1. Introduction
A great deal of effort has been expended in theoretical studies aimed at explaining the observed stationary wave response of the atmosphere to orographic forcing, and observed teleconnection patterns; see, for example, the review by Frederiksen and Webster (1988). One of the most fundamental approaches is to linearize the equations of motion about an axisymmetric flow and investigate the response to an isolated forcing (Hoskins et al. 1977; Grose and Hoskins 1979; Hoskins and Karoly 1981; Webster 1981). These studies, and others, showed that planetary-scale Rossby waves tend to propagate approximately along the arcs of great circles. However, linear theory does not give insight into the behavior of finite-amplitude wave trains when they reach a critical line, where the flow speed equals the phase speed of the waves (e.g., Killworth and McIntyre 1985).
Recently, Waugh et al. (1994, hereafter WPP) and Brunet and Haynes (1996, hereafter BH), have used a single-layer model to investigate the nonlinear behavior of finite-amplitude stationary waves as they approach the zero wind line in low latitudes. Both studies showed a strong tendency for the wave train to be reflected back into midlatitudes from the longitude where the waves begin to break. Similar results have been also been obtained in a realistic three-dimensional flow by Magnusdottir and Haynes (1999). It is certainly an open question as to whether such reflection can be observed in the atmosphere, but typical pseudomomentum flux analyses of stationary wave patterns (e.g., Karoly et al. 1989) seem to indicate that waves often propagate from the extratropics and are absorbed at low latitudes, at least in the case when the flow there is easterly. Both WPP and BH emphasize the fact that their respective models did not contain a representation of the Hadley circulation, which might change the behavior of the low-latitude critical layer from reflecting to absorbing. In this paper we aim to address this issue, by comparing the behavior of waves excited in simple single-layer flows that are closely comparable except for the presence or absence of a representation of the Hadley circulation. Partly because of the limitations of the model chosen, we concentrate on the winter hemisphere case.
There have been several studies of the effect of the Hadley circulation on linear wave propagation characteristics. Schneider and Watterson (1984) showed that southward linear propagation across an easterly layer is possible in a shallow-water model provided the mean meridional flow across the layer is also to the south. They also developed an analysis of the quasigeotrophic β-plane case that was further extended by Farrell and Watterson (1985). They showed that when the mean meridional flow was to the north, the critical line “barrier” to propagation was shifted to the north (see also the discussion in section 4a below). Finally, a three-dimensional primitive-equation model was studied by Watterson and Schneider (1987), who showed that a Hadley circulation could allow cross-hemispheric propagation from the summer hemisphere in a more realistic flow.
Held and Phillips (1990, hereafter HP) are the only authors who, to our knowledge, have investigated the nonlinear interaction of the Hadley circulation and Rossby wave breaking. They used a hybrid model in which a zonally periodic, barotropic Rossby wave encounters the mean meridional flow generated in a zonally averaged, forced shallow-water system with a layer depth of 1 km. This value is chosen as an appropriate equivalent depth for the upper-tropospheric branch of the Hadley circulation, when the effects of moisture are taken into account. Although they were motivated primarily by the need to understand the zonal mean angular momentum budget, their results seem to indicate that significant long-term, low-latitude absorption of finite-amplitude waves can be maintained, in contrast to the situation when the low-latitude critical layer is entirely inviscid.
Based on these earlier results, two key questions to be addressed in this study are the following.
What effect, if any, does the shift in the position of the “barrier” due to the Hadley circulation have on the propagation, breaking, and nonlinear reflection of finite-amplitude waves?
To what extent does the Hadley circulation inhibit the nonlinear reflection observed in “no-Hadley circulation” experiments and allow prolonged absorption of incoming wave activity?
In section 2 we describe the single-layer shallow-model used for our experiments. The model can be thought to represent an isentropic layer spanning the upper troposphere—and some advantages and limitations of this view are described in HP. Following BH, in order to use a single consistent model for our experiments, we choose a compromise layer depth (5 km) between that thought appropriate for extratropical barotropic Rossby waves (≈7–10 km) and that suggested by HP for the Hadley circulation (1 km). Although this choice has its disadvantages, discussed below, we believe that the advantages conferred by the simplicity of our model, together with the fact that direct dissipation of the eddies by this forcing is shown not to be the most important component in the wave activity budgets, justify such an approach. In section 3a we describe experiments in which the waves are forced by a zonally periodic wave-3 topography (as in HP), while in section 3b we use an isolated mountain to generate solitary wave trains (as in WPP and BH). In the discussion of section 4a we revisit the linear theory of wave propagation in the presence of a mean meridional current by reviewing the quasigeostrophic β-channel analysis of Schneider and Watterson (1984). Wave activity budgets are then calculated for both the zonally symmetric wave-3 experiments and the isolated mountain experiments in section 4b. In section 5 the effect of the presence of the Hadley circulation on the time mean stationary wave structure is investigated, along with its effect on the Simmons–Wallace–Branstator instabilities reported for this type of model by Polvani et al. (1999). The summary and conclusions are presented in section 6.
2. The model and experiments
a. Experiments with a Hadley circulation
If the equations (1) are integrated for 10–20 relaxation timescales ≈max{
These values of the damping coefficients generate a circulation that is far from nonlinear, a nonlinear circulation in this sense being one associated with a region in the Tropics of constant angular momentum, and therefore zero potential vorticity. Although the zonal mean angular momentum field is not close to being constant in the tropical atmosphere, Held and Hou (1980) have shown that a nonlinear circulation is generated in an inviscid, axisymmetric atmosphere in the absence of any wave forcing due to either transients or stationary waves. One way to justify the high value of momentum drag in our model, and thus the dissipative nature of our Hadley circulation, is to consider the momentum drag to be a crude parameterization of the effects of transients propagating from the lower troposphere, rather than, for example, surface friction. (By experimenting with shallower layer depths, one might generate a more nonlinear circulation, but in our model this would be at the expense of having less realistic Rossby wave propagation from the extratropics.)
The wave activity budget of section 4b below shows that direct dissipation of the zonal mean part of the eddies by the damping terms does not prove to be an important difference between experiments in which the Hadley circulation is present and absent. This means that from the point of view of wave propagation, one need not be too concerned as to whether the damping values can be justified for the upper troposphere.
The Hadley circulation basic state, denoted by
b. Experiments without a Hadley circulation
c. Topographic forcing
d. Wave activity diagnostics
3. Results from the numerical experiments
a. Zonally symmetric forcing
Experiments were carried out with four different amplitudes of wave-3 forcing: ϵ = 0.001 (the linear case), 0.02, 0.05, and 0.2. Figure 2 shows PV snapshots from the ϵ = 0.05 experiments. The Hadley circulation experiment is on the left, and the no-Hadley circulation experiment on the right. This figure shows three main differences between the two experiments.
In the Hadley circulation case the waves are observed to break in a region displaced 5°–10° north of the wave-breaking region in the no-Hadley circulation case.
A mean gradient of PV is present in the wave-breaking region of the Hadley circulation experiment, in every snapshot, whereas in the no-Hadley circulation case the PV has become largely homogenized throughout the breaking region by day 20.
In the Hadley circulation case there appears to be a continuous cascade of enstrophy toward small scales (also noted in the experiments of HP), whereas once the PV has become well mixed in the critical layer region, the cascade slows dramatically when the Hadley circulation is absent.
Figure 3 shows the same snapshots for the ϵ = 0.2 experiments. These exhibit much the same differences as above. Greater wave amplitudes have led to a much broader wave-breaking region that becomes very well mixed when there is no Hadley circulation present. In the Hadley circulation case the wave-breaking region has an even broader extent, and there is continuous mixing of PV throughout it. A very strong gradient of PV develops in the extratropics by day 20, and there is breaking to the north of this gradient by day 30. Further snapshots reveal sporadic wave-breaking events when high PV fluid is stretched and mixed into the low-latitude turbulent region. The Hadley circulation experiment exhibits much more turbulent, oscillatory behavior compared with the no Hadley circulation case, which settles into a small-amplitude oscillation, with weak wave breaking into the well-mixed region.
Figure 4 shows the evolution in latitude and time of the zonal mean meridional wave activity flux
When ϵ = 0.02, Fig. 4 shows a clear qualitative effect of the presence of the Hadley circulation. When it is present (left panels), the flux is almost directly proportional to the linear flux, indicating near-complete absorption. When it is absent, the momentum fluxes first grow linearly, then decay away almost to zero, indicating that reflection is taking place from the critical layer region [see, e.g., the review in Haynes (1989)]. Note that the friction acting on the mean flow (4) is able to restore the PV gradients in the critical layer, allowing further absorption to take place. In the ϵ = 0.05 experiments the momentum fluxes in the Hadley circulation case remain reasonably close to the linear predictions, but with no Hadley circulation the reflection stage is reached more quickly (by day 18), and the critical layer is then observed to go through another cycle of absorption and reflection. At ϵ = 0.2 the presence of the Hadley circulation is not sufficient to absorb all of the incoming wave activity flux. In fact up until day 15 both experiments go through a similar-looking absorption–reflection cycle. After that, absorption is relatively weak in both experiments, but in the Hadley circulation case there are brief bursts of stronger absorption that correspond to wave-breaking events associated with the strong gradient of PV observed in the snapshots of Fig. 3.
b. Isolated mountain forcing
In the following set of experiments, the waves were forced by an isolated mountain situated at 45°N, 90°W. Amplitudes of ϵ = 0.001 (linear case), 0.05, 0.2, and 0.5 were used. In the linear experiments (not shown), wave activity fluxes showed two wave trains generated by the mountain. One propagated northeastward and remained trapped at high latitudes by the layer of strong westerlies. The other much stronger wave train propagated southeastward and was absorbed at the linear critical line appropriate to each experiment, as in the wave-3 case.
Figure 6a shows snapshots of both PV (left panels) and wave activity density and flux (right panels) for days 8, 16, and 24 of the ϵ = 0.05 Hadley circulation experiment. In the extratropics the patterns are much as they are in the linear case. There is no evidence of any significant reflection from the critical layer region. However, there is a strong westerly flux of wave activity within the critical layer itself. This westerly flux of wave activity is most easily explained by advection due to the mean flow, as the waves now break in a region of significant westerly flow due to the northward shift in the position of the linear critical line. This effect greatly influences the wave activity budget in low latitudes for these experiments (see section 4b) and is one of the most substantial differences between the Hadley circulation and no Hadley circulation cases. Figure 6b shows the same pictures for the no Hadley circulation case. In this case there is strong evidence for a reflected wave train that is completely absent in the corresponding linear experiment. This wave train returns to midlatitudes where it is refracted back southward until it propagates back into the critical layer region around 90° farther eastward. This is exactly the behavior reported in similar experiments by BH, which were entirely inviscid except for hyperdiffusion. Unlike the Hadley circulation experiment, wave activity becomes concentrated at the longitude where the wave train reaches the critical layer, with relatively little downstream advection, instead of being stretched out across a broad region.
Figure 7 shows the evolution of the perturbation streamfunction ψe in the ϵ = 0.05 experiments. The Hadley circulation case is on the left and the no Hadley circulation case is on the right. These snapshots are consistent with the wave activity snapshots of Fig. 6 and show a reflected wave train emerging from low latitudes in the no Hadley circulation case: this reflected wave train is entirely absent in the Hadley circulation case. (Note that the pictures are complicated slightly by the presence of the weaker wave train propagating northeast away from the mountain.) The reflected wave train propagates across midlatitudes before returning to low latitudes around 90° eastward of the point of entry of the incident wave train.
Figures 8a and 8b show snapshots from days 6, 12, and 18 of the ϵ = 0.2 experiments. Earlier times have been chosen as nonlinearity in both experiment sets is earlier due to the increased wave amplitude. Otherwise, the main effect of increasing the wave amplitude in the Hadley circulation case is to cause a weak reflected wave train to propagate out of the critical layer region. This wave train is noticeably stronger in the ϵ = 0.5 experiment (not shown), and this suggests that there is a limit to the magnitude of the incoming flux of wave activity that can be dissipated by the Hadley circulation or advected downstream before reflection must occur, as in the wave-3 case. In the no Hadley circulation case the wave activity flux pictures are similar to the ϵ = 0.05 experiment, although the reflected wave train is somewhat stronger. However, it is worth remarking that by day 18 there is considerable mixing to the east of the initial breaking region noticeable in the PV plot. This appears to be due to the reflected wave train reentering the critical layer at a longitude that is distinct from where the original wave train enters.
4. Discussion
a. Linear theory
Figure 9 shows how the solutions to (18) change as the zonal wind is varied. Figures 9a and 9b show γr = Reγ for the three cases
Consider first the case
In the case of southerly mean meridional wind
For the numerical experiments with the shallow-water model in the linear regime (ϵ = 0.001), using
Finally, in the northerly wind case
b. Nonlinear behavior
In order to gain a deeper understanding of the differences between the individual pairs of experiments, a wave activity budget was calculated for both the ϵ = 0.05 wave-3 and the ϵ = 0.2 isolated mountain sets of experiments. These two sets of experiments are chosen as the total wave activity in each is comparable. In the wave-3 case the budget was taken for the region south of 25°N. From the expression for
Figure 10b shows the same budget for the experiment without the Hadley circulation. In this experiment the total wave activity in the critical layer oscillates with a period of around 15 days. It is clear from the budget that this oscillation is entirely due to variations in the nonadvective component of the incoming flux, which can also been seen in Fig. 4. These variations in the incoming flux are consistent with periods of nonlinear reflection from the critical layer, and these result in the incoming flux falling far short of the linear prediction of 4.75 × 10−5 HΩ2a2. Dissipation is fairly constant and is due to the direct effect of friction on the eddies, and to hyperdiffusion, although the latter term is much weaker than in the Hadley circulation case. It appears that the rate of total dissipation is too slow to prevent the critical layer from becoming “saturated” with wave activity, thereby forcing reflection to take place, as must happen in the case of the simple, inviscid, barotropic critical layer discussed by Killworth and McIntyre (1985, section 6). By contrast, in the Hadley circulation case the rate of dissipation is sufficient to prevent such saturation. Perhaps surprisingly, most of this extra dissipation is achieved not through the direct effect of the Hadley cell forcing acting on the basic state, but instead it is due to the more vigorous cascade of enstrophy to small scales that leads to greater dissipation by the hyperdiffusion term. The proof in Killworth and McIntyre (1985, section 7) states that if dissipation terms act only as a flux of PV within a barotropic critical layer and do not diffuse vorticity into it, then the time-integrated absorptivity of the critical layer will remain bounded. The hyperdiffusion terms taken alone come close in practice to satisfying this condition in the shallow-water experiments.3 This suggests that the unbounded time-integrated absorptivity due to the hyperdiffusion terms in our experiments is entirely a consequence of the presence of the Hadley cell forcing and the eddy friction.
Figures 11a and 11b show the budgets for the ϵ = 0.2 isolated mountain experiments, for the Hadley circulation and no-Hadley circulation cases, respectively. In this case the budgets are for a region bounded to the north by the 25°N latitude circle and to the west and east by the 135°W and 45°E meridians, respectively. This allows us to assess the relative importance of downstream radiation of wave activity in the budgets of the two experiments.4 In the Hadley circulation experiment, after day 20, 50%–60% of the incoming wave activity flux is radiated away at the eastern boundary, making it by far the most important term in the budget. The eastern boundary flux was found to be dominated by the advective component, with the nonadvective component in the westward direction and equal to about 40% of the magnitude of the advective component. In the no-Hadley circulation experiment the total wave activity in the region was found to increase to almost double that in the Hadley circulation case, despite a weaker incoming flux at the northern boundary due to reflection. At the eastern boundary there is no westerly flux due to downstream advection; in fact, there is an easterly flux as the reflected wave train is refracted around, and enters, the region from the east (this can be seen in the day-18 panel of Fig. 8b).
In conclusion, therefore, the main effect of the Hadley circulation in the wave-3 case is the increase in the absorption of wave activity, primarily by increased dissipation due to an increase in the strength of the enstrophy cascade. In the isolated mountain case this increased dissipation, while still significant, is swamped by the larger effect of the downstream advection of wave activity in the Hadley circulation case, which can be explained by the waves breaking in a region of westerly winds.
5. Long-time behavior
The effect of the Hadley circulation on the longtime equilibrium behavior of this simple model is also of interest. Recently, Polvani et al. (1999) showed that, under certain conditions, a shallow-water flow without a Hadley circulation, in which waves are forced by an isolated mountain, can exhibit spontaneous low-frequency variability. If the mountain size is sufficiently large, then this variability typically involves repeated vortex shedding from the northern flank of the midlatitude jet, at some distance downstream of the mountain. The structure of the streamfunction anomaly associated with this vortex shedding is that of a wave train eminating from the Tropics. This wave train is qualitatively similar in structure to both atmospheric teleconnection patterns such as the PNA pattern (see, e.g., Frederiksen and Webster 1988) as well as the most unstable normal modes calculated when the equations of motion are linearized about the 300-mb time mean flow (see, e.g., Simmons et al. 1983).
The effect of the Hadley circulation on this type of oscillation was therefore investigated by integrating the ϵ = 0.5 isolated mountain experiments for 800 days and then analyzing the final 100 days of these integrations. For these long runs, the location of the top of the mountain was moved southward to 36°N, in order to cause the maximum possible interaction between the jet and the mountain. It was also found to be necessary to increase the magnitude of the hyperdiffusivity to 1 × 10−11 a6Ω, in order to maintain numerical stability. Otherwise, the experiments were identical to those described in sections 2–4.
In both experiments the flow remains unsteady throughout the period of integration. Figure 12 shows the time mean PV, averaged over the final 100 days, in both the Hadley circulation and the no-Hadley circulation cases. The main differences between these plots are that in the Hadley circulation experiment the PV north of 60°N is relatively well mixed, and south of 30°N in the critical layer region, wave breaking appears to be a much stronger ongoing process, especially between 90°W and 130°E. Both of these differences are consistent with those seen in the PV snapshots of the initial-value wave-3 experiments (Figs. 2 and 3).
Figure 13 shows the leading complex empirical orthogonal functions (EOFs) of the perturbation streamfunction field ψe (defined as the deviation from the time mean) as well as their complex principal components. The reader is referred to Polvani et al. (1999) for a description of how this method of analyzing the leading order variability of the streamfunction field is applied in these experiments, and to Horel (1984) for a more general review of complex EOFs. Figure 13a shows the Hadley circulation case. The leading order variability has the structure of a zonal wavenumber-4 wave train, which is fairly constant in amplitude around the globe, except for a minimum in the strong jet region just downstream of the mountain. The principal components indicate that the phase propagation is to the east and the period is around 10 days. The PV snapshots (not shown) show large waves in the tight midlatitude PV gradient that extend around the globe. This is in sharp contrast to Fig. 13b, showing the no Hadley circulation case, where a wave train extends from the Tropics at the opposite side of the globe, across the Pole to the mountain. In this case the phase propagation is out of the Tropics toward the mountain, and the period appears to be approximately 80 days. This type of wave train is very similar to that in several experiments reported in Polvani et al. The PV snapshots (not shown) show a small vortex being shed from the northern flank of the jet over this time. The total amplitude of this oscillation is substantially smaller than in the Hadley circulation case.
In the zonally symmetric wave-3 experiments of HP, the flow settles into a periodic oscillation above a certain forcing amplitude. They describe this oscillation as being due to an instability associated with the wave breaking. It seems likely that the strength of such an instability is proportional to the strength of the wave breaking itself, which in the longtime limit has been shown to be much stronger when the Hadley circulation is present. This is a possible explanation for the much greater longtime variability present in the Hadley circulation experiment, and this brief study seems to motivate a detailed investigation into this effect in a more realistic model.
6. Summary and conclusions
In this paper we have addressed the question of how a representation of the Hadley circulation affects the propagation and nonlinear reflection of planetary waves in the winter hemisphere of a shallow-water layer. This was achieved by comparing results from numerical experiments with initial flows that contained a forced Hadley circulation, with those from initial flows that were similar in every other detail but with no Hadley circulation.
Firstly, the absorption–reflection characteristics of the two flows were compared by forcing waves in the midlatitudes with a zonal wave-3 topography. In the linear limit, the presence of the Hadley circulation is found to displace the critical line where the waves are absorbed poleward, an effect that has been previously discussed by Schneider and Watterson (1984) and Farrell and Watterson (1985). At finite forcing amplitudes the waves are found to break in critical layer regions around their respective critical lines. The Hadley circulation flows remain in a regime of near-complete absorption of wave activity at the same forcing amplitudes at which the no Hadley circulation flows exhibited substantial nonlinear reflection. At high forcing amplitudes, the Hadley circulation flows would also reflect wave activity, but less efficiently than the corresponding no Hadley circulation flows. An analysis of the wave activity budgets in the critical layer regions shows that in the Hadley circulation flows, absorption is maintained through increased dissipation of wave activity in the wave-breaking region. This increased dissipation occurred principally due to an increased cascade of enstrophy to small scales, consistent with more sustained wave breaking. The wave breaking is sustained by the continual restoration of the basic-state PV gradient in the wave-breaking region due to the presence of the Hadley circulation, and this prevents the critical layer region from becoming well mixed and, hence, from becoming an efficient reflector.
The behavior of wave trains generated in the extratropics by an isolated mountain was then investigated. In the absence of the Hadley circulation, and at sufficient forcing amplitudes, zonally localized wave breaking led to the wave train being reflected at the same longitude at which it entered the critical layer region. Similar experiments showing localized nonlinear reflection have been previously reported by BH and WPP. In the Hadley circulation case the behavior is quite different. Not only was there increased dissipation of wave activity in the critical layer region, as in the wave-3 experiments, but there was also a considerable westerly flux of wave activity within the critical layer. Further analysis has shown that this flux is largely advective and, therefore, is largely due to the fact that the waves tend to break in relatively strong westerly flow when the Hadley circulation is present. Wave activity budgets of a longitudinally localized region show that the westerly flux is the main contributor to the enhanced absorptivity, at least before the waves have propagated around the globe and entered the region from the west. The importance of the location of the critical line to the wave activity budget motivated the analysis of section 4a where we obtained a simple expression for the minimum value of the zonal mean wind that allowed southward linear propagation in a quasigeostrophic β-plane flow when
The two effects of enhanced dissipation due to the Hadley circulation and the northward displacement of the critical layer into the westerly flow lead us to conclude that the reflection of planetary wave trains from the Tropics is unlikely in the winter hemisphere, at least at those longitudes where conditions do not greatly differ from the zonal mean. Of course, conditions in the real atmosphere are greatly complicated by further mixing due to transient waves, as well as by numerous other nonconservative effects, and GCM experiments with a realistic flow would be necessary to accurately quantify the importance of the Hadley cell in the planetary wave activity budget (see, e.g., Cook and Held 1992). Such experiments could also be used to investigate the summer hemisphere case, as a reasonable summer hemisphere basic state was not obtainable in our model using simple forcing of the form (2).5 The summer hemisphere problem is interesting as the meridional wind then allows linear propagation across easterly layers into the opposite hemisphere (e.g., Schneider and Watterson 1984; Watterson and Schneider 1987). Determining the extent to which this propagation can take place when the waves have finite amplitude is an important theoretical point that could further underpin our understanding of observed cross-hemispheric Rossby wave propagation (e.g., Tomas and Webster 1994). WPP showed that finite-amplitude waves can break even when there are weak westerlies at the equator, and it seems likely that the amplitude at which this breaking occurs would be strongly influenced by the direction and magnitude of the meridional winds at low latitudes.
Acknowledgments
This work was supported by the National Science Foundation, under Grant ATM9528471 to MIT and Grant ATM9527315 to Columbia University.
REFERENCES
Abramowitz, M., and I. E. Stegun, 1972: Handbook of Mathematical Functions. National Bureau of Standards, 106 pp.
Brunet, G., and P. H. Haynes, 1996: Low-latitude reflection of Rossby wave trains. J. Atmos. Sci.,53, 482–496.
Cook, K. H., and I. M. Held, 1992: The stationary response to large-scale orography in a general circulation model and a linear model. J. Atmos. Sci.,49, 525–539.
Farrell, B., and I. G. Watterson, 1985: Rossby waves in opposing currents. J. Atmos. Sci.,42, 1746–1755.
Frederiksen, J. S., and P. J. Webster, 1988: Alternative theories of atmospheric teleconnections and low-frequency fluctuations. Rev. Geophys.,26, 459–494.
Grose, W. L., and B. J. Hoskins, 1979: On the influence of orography on large-scale atmospheric flow. J. Atmos. Sci.,36, 223–234.
Hack, J. J., and R. Jacob, 1992: A global shallow water model based on the spectral transform method. NCAR Tech. Note NCAR/TN-343+STR, 39 pp.
Haynes, P. H., 1988: Forced, dissipative generalizations of finite-amplitude wave-activity conservations for zonal and nonzonal basic flows. J. Atmos. Sci.,45, 2352–2362.
——, 1989: The effect of barotropic instability on the nonlinear evolution of a Rossby wave critical layer. J. Fluid Mech.,207, 231–266.
Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci.,37, 515–533.
——, and P. J. Phillips, 1990: A barotropic model of the interaction between the Hadley cell and a Rossby wave. J. Atmos. Sci.,47, 856–869.
Horel, J. D., 1984: Complex principal component analysis: Theory and examples. J. Climate Appl. Meteor.,23, 1660–1673.
Hoskins, B. J., and D. J. Karoly, 1981: The steady linear response of a spherical atmosphere to thermal and orographic forcing. J. Atmos. Sci.,38, 1179–1196.
——, A. J. Simmons, and D. G. Andrews, 1977: Energy dispersion in a barotropic atmosphere. Quart. J. Roy. Meteor. Soc.,103, 553–567.
Karoly, D. J., R. A. Plumb, and M. Ting, 1989: Examples of the horizontal propagation of quasi-stationary waves. J. Atmos. Sci.,46, 2802–2811.
Killworth, P. D., and M. E. McIntyre, 1985: Do Rossby-wave critical layers absorb, reflect or overreflect? J. Fluid Mech.,161, 449–492.
Lindzen, R. S., and A. Y. Hou, 1988: Hadley circulations for zonally averaged heating centered off the equator. J. Atmos. Sci.,45, 2416–2427.
Magnusdottir, G., and P. H. Haynes, 1999: Reflection of planetary waves in three-dimensional tropospheric flows. J. Atmos. Sci.,56, 652–670.
Plumb, R. A., 1985: On the three-dimensional propagation of stationary waves. J. Atmos. Sci.,42, 217–229.
Polvani, L. M., J. G. Esler, and R. A. Plumb, 1999: Time variability and Simmons–Wallace–Branstator instability in a simple nonlinear one-layer model. J. Atmos. Sci.,56, 1445–1460.
Schneider, E. K., and I. G. Watterson, 1984: Stationary Rossby wave propagation through easterly layers. J. Atmos. Sci.,41, 2069–2083.
Simmons, A. J., J. M. Wallace, and G. W. Branstator, 1983: Barotropic wave propagation and instability, and atmospheric teleconnection patterns. J. Atmos. Sci.,40, 1363–1392.
Tomas, R. A., and P. J. Webster, 1994: Horizontal and vertical structure of cross-equatorial wave propagation. J. Atmos. Sci.,51, 1417–1430.
Watterson, I. G., and E. K. Schneider, 1987: The effect of the Hadley circulation on the meridional propagation of stationary waves. Quart. J. Roy. Meteor. Soc.,113, 779–813.
Waugh, D. W., L. M. Polvani, and R. A. Plumb, 1994: Nonlinear, barotropic response to a localized topographic forcing: Formation of a “tropical surf zone” and its effect on interhemispheric propagation. J. Atmos. Sci.,51, 1401–1416.
Webster, P. J., 1981: Mechanisms determining the atmospheric response to sea surface temperature anomalies. J. Atmos. Sci.,38, 544–571.
However, Magnusdottir and Haynes (1999) have recently shown that nonlinear reflection is not inhibited when thermal damping with a timescale as low as 5 days is acting on the waves, in the context of more realistic three-dimensional flows.
Note that we are using a different sign convention from BH for consistency with other pseudomomentum fluxes, for example, Plumb (1985).
Although it should be noted for this argument to strictly apply it must be assumed that under the conditions of balanced flow discussed in BH, that the Killworth and McIntyre proof applies to the shallow-water system that we are considering.
Note that in this budget we are showing the sum of the advective and nonadvective components of the meridional flux, rather than showing them separately as in Fig. 10.
However, one could take the approach of, for example, Schneider and Watterson (1984) and specify the desired flow, from which the forcing can then be calculated.