1. Introduction
Mountain drag parameterizations estimate the total transfer of momentum across the topographic boundary (the base flux) as well as the convergence of the pseudomomentum flux as a function of height above the boundary. Previous schemes for estimating the base flux have relied entirely on statistical measures of the height, shape, and orientation of the unresolved terrain, guided by dimensional reasoning (e.g., Baines and Palmer 1990; Lott and Miller 1997; Scinocca and McFarlane 2000). Here I propose a scheme that uses an analytical estimate of the drag vector. This estimate is exact in the limit where the mountain waves are linear, inviscid, and hydrostatic, and the medium is nonrotating and uniform in time and space.
I also propose a correction for nonlinearity at the source. It is similar in many ways to existing schemes that allow a gradual transition between linear and nonlinear dimensional drag laws (e.g., Kim and Arakawa 1995, and references therein; Lott and Miller 1997; Scinocca and McFarlane 2000). However, the proposed treatment exploits the correlation of mountain height with mountain width together with the known range of mountain heights within the grid cell. The motivation for adding this complexity is to produce a closer match to the analytical result in the linear limit as well as a less arbitrary transition to nonlinearity. A beneficial side effect is to make the overall scheme less tunable.
The impact of the subgrid topography on the resolved flow is due to the convergence of the pseudomomentum flux. A straightforward extension of the nonlinear base-flux correction yields an algorithm for clipping the flux to the saturation value as a function of height so as to determine this impact on the resolved flow. This part of the scheme adheres to traditional assumptions going back to Lindzen (1981), Palmer et al. (1986), and Pierrehumbert (1987).
2. Base-flux closure
Drag parameterizations for the atmosphere generally assume that the disturbance is steady and contains only upward-propagating waves. If we add to these assumptions certain constraints on the horizontal scale of the forcing, such that the waves are broad enough to be hydrostatic but narrow enough to be unaffected by sphericity, rotation, horizontal shear, or baroclinicity, we can calculate the total drag in the linear limit as accurately as we know the shape of the terrain. This section begins with the linear analysis and then extends the scheme to deal with nonlinearity at the source.
a. Linear drag









The integrals in (5) and (6) diverge if ĥ does not decrease as fast as |k| for small |k| (large scales). However, the result is already physically inconsistent at large scales. For both mathematical convergence and physical consistency, the input topography must be filtered to retain only the scales that force stationary, nonrotating gravity waves. Then, as a practical matter, (6) need not be integrated over the entire globe for each x, but only over a radius equal to a small multiple of the high-pass filter scale. It is simplest to allow the topographic dataset to establish the low-pass scale. The problem of filtering is discussed by most drag-scheme developers, including Baines and Palmer (1990), Lott and Miller (1997), and Scinocca and McFarlane (2000). More is said about it below.

Because of the filtering, it should be consistent to regard (5) as an integral over spherical harmonics and to interpret |k| as the total spherical-harmonic wavenumber. Spherical-harmonic analysis of h is also a practical way to obtain the field A(x, y) and the various second partial derivatives of A needed for 𝗧. However, for the examples to follow, the flow disturbance was obtained from (6) and the small contribution from χ1 was ignored. The topography was preprocessed using a spatial filter that passes only scales of less than about 1.5° of longitude and latitude in the extratropics. The filtering scale, d, was allowed to vary with latitude so as to keep the Rossby number, V/fd, of order unity, where V(y) is a profile of typical surface wind speeds, and f(y) is the Coriolis parameter. The horizontal derivatives in (8) make the result fairly insensitive to the choice of scales retained by a high-pass filter, and this allows a time-independent choice for d(y).
Using a physical filter instead of a numerical one creates the danger of systematically overcounting the drag when the model grid spacing is smaller than d. With a physical filter, the subgrid drag has to be defined as the difference between the total parameterized drag and the model-dependent resolved drag, that is, the subgrid contribution is not known without diagnosing the resolved drag. In many climate-model applications, resolved drag can be neglected. This approach does not solve the problem of representing strong scale interactions near the grid scale—a generic problem of closures that also applies, for example, to moist convection schemes in both the time and space domain.
From (6), the velocity scales as Nrhr, where hr is a measure of the mesoscale component of the topographic height. Therefore, mountain wave velocities have an order of magnitude of V ∼ (10−2 s−1)(103 m) = 10 m s−1, given a typical mesoscale relief of 1000 m. If the scale of the topography is 100 km and the ambient wind is of order 10 m s−1, the vertical velocity is of order 0.1 m s−1. It follows that the drag is characteristically D ∼ ρr(10 m s−1) (0.1 m s−1) ∼ 1Pa. The tensor 𝗧 is dominated by the diagonal elements, which are of the order of ρr(10 m s−1).
Shown in Fig. 1 is a plot of the linear drag 〈τ〉 over the Western Hemisphere for an assumed large-scale wind that is purely zonal at −7 m s−1 in the Tropics and 13 m s−1 in the extratropics. A constant buoyancy frequency, Nr = 0.01 s−1, and constant density, ρr = 1 kg m−3, are also assumed. The same calculation for the Asian continent, shown in Fig. 2, assumes a large-scale wind from the west at 10 m s−1 everywhere. The averaging in both Fig. 1 and Fig. 2 is over latitude–longitude cells of about 1.5° on a side. The topographic dataset has a resolution of 1/30°.
b. Nonlinear extension
The plan is to evaluate the nonpropagating drag associated with blocking mountains by resorting to dimensional analysis and assuming a fast orographic adjustment process (Pierrehumbert and Wyman 1985; Pierrehumbert 1987; Olaffson and Bougeault 1996). What follows is a fairly standard treatment along these lines (e.g., Lott and Miller 1997), except that it allows for a range of mountain heights within the grid cell. As mentioned briefly at the beginning, the motivation for adding complexity here is 1) to produce a better match between the linear drag and the dimensional estimate in the limit of small terrain and 2) to set up a more rational, less tunable, transition to partially blocked flow. Thus, the analytical result will be used to constrain the drag coefficient for propagating components, and this scaling, together with some assumptions about the mountain ensemble, will determine the transition to partially blocked flow. Since the excess mountain height responsible for blocking cannot be straightforwardly attributed to spectral components, the key assumptions here will be 1) that the topography is characterized by well-defined features, each with a well-defined areal extent, that can be binned into height ranges, and 2) that the flow disturbances induced by individual features do not interact strongly.
1) Drag model for individual features
As the flow becomes blocked by topography, the drag law in the blocked layer changes from Dl ∼ ρNVh2/L to Dnl ∼ ρV2(h − hc)/L, where V and L are the ambient wind component and length of the mountain, respectively, in the same direction. Orographic adjustment essentially means that the nondimensional depth h̃ = h(
2) Drag model for the grid-cell ensemble
Linear drag closures typically take Lb in (13) to be a universal constant and estimate h̃2 as the mean-squared height of the filtered topography within the grid cell. However, we have a better chance of matching the analytical drag if h and Lb are allowed to covary when averaging (13) over the grid cell. In the general case, with both Dnp ≠ 0 and Dp ≠ 0, it is impractical to average explicitly over the subgrid height distribution because while h̃c is assumed constant, h̃ is time dependent.
If we can neglect overlaps between mountains and ignore any correlation between anisotropy and mountain height, the areal coverage dA ∼ LdL of features in the range from h̃ to h̃ + dh̃ will be proportional to n(h̃)h̃2γ−1 dh̃, where n(h̃) is the number of features in the range. For this number density we assume another power law, namely, n(h) = n1(h/h1)−ε. We introduce n(h̃) mainly to acknowledge that it need not be unity: the value of ε used in the examples to follow is only a guess. The reasons for the specific functional form of n(h̃) are the same as for the relation (14), namely, monotonicity and integrability. Because of area averaging, the constants n1 and h1 will not appear in any results.
3) Scaling of the analytical flux
Although the linear drag coefficient a0 is no longer explicit, the strategy for choosing V in the dimensional analysis should still be to minimize its variability in time and space. The direction of τ* varies between that of −
c. Illustrative examples

For hmin and hmax, the actual minimum and maximum from the set of local extrema of high-pass topography within each cell were not used. There is a great deal of scatter in the height distribution, as shown in the appendix. To produce a less erratic estimate of the height limits, the small-amplitude limit of Dp in (13) was averaged over the actual topographic data and equated to the small-amplitude limit of 〈Dp〉 in (15). This yielded the desired hmax after assuming hmin = μhmax for some fixed μ. This way of handling hmin makes the linear drag estimate independent of ε. Since hmin ≪ hmax almost everywhere, the result is insensitive to μ for realistic choices. To integrate (13) over grid cells with bandpassed data, the integrand was taken as proportional to h2−γ and summed directly over area instead of bins. Integrating with respect to area fails to hold h and L fixed across individual mountains, but the consequence, according to present assumptions, is a constant multiplicative factor that depends on the mountain shape and does not change the normalized drag. The condition h > 0 was enforced by subtracting the minimum value within a bandpass radius, with h nudged to zero in the far field to avoid discontinuities. The same thing was done to obtain the rms result. Although this is not exactly the traditional rms result, it is a better match with the analytical drag and therefore shows the separate impact of γ. For the result in the bottom panel of Fig. 4, μ = 0.
The transition from linear flux to saturation flux for the two extreme cases of h̃min/h̃max is shown in Fig. 5, in which 〈Dp〉/D* is graphed as a function of h̃max/h̃c with γ = 0.4, β = 0.5, and ε = 0. Also shown is the total normalized drag (〈Dp〉 + 〈Dnp〉)/D* based on the additional assumption that a1/a0 = 9.0h̃c. This choice for a1 produces a maximum total drag of approximately 2D*, which is a compromise between the maximum drags obtained in two- and three-dimensional nonlinear simulations, as summarized by Lott and Miller (1997) and Scinocca and McFarlane (2000). The sharpest transitions in both Dp and Dnp occur for h̃max ≈ h̃min. Results are less sensitive to γ, but large values of this parameter make the transitions sharper (not shown).
Shown in Fig. 6 is the bulk-dimensional drag for North America based on the same parameter choices as in Fig. 5. The assumed large-scale wind, needed for determining the range of h̃, is purely zonal at 10 m s−1, and the static stability is
3. Level-by-level determination of momentum forcing


The change to U in (20) is a z-dependent linear transformation of the original integration variable δ. According to the theorem of Eliassen and Palm (1960), the value of U associated with a particular feature is independent of z until that part of the field breaks. The unbroken components are the only parts of the flux that depend on the amplitude of the source feature, or U. This allows us to fix the integration limits, Umin and Umax, and move all z dependence in Dp to the saturation threshold Uc. The threshold is already z dependent because the saturation value of δ̃ depends on the minimum value of hc = (
Figure 7 is a graph of 〈Dp(z)〉/D*, showing the transition to saturation as a function of height above the mountains for two different ranges of h̃ and an assumed environmental column described in the caption. The dotted curves show the momentum forcing, which is proportional to the derivative of 〈Dp(z)〉. Pseudomomentum is deposited in the layers where Dp is decreasing with height, which, in this example, occur near the ground and just above the two assumed jets. The forced layers are broadened and full saturation is delayed when the terrain features within grid cells vary most widely in height (h̃min/h̃max = 0).
4. Time-varying resolved flow




For the earth’s semidiurnal tides, ω0 ≈ 1.4 × 10−4 s−1. Then since the deep ocean tidal amplitude is
5. Summary
Linear analysis gives the vector drag due to stationary, hydrostatic, nonrotating internal waves forced by arbitrary topography. Therefore, statistical or dimensional characterizations of the subgrid terrain are not required for the linear base-flux computation at most of the relevant spatial scales. It turns out that the linear base-flux formula in the case of an unsteady resolved flow is virtually identical to that for a steady background flow if the dominant frequencies satisfy ω ≫
The analytical drag can be made part of a practical closure for atmospheric general circulation models. The analysis in the unsteady case could be used for parameterizing topographic drag due to semidiurnal tidal oscillations in the ocean. The correction for nonlinearity and the related formula for saturation flux continue to require an ad hoc treatment. I have suggested a dimensional treatment of the nonlinearity that takes maximal advantage of the linear solutions. Dimensional analysis indicates that most of the nonlinearity in the semi-diurnal ocean tide is captured by the steady-state limit.
The present approach to the problem of drag closure makes it possible to use statistics much more sparingly than in previous schemes. The use of statistics has been reduced to a one-parameter height–width relation used to separate the nonpropagating from the propagating drag, and this is tightly constrained by the analytical result. For base-flux estimates in strongly blocking regimes, it is not known whether the nonpropagating drag estimate improves on methods that rely more fundamentally on statistical representations of the unresolved topography. However, the propagating part of the base flux estimate is preferable because the linear analysis extracts the most relevant information from the terrain data. The improvement almost certainly extends to mountains that are marginally blocking.
Above the mountains, the sharpness of the transition to saturation determines the depth of the layers subjected to momentum forcing. The forcing is therefore broadened by the proposed ensemble integration. Broadening due to nonparallel vertical shear and subgrid variation of drag orientation (Shutts 1995) is missing from the scheme. For real topography, there is no universal relationship between the subgrid drag orientation and subgrid mountain height, which would be needed to make a simple correction to (21) for rotational shear. A credible solution may be to introduce a probability distribution for drag angles analogous to, but independent of, (14). The ensemble integration would then be two-dimensional.
Acknowledgments
I am grateful to Brian Arbic for checking the analysis. I also benefited from discussions with Isaac Held, Jonas Nycander, and John Scinocca about implementation and applicability. I thank Amy Braverman for some good advice on analyzing the scatterplots.
REFERENCES
Arbic, B K., S T. Garner, R W. Hallberg, and H L. Simmons, 2004: The accuracy of surface elevations in forward global barotropic and baroclinic tide models. Deep-Sea Res., 51B , 3069–3101.
Bacmeister, J T., 1993: Mountain-wave drag in the stratosphere and mesosphere inferred from observed winds and a simple mountain-wave parameterization scheme. J. Atmos. Sci., 50 , 377–399.
Bacmeister, J T., and M R. Schoeberl, 1989: Breakdown of vertically propagating two-dimensional gravity waves forced by orography. J. Atmos. Sci., 46 , 2109–2134.
Baines, P G., and T N. Palmer, 1985: Rationale for a new physically based parametrization of subgridscale orographic effects. Tech. Memo. 169, European Centre for Medium-Range Weather Forecasts, 11 pp.
Bannon, P R., and J A. Zehnder, 1985: Surface pressure and mountain drag for transient airflow over a mountain ridge. J. Atmos. Sci., 42 , 2454–2462.
Bell, T H., 1975: Lee waves in stratified flows with simple harmonic time dependence. J. Fluid Mech., 67-4 , 705–722.
Eliassen, A., and E. Palm, 1960: On the transfer of energy in stationary mountain waves. Geofys. Publ., 22 , 1–23.
Holton, J R., 1973: An Introduction to Dynamical Meteorology. Academic Press, 361 pp.
Kim, Y-J., and A. Arakawa, 1995: Improvement of orographic gravity wave parameterization using a mesoscale gravity wave model. J. Atmos. Sci., 52 , 1875–1902.
Lindzen, R S., 1981: Turbulence and stress due to gravity wave and tidal breakdown. J. Geophys. Res., 86 , 9707–9714.
Lindzen, R S., 1988: Supersaturation of vertically propagating internal gravity waves. J. Atmos. Sci., 45 , 705–711.
Lott, F., and M J. Miller, 1997: A new subgrid-scale orographic drag parametrization: Its formulation and testing. Quart. J. Roy. Meteor. Soc., 123 , 101–127.
Miranda, M A., and I N. James, 1992: Non-linear three-dimensional effects on gravity-wave drag: Splitting flow and breaking waves. Quart. J. Roy. Meteor. Soc., 118 , 1057–1081.
Olaffson, H., and P. Bougeault, 1996: Nonlinear flow past an elliptic mountain ridge. J. Atmos. Sci., 53 , 2465–2489.
Palmer, T N., G J. Shutts, and R. Swinbank, 1986: Alleviation of a systematic westerly bias in general circulation and numerical weather prediction models through an orographic gravity wave drag parameterization. Quart. J. Roy. Meteor. Soc., 112 , 1001–1039.
Pierrehumbert, R T., 1987: An essay on the parametrization of orographic gravity wave drag. Proc. Seminar/Workshop on Observations, Theory and Modeling of Orographic Effects, Vol. 1, Shinfield Park, Reading, United Kingdom, ECMWF, 251–282.
Pierrehumbert, R T., and B. Wyman, 1985: Upstream effects of mesoscale mountains. J. Atmos. Sci., 42 , 977–1003.
Scinocca, J F., and N A. McFarlane, 2000: The parameterization of drag induced by stratified flow over anisotropic orography. Quart. J. Roy. Meteor. Soc., 126 , 2353–2393.
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., 1977: The steepening of hydrostatic mountain waves. J. Atmos. Sci., 34 , 1634–1654.
Stern, W F., and R T. Pierrehumbert, 1988: The impact of an orographic gravity-wave drag parameterization on extended range prediction with a GCM. Preprints, Eight Conf. on Numerical Weather Prediction, Baltimore, MD, Amer. Meteor. Soc. 745–750.
APPENDIX
Constraint on the h(L) Relationship
Whether the relationship assumed in (14) is reasonable can be judged from scatterplots of log height versus log width of actual terrain features. A naive approach to generating these values from the bandpassed data is as follows. A terrain feature is considered to be any local maximum. Local minima are ignored because they would presumably share a width scale with a nearby maximum. This allows us to enforce h > 0, as was done in the nonlinear drag analysis, by referring heights to the minimum value observed within the local bandpass radius. The widths are then estimated from the finite-difference Laplacian of the terrain height normalized by the height itself.
The result of using the topographic data directly in these calculations is an essentially flat distribution of width that is dominated by the scale of the input grid. This is because the derivatives in the Laplacian operator strongly favor the smallest scales. To get a better picture of the mountain shape distribution, I instead perform the same analysis on the velocity potential (6), which is a horizontal integral of height.
The plots in Fig. A1 focus on six geographical regions. The abscissa is the logarithm of normalized velocity potential, and the ordinate is the logarithm of the aforementioned function of velocity potential that can be interpreted as the width scale, namely the square root of A = −χ/∇2χ. The clustering of points near the mean is exaggerated by the filtering. Although the positive sign of the correlation is unambiguous, the scatter is considerable (with correlation coefficients of only 0.2 to 0.3) and difficult to distill.
With this in mind, I have drawn, for each of the six regions, the principal component axis for a subset of the distribution formed by ignoring the largest scales. The justification for leaving out the largest scales is that the width computation at that end is strongly affected by interference from smaller features. When features are not cleanly separated, the distribution becomes flat. The slopes of the principal component axes for the full distributions are generally 0.1 to 0.2, but when we exclude the mountains with dimension exceeding twice the mean (measured along the principal axis) one gets the steeper slopes recorded in the plots. These range from about 0.3 to 0.5. Based on this analysis, I would settle on γ = 0.4 ± 0.1 for the parameter in (14), with the largest values for the Andes and the Alps, and the smallest values for the Himalayas.
The drag associated with stationary linear mountain waves over North and South America and western Antarctica. The assumed surface wind is purely zonal at −7 m s−1 in the Tropics and 13 m s−1 elsewhere. The assumed surface static stability and density are 0.01 s−1 and 1.0 kg m−3, respectively. The arrow below the plot shows the scale for 2 Pa.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
The drag associated with stationary linear mountain waves over Asia. The assumed surface wind is purely zonal at 10 m s−1 everywhere. The assumed surface static stability and density are as in Fig. 1. The scale for 2 Pa is shown by the arrow below the plot.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
Details of the scheme for separating the propagating part of the base flux from the nonpropagating (blocked or deflected) part. The hatched area is the cross section presented to the blocked or deflected part of the flow.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
The distribution over the Northern Hemisphere of the linear drag computed in three ways: (top) the analytical drag reduced to a scalar, (middle) the dimensional drag based on topographic variance, and (bottom) the dimensional drag based on (17) with γ = 0.5. Values are normalized by the respective global maxima, which occur in the Himalayas. The output grid has a resolution of 2.0° in latitude by 2.5° in longitude. The input grid is 1/30° in both directions.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
Normalized estimates, based on dimensional analysis, of the propagating drag, 〈Dp〉/D* (dashed), and total drag, 〈Dp + Dnp〉/D* (solid), as a function of normalized maximum mountain height, h̃max/h̃c, for the two extreme cases h̃min = 0 and h̃min = h̃max with γ = 0.4, β = 0.5, and ε = 0. The drag is normalized by the linear dimensional estimate D* defined by (16) and (17). The nonpropagating part assumes a1/a0 = 9.0h̃c.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
The distribution over North and South America of the normalized (left) total base flux and (right) nonpropagating fraction, according to the bulk dimensional analysis, with γ = 0.4, β = 0.5, and ε = 0. The total flux is normalized by D* + 0.05, where D* is the small-amplitude limit and the small increment (in units of Pa) serves to mask out regions of weak forcing. The assumed large-scale surface wind is purely zonal at 10 m s−1 everywhere and the assumed static stability and density are as in Fig. 1. The input and output grids are the same as in Fig. 4. The critical mountain height is set to h̃c = 0.7 to make the drag fully nonpropagating (blocked) only in the highest part of the Andes.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
The normalized pseudomomentum flux 〈Dp〉/D* (solid) and normalized momentum forcing (dashed) as a function of height for the two extreme cases h̃min = 0 and h̃min = h̃max with γ = 0.4, β = 0.5, ε = 0, and h̃max = 1.2h̃c. The assumed wind profile (shown at right) has jets of 38 and 58 m s−1 centered at heights of 9 and 25 km, respectively. The static stability increases from 0.011 to 0.022 s−1 across z = 11 km, and the assumed density scale height is 8 km. The momentum forcing is normalized by D*/h.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
Fig. A1. Scatterplots of mountain width vs mountain height on log–log axes for six geographical regions. Width (ordinate) is estimated from the finite-difference Laplacian of the velocity potential, and height (abscissa) is identified with the velocity potential itself, referred to the local minimum value. Correlation coefficients range from 0.2 to 0.3. The line segments show the orientation of the principal components of the distributions excluding the largest features (see text), with the slope values indicated at the bottom right of each plot.
Citation: Journal of the Atmospheric Sciences 62, 7; 10.1175/JAS3496.1
The high-frequency limit is determined by the terms n = ±1 in the solution by Bannon and Zehnder (1985). Note, however, that their passing remark about the surface pressure in this limit is incorrect: the pressure perturbation does not vanish but is given by (26) below.