solutions of (5)–(6) for negatively sheared mean flows with critical levels.
b. Integral solutions
In the limit of zero shear, zc → ∞,
Unlike the 3D linear mountain waves in an atmosphere with constant wind and stability, the linear solution of (1a)–(1e) contains nonzero vertical vorticity. The vertical component of vorticity is generated by the tilting of the mean shear vector Uzŷ in the linear wave field while the total vorticity vector remains tangential to the isopycnal surfaces. This mechanism is already captured in the linearized vorticity equation implied by(1a)–(1c). Figure 3 shows the vertical vorticity field corresponding to the wave solution in Fig. 1a.
In the following two sections we examine the asymptotic solutions describing the displacement of isopycnal surfaces at large distances from the mountain and close to the critical level, and the wave drag, a force exerted on the mountain by the flow.
c. Asymptotic solutions
d. Surface pressure and mountain wave drag
In the remainder of this paper we compare numerical and analytical solutions in order to assess the realizability of the linear solutions as well as the impact of local nonlinearities and dissipation on the singularities predicted by the linear model.
3. Numerical model
a. Model description and experimental design
The numerical model used in this study has been described in Smolarkiewicz and Margolin (1997). It is representative of a class of nonhydrostatic atmospheric models (Clark 1977; Kapitza and Eppel 1992) that solve the anelastic equations of motion (Lipps 1990; Lipps and Hemler 1982) in nonorthogonal terrain-following coordinates (Gal-Chen and Somerville 1975). The distinctive aspect of our model is its numerical design, which incorporates two-time-level, either semi-Lagrangian (Smolarkiewicz and Pudykiewicz 1992) or Eulerian (Smolarkiewicz and Margolin 1993), nonoscillatory forward-in-time (NFT), at least second-order-accurate finite-difference approximations to, respectively, point-wise andtrajectory-wise integrals of the governing fluid dynamic equations. Also, the model admits two optional lattice structures: a co-located A grid, and a staggered B grid (section 4 in Smolarkiewicz and Margolin 1994). All results reported in this paper have been generated with the semi-Lagrangian, A-grid variant of the model; however, selected experiments were repeated in other model configurations as well, to verify the overall accuracy of numerical results.
The numerically simulated problem is that posed in section 2. The mean velocity profile and obstacle shape are specified by (7) and (15), respectively. Given all the assumptions adopted, the parameter space of this problem is spanned by only two parameters: the nondimensional mountain height
The physical domain, 20a × 20a × 3zc, with the mountain centered in z = 0 plane, is covered with 81 × 81 × 91 grid points. Note that the actual vertical grid increment depends on the height of the critical level itself. In the vertical, the upper third of the domain is occupied by a gravity wave absorber (cf. Smolarkiewicz and Margolin 1997), while the critical level is placed in the middle of the remaining portion of the domain. Periodic boundaries are assumed in the y direction, whereas open boundaries in x are simulated using 10Δx-wide lateral absorbers (Davies 1983; Kosloff and Kosloff 1986). The model was initialized by elevating the mountain gradually over the first T0 = t0U0/a = 8 time units. Since the model formulation admits a generalized time-dependent “terrain-following” coordinate transformation (Prusa et al. 1996), this initialization is physically realizable (i.e., free of zeroth-order errors due to neglect of time-dependency of the geometry in the governing fluid equations). All experiments were continued until T = tU0/a = 18—this was sufficient to reach the steady state (if such existed)—and selected experiments were carried out to T = 30. The integration time step was chosen such that the Lipschitz number ∥ Δt∂v/∂x ∥, a semi-Lagrangian counterpart of the Courant number familiar from Eulerian computations, was always less than 0.5 [see Smolarkiewicz and Pudykiewicz (1992) for a discussion].
The accuracy of the model in reproducinglinear hydrostatic steady-state solutions for flows past an axisymmetric obstacle has been tested using the reference problem of a constant mean wind and stability (Smith 1980, 1988). The experiments were conducted for
b. Results and discussion
In order to assess realizability of the linear solutions, that is, to establish a region of the (
Figure 8 is a regime diagram of the (
Even well within the linear regime, the analytic and numerical solutions differ considerably in the vicinity of the critical level. Figure 13 shows the isentropes in the central x–z plane for the LS2 together with the corresponding linear result. Respective regularity and singularity of the numerical and analytic results are apparent. Guided by the 2D viscous linear theory of Hazel (1967), we anticipate that the difference between the two solutions is primarily due to the implicit viscosity of the numerical schemes becoming nonnegligible in the vicinity of the critical level. There, our linear theory predicts sharp gradients of the primary variables, and the nonoscillatory finite-difference schemes employed in the model activate minimized
Experiments LS3-LS5 (Ri = 1;
4. Summary
Using linear theory and numerical simulations, we have examined small-amplitude 3D gravity waves in a vertically sheared, unidirectional flow past an isolated axisymmetric mountain. We have assumed linear ambient wind profiles and constant environmental stability. This is perhaps the simplest scenario that provides a critical level for all wave components at the height where the mean flow vanishes. Our linear theory, formally valid for Richardson numbers greater than one quarter, is hydrostatic, Boussinesq, irrotational, and inviscid. In order to verify realizability of the analytical results, we have conducted a series of experiments with a numerical model suitable for simulating natural stratified flows.
Below the critical level, the linear theory predicts 3D wave pattern similar, in some respect, to that characteristic of a constant mean flow. The asymptotic far-field solutions yield waves confined to paraboloidal envelopes [Eq. (21)] that widen quickly with height. As a result, the wave fronts become normal to the mean flow as z ↗ zc (rather than as z → ∞, a behavior familiar from the constant mean flow case). At z = zc the linear solutions are singular; the vertical wavenumber, isopycnal displacements, and horizontal velocity become infinite. Above the critical level, the waves are strongly attenuated, as in the equivalent 2D problem. In 3D, however, the attenuation factor depends on the horizontal wavenumber. Modes with wave fronts parallel to the mean flow are filtered more effectively than those with the fronts perpendicular to the mean flow, resulting in a solution above the critical level reminiscent of the flow past a ridge. Similar to the equivalent 2D problem, mountain wave drag [Eq. (29)] decreaseswiththe Richardson number but, in contrast to the 2D case, it does not vanish as Ri d⃗ ¼. Zero drag in 2D flows is a consequence of a fore-and-aft symmetry of the linear solutions at strong shears. In 3D, this symmetry is maintained only for the modes with wave fronts normal to the mean flow, whereas all other modes induce asymmetric pressure perturbations leading to a nonzero drag. The linear theory prediction of drag is the same for positively and negatively sheared flows.
The numerical simulations with a nonhydrostatic nonlinear model confirm the theoretical predictions for a range of
Acknowledgments
We thank Andrew Crook, Teddie Keller, Robert Sharman, Glenn Shutts, and Ronald Smith for their personal reviews of the manuscript and illuminating discussions. This work originated during the lead author’s graduate studies at Yale University and continued during her fellowship with the Advanced Study Program at NCAR. The work at Yale was supported in part by the National Science Foundation, Division of Atmospheric Science, Grant ATM-8914138.
REFERENCES
Bacmeister, J. T., and R. T. Pierrehumbert, 1988: On high-drag states of nonlinear stratified flow over an obstacle. J. Atmos. Sci.,45, 63–80.
Baines, P. G., 1995: Topographic Effects in Stratified Flows. Cambridge University Press, 482 pp.
Booker, J. R., and F. P. Bretherton, 1967: The critical layer for internal gravity waves in a shear flow. J. Fluid Mech.,27, 513–539.
Bretherton, F. P., 1966: The propagation of groups of internal gravity waves in a shear flow. Quart. J. Roy. Meteor. Soc.,92, 466–480.
Broad, A. S., 1995: Linear theory of momentum fluxes in 3-D flows with turning of the mean wind with height. Quart. J. Roy. Meteor. Soc.,121, 1891–1902.
Case, K. M., 1960: Stability of an idealized atmosphere. I. Discussion of results. Phys. Fluids,3, 149–154.
Clark, T. L., 1977: A small-scale dynamic model using a terrain-following coordinate transformation. J. Comput. Phys.,24, 186–215.
——, and W. R. Peltier, 1977: On the evolution and stability of finite-amplitude mountain waves. J. Atmos. Sci.,34, 1715–1730.
Crook, N. A., T. L. Clark, and M. W. Moncrieff, 1990: The Denver cyclone. Part I: Generation in low Froude number flow. J. Atmos. Sci.,47, 2725–2741.
Davies, H. C., 1983: Limitations of some common lateral boundary schemes used in regional NWP models. Mon. Wea. Rev.,111, 1002–1012.
Drazin, P. G., and W. H. Reid, 1971: HydrodynamicStability. Cambridge University Press, 527 pp.
Durran, D. R., 1992: Two-layer solutions to Long’s equation for vertically propagating mountain waves: How good is linear theory? Quart. J. Roy. Meteor. Soc.,118, 415–433.
——, and J. B. Klemp, 1987: Another look at downslope windstorms. Part II: Nonlinear amplification beneath wave-overturning layers. J. Atmos. Sci.,44, 3402–3412.
Eliassen, A., and E. Palm, 1960: On the transfer of energy in stationary mountain waves. Geofys. Publ.,22, 1–23.
Ertel, H., 1942: Ein neuer hydrodynamischer Wirbesatz. Meteor. Z.,59, 277–281.
Fritts, D. C., J. R. Isler, and O. Andreassen, 1994: Gravity-wave breaking in two and three dimensions. Part II: Three-dimensional evolution and instability structure. J. Geophys. Res.,99, 8109–8123.
Gal-Chen, T., and R. C. J. Somerville, 1975: On the use of a coordinate transformation for the solution of the Navier–Stokes equations. J. Comput. Phys.,17, 209–228.
Gradshteyn, I. S., and I. M. Ryzhik, 1980: Table of Integrals, Series, and Products. Academic Press, 1160 pp.
Grubis̄ić, V., 1995: Mountain waves and mountain wakes in stratified airflows past three-dimensional obstacles. Ph.D. dissertation, Yale University, 163 pp. [Available from University Microfilm, 300 North Zeeb Rd., Ann Arbor, MI 48106.].
Hazel, P., 1967: The effect of viscosity and heat conduction on internal gravity waves at a critical level. J. Fluid Mech.,30, 775–783.
Hunt, J. C. R., and W. H. Snyder, 1980: Experiments on stably and neutrally stratified flow over a model three-dimensional hill. J. Fluid Mech.,96, 671–704.
Kapitza, H., and D. P. Eppel, 1992: The non-hydrostatic mesoscale model GESIMA. Part I: Dynamical equations and tests. Beitr. Phys. Atmos.,65, 129–146.
Keller, T. L., 1994: Implications of the hydrostatic assumption on atmospheric gravity waves. J. Atmos. Sci.,51, 1915–1929.
Kelly, R. E., 1977: Linear and nonlinear critical layer effects on wave propagation. Geofluidynamical Wave Mathematics, National Science Foundation Regional Conf. on Mathematics, Applied Mathematics Group, University of Washington, 194–210.
Kosloff, R., and D. Kosloff, 1986: Absorbing boundaries for wave propagation problems. J. Comput. Phys.,63, 363–376.
Kuo, H. L., 1963: Perturbations of plane Couette flow in stratfied fluid and origin of cloud streets. Phys. Fluids,6, 195–211.
Lighthill, J. M., 1980: Waves in Fluids. Cambridge University Press, 504 pp.
Lilly, D., and E. J. Zipser, 1972: The front range windstorm of 11 January 1972—A meteorological narrative. Weatherwise,25, 56–63.
Lipps, F. B., 1990: On the anelastic approximation for deep convection. J. Atmos. Sci.,47, 1794–1798.
——, and R. S. Hemler, 1982: A scale analysis of deep moist convection and some related numerical calculations. J. Atmos. Sci.,39, 2192–2210.
Maslowe, S. A., 1986: Critical layers in shear flows. Annu. Rev. Fluid Mech.,18, 405–432.
McIntyre, M. E., and W. A.Norton, 1990: Dissipative wave-mean interactions and the transport of vorticity and potential vorticity. J. Fluid Mech.,212, 403–435.
Miles, J. W., 1961: On the stability of heterogeneous shear flows. J. Fluid Mech.,10, 496–508.
Miranda, P. M., and I. N. James, 1992: Non-linear three-dimensional effects on the wave drag: Splitting flow and breaking waves. Quart. J. Roy. Meteor. Soc.,118, 1057–1081.
Nappo, C. J., and G. Chimonas, 1992: Wave exchange between the ground surface and a boundary-layer critical level. J. Atmos. Sci.,49, 1075–1091.
Peltier, W. R., and T. L. Clark, 1979: The evolution and stability of finite-amplitude mountain waves. Part II: Surface wave drag and severe downslope windstorms. J. Atmos. Sci.,36, 1498–1529.
Phillips, D. S., 1984: Analytical surface pressure and drag for linear hydrostatic flow over three-dimensional elliptical mountains. J. Atmos. Sci.,41, 1073–1084.
Prusa, J. M., P. K. Smolarkiewicz, and R. R. Garcia, 1996: On the propagation and breaking at high altitudes of gravity waves excited by tropospheric forcing. J. Atmos. Sci.,53, 2186–2216.
Reisner, J. M., and P. K. Smolarkiewicz, 1994: Thermally forced low Froude number flow past three-dimensional obstacles. J. Atmos. Sci.,51, 117–133.
Schär, C., and D. R. Durran, 1997: Vortex formation and vortex shedding in continuously stratified flows past isolated topography. J. Atmos. Sci.,54, 534–554.
Shutts, G., 1995: Gravity-wave drag parametrization over complex terrain: The effect of critical-level absorption in directional wind-shear. Quart. J. Roy. Meteor. Soc.,121, 1005–1021.
Smith, R. B., 1980: Linear theory of stratified hydrostatic flow past an isolated obstacle. Tellus,32, 348–364.
——, 1985: On severe downslope winds. J. Atmos. Sci.,42, 2597–2603.
——, 1986: Further development of a theory of lee cyclogenesis. J. Atmos. Sci.,43, 1582–1602.
——, 1988: Linear theory of hydrostatic flow over an isolated mountain in isosteric coordinates. J. Atmos. Sci.,45, 3889–3896.
——, 1989: Comment on “Low Froude number flow past three-dimensional obstacles. Part I: Baroclinically generated lee vortices.” J. Atmos. Sci.,46, 3611–3613.
——, and S. Grønås, 1993: Stagnation points and bifurcation in 3-D mountain airflow. Tellus,45A, 28–43.
Smolarkiewicz, P. K., and R. Rotunno, 1989: Low Froude number flow past three-dimensional obstacles. Part I: Baroclinically generated lee vortices. J. Atmos. Sci.,46, 1154–1164.
——, and ——, 1990: Low Froude number flow past three-dimensional obstacles. Part II: Upwind flow reversal zone. J. Atmos. Sci.,47, 1498–1511.
——, and J. A. Pudykiewicz, 1992: A class of semi-Lagrangian approximations for fluids. J. Atmos. Sci.,49, 2082–2096.
——, and L. G. Margolin, 1993: On forward-in-time differencing for fluids: Extension to curvilinear framework. Mon. Wea.Rev.,121, 1847–1859.
——, and ——, 1994: On forward-in-time differencing for fluids: Variational solver for elliptic problems in atmospheric flows. Appl. Math. Comput. Sci.,4, 527–551.
——, and ——, 1997: On forward-in-time differencing for fluids: An Eulerian/semi-Lagrangian nonhydrostatic model for stratified flows. Atmos.–Ocean,34, in press.
Winters, K. B., and E. A. D’Asaro, 1989: Two-dimensional instability of finite amplitude internal gravity wave packets near a critical level. J. Geophys. Res.,94, 12 709–12 719.
——, and J. J. Riley, 1992: Instability of internal waves near a critical level. Dyn. Atmos. Oceans,16, 249–278.
Worthington, R. M., and L. Thomas, 1996: Radar measurements of critical-layer absorption in mountain waves. Quart. J. Roy. Meteor. Soc.,122, 1263–1282.
Wurtele, M. G., A. Datta, and R. D. Sharman, 1996: The propagation of gravity–inertia waves and lee waves under a critical level. J. Atmos. Sci.,53, 1505–1523.
APPENDIX A
Asymptotic Solutions: Further Details
APPENDIX B
Mountain Wave Drag: Further Details
FFT representations of the displacement fields of the isopycnals with the undisturbed heights zc (1 ± 0.15) below (a) and above (b) the critical level, respectively. The mean flow (with Ri = 1 and
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
FFT representations of isopycnal surfaces in the central x–z plane for the same flow as in Fig. 1.
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
FFT representations of the vertical vorticity field(× 10−4 s−1) for the wave solution in Fig. 1a.
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Asymptotic solution (20b) corresponding to that shown in Fig. 1a. The thick solid line represents the wave envelope defined by (21).
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Asymptotic wave envelopes for mean flows with a linear negative shear (Ri = 1,
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Linear prediction for the pressure perturbation at z = 0 in a mean flow with
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Fig. 7.The mountain wave drag (normalized by D0 = π/4 ρ0 NU0a
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Regime diagram of the critical-level flow past an axisymmetric obstacle for linearly sheared ambient wind. The parameter space is spanned by the nondimensional mountain height
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
The history of drag in the small-amplitude critical-level numerical experiments with linear negative shear. Numbers on the curves correspond to the experiment labels in Table 1.
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
The surface pressure perturbation for the numerical experiments: (a) LS1 (Ri = 9) and (b) LS7 (Ri = 0.49). The display convention is the same as in Fig. 6.
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
The perturbation u-velocity component in the central x–z plane for the numerical experiments LS1 (a) and LS7 (b). The vectors represent the total velocity.
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Profiles of the normalized vertical flux of the horizontal momentum 〈uw〉 from numerical experiment LS7.
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Isentropic surfaces in the central x–z plane for (a) numerical experiment LS2 (the linear regime, Ri = 1.0,
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Vertical vorticity (×10−4 s−1) from numerical experiment LS2 (Ri = 1.0,
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Normalized potential vorticity (×10−5 s−1) from numerical experiment LS2 (Ri = 1.0,
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Numerical solutions from experiments LS3–LS5 (the nonlinear regime): (a) isentropes in the central x–z plane from the experiment LS3 (Ri = 1.0,
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Normalized potential vorticity (×10−4 s−1) from numerical experiment LS5 (Ri = 1.0,
Citation: Journal of the Atmospheric Sciences 54, 15; 10.1175/1520-0469(1997)054<1943:TEOCLO>2.0.CO;2
Parameters of the selected critical-levelnumerical experiments LS1–LS9. The second and third columns contain values of the Richardson number and nondimensional mountain height, respectively. The fourth column displays the total integration time normalized by the advective time scale a/U0 based on the mountain half-width and ambient wind at the surface. The fifth and sixth columns contain, respectively, analytic and numerical values of the mountain wave drag normalized by the hydrostatic drag for uniform ambient flow D0 = π/4ρ0 N U0 a h20.
For a linear shear and unbounded domain, such flows are stable actually for all R̄i ≡ Ri(κ/k)2 > 0 (Case 1960; Kuo 1963). For 0 < R̄i < ¼, however, the stability is only asymptotic with the amplitude of perturbations decaying as a fractional power of time, unlike the exponential decay for R̄i > ¼.
In the zero shear limit, c → ∞, r̂c → r̂,
The pressure field for the positively sheared mean flow can be obtained from Figs. 6a, b by rotating the mean flow by π and reversing the signs of the pressure anomalies.
Both (30) and (31) can be obtained directly by evaluating (28) or alternately from (29) using power series expansions of K and E (Grubis̄ić 1995). The latter approach provides an explicit form of D(Ri) in the vicinity of the limits, apparent in Fig. 7.
For Ri < 0.5 (experiments LS7 to LS9), an accurate steady state is difficult to define as the drag oscillates slowly, with amplitude less than 4%, around the values listed in Table 1.
Radical departures from the linear theory for small mountain heights are not limited to critical-level flows. For a model atmosphere with uniform wind and a two-layer stability profile, Durran (1992) found a similar behavior for