Upstream-Propagating Wave Modes in Moist and Dry Flow over Topography

Similar documents
ESCI 343 Atmospheric Dynamics II Lesson 10 - Topographic Waves

Downslope Wind Storms

Gravity waves in stable atmospheric boundary layers

Effects of Moist Froude Number and CAPE on a Conditionally Unstable Flow over a Mesoscale Mountain Ridge

Large-amplitude internal wave generation in the lee of step-shaped topography

Super-parameterization of boundary layer roll vortices in tropical cyclone models

DUE TO EXTERNAL FORCES

The total precipitation (P) is determined by the average rainfall rate (R) and the duration (D),

Mountain Forced Flows

Gravity Waves in Shear and

The impacts of explicitly simulated gravity waves on large-scale circulation in the

Thorsten Mauritsen *, Gunilla Svensson Stockholm University, Stockholm, Sweden

An experimental study of internal wave generation through evanescent regions

Modeling of Mountain Waves in T-REX

Xiaoli Guo Larsén,* Søren Larsen and Andrea N. Hahmann Risø National Laboratory for Sustainable Energy, Roskilde, Denmark

10.6 The Dynamics of Drainage Flows Developed on a Low Angle Slope in a Large Valley Sharon Zhong 1 and C. David Whiteman 2

Lecture 7. More on BL wind profiles and turbulent eddy structures. In this lecture

Estimating Topographic Blocking Using a Froude Number When the Static Stability Is Nonuniform

Goal: Develop quantitative understanding of ENSO genesis, evolution, and impacts

Lee Waves and Mountain Waves

2.4. Applications of Boundary Layer Meteorology

WAVE BREAKING OVER LOCAL TOPOGRAPHY DURING THE MAP IOP 15 MISTRAL EVENT: OBSERVATIONS AND HIGH-RESOLUTION NUMERICAL SIMULATIONS

13.1 EFFECTS OF DIABATIC COOLING ON THE FORMATION OF CONVECTIVE SYSTEMS UPSTREAM OF THE APENNINES DURING MAP IOP-8

VI. Static Stability. Consider a parcel of unsaturated air. Assume the actual lapse rate is less than the dry adiabatic lapse rate: Γ < Γ d

Aerosol-cloud interactions and the effects on orographic precipitation

Selected Papers on Downslope Windstorms

Forest Winds in Complex Terrain

Generation of mountain wave-induced mean flows and turbulence near the tropopause

Gravity wave breaking, secondary wave generation, and mixing above deep convection in a three-dimensional cloud model

Chapter 4. Convec.on Adiaba.c lapse rate

A Theory for Strong Long-Lived Squall Lines Revisited

Precipitation over concave terrain

Atmospheric Waves James Cayer, Wesley Rondinelli, Kayla Schuster. Abstract

3.3 USING A SIMPLE PARCEL MODEL TO INVESTIGATE THE HAINES INDEX

The Dissipation of Trapped Lee Waves. Part I: Leakage of Inviscid Waves into the Stratosphere

Modelling atmospheric stability with CFD: The importance of tall profiles

A R e R v e iew e w on o n th t e h e Us U e s s e s of o Clou o d u - (S ( y S s y t s e t m e )-Re R sol o ving n Mod o e d ls Jeff Duda

Influence of rounding corners on unsteady flow and heat transfer around a square cylinder

A numerical study of moist stratified flow regimes over isolated topography

A new theory for downslope windstorms and trapped lee wave François Lott Lab. Météorologie Dynamique, Ecole Normale Supérieure, Paris

Wave Motion. interference destructive interferecne constructive interference in phase. out of phase standing wave antinodes resonant frequencies

Numerical Approach on the Mechanism of Precipitation-Topography Relationship in Mountainous Complex Terrain

Analysis of NAM Forecast Wind Shear for Dissipation of Mesoscale Convective Systems

Chapter 2. Turbulence and the Planetary Boundary Layer

A flow regime diagram for forecasting lee waves, rotors and downslope winds

Proceedings of Meetings on Acoustics

Gravity waves and bores. Material kindly provided by Dr. Steven Koch GSD NOAA (Boulder, CO)

Atmospheric Factors Governing Banded Orographic Convection

Topographic Effects on the Tropical Land and Sea Breeze

ISOLATION OF NON-HYDROSTATIC REGIONS WITHIN A BASIN

3. Observed initial growth of short waves from radar measurements in tanks (Larson and Wright, 1975). The dependence of the exponential amplification

P2.17 OBSERVATIONS OF STRONG MOUNTAIN WAVES IN THE LEE OF THE MEDICINE BOW MOUNTAINS OF SOUTHEAST WYOMING

Effect of Orography on Land and Ocean Surface Temperature

Abrupt marine boundary layer changes revealed by airborne in situ and lidar measurements

Vertical Motion and Atmospheric Stability

INTRODUCTION * Corresponding author address: Michael Tjernström, Stockholm University, Department of Meteorology, SE-

Idealized Numerical Modeling of a Land/Sea Breeze

CFD Simulation and Experimental Validation of a Diaphragm Pressure Wave Generator

AIRFLOW GENERATION IN A TUNNEL USING A SACCARDO VENTILATION SYSTEM AGAINST THE BUOYANCY EFFECT PRODUCED BY A FIRE

Mountain Waves, Downslope Winds, and Low-Level Blocking Forced by a Midlatitude Cyclone Encountering an Isolated Ridge

SUPPLEMENTARY INFORMATION

Flow over mountains: Coriolis force, transient troughs and three dimensionality

18 Flight Hazards over High Ground

Exploring wave-turbulence interaction through LES modeling

ASSESSMENT OF SEA BREEZE CHARACTERISTICS FROM SODAR ECHOGRAMS

Waves. harmonic wave wave equation one dimensional wave equation principle of wave fronts plane waves law of reflection

Flow and Mixing in the Liquid between Bubbles

Conditions for Offshore Wind Energy Use

Geophysical Fluid Dynamics of the Earth. Jeffrey B. Weiss University of Colorado, Boulder

The dryline is a mesoscale phenomena whose development and evaluation is strongly linked to the PBL.

RECTIFICATION OF THE MADDEN-JULIAN OSCILLATION INTO THE ENSO CYCLE

An Undular Bore and Gravity Waves Illustrated by Dramatic Time-Lapse Photography

Quantification of the Effects of Turbulence in Wind on the Flutter Stability of Suspension Bridges

Hong Kong International Airport North Runway South Runway

Atmospheric Rossby Waves in Fall 2011: Analysis of Zonal Wind Speed and 500hPa Heights in the Northern and Southern Hemispheres

Gravity waves above a convective boundary layer: A comparison between wind-profiler observations and numerical simulations

Scott Denning CSU CMMAP 1

Stratified Flow Separation in the Lee of the Knight Inlet Sill

Mesoscale Meteorology

Morning-Glory Disturbances and the Environment in which They Propagate

1.6 HIGH RESOLUTION SIMULATIONS OF THE ISLAND-INDUCED CIRCULATIONS FOR THE ISLAND OF HAWAII DURING HaRP. Yang Yang and Yi-Leng Chen*

PHSC 3033: Meteorology Stability

Orographic effects on a conditionally unstable flow over an idealized three-dimensional mesoscale mountain

Impact of Sea Breeze Fronts on Urban Heat Island & Air Quality in Texas

The Influence of Ocean Surface Waves on Offshore Wind Turbine Aerodynamics. Ali Al Sam

Study on Fire Plume in Large Spaces Using Ground Heating

Internal Tides and Solitary Waves in the Northern South China Sea: A Nonhydrostatic Numerical Investigation

Chapter 3 Atmospheric Thermodynamics

Investigation of Upstream Boundary Layer Influence on Mountain Wave Breaking and Lee Wave Rotors Using a Large-Eddy Simulation

Impact of Typhoons on the Western Pacific: Temporal and horizontal variability of SST cooling Annual Report, 2011 James F. Price

Lecture 7. The Indian monsoon: is it a gigantic land-sea breeze?

6.28 PREDICTION OF FOG EPISODES AT THE AIRPORT OF MADRID- BARAJAS USING DIFFERENT MODELING APPROACHES

THE EFFECT OF THE OCEAN EDDY ON TROPICAL CYCLONE INTENSITY

Self-Acceleration in the Parameterization of Orographic Gravity Wave Drag

10.2 ATMOSPHERIC ROTORS: AIRCRAFT IN SITU AND CLOUD RADAR MEASUREMENTS IN T-REX

Dynamics and variability of surface wind speed and divergence over mid-latitude ocean fronts

Supplementary Material for Satellite Measurements Reveal Persistent Small-Scale Features in Ocean Winds Fig. S1.

Effects of an Inclined Land Surface on the Land and Sea Breeze Circulation: A Numerical Experiment. By Tomb Asai

MODELING INDIAN OCEAN CIRCULATION: BAY OF BENGAL FRESH PLUME AND ARABIAN SEA MINI WARM POOL

INTERACTION BETWEEN WIND-DRIVEN AND BUOYANCY-DRIVEN NATURAL VENTILATION Bo Wang, Foster and Partners, London, UK

Transcription:

3060 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 69 Upstream-Propagating Wave Modes in Moist and Dry Flow over Topography TEDDIE L. KELLER, RICHARD ROTUNNO, MATTHIAS STEINER, AND ROBERT D. SHARMAN National Center for Atmospheric Research,* Boulder, Colorado (Manuscript received 4 January 2012, in final form 25 April 2012) ABSTRACT Previous studies have observed upstream-propagating modes in two-dimensional numerical simulations of idealized flow over topography with moist, nearly neutral conditions in the troposphere, topped by a stable stratosphere. The generation and propagation mechanisms for these modes were attributed to localized and dramatic changes in stability induced by the desaturation of the flow impinging on the mountain. In the present paper it is shown that these modes are transient upstream-propagating gravity waves, which are a fundamental feature of both moist and dry flow over topography of a two-layer troposphere stratosphere atmospheric profile impulsively started from rest. The mode selection and propagation speeds of these transient waves are highly dependent on the tropospheric stability, as well as the wind speed and tropopause depth. In the moist case these modes appear to propagate according to an effective static stability that is intermediate to the normal dry stability and the lower moist stability. Comparisons with the linear, timedependent, hydrostatic analytic solution show that these modes are similar to the transients observed in flow of a constant wind and stability layer over topography with a rigid upper boundary. 1. Introduction Orographic precipitation patterns can be significantly affected by terrain-induced flow patterns due to a variety of mechanisms, such as orographic uplift (Smith and Barstad 2004; Smith 1979, among others), windward flow blockage (e.g., Jiang 2003), and gravity wave modification of flow patterns aloft (Colle 2004, among others). The identification of moist, nearly neutral atmospheric conditions with alpine orographic rainfall events (e.g., Rotunno and Ferretti 2001, 2003) prompted Miglietta and Rotunno (2005) to conduct a series of idealized numerical simulations with constant wind and a two-layer atmospheric stability structure, that is, a saturated, nearly neutral troposphere capped by a stable stratosphere. The atmospheric profile used in their study is shown in Fig. 1. * The National Center for Atmospheric Research is sponsored by the National Science Foundation. Corresponding author address: Teddie L. Keller, National Center for Atmospheric Research, Research Applications Laboratory, P.O. Box 3000, Boulder, CO 80307 3000. E-mail: tkeller@ucar.edu A striking and unexpected feature appearing in several simulations of Miglietta and Rotunno (2005) was a region of desaturated air expanding upwind of the mountain in conjunction with an upstream-propagating wave mode (their Fig. 5, reproduced in Figs. 2a and 2b). The authors hypothesized that buoyancy forces induced by local changes in stability resulting from desaturation of the air as it moved over the mountain was the driving force in the upwind progression of this disturbance, as no such disturbance occurred with a very small amplitude mountain in which desaturation did not occur. Using similar background atmospheric flow conditions to Miglietta and Rotunno, Steiner et al. (2005) conducted a series of both moist and dry numerical simulations using the Weather Research and Forecasting Model (WRF). Although a windward expanding region of desaturated air was cited as a causal factor in the reduced upwind flow deceleration observed in some of the moist simulations, Steiner et al. did not explicitly investigate the upstream modes, and only long-time statistics were presented. Recently we analyzed the temporal evolution of the flow in these same simulations and found upstream-propagating modes in both moist and dry simulations in some, but not all, cases. An important aspect of the background atmospheric conditions in these simulations is the large stability jump DOI: 10.1175/JAS-D-12-06.1 Ó 2012 American Meteorological Society

OCTOBER 2012 K E L L E R E T A L. 3061 FIG. 1. Atmospheric structure used in the Miglietta and Rotunno (2005) idealized numerical simulations of moist nearly neutral flow over a 700-m-high and 10-km-wide Witch of Agnesi mountain (adapted from their Fig. 3); N m is the moist static stability. between the troposphere and stratosphere. It is well known that the stability jump at the tropopause can result in significant wave reflections for steady-state mountain waves in dry flow (e.g., Eliassen and Palm 1961; Klemp and Lilly 1975; Durran 1992; Barstad and Schuller 2011). The tropopause acting as a leaky lid can be important for lee waves in sheared flow (e.g., Keller 1994) and for propagating gravity waves generated by convection (Lane and Zhang 2011). As the troposphere 2 stratosphere stability difference becomes greater, less energy leaks through and the tropopause begins to act more like a rigid lid (e.g., Berkshire and Warren 1970). Stable layers functioning as reflective surfaces similar to rigid lids have also been noted in density current simulations (Xue 2002) and for elevated inversions in downslope windstorm simulations (e.g., Smith and Skyllingstad 2011). In fact, as will be shown in section 2, characteristics of the wave motion in Fig. 2a resemble the transient waves studied extensively for channel flow that is, a constant wind and stability layer flowing over an obstacle with a rigid upper boundary in place of the tropopause. These transient modes are generated as a consequence of the impulsive startup of the flow from rest. Similar modes are generated by a localized heat source representing convection in an atmosphere at rest (e.g., Nicholls et al. 1991; Mapes 1993; Lane and Reeder 2001; Lane and Zhang 2011). Transient propagating waves for stratified channel flow over topography have been studied extensively using towing tank experiments (e.g., Baines 1977, 1979; Castro and Snyder 1988), analytic theory (Janowitz 1981; McIntyre 1972, among others), and numerical modeling (e.g., Lamb 1994; Hanazaki 1989; Uchida and Ohya 2001). Nonlinear transcritical flow near resonance, when the incoming flow speed is approximately equal to the speed of an upstream-propagating mode, has been investigated numerically (e.g., Rottman et al. 1996; Hanazaki 1992, 1993) and using various forms of the Korteweg de Vries equation (Grimshaw and Yi 1991, among others). These studies demonstrate the ubiquitous occurrence of transient upstream-propagating modes in stratified channel flow over topography, and confirm that only the waves with an upstream propagation speed greater than the background flow speed actually appear upwind of the obstacle. Here we show that transient modes similar to those appearing in constant wind and stability flow over terrain with a rigid upper boundary also dominate in both moist and dry simulations of idealized two-layer troposphere stratosphere flow. For certain parameter ranges the upstream waves propagate slowly enough to induce vertical velocity perturbations upstream of the mountain that may last for many hours. For moist flow upstream-propagating modes alter the relative humidity of the air impinging on the mountain and, if the amplitude of the transient wave is large enough, these modes may modify both the intensity and location of terrain-induced rainfall rates. And, the vertical velocity perturbations associated with the rapidly moving upstream modes could trigger convection far upwind of the topography, just as gravity wave modes with similar vertical structure generated by convective heating can induce new convection (e.g., Mapes 1993; Lane and Zhang 2011). Thus it is important to understand the mechanisms associated with the generation and maintenance of these modes. Furthermore, it is important to determine if these upstream modes are a common feature and, if so, to what extent they may be contributing to both observed and numerically simulated orographic precipitation. The paper is organized as follows. Fundamental features of the analytic solution for the linear, timedependent, hydrostatic rigid-lid case are discussed in section 2. Dry numerical simulations of these upstreampropagating modes for both the single-layer and a twolayer troposphere stratosphere atmospheric profile using a gravity wave model are provided in section 3. The temporal evolution of the upstream-propagating modes in both moist and dry WRF simulations from Steiner et al. (2005) is discussed in section 4. Summary and conclusions are provided in section 5.

3062 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 69 FIG. 2. (a) Vertical velocity w and (b) cloud water content q c at 5 h from the idealized numerical simulations of moist nearly neutral flow over topography [reproduced from Miglietta and Rotunno (2005), Figs. 5a,d]. Contour interval in (a) is 0.5 m s 21. Shading in (b) is q c, 0.01 (white), 0.01, q c, 0.1 (light gray), 0.1, q c, 0.5 (medium gray), and q c. 0.5 g kg 21 (dark gray). 2. Linear transient analytic solution This section discusses fundamental features of the upstream modes in the two-dimensional, time-dependent, hydrostatic, linear analytic solution for constant wind and stability flow over topography in which the tropopause has been replaced by a rigid lid (channel flow). These modes will then be compared with the transient modes observed in the nonhydrostatic numerical simulations of idealized troposphere stratosphere flow in sections 3 and 4. The linear, time-dependent initial value problem for two-dimensional constant wind and stability flow in a channel over arbitrary topography was solved in integral form by Engevik (1971) using a Fourier transform in the horizontal (x) and a Laplace transform in time (t). Beginning with the Engevik Eq. (3.6), making the hydrostatic assumption, taking the inverse transform, and solving for the vertical velocity w yields h sin[pk(1 2 z/z w(x, z, t) 5 U t )] 0 x sin(pk) 2 U 0 K p å n51 sin(npz/z t ) n 1 n 2 K h(x 2 c 2 t) x 2 1 h(x 2 c 1 t), (1) n 1 K x where the background wind speed, U 0, is positive (i.e., flow is from left to right), h (x) is the topography profile, Z t the depth of the channel, N the Brunt Väisälä frequency, and the nondimensional parameter K5NZ t / pu 0 is a measure of the speed of transient mode one relative to the background wind. Equation (1) is valid for n 6¼ K. Similar solutions for the streamfunction can be found in Baines and Guest (1988) and for the vertical displacement in Rottman et al. (1996). The first term on the rhs of Eq. (1) corresponds to the steady-state solution, while the second and third terms are sums over the infinite number of transient modes n. For a given n there are two transient waves, one propagating downstream (i.e., to the right) and the other upstream (to the left), with speeds c 6 5 U 0 6 NZ t np 5 U 0 1 6 K n (2) and vertical wavelength l n 5 2Z t /n. The leftward moving modes propagating against the background flow emerge upwind of the mountain only when their propagation speed exceeds the ambient wind speed; that is, c 2, 0. Thus for a given K only n k modes actually appear upstream, where n k is the largest integer less than K. All other modes are swept downstream.

OCTOBER 2012 K E L L E R E T A L. 3063 Note that the transient modes in these hydrostatic solutions translate the initial disturbance away from the mountain with no change in amplitude; however, in the nonhydrostatic solution the wave amplitude falls off as t 21/2 as t / [Engevik (1971), see Eq. (4.5)] due to dispersive effects. There are several important points to note about Eq.(1).First,foragivenmoden the amplitude of the upstream, leftward moving wave (second term) is larger than the corresponding downstream, rightward-propagating mode (third term). Second, this linear solution is invalid for K [ n, which occurs when the incoming flow speed is equal to the linear long wave phase (and group) velocity of mode n. In this resonant case the amplitude of the upstream-propagating mode becomes infinite [second term in Eq. (1)]. A number of papers have focused on nonlinear effects when K n, that is, near resonance, using nonlinear numerical modeling (e.g., Rottman et al. 1996) or solving the Korteweg de Vries equation (Grimshaw and Yi 1991, among others), but these effects are not considered here. The temporal evolution of the vertical velocity upstream of the mountain [from Eq. (1)] for K 5 0.85, 1.27, and 2.55, which have zero, one, and two modes, respectively, penetrating upwind of the topography against thebackgroundflow,isshowninfig.3.hereboththe channel depth and atmospheric stability are fixed, with Z t 5 10 km and N 5 0.004 s 21, and the background wind speed is varied to change K. The mountain profile is a Witch of Agnesi, h(x) 5 h 0 /[1 1 (x/a) 2 ], with a mountain height h 0 of 10 m and half-width, a, of 20 km. For reference, the steady-state solution for these cases is shown in Fig. 4. Because the vertical wavelength of the steady-state part of the hydrostatic solution is also a function of K [the first term in Eq. (1)], the steady-state wave pattern above the mountain crest is indicative of the flow regime; for example, for 1, K, 2 (Fig. 4b) the hydrostatic vertical wavelength over the topography is between 0.5 and 1.0 Z t. Note that for K 5 0.85 all upstream-propagating modes are swept downstream (Figs. 3a c) and the vertical velocity pattern upwind of the mountain peak differs little from the steady state (Fig. 4a) throughout the entire time period. In contrast, as K increases above one (Figs. 3d i), the vertical velocity pattern upwind deviates considerably from the steady-state solution (Figs. 4b,c) owing to relatively slow-moving transient modes. In the flow regime where 1, K, 2 only one transient mode has a leftward propagation speed exceeding the background wind speed and is able to penetrate upwind (K 5 1.27; Figs. 3d f). This mode has a vertical wavelength twice the channel depth, with a cell of upward vertical velocity followed by a cell of downward vertical velocity filling the depth of the troposphere upwind of the topography. Also, in this regime the steady-state hydrostatic gravity wave directly over the mountain has a vertical wavelength between 0.5 and 1.0 times the channel depth. This is similar to the wave pattern seen in Miglietta and Rotunno (2005) (reproduced in Fig. 2a). Note, however, that using their background environmental moist stability to calculate K results in a value of K, 1; that is, all modes should have been swept downstream. This is discussed further in section 4. For K. 2 the first mode moves rapidly upstream (Fig. 3g) and transient mode 2, with positive vertical velocity at lower levels and negative w aloft at the leading edge, dominates directly upwind of the mountain at 8 and 12 h (Figs. 3h,i). Leftward-moving mode 3, which has a propagation speed less than the background wind speed, is apparent downwind of the mountain in Figs. 3h and 3i. A plot of the nondimensional wave speeds c 2 /U 0 versus K for leftward-propagating modes traveling faster than the background wind that is, c 2, 0 is shown in Fig. 5. The black lines are comprised of data points representing transient wave speeds calculated from the analytic solution. This clearly shows that for K, 1all leftward-moving wave modes propagate with speeds less than the background wind and hence are swept downstream. As K increases, the number of modes able to penetrate upwind increases and their nondimensional propagation speeds increase as well. Also plotted are points corresponding to the rigid lid and idealized twolayer troposphere stratosphere dry numerical simulations discussed in section 3. Note that the mode number and propagation speed of the transients for the numerical simulations correspond closely with the linear analytic rigid-lid solution. 3. Dry numerical simulations a. Model description The time-dependent, nonhydrostatic gravity wave model used for the dry simulations discussed in this section is described in detail in Sharman and Wurtele (1983). This model integrates the two-dimensional Boussinesq equations of motion, and the topography is introduced by prescribing the vertical velocity at the lower boundary that is, w 5 U 0 dh/dx at z 5 0 rather than using terrainfollowing coordinates. Thus the lower boundary condition is strictly linear, making it consistent with the analytic solution [Eq. (1)], and the mountain can be raised slowly, eliminating effects of a shock start. Sensitivity studies varying the number of time steps over which the mountain was raised were conducted, and it was found that slowly increasing the mountain height had only a slight

3064 JOURNAL OF THE ATMOSPHERIC SCIENCES VOLUME 69 FIG. 3. Temporal evolution of the vertical velocity [Eq. (1)] as a function of K for constant wind and stability flow over topography with a rigid lid at 4, 8, and 12 h. Domain height is 10 km and the Brunt Va isa la frequency 0.004 s21. Background wind speed U0 is (a) (c) 15, (d) (f) 10, and (g) (i) 5 m s21, yielding K 5 0.85, 1.27, and 2.55. The mountain profile is a Witch of Agnesi with a half-width 20 km and height 10 m. Contour interval is 0.05 m s21, the color scale is given at the bottom, and vertical velocity is multiplied by 40 in (a) (c), 100 in (d) (f), and 200 in (g) (i). effect on the amplitude of the transient mode. Thus, the mountain was impulsively raised at t 5 0 in the simulations presented here. The linear lower boundary formulation in this model has the added advantage that the channel depth between the lower boundary and the tropopause does not change due to a finite mountain height, making direct comparison with analytic theory more consistent. As in section 2, the mountain profile is

OCTOBER 2012 K E L L E R E T A L. 3065 FIG. 4. As in Fig. 3, but only the steady-state part of the analytic solution [Eq. (1)]. a Witch of Agnesi with a half-width of 20 km and a height of 10 m for the linear simulations. A useful feature of this model is the ability to turn off the nonlinear terms in the equations of motion, that is, to use advection terms in linearized form. This makes it an ideal choice for comparing numerical simulations with linear analytic solutions, as well as comparing linear and nonlinear simulations. b. Background atmospheric profile The numerical model was used to simulate both a finitedepth tropospheric layer capped by a rigid lid, to compare with the analytic solution, and two-layer troposphere stratosphere flow. The background atmospheric flow for the troposphere stratosphere simulations follows that used in Miglietta and Rotunno (2005) and Steiner et al. (2005) that is, a two-layer stability profile with constant wind U 0 (Fig. 1) except the flow is dry. The background wind speed U 0, tropopause depth Z t,and tropospheric stability N 1 were varied, with particular emphasis on values yielding a similar dynamical regime to that in Fig. 2a, which has one transient mode penetrating upwind; that is, K 5 NZ t /pu 0 primarily between 1 and 2. c. Transient wave modes A comparison of transient modes in the linear singlelayer hydrostatic analytic solution and the nonhydrostatic numerical simulations for both the troposphere-only and troposphere stratosphere atmospheric profile at 5, 10, and 15 h is shown in Fig. 6. Here U 0 5 10 m s 21, N 1 5 0.004 s 21, N 2 5 0.0228 s 21, and the tropopause depth/ rigid-lid height is 9 km, yielding K 5 1.15. In Fig. 6 transient mode 1, with a single cell of upward vertical velocity filling the depth of the troposphere, moves upstream at about the same speed in the analytic solution and in both simulations. This slow-moving mode, with a vertical wavelength twice the tropopause depth and a maximum velocity in the midtroposphere, creates a broad region of uplift upstream of the mountain which lasts the entire 15 h of simulation. The region of maximum uplift of the slowly upstream-propagating vertical velocity cell is located at about x 5240, 265, and 290 km upstream of the mountain peak at 5, 10, and 15 h, respectively. Closer to the mountain a cell of downward velocity can be seen expanding upstream with time. Nonhydrostatic dispersion, which is included in the numerical model but not the hydrostatic analytic solution, results in a slight decrease in amplitude of the transient mode as it propagates upwind in the rigid-lid simulations (Figs. 6d f). In addition, in the troposphere stratosphere FIG. 5. Nondimensional wave speed (c 2 /U 0 ) of modes traveling upstream faster than the background wind speed as a function of K. Black lines calculated from the linear analytic solution [Eq. (1)]. Symbols indicate wave speeds from the gravity wave numerical model for a linear, single-layer with rigid lid (open diamonds), and a linear (gray filled circles) and nonlinear (asterisk) troposphere stratosphere atmospheric profile.

3066 JOURNAL OF THE ATMOSPHERIC SCIENCES VOLUME 69 FIG. 6. Vertical velocity for K 5 1.15 at 5, 10, and 15 h for (a) (c) the linear, hydrostatic analytic rigid-lid solution [Eq. (1)], (d) (f) the linear nonhydrostatic numerical model with rigid lid at the tropopause, and (g) (i) the linear nonhydrostatic numerical model with the two-layer troposphere stratosphere atmospheric profile. Background wind speed is U0 5 10 m s21, tropopause depth is 9 km, and tropospheric stability is 0.004 s21. Contour interval is 0.05 m s21 and color scale is on the bottom. Vertical velocity is scaled by a factor of 50 in (a) (f) and 100 in (g) (i). simulation the amplitude of the upstream-propagating mode visibly decreases further as the wave moves upwind owing to leakage of energy into the stratosphere (Figs. 6g i). Leftward-propagating transient mode 2, which has a propagation speed less than the incoming wind speed, is clearly evident downstream at x 5 190 km at 5 h in the troposphere stratosphere simulation (Fig. 6g). This mode has vertical wavelength Zt and upward (downward) motion in the lower (upper) troposphere. The other transient modes have much smaller amplitudes and do not contribute significantly to the overall wave pattern in the region shown.

OCTOBER 2012 KELLER ET AL. 3067 FIG. 7. Vertical velocity as a function of K for the linear, nonhydrostatic simulations of troposphere stratosphere atmospheric profile over topography at 4, 8, and 12 h. Tropopause depth is 10 km and tropospheric Brunt Va isa la frequency 0.004 s21. Background wind speed U0 is (a) (c) 10 m s21 and (d) (f) 5 m s21, yielding K 5 1.27 and 2.55, respectively. Contour interval is 0.05 m s21, color scale is at the bottom, and vertical velocity is multiplied by 100 in (a) (c) and 200 in (d) (f). To further explore the behavior of the transient modes for the idealized troposphere stratosphere atmospheric profile, K was varied by altering both the tropopause depth and the background wind speed. As in the single-layer analytic solution, the speed of the upstreampropagating mode in the troposphere stratosphere simulations increases as K increases. In Fig. 7 the atmospheric stability is fixed (N1 5 0.004 s21, N2 5 0.0228 s21), the tropopause depth Zt 5 10 km, and U0 is 10 m s21 (Figs. 7a c) and 5 m s21 (Figs. 7d f), yielding K 5 1.27 and 2.55, respectively. The corresponding single-layer analytic solution for these K values was presented in Figs. 3d i. Although the atmospheric wind and stability in Figs. 7a c are the same as in Figs. 6g i, the minor increase in tropopause depth (from 9 to 10 km) results in a slight increase in K and a concomitant increase in speed of the transient waves. Now downward vertical velocity dominates the troposphere between 250 and 2100 km upwind of the mountain by 12 h (Fig. 7c), whereas in Figs. 6h and 6i upward vertical velocity persists in this region throughout the time period from 10 to 15 h. As K increases above 2 (Figs. 7d f) transient mode 1, with vertical wavelength 2 Zt, moves quickly upstream and the second leftward-moving mode, with vertical wavelength Zt, dominates just upwind of the mountain, as in the analytic solution for channel flow presented in section 2 (Figs. 3g i). This creates a persistent pattern of upward velocity in the lower troposphere and downward velocity in the upper troposphere upstream of the topography for many hours. Leftward-propagating mode 3, with vertical wavelength (2/3)Zt, is swept downstream and is clearly visible downwind at 8 and 12 h (Figs. 7e,f). Thus, the linear numerical simulations show that for this idealized two-layer troposphere stratosphere atmospheric profile the stability jump at the tropopause, acting like a leaky lid, selects the same transient modes in the troposphere as the rigid lid does in channel flow. And, as with channel flow, minor changes in atmospheric parameters that result in slight changes in K values can

3068 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 69 markedly alter the horizontal extent and persistence of the vertical velocity pattern upstream, which in turn could have substantial effects on the upstream relative humidity profile when moist air is impinging on the mountain. The speed of the transient modes moving upwind in the numerical simulations agrees well with linear analytic theory. As discussed in section 2, Fig. 5 shows a plot of nondimensional wave speed c 2 /U 0 versus K. Transient speeds calculated from the analytic solution comprise the black lines, and wave speeds measured in the linear numerical simulations for a single layer with a rigid lid and the troposphere stratosphere atmospheric profile are indicted by the open diamonds and gray filled circles, respectively. To determine how nonlinearities affect the transient wave speeds, further simulations were run with similar atmospheric conditions but increasing the mountain height and switching on the nonlinear advection terms in the model. The same modes dominate upstream, and propagate at speeds close to the speed predicted by linear theory, as indicated by the asterisk symbols in Fig. 5. Vertical velocity patterns from simulations using the nonlinear version of the gravity wave model and increasing the mountain height to 1 km for K 5 2.55, which has two transients penetrating upwind, are shown in Fig. 8. In these simulations tropospheric stability is 0.008 s 21, U 0 5 10 m s 21, and the tropopause depth is 10 km. Also shown in Fig. 8 is the effect of varying the stratospheric stability; the stratospheric stability is increased from 0.0114 s 21 in Figs. 8a and 8b, to 0.0228 s 21 in Figs. 8c and 8d, to 0.0456 s 21 in Figs. 8e and 8f. Note that the amplitude of the transient modes increases as the stability jump at the tropopause increases, as would be expected. d. Nonhydrostatic effects In contrast to the hydrostatic linear analytic solution presented in section 2, the numerical simulations, even with this gently sloped and broad 20-km half-width mountain, include nonhydrostatic effects that impact both the steady-state and transient wave solutions. Nonhydrostatic effects on the steady-state wave solution produce lee waves downstream of the topography in the numerical simulations (contrast Figs. 6b,c with Figs. 6e,f and Figs. 6h,i for 0, x,1100 km). Linear stationary lee waves have a horizontal wavelength determined by the vertical variations in wind and stability (Wurtele et al. 1987, 1996; Keller 1994, among others) and amplitude dependent on the mountain forcing. In our case the lee waves are trapped by the rigid lid in the single-layer simulations and partially trapped by the large stability jump at the tropopause in the two-layer simulations. The amplitude of the downstream lee wave is reduced due to leakage into the stratosphere in the troposphere stratosphere simulation. Note that transient modes swept downstream can constructively or destructively interfere with the stationary lee wave pattern, resulting in temporal variability in the standing wave pattern. Nonhydrostatic effects on the transient modes create a train of dispersive traveling waves. Since the magnitude of the group velocity is smaller than the phase velocity in the nonhydrostatic case, the wave pattern in the nonhydrostatic transient solution has an additional oscillatory behavior that moves more slowly upstream of the mountain, trailing behind the hydrostatic upwindpropagating disturbance (e.g., Hanazaki 1989). Pandya et al. (1993) have nicely illustrated these nonhydrostatic effects for thermally generated gravity waves in an atmosphere at rest. Note that the hydrostatic approximation is valid for a disturbance when the horizontal scale is much larger than the vertical scale (e.g., Lin 2007, p. 19). For the atmospheric structure investigated here the tropopause depth, which plays a key role in the wave development, is the appropriate vertical length scale. Using the mountain half-width for the horizontal scale, the flow response is approximately hydrostatic when the parameter a/z t is large, as noted in previous channel flow studies (e.g., Rottman et al. 1996). This is in contrast to an infinitely deep, constant wind and stability atmosphere where the vertical scale is U 0 /N, related to the vertical wavelength, and the parameter Na/U 0, a measure of the advective time scale compared to the period of a buoyancy oscillation (e.g., Smith 1979), is used to determine hydrostaticity. Various authors have used different cutoff values of Na/U 0 for the validity of the hydrostatic assumption; for example, in Lin and Wang (1996) Na/U 0. 7.5 and in Laprise and Peltier (1989) Na/U 0. 10. In all simulations discussed in this section Na/U 0 is greater than or equal to 8. Nonhydrostatic dispersion producing a train of transient waves is illustrated in Fig. 9 for rigid-lid numerical simulations. Here K is fixed at 1.5 and U 0 5 10 m s 21,so the one transient mode penetrating upwind propagates at the same speed according to hydrostatic theory in all three simulations. The mountain half-width is 20 km, and the layer depth is changed to yield an increasingly more nonhydrostatic solution; that is, a/z t 5 6.7, 3.3, and 1.7 in Figs. 9a c, respectively. Nonhydrostatic effects are particularly evident in Fig. 9c, where a train of waves can be seen propagating upwind, with alternating up and downdrafts. Note that the position of the maximum amplitude of the first cell advances slightly more slowly as nonhydrostatic effects become more dominant. Although the stability is also changing as Z t varies, in order to hold K fixed, Na/U 0 $ 8 in Fig. 9.

OCTOBER 2012 KELLER ET AL. FIG. 8. Vertical velocity at 1.5 and 5 h for troposphere stratosphere simulations using the nonlinear gravity wave model and varying stratospheric stability. Mountain height and half-width are 1 km and 20 km, respectively. Tropospheric K 5 2.55 (N1 5 0.008 s21, U0 5 10 m s21, and tropopause depth is 10 km). Stratospheric stability is (a),(b) 0.0114, (c),(d) 0.0228, and (e),(f) 0.0456 s21. Contour interval is 0.05 m s21. 3069

3070 JOURNAL OF THE ATMOSPHERIC SCIENCES VOLUME 69 Interestingly, in the regime K, 1, in which all leftward-moving transient modes have speeds less than the background wind and hence are swept downstream, there are no stationary nonhydrostatic lee waves possible (e.g., Lamb 1994). Now leftward-propagating mode one is moving the most slowly downstream and may dominate downwind of the mountain for hours. When nonhydrostatic effects are significant, the pattern produced by this dispersive packet of waves associated with this transient mode may resemble a lee wave train (Fig. 10). Since the nonhydrostatic waves travel more slowly against the flow than the hydrostatic mode, they are swept downwind more quickly and actually precede the hydrostatic component of the transient wave downstream. Thus, nonsteady lee wave patterns may be generated not only by mechanisms such as nonlinearities or unsteadiness in the background flow (e.g., Nance and Durran 1997, 1998; Ralph et al. 1997), but also by these slowly moving nonhydrostatic transient wave packets. In summary, for the two-layer dry simulations the stability jump at the tropopause acts as a leaky lid in the troposphere stratosphere simulations, selecting the same transient modes in the troposphere as the rigid lid in channel flow. The wave amplitude depends on the stability jump between troposphere and stratosphere, with less energy leakage for larger jumps. The same modes dominate in both linear and nonlinear numerical simulations. When nonhydrostatic effects become important, the transient mode becomes a traveling wave packet and, when these modes are swept downstream, they may resemble an unsteady lee wave. 4. Moist numerical simulations Steiner et al. (2005) conducted a suite of both moist and dry numerical simulations with background atmospheric flow conditions similar to those of Miglietta and Rotunno (2005) using the Weather Research and Forecasting model (Skamarock et al. 2005), version 1.3. However, in that study the upstream modes were not explicitly investigated and only long-time (30 h) statistics were presented. We analyzed the temporal evolution of FIG. 9. Nonhydrostatic effects on upstream-propagating transient wave mode 1 in linear numerical simulations of a single layer capped by a rigid lid for K 5 1.5 at 11 h; U0 is 10 m s21. Layer height and stability are (a) 3 km, 0.016 s21, (b) 6 km, 0.008 s21, and (c) 12 km, 0.004 s21. Thus a/zt is (a) 6.7, (b) 3.3, and (c) 1.7. Contour interval is 0.05 m s21, color scale is at the bottom, and vertical velocity is multiplied by 100.

OCTOBER 2012 K E L L E R E T A L. 3071 FIG. 10. Vertical velocity contours for a single-layer rigid-lid simulation showing the nonhydrostatic wave packet associated with leftward-moving transient mode 1 being swept downstream at (a) 20 and (b) 35 h. Tropopause depth Z t is 7 km, Brunt Väisälä frequency N 5 0.004 s 21, and wind speed U 0 is 10 m s 21, resulting in K 5 0.9. Contour interval is 0.05 m s 21, and vertical velocity is multiplied by 50. the flow in these simulations and found upstreampropagating modes in both moist and dry simulations in some, but not all, cases. Representative simulations of this two-layer flow over a 500-m-high Witch of Agnesi mountain with a half-width of 20 km are discussed here. Upstream-propagating modes in the dry troposphere stratosphere WRF simulations have the same characteristics as the modes discussed in section 3, with the number of upwind modes and relative propagation speeds increasing with increasing tropospheric K (Fig. 11). Here tropospheric stability N 1 5 0.004 s 21, the tropopause depth is 11 km, and the wind is decreased from 20 m s 21 in the top panel, to 10 m s 21 in the middle panel, and down to 5 m s 21 in the bottom panel, yielding a K of 0.7, 1.4, and 2.8, respectively. Thus there are zero, one, and two transient modes able to propagate upwind of the mountain in Figs. 11a,b, 11c,d, and 11e,f, respectively. Higher transient modes that are swept downstream are evident downwind of the mountain in Figs. 11c,d (mode 2) and Fig. 11f (mode 3). Thus the existence of these transient modes is model independent that is, despite differences in topography representation and model initialization procedures, traveling wave modes are a persistent feature in dry simulations for the two-layer troposphere stratosphere atmospheric profile. Similar transient modes were found in the moist-saturated troposphere stratosphere simulations. Figure 12 shows vertical velocity (contour lines) and relative humidity (color) at hours 4 and 8 from three moist flow simulations in which the moist tropospheric stability N m, calculated following Miglietta and Rotunno (2005), is 0.002 s 21 and the tropopause depth is 11 km. As in the dry simulations above, the background wind speed is decreased from 20 m s 21 in the top panel, to 10 m s 21 in the middle panel, and down to 5 m s 21 in the bottom panel, resulting in an increasing K m (5N m Z t /U 0 ) from top to bottom. Thus Figs. 11 and 12 are comparable in the sense that the flow regimes have zero, one, and two modes propagating upwind in panels (a) and (b), (c) and (d), and (e) and (f), respectively. In Figs. 12a and 12b all transient modes are swept downstream. There is no region of reduced relative humidity expanding upstream and the upward vertical velocity cell along the upwind slope of the mountain reaches steady state very quickly. The rapidly moving transient wave swept downstream is associated with local changes in humidity and moves out of the domain by 8 h. When the background wind speed is decreased to 10 m s 21 leftward-propagating mode one is traveling faster than the background flow and is associated with a broad region of reduced relative humidity extending upwind (Figs. 12c,d). By 8 h this region has expanded upwind as far as 2200 km (Fig. 12d). Vertical velocity perturbations associated with leftward-propagating mode 2, which is swept downstream, are also associated with changes in relative humidity in the upper troposphere as it moves downwind from the mountain. This flow regime is analogous to dry flow with 1, K, 2 (e.g., Figs. 6g i, 7a c, 9, and 11c,d), with only transient mode 1 penetrating upwind, mode two clearly visible downstream,

3072 JOURNAL OF THE ATMOSPHERIC SCIENCES FIG. 11. Vertical velocity contours from dry troposphere stratosphere WRF simulations described in Steiner et al. (2005). Mountain is a 500-m Witch of Agnesi with a 20-km half-width. Tropospheric stability N1 5 0.004 s21, tropopause depth 5 11 km, and U0 is (a),(b) 20, (c),(d) 10, and (e),(f) 5 m s21, yielding K 5 0.7, 1.4, and 2.8, respectively. Simulation times are 6 and 15 h in (a) (d) and 6 and 20 h in (e) and (f). Contour interval is 0.01 m s21 in (a) (d) and 0.005 m s21 in (e) and (f). VOLUME 69

OCTOBER 2012 KELLER ET AL. FIG. 12. Vertical velocity (lines) and relative humidity (color) from moist WRF simulations at 4 and 8 h. Mountain is a 500-m Witch of Agnesi with a 20-km half-width. Moist stability Nm 5 0.002 s21, tropopause depth 5 11 km, and U0 is (a),(b) 20, (c),(d) 10, and (e),(f) 5 m s21. Vertical velocity contour interval is 0.05, 0.01, and 0.005 m s21 in (a) and (b),(c) and (d), and (e) and (f), respectively, and relative humidity color scale is on the bottom. 3073

3074 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 69 and a steady-state hydrostatic wave directly above the mountain with a vertical wavelength between 0.5 and 1 times the tropopause depth. This is similar to the wave pattern seen in Miglietta and Rotunno (2005). As the wind is decreased further to 5 m s 21, transient mode 1 moves more quickly upstream of the mountain, and has traveled out of the domain by 8 h (Fig. 12f). At this time mode 2 is visible upwind at approximately 2120 km. Note that simply substituting the moist static stability directly into the nondimensional number K to calculate a corresponding K m, as well as the propagation speed of the upstream modes, does not result in the same cutoff values as in the dry case. In fact, from analyzing a number of moist simulations it appears that the effective stability determining the speed of the modes is actually a value somewhere between the moist and dry stability. This is likely due to the mechanism discussed in Miglietta and Rotunno (2005), in which perturbations induced by the air flowing over the mountain result in regions of subsaturated air that have much larger static stability locally. This increased stability plays a role in conditioning the background flow, allowing the flow to transition between a regime in which there are no modes with a leftward propagation speed exceeding the background flow to a regime conducive to upwind penetration of transient mode one. This effect explains why the wave pattern for the small 50-m-high mountain presented in Miglietta and Rotunno (2005) (their Fig. 4) differs from the wave pattern for the 700-m-high mountain (reproduced in our Fig. 2a). In the moist-saturated simulations shown in Fig. 12 vertical motions associated with the transient modes induce expanding regions of desaturated air. For the same background flow conditions as in Figs. 12a,b and 12c,d, if the mountain height is increased to 1.5 km, the more vigorous vertical velocities associated with upstreampropagating modes induces areas of precipitation that also move upwind (Fig. 13). In Fig. 13a the background flow is U 0 5 20 m s 21 and precipitation remains concentrated on the mountain slope. In contrast, when the background wind is decreased to 10 m s 21, transient mode 1 moves upwind and creates precipitation moving along with the wave (Fig. 13b). Here lines represent model precipitation accumulated over the last 30 min, plotted for hours 2 6, with the accumulation peak shifting leftward with each hour. Furthermore, preliminary investigation of simulations in which the background flow is initially subsaturated for example, when the relative humidity of the incoming flow is 70% 90% (not shown) vertical velocity perturbations associated with these modes increases the relative humidity upwind of the mountain, potentially triggering precipitation. Thus, the upstream modes in the saturated case are similar to the upstream modes in the dry case in that 1) FIG. 13. Spatial distribution of 30-min accumulated precipitation as a function of time for WRF simulations with (a) no upstreampropagating mode (U 0 5 20 m s 21 ) and (b) one upstream-propagating mode (U 0 5 10 m s 21 ). Background atmospheric conditions as in Figs. 12a,b and 12c,d. Mountain is centered at x 5 0, with a height of 1.5 km and half-width of 20 km. Lines correspond to accumulated precipitation over the past 30 min for hours 2 6. both are transient perturbations induced by the impulsive start; 2) the number of leftward-moving modes appearing upwind depends on the background stability, wind, and tropopause depth; and 3) either trapping by a rigid lid or partial trapping by the large stability jump at the tropopause is essential. However, moisture alters the propagation of the transient modes in fundamental ways so that the number and propagation speeds of the upwind penetrating transient waves cannot be calculated by simply replacing the dry stability by the moist stability in K. In addition, for the gravest mode propagating into its saturated environment, downward vertical velocity leads the disturbance. These important differences are currently under investigation and will be elucidated in an upcoming paper by the second author and a colleague. 5. Summary and conclusions In the idealized two-dimensional simulations of saturated, nearly neutral flow over topography, Miglietta and Rotunno (2005) found a wave mode and region of desaturated air expanding upwind of the mountain with time (reproduced in our Fig. 2). The generation and propagation mechanism of this upstream-moving mode was attributed to localized buoyancy variations due to mountain-induced flow patterns creating localized desaturated regions and altered stability. Here we have shown that this wave mode is one of the transient modes excited by the initialization of the

OCTOBER 2012 K E L L E R E T A L. 3075 troposphere stratosphere flow over topography, which induces transient upstream and downstream-propagating wave modes in both moist and dry two-dimensional flow. Modes traveling with speeds faster than the background wind appear upstream of the topography and may at times move slowly enough that their effects persist upwind for hours. These transient modes are similar to the modes observed in dry channel flow that is, a single constant wind and stability layer in which the tropopause is replace by a rigid lid (Baines 1977; Hanazaki 1989, among others). In moist-saturated air, these traveling waves induce a concomitant upwind-progressing desaturated zone. Also, waves with sufficiently large vertical velocities induce a precipitation pattern that moves upstream with the mode, even if the incoming flow is subsaturated (not shown). In the moist simulations of Miglietta and Rotunno (2005) the tropospheric depth, moist stability, and wind velocity, which determine the propagation speed of the modes, were such that that no upwind propagation was possible for a very small mountain; however, some desaturation of the air encountering the 700-m-high mountain increased the effective stability to a value between moist and dry stability so that the flow became ideally suited for the generation of a slow-moving mode 1 upwind of the mountain. In addition, the large stability jump at the tropopause reduced the leakage of wave energy into the stratosphere. Although the focus herein has been on a deep, low-stability troposphere, topped by a stable stratosphere, the regime associated with the waves explored here would also include a lower-stability boundary layer capped by a stable layer aloft. Transient modes, which are a fundamental feature of the time evolution of flow over topography, are part of the solution in any numerical simulation in which the flow is impulsively started from rest at time zero. Often transient features propagate out of the domain during the spinup time of the numerical simulation and do not significantly affect the phenomena being simulated. However, when atmospheric conditions are such that one or more modes propagate slowly upstream of the mountain, their vertical velocity cells may persist upwind for hours, significantly altering the moisture content of the airstream impacting the mountain, as in Miglietta and Rotunno (2005). For moist but unsaturated flow, upstreampropagating waves can increase the relative humidity upwind of the mountain. Thus, just as the propagating wave modes generated by convection with similar vertical structure can trigger new convection (e.g., Mapes 1993; Lane and Zhang 2011), these modes can induce precipitation upwind of topography under certain conditions. This type of feedback loop between the terrain-induced flow dynamics and thermodynamics was diagramed in Steiner and Rotunno (2004). Whether these modes play a significant role in nature needs to be determined. It is possible that upstreampropagating transient gravity waves could be excited whenever there is a rapid change in the background flow, such as the passage of a front, as this would be similar to an impulsive start. Also, for saturated flow, transients can be generated when orographic flow perturbations result in moisture changes that alter the effective stratification, resulting in a transition to a dynamical regime conducive to transient gravity waves penetrating upwind of the mountain. The findings from this study suggest that caution must be taken when simulating moist flow over topography, as transient wave modes can alter rain patterns and relative humidity upwind of the mountain. To what extent these transient modes contribute to rainfall rates observed upwind of topography both in nature and in numerical simulations needs to be investigated further. Acknowledgments. The initial research for this study was supported by the National Science Foundation (NSF) under Grants ATM-9906012 and ATM-0223798 to Princeton University. We thank Dr. Todd Lane for helpful comments on an earlier version of this paper. We also thank the anonymous reviewers whose comments on an earlier version of this paper helped improve the presentation. REFERENCES Baines, P. G., 1977: Upstream influence and Long s model in stratified flows. J. Fluid Mech., 82, 147 159., 1979: Observations of stratified flow over two-dimensional obstacles. Tellus, 31, 351 371., and F. Guest, 1988: The nature of upstream blocking in uniformly stratified flow over long obstacles. J. Fluid Mech., 188, 23 45. Barstad, I., and F. Schuller, 2011: An extension of Smith s linear theory of orographic precipitation: Introduction of vertical layers. J. Atmos. Sci., 68, 2695 2709. Berkshire, F. H., and F. W. G. Warren, 1970: Some aspects of linear lee wave theory for the stratosphere. Quart. J. Roy. Meteor. Soc., 96, 50 66. Castro, I. P., and W. H. Snyder, 1988: Upstream motions in stratified flow. J. Fluid Mech., 187, 487 506. Colle, B. A., 2004: Sensitivity of orographic precipitation to changing ambient conditions and terrain geometries: An idealized modeling perspective. J. Atmos. Sci., 61, 588 606. 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. Eliassen, A., and E. Palm, 1961: On the transfer of energy in stationary mountain waves. Geofys. Publ., 22 (3), 1 23. Engevik, L., 1971: On the flow of a stratified fluid over a barrier. J. Eng. Math., 5, 81 88. Grimshaw, R., and Z. Yi, 1991: Resonant generation of finiteamplitude waves by the flow of a uniformly stratified fluid over topography. J. Fluid Mech., 229, 603 628.

3076 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 69 Hanazaki, H., 1989: Upstream advancing columnar disturbances in two-dimensional stratified flow of finite depth. Phys. Fluids A, 1, 1976 1987., 1992: A numerical study of nonlinear waves in a transcritical flow of stratified fluid past an obstacle. Phys. Fluids A, 4, 2230 2243., 1993: On the nonlinear internal waves excited in the flow of a linearly stratified Boussinesq fluid. Phys. Fluids A, 5, 1201 1205. Janowitz, G. S., 1981: Stratified flow over a bounded obstacle in a channel of finite depth. J. Fluid Mech., 110, 161 170. Jiang, Q., 2003: Moist dynamics and orographic precipitation. Tellus, 55A, 301 316. Keller, T. L., 1994: Implications of the hydrostatic assumption on atmospheric gravity waves. J. Atmos. Sci., 51, 1915 1929. Klemp, J. B., and D. K. Lilly, 1975: The dynamics of wave-induced downslope winds. J. Atmos. Sci., 32, 320 339. Lamb, K. G., 1994: Numerical simulations of stratified inviscid flow over a smooth obstacle. J. Fluid Mech., 260, 1 22. Lane, T. P., and M. J. Reeder, 2001: Convectively generated gravity waves and their effect on the cloud environment. J. Atmos. Sci., 58, 2427 2440., and F. Zhang, 2011: Coupling between gravity waves and tropical convection at mesoscales. J. Atmos. Sci., 68, 2582 2598. Laprise, R., and W. R. Peltier, 1989: On the structural characteristics of steady finite-amplitude mountain waves over bellshaped topography. J. Atmos. Sci., 46, 586 595. Lin, Y. L., 2007: Mesoscale Dynamics. Cambridge University Press, 630 pp., and T. A. Wang, 1996: Flow regimes and transient dynamics of two-dimensional stratified flow over an isolated mountain ridge. J. Atmos. Sci., 53, 139 158. Mapes, B. E., 1993: Gregarious tropical convection. J. Atmos. Sci., 50, 2026 2037. McIntyre, M. E., 1972: On Long s hypothesis of no upstream influence in uniformly stratified or rotating fluid. J. Fluid Mech., 52, 209 243. Miglietta, M. M., and R. Rotunno, 2005: Simulations of moist nearly neutral flow over a ridge. J. Atmos. Sci., 62, 1410 1427. Nance, L. B., and D. R. Durran, 1997: A modeling study of nonstationary trapped mountain lee waves. Part I: Mean-flow variability. J. Atmos. Sci., 54, 2275 2291., and, 1998: A modeling study of nonstationary trapped mountain lee waves. Part II: Nonlinearity. J. Atmos. Sci., 55, 1429 1445. Nicholls, M. E., R. A. Pielke, and W. R. Cotton, 1991: Thermally forced gravity waves in an atmosphere at rest. J. Atmos. Sci., 48, 1869 1884. Pandya, R. E., D. R. Durran, and C. S. Bretherton, 1993: Comments on Thermally forced gravity waves in an atmosphere at rest. J. Atmos. Sci., 50, 4097 4101. Ralph, F. M., P. J. Neiman, T. L. Keller, D. Levinson, and L. Fedor, 1997: Observations, simulations, and analysis of nonstationary trapped lee waves. J. Atmos. Sci., 54, 1308 1333. Rottman, J. W., D. Broutman, and R. Grimshaw, 1996: Numerical simulations of uniformly stratified fluid flow over topography. J. Fluid Mech., 306, 1 30. Rotunno, R., and R. Ferretti, 2001: Mechanisms of intense alpine rainfall. J. Atmos. Sci., 58, 1732 1749., and, 2003: Orographic effects on rainfall in MAP cases IOP2b and IOP8. Quart. J. Roy. Meteor. Soc., 129, 373 390. Sharman, R. D., and M. G. Wurtele, 1983: Ship waves and lee waves. J. Atmos. Sci., 40, 396 427. Skamarock, W. C., J. B. Klemp, J. Dudhia, D. O. Gill, D. M. Barker, W. Wang, and J. G. Powers, 2005: A description of the Advanced Research WRF version 2. NCAR/TN-4681STR, 88 pp. [Available online at http://www.mmm.ucar.edu/wrf/ users/docs/arw_v2.pdf.] Smith, C. M., and E. D. Skyllingstad, 2011: Effects of inversion height and surface heat flux on downslope windstorms. Mon. Wea. Rev., 139, 3750 3764. Smith, R. B., 1979: The influence of mountains on the atmosphere. Advances in Geophysics, Vol. 21, Academic Press, 87 230., and I. Barstad, 2004: A linear theory of orographic precipitation. J. Atmos. Sci., 61, 1377 1391. Steiner, M., and R. Rotunno, 2004: Airflow within complex terrain. Bull. Amer. Meteor. Soc., 85, 1059.,, and W. C. Skamarock, 2005: Examining the moisture effects on idealized flow past 2D hills. Preprints, 11th Conf. on Mesoscale Processes, Albuquerque, NM, Amer. Meteor. Soc., 5M.7. [Available online at https://ams.confex.com/ams/ 32Rad11Meso/techprogram/paper_95767.htm.] Uchida, T., and Y. Ohya, 2001: Numerical study of stably stratified flows over a two-dimensional hill in a channel of finite depth. Fluid Dyn. Res., 29, 227 250. Wurtele, M. G., R. D. Sharman, and T. L. Keller, 1987: Analysis and simulations of a troposphere stratosphere gravity wave model, Part I. J. Atmos. Sci., 44, 3269 3281.,, and A. Datta, 1996: Atmospheric lee waves. Annu. Rev. Fluid Mech., 28, 429 476. Xue, M., 2002: Density currents in shear flows: Effects of rigid lid and cold-pool internal circulation, and application to squall line dynamics. Quart. J. Roy. Meteor. Soc., 128, 47 73.