1. Introduction
The inverse cascade, or negative viscosity, is a characteristic feature of two-dimensional turbulence. Our interest here is the arrest of this cascade on a β plane by the spontaneous formation of zonal jets. The early turbulence simulations of Rhines (1975) and Williams (1978) showed that the β effect prevents large meridional (i.e., “cross-β”) fluid excursions. Instead, zonally elongated, persistent flows extend unimpeded across the computational domain. An important and realistic feature of these anisotropic velocity fields is that there is also a strong asymmetry between eastward and westward flow: there are narrow eastward jets separated by broader regions of slower westward flow. This effect is particularly striking in the simulations of baroclinic turbulence reported by Panetta (1993) [see also the review by Rhines (1994)].
Another interesting feature of Panetta’s simulations is that, although arising from active baroclinic turbulence, these zonal jets have an “almost barotropic” character. It seems plausible that the baroclinic eddies are effectively a random, small-scale forcing for the barotropic mode (as envisaged by Salmon 1980 and Williams 1978). Recently, several investigators have taken this approach either on a β plane (Vallis and Maltrud 1993) or on a sphere (Nozawa and Yoden 1997; Huang and Robinson 1998). A remarkable aspect of the simulations in these three papers is that the large-scale jets are virtually steady even though the system is forced stochastically at small scales. Huang and Robinson emphasize another equally remarkable fact: the energy transfer from the small-scale forcing to the jets is spectrally nonlocal; there is no evidence of a local upscale energy cascade in wavenumber space.
In this article we prescribe deterministic small-scale forcing, which drives a steady sinusoidal shear in the meridional direction, and view the problem as one of weakly nonlinear hydrodynamic stability rather than turbulence phenomenology.
An analytic path is provided by studying the nonlinear evolution of the first instabilities that develop on this small-scale flow as the Reynolds number is increased. Nepomnyashchy (1976) and Sivashinsky (1985) realized that these instabilities have a much larger length scale than that of the forced sinusoidal flow; the nonlocal energy transfer reported by Huang and Robinson (1998) is equivalent to this scale separation. Indeed, Sivashinsky’s expansion accesses an interesting nonlinear regime because it is primarily based on scale separation rather than on weak nonlinearity.
We follow this path, incorporating three additional features of geophysical interest: the β effect, a uniform mean flow, and bottom drag. The special case in which the basic sinusoidal flow is a stationary (but inviscid) Rossby wave has been studied by Lorenz (1972) and Gill (1974). The inclusion of viscosity makes the problem easier because one can control the instability by operating just past the stability boundary with a slightly supercritical Reynolds number.
The studies by Vallis and Maltrud (1993), Nozawa and Yoden (1997), and Huang and Robinson (1998) all use stochastic forcing, while in our model the forcing is deterministic and steady. The advantage of the deterministic approach is that many aspects of the problem are captured by a tractable and mechanistic model. We show how large-scale instabilities lead to the formation of slowly evolving zonal jets, which are similar to those observed in the turbulence simulations described above and in a number of geophysical circumstances. Specifically, there are narrow and fast eastward jets separated by broader and slower regions of westward flow. We also stress the importance of the bottom drag in the ultimate selection of the spatial scale of the zonal jets. With no bottom friction, the zonal jets very slowly migrate and merge until only one eastward jet remains in the domain of integration. This scale increase is a type of “one-dimensional” inverse cascade. With nonzero bottom drag, the scale increase is arrested and the typical distance between eastward jets of this asymptotic state is proportional to (bottom drag)−1/3.
In section 2 we introduce the analytical model and derive the amplitude equation for the leading order perturbation. In section 3 we obtain analytical results for zonally uniform solutions of the amplitude equations. Through various numerical computations we study the formation and evolution of zonal flows in section 4. Section 5 is a discussion of the results and their generality.
2. The amplitude equation


If (
Frisch et al. (1996) have recently studied the case in which the sinusoidal part of the base-state flow is parallel to the planetary vorticity contours (i.e., α = π/2 and


The advection term involving




Although our itinerary largely follows that of Sivashinsky, there are some interesting differences in the scenery, including the appearence of a subcritical transition. We sketch the development and relegate the details to appendix A.
One substitutes ψ = ψ0 + ϵψ1 + · · · into (2.3) and collects powers of ϵ. It is important to note that the leading term ψ0 is not small relative to the basic-state streamfunction Ψ0. The expansion works because of the spatial-scale separation of the basic flow from the perturbation.
The alternative to (2.9b) is to admit evolution on the relatively fast timescale ϵ3t. But the resulting amplitude equation is an ill-posed nonlinear diffusion equation. With these considerations we are glimpsing the possibility of subcritical instability. The parameter restriction in (2.9b) is required to control this subcriticality so that it is described by the amplitude equation in (2.12) below.
If one wants to consider a basic state with
Although the expansion at first looks more complicated than that of Sivashinsky (1985), because of (2.9b) many of the expressions simplify and the calculation is comparably simple. It is also remarkable that the condition in (2.9b) ensures that the basic state in (2.2) is, to leading order, the stationary Rossby wave whose inviscid stability was considered by Lorenz (1972) and Gill (1974).




3. Zonally uniform solutions of the amplitude equation


The amplitude equation for the perturbation streamfunction A in (3.2) has a negative viscosity term, −rAηη, and also a stabilizing hyperviscosity, −3Aηηηη. The parameter γ (the effective planetary vorticity gradient) in (3.1a) plays two roles. First, γ raises the critical Reynolds number so that if γ2 > 2 (i.e., if r < 0, so the viscosity is positive) then A = 0 is a linearly stable solution. Second, because of the quadratic term 2γ(
In (3.1) the effects of the net advection
a. The Lyapunov functional






b. Linear stability of the trivial solution A = 0




c. Weakly nonlinear development of the instabilities




If γ <
d. Expansion around the critical point
The Lyapunov functional, V[A] in (3.3), ensures that U(η, τ) cannot grow indefinitely, even if ℓ < 0; it is just that the saturation is beyond the scope of the approximation in (3.10a). One can capture the saturation of the subcritical instability by expanding around the critical value γ =
4. An inverse cascade through merger of zonal jets
a. The case with no bottom drag: μ = 0




The solution in Fig. 3 tends to increase its dominant length scale. This inverse cascade occurs by the successive mergers of westward flow regions with the absorption of the eastward jet that separates them. An example of this can clearly be seen between τ = 500 and τ = 600. The next eastward jet that will be absorbed is the one centered at around η = 20. This is an increasingly slow process because the eastward jets interact through exponentially small tails. So, the larger the spacing of the jets, the smaller the interaction and the longer it will take for the next merger. In Fig. 3, it is very difficult to notice that the jet centered around η = 20 has indeed decreased in size between τ = 600 and τ = 2000. Figure 4a is in this respect more clear: it shows the time evolution of U as a contour plot in the (τ, η) plane. In this figure one can also clearly see the disappearance of the jet centered around η = 75 at times between τ = 500 and τ = 600.




During these merger events, the final term in (4.3) is greatly reduced because two jet boundaries with large shear, Uη ≫ 1, are eliminated. Intuitively, one can view the final term in (4.3) as a penalty associated with having a boundary or interface between eastward and westward flow. By merging jets and eliminating these shear zones, one is reducing this penalty. The best that one can do in this respect, while observing the momentum constraint in (4.2), is to produce one contiguous region of eastward flow (bounded by two shear zones) in each period; that is the ultimate state, with one eastward jet, toward which the solution in Figs. 3 and 4 is very slowly heading.




As remarked by an anonymous reviewer, the substitution Ũ ≡ U − γ eliminates the quadratic nonlinearity and so puts (4.1) in the form of a “Cahn–Hilliard equation” with symmetric potential. Then the perturbation schemes developed by Kawasaki and Ohta (1982) and Fraerman et al. (1997) can be used to obtain analytic expressions for the motion of “kinks” (which correspond to the boundaries between easterlies and westerlies in this work).
b. The case with small bottom friction






c. The case with “strong” bottom friction
The scaling law L ∝ μ−1/3 is difficult to verify numerically because as μ is decreased the adjustment times become very long, and because the domain size Λ eventually intrudes, since in our numerical integrations the condition L ≪ Λ is not satisfied. However, numerical calculations clearly show that the ultimate spacing of the jets decreases as μ is increased with γ fixed; the story is complicated by the qualitative changes that occur depending on whether γ is fixed at a value for which the instability is supercitical (γ <
Figure 8 shows the profile of the zonal velocity U at τ = 2000 for various runs with γ = 0.5 <
Increasing the bottom drag μ with γ = ½ in Fig. 2 means that one eventually passes over the stability boundary μc in (3.6) (i.e., the solid curve in Fig. 2). For γ = 0.5, μc ≈ 0.255 21, and Fig. 8 shows how the jets smoothly disappear by reducing their amplitude as μ → μc. Indeed, from (3.8a) and (3.8b) we see that the jet amplitude scales as
The above results are for a value of γ that is less than the critical value γ∗ =






Even though μc = 0.083 33, the case with μ = 0.1 in Fig. 10 is not qualitatively different from the one with μ = 0.08. In these cases the finite amplitude solution in Fig. 10 is globally stable (i.e., μm > 0.1). For μ = 0.12 and 0.14 the zonal velocity has a pulselike structure with localized regions of eastward and westward flow separated by regions of zero velocity. Computations show that the number and position of these structures depend on the initial conditions. Therefore, from solutions of the initial value problem it is not obvious if the finite amplitude solution is globally stable or metastable in these cases. An answer is obtained by numerically evaluating the Lyapunov functional in (4.8). The solution A = 0 obviously has V[0] = 0. So if the finite amplitude solution has V[A] < 0, then it is globally stable (i.e., it is more stable than A = 0). And if V[A] > 0, then the finite amplitude solution is metastable. In Fig. 10, the Lyapunov functional has a negative value for the run with μ = 0.12, which shows that the finite amplitude solution is globally stable. On the other hand, the Lyapunov functional is positive for the run with μ = 0.14, which shows that the finite amplitude solution is metastable in this case. Therefore, for γ = 1, we have 0.12 < μm < 0.14. The different sign of the Lyapunov functional for these two values of μ has been verified by various runs with different initial conditions.
d. Reconstruction of the flow
To this point we have visualized the solution by dealing mostly with the zonal velocity U = −Aη. However, this is only the first term in the amplitude expansion. It is interesting to reconstruct the two-dimensional field by calculating Ψ + ψ0 + ϵψ1 + ϵ2ψ2 + ϵ3ψ3 using the expressions for ψn in appendix A and with Ψ given by (2.2). The supercriticality parameter ϵ defined in (2.4) does not appear in the amplitude equation. However, in order to reconstruct the total streamfunction in Fig. 12, it is necessary to specify a value of ϵ. We take ϵ = ½; this is a tradeoff between a convincingly small ϵ and a value that is large enough to show the main effects of the instability. (If ϵ is too small, the large separation of scale between the zonal and meridional directions gives even narrower zonal oscillations than in Fig. 12.)
Figure 12 shows a case with basic flow Ψ = −Ψ0 cosmx. Because
5. Discussion and conclusions
Earlier studies have shown that stochastically forced β plane motion evolves into strong, quasi-steady zonal flows. The goal of the present study has been to develop a mechanistic model of these jets. Because the model in (2.12) involves at least three interesting nondimensional parameters (viz., the effective β parameter, the bottom drag μ, and the domain size Λ), we do not claim to have exhausted this problem even in the special case (3.2) in which there are no zonal variations.
Our approach, which is based on the parametric assumption that the small-scale forcing is only slightly supercritical, cannot access the eddying regime that is the focus of recent works such as Vallis and Maltrud (1993), Nozawa and Yoden (1997), and Huang and Robinson (1998). Yet our results are similar to those found in these earlier studies; our system develops a number of very slowly evolving eastward jets separated by wider regions of westward flow.
Another point of comparison with simulations is that, in the expansion scheme, the relative vorticity gradient of the zonal mean flow is O(ϵ2) smaller than the planetary vorticity gradient β. (And so the jets we describe are all strongly stable by the Rayleigh–Kuo condition.) An analogous inequality is characteristic of the numerical simulations referenced above. In all three cases the β effect is stronger than the vorticity gradients of the zonal mean flow. (For instance, Fig. 10 of Vallis and Maltrud shows that
One qualitative effect that has not been seen in numerical simulations, but that is strongly suggested by our results, is that jets with the characteristic east–west asymmetry should form if β = 0 and
The very slow evolution of the zonal jets is a problem for direct numerical simulations of the β plane turbulence problem. We are fortunate that there is a Lypunov functional that shows that (3.2) is evolving to a steady state, that is, the long time behavior of (3.2) is easily understood. For example, we know in advance that the jets cannot perpetually migrate back and forth across the domain. [Panetta (1993) suggests that this happens in his baroclinic system.] Rather, the final location of a jet in our model is merely an accident of the initial conditions. One is left to wonder how (or if) this variational structure is lost as more physics is systematically included. For instance, we have not been able to obtain a variational principle for the system (2.12), which retains the zonal structure of the perturbation (Aξ ≠ 0). And one expects that increasing the Reynolds number of the small-scale forcing will surely result in more interesting β plane–jet dynamics.
Another possibility is that, even if the deterministic part of the dynamics is variational, the small-scale turbulence also provides a fluctuating force that may cause the jets to meander in a Brownian fashion; when the jets are widely separated they interact only through the overlap of exponentially small tails.3 The weak interaction of well-separated jets could be overwhelmed by small noise. This does not seem to be the case in the simulations by Vallis and Maltrud (1993), Nozawa and Yoden (1997), and Huang and Robinson (1998), who observe quasi-steady jets despite the stochastic forcing. In fact, these authors note that the observed jets are steadier when the separation between the forcing scale and the jet scale is larger.
We have also found that bottom drag is important in selecting the spatial separation of the jets in our model. If the bottom drag is zero, then the jets very slowly migrate and merge until only a single region of eastward flow remains in the domain. Increasing the bottom drag arrests this one-dimensional inverse cascade and results in a periodic array of jets; the stronger the bottom drag, the smaller the jet spacing. Because increasing bottom drag also decreases the zonal velocities, weaker zonal flows are associated with smaller jet spacing (e.g., see Fig. 8). However, this is not an endorsement of Rhines’s
Acknowledgments
We would like to thank Neil Balmforth and Paola Cessi for useful conversations on this problem. This research is supported by the National Science Foundation Grant OCE-9529824.
REFERENCES
Balmforth, N. J., 1995: Solitary waves and homoclinic orbits. Annu. Rev. Fluid Mech.,27, 335–373.
Chapman, C. J., and M. R. E. Proctor, 1980: Nonlinear Rayleigh-Bénard convection with poorly conducting boundaries. J. Fluid Mech.,101, 759–782.
Fraerman, A. A., A. S. Mel’nikov, I. M. Nefedov, I. A. Shereshevskii, and A. V. Shpiro, 1997: Nonlinear relaxation dynamics in decomposing alloys: One-dimensional Cahn-Hilliard model. Phys. Rev. B,55, 6316–6323.
Frisch, U., B. Legras, and B. Villone, 1996: Large-scale Kolmogorov flow on the beta-plane and resonant wave interactions. Physica D,94, 36–56.
Gill, A. E., 1974: The stability of planetary waves on an infinite beta-plane. Geophys. Fluid Dyn.,6, 29–47.
Huang, H. P., and A. Robinson, 1998: Two-dimensional turbulence and persistent zonal jets in a global barotropic model. J. Atmos. Sci.,55, 611–632.
Kawasaki, K., and T. Ohta, 1982: Kink dynamics in one-dimensional nonlinear systems. Physica,116A, 573–593.
Lorenz, E. N., 1972. Barotropic instability of Rossby wave motion. J. Atmos. Sci.,29, 258–269.
Meshalkin, L. D., and I. G. Sinai, 1961: Investigation of the stability of a stationary solution of a system of equations for the plane movement of an incompressible viscous fluid. Appl. Math. Mech.,24, 1700–1705.
Nepomnyashchy, A. A., 1976: On the stability of the secondary flow of a viscous fluid in an infinite domain. Appl. Math. Mech.,40, 836–841.
Nozawa, T., and S. Yoden, 1997: Formation of zonal band structures in forced two-dimensional turbulence on a rotating sphere. Phys. Fluids,9, 2081–2093.
Panetta, R. L., 1993: Zonal jets in wide baroclinically unstable regions: Persistence and scale selection. J. Atmos. Sci.,50, 2073–2106.
Rhines, P. B., 1975: Waves and turbulence on a β-plane. J. Fluid Mech.,69, 417–443.
——, 1994: Jets. Chaos,4, 313–339.
Salmon, R., 1980: Baroclinic instability and geostrophic turbulence. Geophys. Astrophys. Fluid Dyn.,15, 167–211.
Sivashinsky, G. I., 1985: Weak turbulence in periodic flows. Physica D,17, 243–255.
Thual, O., and S. Fauve, 1988: Localized structures generated by subcritical instabilities. J. Phys.,49, 1829–1833.
Vallis, G. K., and M. Maltrud, 1993. Generation of mean flow and jets on a beta-plane and over topography. J. Phys. Oceanogr.,23, 1346–1362.
Williams, G. P., 1978: Planetary circulations: I. Barotropic representation of Jovian and terrestrial turbulence. J. Atmos. Sci.,35, 1399–1426.
APPENDIX A
Derivation of the Amplitude Equation
























APPENDIX B
Weakly Nonlinear Analysis























The linear dispersion relation (3.5a). Wavenumbers between k1 and k2 have positive growth rate (σ > 0) and are unstable. The most unstable wavenumber is kc =
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The linear dispersion relation (3.5a). Wavenumbers between k1 and k2 have positive growth rate (σ > 0) and are unstable. The most unstable wavenumber is kc =
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
The linear dispersion relation (3.5a). Wavenumbers between k1 and k2 have positive growth rate (σ > 0) and are unstable. The most unstable wavenumber is kc =
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The area below the solid curve is the region of the (γ, μ) plane in which the trivial solution A = 0 is linearly unstable. For γ >
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The area below the solid curve is the region of the (γ, μ) plane in which the trivial solution A = 0 is linearly unstable. For γ >
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
The area below the solid curve is the region of the (γ, μ) plane in which the trivial solution A = 0 is linearly unstable. For γ >
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The zonal velocity U = −Aη at six different times indicated near the upper right-hand corner of each panel. U is obtained by numerically integrating (4.1) with periodic boundary conditions. The domain size Λ is 100 and γ = 1. UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The zonal velocity U = −Aη at six different times indicated near the upper right-hand corner of each panel. U is obtained by numerically integrating (4.1) with periodic boundary conditions. The domain size Λ is 100 and γ = 1. UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
The zonal velocity U = −Aη at six different times indicated near the upper right-hand corner of each panel. U is obtained by numerically integrating (4.1) with periodic boundary conditions. The domain size Λ is 100 and γ = 1. UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

(a) The history of the solution shown in Fig. 3; contours of constant zonal velocity U in the (τ, η) plane. (b) The evolution of the Lyapunov functional. The sudden drop near τ = 550 corresponds to the elimination of an eastward jet.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

(a) The history of the solution shown in Fig. 3; contours of constant zonal velocity U in the (τ, η) plane. (b) The evolution of the Lyapunov functional. The sudden drop near τ = 550 corresponds to the elimination of an eastward jet.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
(a) The history of the solution shown in Fig. 3; contours of constant zonal velocity U in the (τ, η) plane. (b) The evolution of the Lyapunov functional. The sudden drop near τ = 550 corresponds to the elimination of an eastward jet.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The potential W(U) must adjust so that the two wells at UE and UW have equal depths [as in (b)]. This fixes the value of C as in (4.5). In (a) the value of C is zero.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The potential W(U) must adjust so that the two wells at UE and UW have equal depths [as in (b)]. This fixes the value of C as in (4.5). In (a) the value of C is zero.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
The potential W(U) must adjust so that the two wells at UE and UW have equal depths [as in (b)]. This fixes the value of C as in (4.5). In (a) the value of C is zero.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The zonal velocity U at six different times indicated near the upper right-hand corner of each panel. Here, U is obtained by numerical integration of (4.7) with domain size Λ = 100 and periodic boundary conditions. In this solution γ = 1 and μ = 0.01. The initial condition is similar to the one shown in Fig. 3. In this case the inverse cascade is arrested by bottom drag. The result is a periodically spaced set of eastward jets. Here, UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

The zonal velocity U at six different times indicated near the upper right-hand corner of each panel. Here, U is obtained by numerical integration of (4.7) with domain size Λ = 100 and periodic boundary conditions. In this solution γ = 1 and μ = 0.01. The initial condition is similar to the one shown in Fig. 3. In this case the inverse cascade is arrested by bottom drag. The result is a periodically spaced set of eastward jets. Here, UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
The zonal velocity U at six different times indicated near the upper right-hand corner of each panel. Here, U is obtained by numerical integration of (4.7) with domain size Λ = 100 and periodic boundary conditions. In this solution γ = 1 and μ = 0.01. The initial condition is similar to the one shown in Fig. 3. In this case the inverse cascade is arrested by bottom drag. The result is a periodically spaced set of eastward jets. Here, UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

(a) The complete history of the solution shown in Fig. 6; curves of constant U in the (τ, η) plane. (b) The evolution of the Lyapunov functional. As in Fig. 4, the removal of eastward jets produces sudden drops in the Lyapunov functional.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

(a) The complete history of the solution shown in Fig. 6; curves of constant U in the (τ, η) plane. (b) The evolution of the Lyapunov functional. As in Fig. 4, the removal of eastward jets produces sudden drops in the Lyapunov functional.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
(a) The complete history of the solution shown in Fig. 6; curves of constant U in the (τ, η) plane. (b) The evolution of the Lyapunov functional. As in Fig. 4, the removal of eastward jets produces sudden drops in the Lyapunov functional.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Five different numerical solutions of (4.7), all with γ = ½, and various values of the bottom drag μ indicated at the top of each panel. The upper-left panel is the initial condition. The results at time τ = 2000 are shown in the other five panels. For μ = 0 the inverse cascade is slowly reducing the number of jets (three at τ = 2000 in the middle left panel). As μ is increased, the number of jets increases. Note that the scale for U in the lower-right panel is different from the other panels. In this last panel the value of μ is slightly below the critical value for linear stability so the amplitude of the solutuion is very small. UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Five different numerical solutions of (4.7), all with γ = ½, and various values of the bottom drag μ indicated at the top of each panel. The upper-left panel is the initial condition. The results at time τ = 2000 are shown in the other five panels. For μ = 0 the inverse cascade is slowly reducing the number of jets (three at τ = 2000 in the middle left panel). As μ is increased, the number of jets increases. Note that the scale for U in the lower-right panel is different from the other panels. In this last panel the value of μ is slightly below the critical value for linear stability so the amplitude of the solutuion is very small. UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
Five different numerical solutions of (4.7), all with γ = ½, and various values of the bottom drag μ indicated at the top of each panel. The upper-left panel is the initial condition. The results at time τ = 2000 are shown in the other five panels. For μ = 0 the inverse cascade is slowly reducing the number of jets (three at τ = 2000 in the middle left panel). As μ is increased, the number of jets increases. Note that the scale for U in the lower-right panel is different from the other panels. In this last panel the value of μ is slightly below the critical value for linear stability so the amplitude of the solutuion is very small. UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

For μ slightly less than μc, the amplitude of the stable solution scales as
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

For μ slightly less than μc, the amplitude of the stable solution scales as
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
For μ slightly less than μc, the amplitude of the stable solution scales as
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Six different solutions for U(η, τ) calculated numerically from (4.7), all with γ = 1 and various values of the bottom drag μ (indicated at the top of each panel) at time τ = 2000. The initial condition is similar to that in Fig. 8. Here, UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Six different solutions for U(η, τ) calculated numerically from (4.7), all with γ = 1 and various values of the bottom drag μ (indicated at the top of each panel) at time τ = 2000. The initial condition is similar to that in Fig. 8. Here, UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
Six different solutions for U(η, τ) calculated numerically from (4.7), all with γ = 1 and various values of the bottom drag μ (indicated at the top of each panel) at time τ = 2000. The initial condition is similar to that in Fig. 8. Here, UE and UW, calculated from (4.6), are indicated by the vertical lines in each panel.
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Bifurcation diagram in the case γ >
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Bifurcation diagram in the case γ >
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
Bifurcation diagram in the case γ >
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Results at τ = 2000 for a run with
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2

Results at τ = 2000 for a run with
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
Results at τ = 2000 for a run with
Citation: Journal of the Atmospheric Sciences 56, 5; 10.1175/1520-0469(1999)056<0784:SEOZJO>2.0.CO;2
We have also included an extra term, CU(η, τ), in (4.4); the constant C is a Lagrange multiplier that is used to enforce the momentum constraint in (4.2). Also, again because of (4.2), CU does not change the value of the functional V[U].
Notice that if γ is negative, then the eastward flow is wider and slower than the westward flow. This effect might be seen in numerical simulations if one added a net advection
This observation is systematically exploited by the method of Kawasaki and Ohta (1982) and Fraerman et al. (1997); see also the review by Balmforth (1995). These techniques allow one to access the slowly evolving regime with modest computational resources.