1. Introduction
Flow past mountains is almost inevitably associated with vertical air motion, giving rise to a whole range of meteorological phenomena (Whiteman 2000) including waves (Durran 2003), strong winds (Durran 1990), clouds (Banta 1990; Houze 1993), and precipitation (Roe 2005). From a theoretical point of view it is interesting to know whether and to what extent these phenomena are associated with specific flow conditions. In this paper we study flow past an isolated mountain in a nonrotating atmosphere and focus on so-called banner clouds (Glickman 2000). The question we will address is “Under which flow conditions can banner clouds be expected to occur?”
Banner clouds are clouds which appear to be attached to the leeward slope of a steep mountain. They have been observed, for instance, at Mount Matterhorn in the Swiss Alps or at Mount Zugspitze in the Bavarian Alps. A systematic definition was provided by Schween et al. (2007) based on time lapse movies taken at Mount Zugspitze. According to their definition a banner cloud has to simultaneously satisfy four criteria: 1) the cloud should exclusively appear in the lee of the mountain, where it is attached to the ground, 2) the cloud should not consist of snow crystals that are blown off the mountain, 3) the cloud should be persistent, and 4) the cloud should not be primarily of convective character. Observations of banner clouds were reported by Wirth et al. (2012), showing (amongst others) characteristic diurnal and seasonal cycles as well as the dependence of banner cloud occurrence on wind strength and direction.
Different hypotheses have been put forward to explain the formation of banner clouds. Voigt and Wirth (2013) tested these hypotheses through numerical simulations. They found that leeside lifting and the associated adiabatic cooling is the main formation mechanism, consistent with earlier suggestions by Schween et al. (2007), Reinert and Wirth (2009), and Wirth et al. (2012). The leeside upwelling was associated with a bow-shaped vortex right behind the mountain [see Fig. 4 in Voigt and Wirth (2013)], implying highly nonlinear flow. However, in all previous studies the model setup was more or less the same, judiciously chosen such as to provide large leeside uplift and, therefore, optimum conditions for banner cloud formation. This leaves open the following questions:
Which aspect of the model setup is essential in order to produce the strong leeside uplift?
What physical mechanism is responsible for the leeside vortex and, hence, for the strong leeside uplift?
As mentioned above, a key ingredient of banner cloud formation is upwelling on the leeward face of the mountain all the way up to its summit. In combination with upward flow on the windward side, this implies the separation of the flow from the ground right at the top of the mountain (Scorer 1955) and, hence, the formation of a lee vortex. In the past there have been two distinct physical mechanisms associated with the formation of vortices in the lee of a mountain. The first one is surface friction, and this mechanism does not require the atmosphere to be stably stratified (Scorer 1955). It is well known in the engineering literature and goes back to the seminal work of Ludwig Prandtl [see, e.g., Anderson (2005)]. The second mechanism relies on the baroclinicity of the flow past a three-dimensional mountain (Smolarkiewicz and Rotunno 1989; Smith 1989; Rotunno and Smolarkiewicz 1991; Rotunno et al. 1999; Epifanio and Rotunno 2005). While this mechanism only operates in a stably stratified atmosphere, it does not require viscosity and surface friction to be present.









Regime diagram for flow past a two-dimensional obstacle, adapted from Fig. 5.8 of Baines (1995). Depending on the combination of aspect ratio
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Although these developments provide some guidance regarding flow separation, they are not directly applicable to our problem. The reason is that the simulations of Schappert and Wirth (2015) suggest that banner cloud formation requires genuinely three-dimensional orography. On the other hand, both the experiments of Baines and Hoinka (1985) and the theory of Ambaum and Marshall (2005) are based on flow past two-dimensional obstacles. Fortunately, it seems that flow past three-dimensional obstacles is similar to flow past two-dimensional obstacles in many aspects. For instance, Hunt and Snyder (1980) carried out towing tank experiments with a three-dimensional hill of moderate slope for different Froude numbers. These experiments indicate overall similar behavior regarding flow separation, although they do not cover the complete parameter space of Baines (1995). We are, therefore, motivated to revisit the regime diagram of Baines (1995) for three-dimensional instead of two-dimensional orography.
Our investigation is based on both two- and three-dimensional simulations of the turbulent flow past an idealized mountain. We will carry out several sets of simulations, analyzing regime behavior and associated flow patterns. It will be shown that many aspects of the two-dimensional regime diagram of Baines (1995) apply also to the three-dimensional simulations. Nevertheless, banner clouds turn out to occur only in a small subspace of the parameter space explored. This means that flow separation is a necessary, but not a sufficient condition for banner cloud formation. In addition, we will elucidate through sensitivity experiments the physical mechanisms which are responsible for flow separation in the different regimes.
The remainder of the paper is organized as follows. Section 2 describes the numerical model, the model setup, as well as the diagnostics used to detect flow separation and banner cloud occurrence. The results are presented in section 3, where we discuss regime behavior and vertical uplift in both two- and three-dimensional simulations and compare them with each other. Finally, section 4 provides some discussion and our conclusions.
2. Model and diagnostics
a. Numerical model and model setup
We use the nonhydrostatic anelastic version of the Eulerian/semi-Lagrangian fluid solver (EULAG; Prusa et al. 2008) in order to simulate the flow of dry air past an idealized mountain in a nonrotating atmosphere. The equations of motion are formulated in a terrain-following coordinate system. The model is run in large-eddy simulation (LES) mode, and the subgrid-scale fluxes of momentum are parameterized through a turbulent kinetic energy (TKE) closure (Sorbjan 1996). The advection scheme is MPDATA (Smolarkiewicz 1983), which contains just enough numerical diffusion so as to keep the numerics stable (Smolarkiewicz 1983; Smolarkiewicz et al. 2007; Beaudoin et al. 2013). For further model details, see also Voigt and Wirth (2013) and Schappert and Wirth (2015).

The model domain for our two-dimensional simulations consists of two parts: an outer part in which we implemented two sponge layers (one at the top and one at the inflow boundary) and an inner part that is free of any wave absorbers (see Fig. 2). In the x direction (streamwise direction), the extent of the inner part of the domain is specified as a function of mountain steepness to be 9.6
Illustration of the model domain for the two-dimensional simulations.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
The requirements on the size of the model domain are less severe for the three-dimensional simulations, because in this case the gravity waves are significantly weaker. Here, the inner part of the model domain extends 4.0
Illustration of the model domain for the three-dimensional simulations.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1

The equations are integrated forward in time using a time step Δt = 0.5 s. Each model run consists of two parts: the spinup run and the ensuing main run. The spinup run allows the model variables to reach a statistically stationary state. Several tests revealed that this was reached to a satisfying approximation after a duration of ts = 10
b. Diagnostics used for the analysis
Although our primary interest is a specific type of orographic clouds, we only simulate flow of dry air past a mountain. At first sight this appears strange, but the approach was chosen quite deliberately. For one thing, earlier simulations have indicated that the release of latent heat, albeit not completely negligible, is of minor importance for banner cloud formation (Reinert and Wirth 2009). It would, therefore, be sufficient to advect a passive tracer q representing specific humidity and to diagnose relative humidity RH from this tracer; a cloud could be said to exist wherever RH > 100% (Voigt and Wirth 2013). However, the need to initialize the tracer q introduces some arbitrariness, since obviously cloud formation depends on the amount of moisture in the air. Instead, it turned out useful to only diagnose the Lagrangian vertical uplift of air parcels, because this can be taken as a proxy for the potential of banner cloud formation: to the extent that uplift is larger on the leeward side than on the windward side, one may expect a banner cloud to occur given suitable moisture conditions.
The flow in our simulations is highly nonstationary, and for some simulations there is quasi-periodic shedding of vortices in the lee of the mountain (Schär and Durran 1997). However, in this paper we are only interested in the major characteristics of the flow that show up in the time-averaged fields. The time average used in our analyses extends over the entire main model run and will be denoted by angle brackets 〈⋅〉.
To identify flow separation from the surface, we use the time-averaged streamwise wind component 〈u〉. The separation point is characterized by a change in 〈u〉 from positive to negative values in the streamwise direction at the surface. Additionally, we use streamlines of the time-averaged wind in order to show the overall behavior of the flow and to delineate the spatial extent of lee vortices (if present).
3. Results
a. Two-dimensional simulations
We start with the two-dimensional simulations. Both
Streamlines of the time-averaged flow for 12 different combinations of
Regime behavior in the two-dimensional simulations. The plots show streamlines of the time-average flow, with the overall flow direction from left to right. Simulations with different values for the parameters
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Regime behavior in the two-dimensional simulations. The color fill represents the time-mean streamwise component of the wind 〈u〉 (m s−1), and the black contour is the zero contour of 〈u〉. Otherwise the figure conventions are as in Fig. 4.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
For shallow obstacles in weakly stratified flow there is no flow separation at all (Figs. 4i and 5i). This behavior is referred to as “no separation.” Keeping the stability parameter
Overall, our results are in a good agreement with the diagram of Baines (1995), which makes us confident that our numerical model is suitable for our purposes. Note that the amount of surface friction can slightly modify the exact location of the transitions between the regimes (cf. Doyle and Durran 2002). To quantify the related sensitivity we carried out the same set of simulations with cd = 0.001 instead of 0.01. The result is shown in Fig. 6. Overall, the regime behavior is unchanged with leeside separation, no separation, and postwave separation all being present quite like in the previous simulations. At the same time there are subtle differences, and the transitions between the regimes have shifted to different values of
As in Fig. 5, but for cd = 0.001.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
It is interesting to note that the theory of Ambaum and Marshall (2005) does not explicitly contain any treatment of surface friction; rather, it only stipulates that there must be some surface friction such as to produce a boundary layer. This is consistent with the idea that the regimes of boundary layer separation in viscous two-dimensional flow are universal, transcending the exact formulation of surface friction. We take the same view in the present paper and focus mostly on the qualitative behavior with the existence of three distinct regimes and the transitions between them.
b. Three-dimensional simulations
We now extend our simulations to three-dimensional flow past three-dimensional orography, allowing us to systematically explore whether and to what extent the above regime behavior remains valid. To cover the regime diagram of Fig. 1 we varied, again,
Regime behavior in the three-dimensional simulations. The panels show streamlines of the time-averaged flow in a vertical cross section through the middle of the domain, with the overall flow direction from left to right. Simulations with different values for the parameters
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Regime behavior in the three-dimensional simulations. The color fill represents the time-mean streamwise component of the wind 〈u〉 (m s−1), and the black contour is the zero contour of 〈u〉. Otherwise, the figure conventions are as in Fig. 7.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Streamlines in a horizontal section are shown in Fig. 9, and the horizontal extent of the recirculation area close to the ground can be gleaned from Fig. 10. For a steep mountain with weak stratification (Fig. 9a), there are two counterrotating vortices with a vertical axis in the immediate lee of the mountain; combining this information with the leeside vortex pattern in the vertical section (Fig. 7a), one recognizes that these features are signatures of a three-dimensional bow-shaped vortex, which has been mentioned and sketched before by Voigt and Wirth (2013, see their Fig. 4). For weak stratification (left column in Fig. 10) the leeside recirculation area decreases with decreasing aspect ratio, and the transition from leeside separation to no separation occurs between
Regime behavior in the three-dimensional simulations. The panels show streamlines of the time-averaged flow in a horizontal cross section at z = 250 m. Otherwise the figure conventions are as in Fig. 7.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Dependence of the recirculation area on aspect ratio and stratification in the three-dimensional simulations. The color fill quantifies the time-mean streamwise component of the wind 〈u〉 (m s−1) in the lowest model layer. The black contour delineates the horizontal extent of the mountain. Otherwise the figure conventions are as in Fig. 7.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Apparently, the best combination of parameters for leeside separation right at the mountain top is a steep mountain in weakly stratified flow (Figs. 7a and 8a). Increasing the stratification for a large aspect ratio (
For all simulations with strong stratification (
As mentioned earlier, there are two distinct physical mechanisms that can generate flow separation and the formation of lee vortices: one based on surface friction and another one based on the baroclinicity of the flow. Since our simulations account for both surface drag and stable stratification, they contain both mechanisms at the same time, and it is not clear which one is primarily responsible for flow separation and vortex formation in the different parts of the parameter space. To separate the two mechanisms, we switched off surface friction by setting cd = 0 and repeated all simulations. The result is shown in Fig. 11. Apparently, for weak stratification (first and second column on the left) the flow does not separate from the surface any longer even for the steep mountains, in distinct contrast with the simulations from Fig. 8. This indicates that leeside separation is essentially due to a viscous mechanism requiring nonzero surface friction. On the other hand, for the runs with strong stratification the flow separates from the surface even in the absence of surface friction (right column in Fig. 11). This is consistent with earlier results where lee vortex formation was found in simulations with a free-slip boundary condition (Smolarkiewicz and Rotunno 1989; Schär and Durran 1997). We also carried out a set of simulations with neutral stratification (i.e., with
As in Fig. 8 except that surface drag has been set to zero.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Regime behavior of vertical uplift in the three-dimensional simulations. The color fill shows the time-mean Lagrangian vertical displacement 〈Δz〉 (m) in a vertical section through the center of the domain. The black contour is the zero contour, and the white contour depicts 〈Δz〉 = +400 m—that is, an uplift almost as large as the mountain height, indicating significant potential for cloud formation. Otherwise, the figure conventions are as in Fig. 7.
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Smolarkiewicz and Rotunno (1989) considered a rather shallow mountain corresponding to
We conclude this section by reemphasizing that the regime behavior in terms of flow separation is quite similar for our two- and three-dimensional simulations, with the same three regimes appearing in the same parts of the phase space.
c. Banner cloud formation
As mentioned before, the occurrence of banner clouds is accompanied by the occurrence of flow separation and lee vortices, but the latter two are only necessary, not sufficient conditions for banner cloud formation. The key diagnostic is vertical uplift Δz, which has to feature a plume of large values in the lee of the mountain. This is what we want to analyze now.
The regime behavior of 〈Δz〉 for our three-dimensional simulations is shown in Fig. 12. Steep mountains in combination with weak stratification (Fig. 12a) show a pronounced windward–leeward asymmetry with a plume of large values in the lee. Of all the panels shown in Fig. 12, this is the situation which is most conducive to banner cloud formation. Keeping stratification fixed and decreasing the aspect ratio leads to a more symmetric distribution of 〈Δz〉 (left column in Fig. 12); the leeward–windward asymmetry and, hence, the potential for banner cloud formation vanishes for shallow mountains. Instead, Fig. 12i suggests that the most likely location of a cloud is right above the mountain, and this cloud would be symmetric with respect to the streamwise direction. These features are usually observed in connection with cap clouds (Glickman 2000). Figure 12 also shows that for intermediate stratification the vertical displacement is downward (rather than upward) in the lee, which is associated with adiabatic warming and the dissolution of clouds, akin to foehnlike conditions (Steinacker 2006; Richner and Hächler 2013). It also shows that optimum conditions for this foehnlike behavior change with changing aspect ratio. Finally, for strong stratification (right column in Fig. 12) the magnitude of vertical displacement in the lee is overall rather small. Of course this is intuitively right, since strong stratification suppresses vertical parcel displacement, and it is consistent with theoretical considerations that predict purely two-dimensional horizontal flow in the limit of very strong stratification (Drazin 1961). In our current context this means that although postwave separation is associated with upward flow in the lee of the mountain, the magnitude of the uplift in this part of the parameter space is rather small such that one would not expect banner clouds to occur.
We noted before that for steep mountains there is a transition from leeside separation to postwave separation as stratification is increased. The top row in Fig. 12 indicates that this transition is accompanied by a gradual change from Lagrangian upward displacement in the lee (i.e., banner cloud conditions; Fig. 12a) to Lagrangian downward displacement producing foehnlike conditions with a tendency for clouds to dissolve in the lee (Fig. 12c). This demonstrates, again, that the existence of flow separation and lee vortex formation in the postwave separation regime is not sufficient to provide conditions conducive to banner cloud formation.
It is also instructive to consider the dependence on stratification for shallow mountains (bottom row in Fig. 12). For weak stratification the displacement is symmetric about the mountain top with its maximum right above the mountain, as one would expect for potential flow. Increasing the stratification renders the uplift asymmetric about the mountain with larger values on the windward side and smaller values on the leeward side (Fig. 12j). Again, this is exactly the opposite of what is required for banner cloud formation. Increasing stratification even further (Figs. 12k,l) continues the tendency for the region of uplift to shift to the windward side, while the pattern on the leeward side gets more complex. Although this more complex behavior is associated with vortexlike flow features (see the bottom-right panels of Figs. 7 and 9) including some uplift, the magnitude of this uplift is smaller than on the windward side. This makes the formation of banner clouds unlikely in this part of the parameter space. It transpires that gently sloping mountains cannot be expected to form banner clouds no matter how small or large the stratification is.



Regime behavior of the banner cloud formation potential P (see text for definition) for three different model configurations. The 12 colored tiles in each plot represent the value of P (m) for the same 12 combinations of
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
Earlier we saw that the two- and three-dimensional simulations are qualitatively similar regarding their regime behavior (cf. Figs. 4 and 7). Does this also mean that the potential for banner cloud formation has a similar parameter dependence for the two- and the three-dimensional simulations? To address this question, we compare the vertical uplift in the three-dimensional simulations (Fig. 12) with the corresponding results in the two-dimensional simulations shown earlier. Here, we focus on the steep mountain in weak stratification (Fig. 12a), which in the case of the three-dimensional simulations gives rise to the characteristic windward–leeward asymmetry conducive to banner cloud formation. For the corresponding two-dimensional simulation (Fig. 14a), there is also some windward–leeward asymmetry. However, below the mountain top the uplift is now larger on the windward rather than the leeward side; above the mountain top there is a (somewhat asymmetric) plume of large values of uplift, but this plume extends to high altitudes way beyond the mountain top. Both features are inconsistent with the observed character of a banner cloud (Schween et al. 2007). Summarizing these results in terms of our metric P gives Fig. 13b. Apparently, all tiles are blue, indicating that for the two-dimensional simulations there is no potential for banner cloud formation in any part of the parameter space.
Vertical uplift for (
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
We return to the three-dimensional simulations and further investigate the mechanisms of lee vortex formation. As mentioned earlier, banner cloud formation is associated with a bow-shaped lee vortex, and this vortex owes its existence to surface friction. This raises the question: how does the field of vertical uplift look like in a run without any surface friction (i.e., for cd = 0). The resulting regime behavior in terms of Δz turns out to be qualitatively similar to that in Fig. 12 except for steep mountains with weak stratification (see Fig. 14b, to be compared with Fig. 12a). In this part of the parameter space the windward–leeward asymmetry is completely lost when surface friction is switched off. This demonstrates, again, that the bow-shaped vortex associated with banner cloud formation arises from flow separation due to surface friction. Condensing this result in terms of our metric P, we obtain Fig. 13c. This plot looks similar as Fig. 13a except in the top-left corner, which indicates that surface friction has a major impact on P in the case of steep mountains and weak stratification; it is only through the inclusion of surface friction that the top-left tiles in the figure turn red. This is but another manifestation of the result that banner cloud formation relies on surface friction.
d. Three-dimensional flow past two-dimensional orography
We considered a third model configuration (not mentioned so far), in which we simulated three-dimensional flow past two-dimensional orography. This setup appears interesting, because two- and three-dimensional flows are known to be fundamentally different regarding their turbulence behavior (e.g., Vallis 2006). To the extent that inertial range arguments apply, kinetic energy is expected to cascade toward larger scales in two-dimensional turbulence but toward smaller scales in three-dimensional turbulence. This can have profound implications for the occurrence of coherent structures (McWilliams 1984). On the other hand, in the present context we do not expect huge differences between two- and three-dimensional flow past two-dimensional orography. This is because the laboratory experiments of Baines and Hoinka (1985) and the theory of Ambaum and Marshall (2005)—both using two-dimensional orography—indicate overall very similar regime behavior despite the fact that the former are based on three-dimensional flow, while the latter is based on two-dimensional flow.
The domain size for the simulations in the present section is like for the two-dimensional simulations in both the streamwise and the vertical direction, while its extent in the spanwise direction is half that from the three-dimensional simulations. The orography is a ridge with translational symmetry in the y direction and with its profile in the streamwise section given by h(x, y = y0) with h taken from (3). We carried out 12 simulations for the same 12 combinations of
Results from an additional set of simulations for three-dimensional flow past two-dimensional orography with fixed
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
We conclude that the dimensionality of the flow is of minor importance for the flow topology in the neighborhood of the mountain and, therefore, for the question which we address in this paper. Arguably this is due to the fact that steep orography represents a major external forcing such that inertial range arguments do not straightforwardly apply in the neighborhood of the orography. Consistent with this interpretation, the only region where dimensionality of the flow seems to make a difference is in the lee of the orography, where the turbulence (which was generated by the flow over the mountain) is allowed to develop freely. We tested this interpretation by ploting the instantaneous spanwise component υ of the wind in the bottom row of Fig. 15. Apparently, significant amplitudes of υ only occur in the lee of the mountain, and they have a clear maximum in Fig. 15h corresponding to that part of the parameter space in which 〈u〉 shows the largest differences owing to the dimensionality of the flow (Fig. 15d versus Fig. 5h).
e. Relevance of the dimensionless parameters 
and 


So far we have explored the two-dimensional parameter space spanned by
For this purpose, we carried out two additional sets of 12 three-dimensional simulations each. In set A we replaced both N and U by half of the value used previously, while in set B we replaced h0, L, and U by half of the value used previously. Both modifications leave the dimensionless numbers
Results from three different simulations with different combinations of dimensional parameters N, L, U, and h0, but such that
Citation: Journal of the Atmospheric Sciences 73, 6; 10.1175/JAS-D-15-0319.1
4. Discussion and conclusions
Previously, the formation of banner clouds has been investigated through numerical simulations, but in these studies the model setup and flow configuration have always been such as to provide conditions that are favorable to banner cloud formation. This left open the question as to which flow conditions actually are conducive to banner cloud formation. It is the goal of the present paper to answer this question through sets of large-eddy simulations, systematically exploring the parameter space spanned by two dimensionless numbers, namely the mountain aspect ratio
In a first step we carried out two-dimensional simulations and were able to reproduce the regime behavior regarding flow separation, which was established some time ago through laboratory experiments (Baines 1995). The three regimes correspond to 1) leeside separation, 2) no separation, and 3) postwave separation. We went on to show that this regime behavior remains qualitatively valid for three-dimensional flow past three-dimensional orography. On the other hand, regarding Lagrangian uplift we found substantial differences between the two- and the three-dimensional simulations. We showed explicitly that these differences are due to the different dimensionality of the orography, not the different dimensionality of the flow: banner clouds require genuinely three-dimensional orography—they cannot occur behind two-dimensional ridges. This conclusion is consistent with the study of Schappert and Wirth (2015), who found that a large fraction of those parcels which end up in a banner cloud have a rather complex and fully three-dimensional path history, flowing around the mountain at rather low altitudes. We also showed that banner clouds require the mountain to be steep and the stratification to be weak such that
It is informative to distinguish hydrostatic from nonhydrostatic flow. Previously, nonhydrostaticity has been diagnosed as NL/U ≡
We discussed two different physical mechanisms that have previously been invoked to explain flow separation from the surface and lee vortex formation: a viscous one associated with surface friction and an inviscid one based on baroclinicity. Our simulations indicate that banner clouds are inevitably associated with boundary layer separation due to surface friction. By contrast, the baroclinic mechanism requires strong stratification, which suppresses vertical uplift and is, therefore, not conducive to banner cloud formation. The fact that friction is required to produce banner-cloud-like conditions was clearly recognized earlier in a conference contribution by Geerts (1992); however, this work seems not to have led to a journal publication that would have allowed one to obtain a more detailed understanding for the line of argument, including the important distinction between upwelling (w > 0) and Lagrangian uplift (Δz > 0) in the lee of the mountain.
Considerations of numerical stability restricted our simulations to aspect ratios
How well are our results supported by available observations? First, we have shown that steep mountains are favorable for banner cloud formation. This is consistent with the general experience that banner clouds typically occur at steep mountains such as Matterhorn, Mount Zugspitze, Fitz Roy, or Mount Everest. Second, we have shown that weak stratification is favorable for banner cloud formation. This is consistent with the observations of Wirth et al. (2012); their measurements indicate that during two specific events at Mount Zugspitze the stratification was weak below the summit of the mountain. Note that for these two events the stratification turned out to be rather strong in the free atmosphere above the mountain. However, this seems to be of little relevance to the phenomenon of banner clouds, as we could show through additional simulations (not presented in this paper).
Quite deliberately we use in our simulations an idealized model setup with a smooth mountain and with inflow profiles characterized by constant wind and constant N. This allows us to explore generic regime behavior, but at the same time it excludes more subtle phenomena that require vertical variations of wind and stratification (e.g., Sheridan and Vosper 2006). We are currently investigating different mountain shapes and inflow profiles and find some nonnegligible sensitivities. For instance, a mountain with salient edges is almost inevitably associated with flow separation (Scorer 1955), but the flow may quickly reattach and the magnitude of the uplift is small—except when
In summary, our results (Figs. 12 and 13a) clearly indicate that there is a general tendency for uplift to be larger on the windward side than on the leeward side except for steep mountains and weak stratification. This is consistent with the intuitive notion that generally a mountain forces the air to rise during the approach, leading to positive vertical parcel displacements and, possibly, cloud formation on the windward side. In this sense the formation of a banner cloud on the leeward side is the exception rather than the rule, and this is possibly why it almost invariably captures the attention of the observer.
Acknowledgments
We sincerely thank the anonymous reviewers for their insightful comments and suggestions, which led to significant enhancements during revision. We thank P. Smolarkiewiocz for his expert advice on numerical modeling and S. Serafin for fruitful discussions on regime behavior. The second author gratefully acknowledges discussions with R. Rotunno about the subtleties of flow past mountains. Computational resources were made available by Deutsches Klimarechenzentrum through support from the Bundesministerium für Bildung und Forschung, as well as by the Zentrum für Datenverarbeitung at the Johannes Gutenberg University Mainz. This work was supported by the German Research Foundation Grant WI 1685/11-1 as well as the Center for Computational Sciences Mainz.
REFERENCES
Ambaum, M. H. P., and D. P. Marshall, 2005: The effects of stratification on flow separation. J. Atmos. Sci., 62, 2618–2625, doi:10.1175/JAS3485.1.
Anderson, J. D., 2005: Ludwig Prandtl’s boundary layer. Phys. Today, 58, 42–48, doi:10.1063/1.2169443.
Baines, P. G., 1987: Upstream blocking of air flow over mountains. Annu. Rev. Fluid Mech., 19, 75–97, doi:10.1146/annurev.fl.19.010187.000451.
Baines, P. G., 1995: Topographic Effects in Stratified Flows. Cambridge University Press, 482 pp.
Baines, P. G., and K. P. Hoinka, 1985: Stratified flow over two-dimensional topography in fluid of infinite depth: A laboratory simulation. J. Atmos. Sci., 42, 1614–1630, doi:10.1175/1520-0469(1985)042<1614:SFOTDT>2.0.CO;2.
Baines, P. G., and R. B. Smith, 1993: Upstream stagnation points in stratified flow past obstacles. Dyn. Atmos. Oceans, 18, 105–113, doi:10.1016/0377-0265(93)90005-R.
Banta, R. M., 1990: The role of mountain flows in making clouds. Meteorological Processes over Complex Terrain, Meteor. Monogr., No. 45, Amer. Meteor. Soc., 229–283.
Beaudoin, P., P. Charbonneau, E. Racine, and P. Smolarkiewicz, 2013: Torsional oscillations in a global solar dynamo. Sol. Phys., 282, 335–360, doi:10.1007/s11207-012-0150-2.
Doyle, J. D., and D. R. Durran, 2002: The dynamics of mountain-wave-induced rotors. J. Atmos. Sci., 59, 186–201, doi:10.1175/1520-0469(2002)059<0186:TDOMWI>2.0.CO;2.
Drazin, P. G., 1961: On the steady flow of a fluid of variable density past an obstacle. Tellus, 13, 239–251, doi:10.1111/j.2153-3490.1961.tb00081.x.
Durran, D. R., 1986: Mountain waves. Mesoscale Meteorology and Forecasting, P. S. Ray, Ed., Amer. Meteor. Soc., 472–492.
Durran, D. R., 1990: Mountain waves and downslope winds. Atmospheric Processes over Complex Terrain, Meteor. Monogr., No. 45, Amer. Meteor. Soc., 59–81.
Durran, D. R., 2003: Lee waves and mountain waves. Encyclopedia of the Atmospheric Sciences, J. R. Holton, J. A. Curry, and J. A. Pyle, Eds., Academic Press, 1161–1169.
Epifanio, C. C., and R. Rotunno, 2005: The dynamics of orographic wake formation in flows with upstream blocking. J. Atmos. Sci., 62, 3127–3150, doi:10.1175/JAS3523.1.
Fröhlich, J., 2006: Large Eddy Simulation Turbulenter Strömungen. Teubner, 414 pp., doi:10.1007/978-3-8351-9051-1.
Geerts, B., 1992: The origin of banner clouds: A potential vorticity perspective. Preprints, Sixth Conf. on Mountain Meteorology, Portland, OR, Amer. Meteor. Soc., 97–98.
Glickman, T. S., Ed., 2000: Glossary of Meteorology. 2nd ed. Amer. Meteor. Soc., 855 pp. [Available online at http://glossary.ametsoc.org/.]
Houze, R. A., Jr., 1993: Cloud Dynamics. Academic Press, 573 pp.
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, doi:10.1017/S0022112080002303.
Klemp, J. B., and D. K. Lilly, 1975: The dynamics of wave-induced downslope winds. J. Atmos. Sci., 32, 320–339, doi:10.1175/1520-0469(1975)032<0320:TDOWID>2.0.CO;2.
McWilliams, J. C., 1984: The emergence of isolated coherent vortices in turbulent flow. J. Fluid Mech., 146, 21–43, doi:10.1017/S0022112084001750.
Mittal, R., and G. Iacarino, 2005: Immersed boundary methods. Annu. Rev. Fluid Mech., 37, 239–261, doi:10.1146/annurev.fluid.37.061903.175743.
Moeng, C.-H., and P. Sullivan, 2003: Large eddy simulation. Encyclopedia of Atmospheric Sciences, J. R. Holton, J. A. Curry, and J. A. Pyle, Eds., Academic Press, 232–240.
Prusa, J. M., P. K. Smolarkiewicz, and A. A. Wyszogrodzki, 2008: EULAG, a computational model for multiscale flows. Comput. Fluids, 37, 1193–1207, doi:10.1016/j.compfluid.2007.12.001.
Reinert, D., and V. Wirth, 2009: A new large-eddy simulation model for simulating air flow and warm clouds above highly complex terrain. Part II: The moist model and its application to banner clouds. Bound.-Layer Meteor., 133, 113–136, doi:10.1007/s10546-009-9419-x.
Richner, H., and P. Hächler, 2013: Understanding and forecasting alpine foehn. Mountain Weather Research and Forecasting, Springer, 219–260, doi:10.1007/978-94-007-4098-3_4.
Roe, G. H., 2005: Orographic precipitation. Annu. Rev. Earth Planet. Sci., 33, 645–671, doi:10.1146/annurev.earth.33.092203.122541.
Rotunno, R., and P. K. Smolarkiewicz, 1991: Further results on lee vortices in low-Froude-number flow. J. Atmos. Sci., 48, 2204–2211, doi:10.1175/1520-0469(1991)048<2204:FROLVI>2.0.CO;2.
Rotunno, R., V. Grubišić, and P. K. Smolarkiewicz, 1999: Vorticity and potential vorticity in mountain wakes. J. Atmos. Sci., 56, 2796–2810, doi:10.1175/1520-0469(1999)056<2796:VAPVIM>2.0.CO;2.
Schappert, S., and V. Wirth, 2015: Origin and flow history of air parcels in orographic banner clouds. J. Atmos. Sci., 72, 3389–3403, doi:10.1175/JAS-D-14-0300.1.
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, doi:10.1175/1520-0469(1997)054<0534:VFAVSI>2.0.CO;2.
Schween, J. H., J. Kuettner, D. Reinert, J. Reuder, and V. Wirth, 2007: Definition of “banner clouds” based on time lapse movies. Atmos. Chem. Phys., 7, 2047–2055, doi:10.5194/acp-7-2047-2007.
Scorer, R. S., 1955: Theory of airflow over mountains: IV-separation of flow from the surface. Quart. J. Roy. Meteor. Soc., 81, 340–350, doi:10.1002/qj.49708134905.
Sheridan, P. F., and S. B. Vosper, 2006: A flow regime diagram for forecasting lee waves, rotors and downslope winds. Meteor. Appl., 13, 179–195, doi:10.1017/S1350482706002088.
Smith, R. B., 1989: Comment on “Low Froude number flow past three-dimensional obstacles. Part I: Baroclinically generated lee vortices.” J. Atmos. Sci., 46, 3611–3613, doi:10.1175/1520-0469(1989)046<3611:COFNFP>2.0.CO;2.
Smolarkiewicz, P. K., 1983: A simple positive definite advection scheme with small implicit diffusion. Mon. Wea. Rev., 111, 479–486, doi:10.1175/1520-0493(1983)111<0479:ASPDAS>2.0.CO;2.
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, doi:10.1175/1520-0469(1989)046<1154:LFNFPT>2.0.CO;2.
Smolarkiewicz, P. K., R. Sharman, J. Weil, S. G. Perry, D. Heist, and G. Bowker, 2007: Building resolving large-eddy simulations and comparison with wind tunnel experiments. J. Comput. Phys., 227, 633–653, doi:10.1016/j.jcp.2007.08.005.
Sorbjan, Z., 1996: Numerical study of penetrative and “solid lid” nonpenetrative convective boundary layers. J. Atmos. Sci., 53, 101–112, doi:10.1175/1520-0469(1996)053<0101:NSOPAL>2.0.CO;2.
Steinacker, R., 2006: Alpiner Föhn—Eine neue Strophe zu einem alten Lied. Promet, 32, 3–10.
Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. Cambridge University Press, 745 pp.
Voigt, M., and V. Wirth, 2013: Mechanisms of banner cloud formation. J. Atmos. Sci., 70, 3631–3640, doi:10.1175/JAS-D-12-0353.1.
Whiteman, C. D., 2000: Mountain Meteorology: Fundamentals and Applications. Reprint ed. Oxford University Press, 376 pp.
Wirth, V., M. Kristen, M. Leschner, J. Reuder, and J. H. Schween, 2012: Banner clouds observed at Mount Zugspitze. Atmos. Chem. Phys., 12, 3611–3625, doi:10.5194/acp-12-3611-2012.