1. Introduction
Zonal flows are banded, anisotropic, weakly fluctuating alternating jets that form spontaneously and persist indefinitely in an otherwise turbulent plasma or planetary fluid (Diamond et al. 2005; Vasavada and Showman 2005). The subject started with Rhines’ discovery that freely evolving barotropic β-plane turbulence transfers energy into zonal shear modes with zero frequency (Rhines 1975). Also in 1975, experiments by Whitehead showed that forcing, without the exertion of azimuthal torque, in a rapidly rotating basin produces prograde jets; in this context the curved upper surface provides an analog of the β effect. We follow Galperin et al. (2006) in referring to the development and persistence of these anisotropic planetary flows as “zonation.” Williams (1978) showed that zonation occurs in statistically steady forced-dissipative flows on the sphere and proposed this as an explanation of the banded structure of the planetary circulations of Jupiter and Saturn.
Figure 1 shows a typical example of fully developed, forced and dissipative zonation obtained by numerical solution of (3) below. The main features of the statistically steady flow, such as the sharp eastward jets, the broader westward return flows, and the sawtooth relative vorticity, are familiar from many earlier studies of statistically steady, stochastically forced, dissipative β-plane turbulence in doubly periodic geometry (Danilov and Gurarie 2004; Danilov and Gryanik 2004; Maltrud and Vallis 1991; Smith 2004; Vallis and Maltrud 1993) and on the sphere (Williams 1978; Nozawa and Yoden 1997; Huang and Robinson 1998; Scott and Polvani 2007).

Nonlinear (NL) zonal jets. (left) A snapshot of the zonally averaged velocity U(y, t) obtained from a solution of (3) in a doubly periodic domain 2πL × 2πL with kfL = 32, where kf is the dominant wavenumber of the forcing ξ. (right) A snapshot of the vorticity ζ, with overlaid zonally averaged vorticity −Uy(y, t) (solid white curve). The parameter values for this run are μ* = 0.018 24 and β* = 1.0. The snapshot is at 2μt = 25 with spinup from rest.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Nonlinear (NL) zonal jets. (left) A snapshot of the zonally averaged velocity U(y, t) obtained from a solution of (3) in a doubly periodic domain 2πL × 2πL with kfL = 32, where kf is the dominant wavenumber of the forcing ξ. (right) A snapshot of the vorticity ζ, with overlaid zonally averaged vorticity −Uy(y, t) (solid white curve). The parameter values for this run are μ* = 0.018 24 and β* = 1.0. The snapshot is at 2μt = 25 with spinup from rest.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Nonlinear (NL) zonal jets. (left) A snapshot of the zonally averaged velocity U(y, t) obtained from a solution of (3) in a doubly periodic domain 2πL × 2πL with kfL = 32, where kf is the dominant wavenumber of the forcing ξ. (right) A snapshot of the vorticity ζ, with overlaid zonally averaged vorticity −Uy(y, t) (solid white curve). The parameter values for this run are μ* = 0.018 24 and β* = 1.0. The snapshot is at 2μt = 25 with spinup from rest.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1








We assume that the forcing, a in (1) and ξ in (3), is a rapidly decorrelating, isotropic, spatially homogeneous, random process. Thus energy and enstrophy are injected into a narrow band of wavenumbers centered on a “forced wavenumber” kf (see appendix A for details of the implementation). This model of exogenous stochastic forcing, first proposed by Lilly (1969), is now a standard protocol used in many barotropic and shallow-water studies of forced-dissipative zonation. The physical interpretation of the forcing and the choice of its spatial structure vary somewhat in literature. Considering ξ to be a representation of baroclinic eddies, Williams (1978) chose the forced wavenumbers to lie in a narrow rectangular band, with the zonal extent of the band equal to the baroclinic deformation radius. Scott and Polvani (2007) and Smith (2004) interpreted the rapidly decorrelating, narrowband, isotropic forcing as a model of small-scale three-dimensional convection. Another possibility is that ξ is a representation of the bubble-cloud forcing used by Whitehead (1975) in the laboratory. Below, in the discussion surrounding (10) we give yet another interpretation of ξ.
We have found no studies that establish any particular forcing protocol as being a reasonable physical representation of three-dimensional small-scale eddies acting on a barotropic flow. However, despite the two strong modeling approximations, namely quasi-geostrophy and the choice of the forcing, some features of Jovian jets, such as the jet width, are approximately captured by simplified models (Smith 2004; Vasavada and Showman 2005). A different model forcing in Showman (2007) uses nonsolenoidal physical space mass forcing to represent moist convection in a shallow water system. Despite the different choice of forcing, Showman’s results on planetary zonal jets are broadly consistent with those obtained by Smith (2004). In light of this fact, and since the barotropic quasigeostrophic (QG) system cannot represent mass forcing, we do not address these issues further.
A common theme in all the studies mentioned above is a separation of scales between the forcing length scale
A striking feature of β-plane zonation is that the translational symmetry, y → y + a, of the equation of motion (3) is spontaneously broken: the locations of the eastward maxima in Fig. 1 are an accident of the initial conditions and of the random number generator used to create ξ. But after the jets form, they remain in the same position, apparently forever. Once these robust quasi-steady jets are in place, their dynamics can be discussed in mechanistic terms using concepts such as potential vorticity (PV) mixing, the resilience of transport barriers at the velocity maxima, radiation stress, and shear straining of turbulent eddies (Rhines and Young 1982; Dritschel and McIntyre 2010). But the primary question addressed here is why the jets form in the first place, given that ξ does not select particular locations. Following earlier investigations of this phenomenon (Farrell and Ioannou 2007; Manfroi and Young 1999), we show that zonation can be understood as symmetry-breaking instability of an isotropic, spatially homogeneous, and jetless β-plane flow.
In section 2 we introduce the eddy-mean decomposition and discuss a statistical method, previously used by Farrell and Ioannou (1993b, 2003, 2007), Marston et al. (2008), and Tobias et al. (2011), which is the basis of our linear stability analysis of zonostrophic instability. This method amounts to forming quadratic averages of the equations of motion and then discarding third-order cumulants. Farrell and Ioannou (2003, 2007) refer to this method as stochastic structural stability theory (SSST), while Marston et al. (2008) call it the second-order cumulant expansion, or CE2. SSST and CE2 are completely equivalent, and only one name is required. We have therefore adopted the more descriptive CE2 terminology of Marston et al. (2008).
In section 3 we present a physical space reformulation of CE2, which has analytic advantages over earlier numerically oriented formulations. Within the context of CE2, section 4 provides a complete analytic description of zonostrophic instability obtained by linearizing around an exact isotropic and homogeneous solution with no jets. As in Farrell and Ioannou (2007), zonation is understood as a linear instability of CE2: part of the linearly unstable eigenmode is a zonal flow. This linear stability problem is characterized by two control parameters, a nondimensional drag parameter μ* and a nondimensional planetary parameter β*, and we determine the CE2 zonostrophic stability boundary in the (β*, μ*)-parameter plane. An important property of CE2 zonostrophic instability is that the most unstable wavenumber, which determines the meridional scale of the exponentially growing jets, is well away from zero. Because the instability unfolds around a nonzero wavenumber, CE2 zonostrophic instability is not properly a negative-viscosity instability. This point is reinforced in section 5 by showing that the CE2 eddy viscosity is identically zero. Section 6 is a comparison between the analytic results and direct numerical simulations of the nonlinear system. Section 7 is the discussion and conclusions. The more technical aspects of the paper are in five appendices.
2. The eddy-mean decomposition and quasilinear dynamics










a. Quasilinear dynamics



Quasilinear (QL) zonal jets. (left) A snapshot of the zonally averaged velocity U(y, t) obtained by integrating the QL system (6), (7), and (9). (right) A snapshot of the QL vorticity ζ, with overlaid zonally averaged vorticity −Uy (solid white curve). The parameters for this run are the same as the nonlinear solution in Fig. 1 (i.e., μ* = 0.0182, β* = 1, and kfL = 32). The snapshot is at 2μt = 40 after spinup from rest.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Quasilinear (QL) zonal jets. (left) A snapshot of the zonally averaged velocity U(y, t) obtained by integrating the QL system (6), (7), and (9). (right) A snapshot of the QL vorticity ζ, with overlaid zonally averaged vorticity −Uy (solid white curve). The parameters for this run are the same as the nonlinear solution in Fig. 1 (i.e., μ* = 0.0182, β* = 1, and kfL = 32). The snapshot is at 2μt = 40 after spinup from rest.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Quasilinear (QL) zonal jets. (left) A snapshot of the zonally averaged velocity U(y, t) obtained by integrating the QL system (6), (7), and (9). (right) A snapshot of the QL vorticity ζ, with overlaid zonally averaged vorticity −Uy (solid white curve). The parameters for this run are the same as the nonlinear solution in Fig. 1 (i.e., μ* = 0.0182, β* = 1, and kfL = 32). The snapshot is at 2μt = 40 after spinup from rest.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Comparing the left panels in Figs. 1 and 2, one sees that the QL jets are faster and wider than NL jets, and the jet profiles are different: QL jets are distinctly more east–west symmetric than NL jets. Nonetheless, we show in section 6 that the QL jets in Fig. 2 do have a small east–west QL asymmetry, and at other points in the (β*, μ*)-parameter space, QL jets are strongly east–west asymmetric.
Because the QL jets are faster, the QL system is more zonostrophically unstable than the NL system. In Figs. 1 and 2, quasi-steady jets evolve spontaneously from an initially jetless state, as shown in the Hovmöller diagram of the zonal mean flow in Fig. 3. Comparing Figs. 3a and b shows that the QL system has significantly longer adjustment times than the NL system.

(a) Hovmöller diagram of the zonal mean velocity U(y, t) obtained by solution of the full NL system in (3). (b) Hovmöller diagram of the zonal mean velocity U(y, t) obtained by solution of the QL system. (c) A comparison of the zonal mean energy fraction, zmf(t) defined in (78), for QL and NL runs. The time-averaged fractions are 〈zmf〉NL = 0.3 and 〈zmf〉QL = 0.51. This figure shows the time evolution of the runs in Figs. 1 and 2.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

(a) Hovmöller diagram of the zonal mean velocity U(y, t) obtained by solution of the full NL system in (3). (b) Hovmöller diagram of the zonal mean velocity U(y, t) obtained by solution of the QL system. (c) A comparison of the zonal mean energy fraction, zmf(t) defined in (78), for QL and NL runs. The time-averaged fractions are 〈zmf〉NL = 0.3 and 〈zmf〉QL = 0.51. This figure shows the time evolution of the runs in Figs. 1 and 2.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
(a) Hovmöller diagram of the zonal mean velocity U(y, t) obtained by solution of the full NL system in (3). (b) Hovmöller diagram of the zonal mean velocity U(y, t) obtained by solution of the QL system. (c) A comparison of the zonal mean energy fraction, zmf(t) defined in (78), for QL and NL runs. The time-averaged fractions are 〈zmf〉NL = 0.3 and 〈zmf〉QL = 0.51. This figure shows the time evolution of the runs in Figs. 1 and 2.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
O’Gorman and Schneider (2007) made the QL approximation (9) in an atmospheric general circulation model and showed by comparison with the full nonlinear version of the model that several important features of the flow are unaffected by complete removal of the eddy–eddy nonlinearity as in (9). Comparing Figs. 1 and 2 we reach a similar conclusion for the more idealized model studied here. This preliminary conclusion is supported by a detailed comparison between NL and QL solutions in section 6.
There are several ways of motivating QL dynamics. The QL system conserves both energy and enstrophy and has the same zonal mean equation and symmetries as the NL system. Thus, arguments based on quadratic integral invariants apply equally to the QL and the NL system (Salmon 1998). Nonetheless, because EENL is discarded, the QL system cannot exhibit a true Batchelor–Kraichnan inverse energy cascade: in the QL model all nonlinear interactions require participation of the zonal mean flow. Because U(y, t) has a larger length scale than the eddy field, all these nonlinear QL interactions are spectrally nonlocal. Fig. 2 shows that the spectrally local Batchelor–Kraichnan inverse cascade is not necessary for zonation.
PV is not materially conserved by the QL system, and consequently nonquadratic functions of PV are not conserved by the nonlinear terms remaining in QL. Thus, Fig. 2 also shows that strict material conservation of PV is not necessary for zonation.
Thus, at the most basic level, the QL system is instructive as an indication of the physically essential processes necessary for zonation.
b. Stochastic closure versus cumulant expansion


However, there is probably no reliable a priori method of determining the right-hand side of (10). Heeding the principle to first do no harm, we prefer the QL alternative (9). This has the advantage that one can then make a specific comparison between QL and NL solutions (e.g., as in Figs. 1 and 2) and assess the role of EENL.
Our point of view, which follows Marston et al. (2008) and Tobias et al. (2011), is to regard the QL system as an approximation to the NL system. In fact, (9) in tandem with the method of Farrell and Ioannou, is precisely the second-order cumulant expansion CE2 of Marston et al. (2008). It is from this perspective that in section 3 we develop a physical-space formulation of CE2, which is suitable for analytic solution.
3. Dynamics of correlations: CE2





a. Correlation functions: Kinematics










Notice that that we have changed notation: undecorated x in (15) is the zonal difference coordinate. We also use the shorthand











b. Correlation functions: Dynamics












c. Collective coordinates




d. The zonal mean flow equation
One advantage of collective coordinates is that mean square quantities, such as the enstrophy, are obtained by evaluating correlation functions at zero separation [i.e., by setting (x, y) = 0]. For example, if one possesses
4. Zonostrophic instability of a spatially homogeneous and isotropic base-state flow
a. The spatially homogeneous basic state






























b. The dispersion relation of inviscid and isotropic flow









Dr. George Carnevale has shown that the dispersion relation in (45) and (46) is also obtained from (5.13) in Carnevale and Martin (1982). The field-theoretic approach of Carnevale and Martin (1982) is different from the approximation used to obtain the CE2 system in (31) and (34); for instance, CE2 contains terms such as
c. Ring forcing
















The growth rate s as a function of m for μ* = 0.15 and five values of β* indicated on the curves. The variables in this figure are nondimensionalized according to (49) and (50). These modes have si = 0 (i.e., s is real). The curves β* = 2.571 and 0.0634 correspond to the marginally stable situation defined by (52).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

The growth rate s as a function of m for μ* = 0.15 and five values of β* indicated on the curves. The variables in this figure are nondimensionalized according to (49) and (50). These modes have si = 0 (i.e., s is real). The curves β* = 2.571 and 0.0634 correspond to the marginally stable situation defined by (52).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
The growth rate s as a function of m for μ* = 0.15 and five values of β* indicated on the curves. The variables in this figure are nondimensionalized according to (49) and (50). These modes have si = 0 (i.e., s is real). The curves β* = 2.571 and 0.0634 correspond to the marginally stable situation defined by (52).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1






(a) The critical curve
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

(a) The critical curve
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
(a) The critical curve
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
d. Approximations to the neutral curve with large and small β*








The lower panel of Fig. 5 shows that linear zonostrophic instability is spectrally nonlocal only in the limit β* → ∞: in that case the most unstable wavenumber is much less than the forced wavenumber kf, implying a scale separation between the scales at which energy is injected and the scale at which jets initially form. In the other limit β* → 0 the linearly unstable wavenumber is close to kf.
e. The small wavenumber structure of the growth rate






In recent work BakIo and Bakas and Ioannou (2011) reach a different conclusion, namely that the antifrictional effect resulting in nonzero Reynolds stress is equivalent to nonzero and positive η2, and that the hyperviscous coefficient η4 is negative and therefore stabilizing. We believe that these differences may result from a different choice of forcing Ξ. Bakas and Ioannou (2011) and BakIo use an anisotropic forcing, while our conclusion above is specifically for isotropic forcing. The importance of isotropy to our conclusion is underscored in the section 5.
5. Isotropy and zero eddy viscosity
In the discussion surrounding (59) we observed that the term in the zonostrophic dispersion relation corresponding to the eddy viscosity is zero. This result emerges from the analysis of a complicated dispersion relation and surely deserves a more fundamental explanation, or at least another explanation. Thus in this section we more directly obtain the eddy viscosity of an isotropically forced QL β-plane shear flow and show that the result is identically zero.





We expect that νe defined above is equal to the coefficient η2 in (58). In the m → 0 limit, the modal solution in (44) varies on the length scale m−1, which is much greater than the length scale of the forcing, namely
a. A solution of the correlation equation












b. The Reynolds stresses






















We remark that the constraints in (42) and (43) are required so that correlation function Ψ on the left of (69) decays as r → ∞, despite the r → ∞ divergence of the Green’s function G(r) in (68). In the convolution integral on the right-hand side of (69), the large r divergence of G is shielded by zero integrals of the vorticity correlation function
There are two important caveats associated with the conclusion that νe = 0: the stochastic forcing is isotropic and dissipation is provided only by Ekman drag. Relaxing either or both of these assumptions might result in nonzero νe.
c. The kinetic energy density





d. Discussion
To a certain extent the result νe = 0 is anticipated in the literature on sheared disturbances. Shepherd (1985) showed that an isotropic initial distribution of Rossby waves maintains a constant energy density, despite shearing by a Couette flow; see also Farrell and Ioannou (1993a) and Holloway (2010). The solution in (67), with the isotropic initial condition in (65), is essentially a time integral of Shepherd’s solution of the sheared-disturbance problem with an isotropic initial condition.
Via direct numerical simulation (but with β = 0), Cummins and Holloway (2010) have recently shown that the eddy–eddy nonlinearity is essential in producing nonzero Reynolds stresses from Couette-sheared eddies. Cummins and Holloway (2010) identify the essential role of EENL as restoration of isotropy at high wavenumbers. Moreover, as a result of nonlinearly restored isotropy, νe is robustly positive and thus cannot serve as an explanation of zonostrophic instability. Whatever the sign of νe, an unfortunate consequence of (9) is that restoration of isotropy at small scales is absent in QL dynamics and not represented in the ensemble-averaged dynamics CE2.
6. Zonation in QL and NL solutions
We now turn to numerical solutions for a comparison of the full nonlinear system, the quasilinear system, and the predictions of CE2. In these calculations the resolution is 512 × 512, and we use the ETDRK4 time-stepping scheme (Cox and Matthews 2005). In addition to the control parameters β* and μ* defined in (49), there is a third control parameter, which is the size of the domain relative to the forced wavenumber kf: in our computations the domain is a doubly periodic square 2πL × 2πL, with kfL = 32. Thus there is scale separation between the forcing and the domain.




a. The onset of zonation in NL and QL solutions



The index 〈zmf〉 is used to classify the flow. Figure 6 summarizes a suite of QL and NL calculations in which the drag parameter is varied at fixed β*. The onset of zonation is indicated by the increase in 〈zmf〉. The dotted lines marked

The time-averaged zonal mean energy fraction 〈zmf〉 as a function of μ*, with β* fixed as indicated in the bottom-right corner of each panel. QL simulations are indicated by a degree sign and NL solutions by an asterisk.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

The time-averaged zonal mean energy fraction 〈zmf〉 as a function of μ*, with β* fixed as indicated in the bottom-right corner of each panel. QL simulations are indicated by a degree sign and NL solutions by an asterisk.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
The time-averaged zonal mean energy fraction 〈zmf〉 as a function of μ*, with β* fixed as indicated in the bottom-right corner of each panel. QL simulations are indicated by a degree sign and NL solutions by an asterisk.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
The onset of zonostrophic instability requires significantly smaller values of μ* in the NL case than in the QL case: in Fig. 6 the ratio
b. Zonostrophically stable NL solutions



Snapshots of the vorticity ζ(x, y, t) with overlaid zonally averaged vorticity −Uy(y, t) (solid white curve) with (a) μ* = 0.309 and (b) μ* = 0.0545. Both snapshots are at nondimensional time 2μt = 25, after spinup from rest, and β* = 1.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Snapshots of the vorticity ζ(x, y, t) with overlaid zonally averaged vorticity −Uy(y, t) (solid white curve) with (a) μ* = 0.309 and (b) μ* = 0.0545. Both snapshots are at nondimensional time 2μt = 25, after spinup from rest, and β* = 1.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Snapshots of the vorticity ζ(x, y, t) with overlaid zonally averaged vorticity −Uy(y, t) (solid white curve) with (a) μ* = 0.309 and (b) μ* = 0.0545. Both snapshots are at nondimensional time 2μt = 25, after spinup from rest, and β* = 1.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Figure 8 compares energy spectra of statistically steady QL and NL solutions. With strong drag (i.e., μ* = 0.309) only the directly forced wavenumbers are significantly excited. As μ* is reduced there is transfer of energy to small wavenumbers. In the NL case the transfer of energy to wavenumbers smaller than kf is the due to the inverse energy cascade. In the QL case the excitation of small wavenumbers is due only to shearing by the zonal mean flow. Comparing QL and NL solutions at the same value of μ*, one sees from Figs. 8b and 8d that there is significantly more low-wavenumber eddy energy in the NL cases. Yet the zonal mean energy is always stronger in the QL case. There is no clear association between the inverse energy cascade and zonation.

(a),(c) The zonal spectrum EZ(ky/kf) for QL and NL solutions with β* = 1. (b),(d) The residual spectrum ER(k/kf), defined as the angularly averaged spectrum after removal of the “zonal modes” with kx = 0. The largest peak in EZ(ky/kf) defines the wavenumber mZ, even if there are no quasi-steady zonal jets [e.g., as in the NL simulation with μ* = 0.0545 in (a)].
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

(a),(c) The zonal spectrum EZ(ky/kf) for QL and NL solutions with β* = 1. (b),(d) The residual spectrum ER(k/kf), defined as the angularly averaged spectrum after removal of the “zonal modes” with kx = 0. The largest peak in EZ(ky/kf) defines the wavenumber mZ, even if there are no quasi-steady zonal jets [e.g., as in the NL simulation with μ* = 0.0545 in (a)].
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
(a),(c) The zonal spectrum EZ(ky/kf) for QL and NL solutions with β* = 1. (b),(d) The residual spectrum ER(k/kf), defined as the angularly averaged spectrum after removal of the “zonal modes” with kx = 0. The largest peak in EZ(ky/kf) defines the wavenumber mZ, even if there are no quasi-steady zonal jets [e.g., as in the NL simulation with μ* = 0.0545 in (a)].
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
The NL solution shown in right panel of Fig. 7 with μ* = 0.0545 has an eddy energy spectrum in Fig. 8b exhibiting the beginning of
c. The jet scale
If zonation occurs, as evinced by significantly nonzero values of 〈zmf〉, then by counting the number of distinct jets one can reliably estimate3 a jet wavenumber mJ. For example, in Fig. 1 there are seven jets and therefore








Figure 9 compares the zonal wavenumber obtained from QL and NL solutions with the Rhines wavenumber on the right-hand side of (83), and with the most unstable wavenumber obtained from the linear stability analysis of section 4. In Fig. 9 we show only the β* = 1 and β* = 0.5 solutions: solutions with other values of β* exhibit a broadly similar dependence of mZ on μ*.

A summary of zonal wavenumbers (jet scales) for solutions with (a) β* = 0.5 and (b) β* = 1. The dot-dashed curve is the Rhines wavenumber defined in (83). The solid curve labeled QLS is most unstable wavenumber calculated from the dispersion relation (51). The NL solutions are indicated by asterisks and the QL solutions by degree signs.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

A summary of zonal wavenumbers (jet scales) for solutions with (a) β* = 0.5 and (b) β* = 1. The dot-dashed curve is the Rhines wavenumber defined in (83). The solid curve labeled QLS is most unstable wavenumber calculated from the dispersion relation (51). The NL solutions are indicated by asterisks and the QL solutions by degree signs.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
A summary of zonal wavenumbers (jet scales) for solutions with (a) β* = 0.5 and (b) β* = 1. The dot-dashed curve is the Rhines wavenumber defined in (83). The solid curve labeled QLS is most unstable wavenumber calculated from the dispersion relation (51). The NL solutions are indicated by asterisks and the QL solutions by degree signs.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
At large values μ* only the directly forced modes are excited, and consequently mZ ≈ kf in both the QL and NL cases. At the critical value
There is an interesting transition at
In Fig. 9 the analytic result QLS agrees with the observed QL jet scale only when μ* is not too far from the linear stability boundary

Hovmöller diagrams for the (a) NL and (b) QL runs with β* = 1.0 and μ* = 0.0545. The NL run corresponds to the vorticity snapshot shown in Fig. 7b and shows zonal “streaks.” In (b) the QL jets initially appear at a wavenumber predicted by linearization of CE2. Then successive mergers result in an increase in jet spacing.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Hovmöller diagrams for the (a) NL and (b) QL runs with β* = 1.0 and μ* = 0.0545. The NL run corresponds to the vorticity snapshot shown in Fig. 7b and shows zonal “streaks.” In (b) the QL jets initially appear at a wavenumber predicted by linearization of CE2. Then successive mergers result in an increase in jet spacing.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Hovmöller diagrams for the (a) NL and (b) QL runs with β* = 1.0 and μ* = 0.0545. The NL run corresponds to the vorticity snapshot shown in Fig. 7b and shows zonal “streaks.” In (b) the QL jets initially appear at a wavenumber predicted by linearization of CE2. Then successive mergers result in an increase in jet spacing.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Figure 10a shows the Hovmöller diagram of the jetless NL solution from Fig. 7b. There is no zonation and U(y, t) shows “streaks” rather than jets. These streaks are not strong relative to the turbulent eddy field (i.e., 〈zmf〉 ≈ 0). The corresponding zonal energy spectrum in Fig. 8a exhibits a strong peak, which is a signature of these transient zonal steaks.
Figure 10b shows the QL case in which jets initially appear with a relatively small meridional spacing predicted by linear theory, followed by a sequence of mergers so that the mature flow has mZ much less than the most linearly unstable wavenumber. The QL jet-merging phenomenology, which is effectively a one-dimensional inverse cascade, is very similar to the “Cahn–Hilliard” solutions obtained by Manfroi and Young (1999) from a model of deterministically forced zonation.
d. The small drag regime
The flows in Figs. 1 and 2 have relatively light damping and both flows have organized jets containing a substantial fraction of the total kinetic energy. Figure 11 shows the time-averaged zonal mean flow 〈U〉 and the corresponding PV gradient β* − 〈Uyy〉. In Fig. 2 the QL jets are almost symmetrical in the zonal direction, in contrast to the NL jets.5 But the QL jets are not perfectly symmetric: the PV gradient in Fig. 11b reveals the QL east–west asymmetry. The NL PV gradient is positive for all y and thus the NL jets are stable according to the Rayleigh–Kuo criterion. The QL PV gradient in Fig. 11b reverses sign on the flanks of the eastward jet, and also at the centers of the westward jets. Nonetheless the QL zonal mean flow shows no indication of barotropic instability [i.e., the deep spikes with β* − 〈Uyy〉 < 0 are permanent features of the QL zonal mean flow even after time averaging].

Comparison of zonal mean velocity profiles of the β* = 1 NL and QL runs in Figs. 1 and 2.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Comparison of zonal mean velocity profiles of the β* = 1 NL and QL runs in Figs. 1 and 2.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Comparison of zonal mean velocity profiles of the β* = 1 NL and QL runs in Figs. 1 and 2.
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1



Comparison of time-averaged zonal mean velocity profiles (thin lines) of QL runs. (a) The solution from Fig. 10b with β* = 1.0 and μ* = 0.054; (b) the strongly forced solution with β* = 0.1 and μ* = 0.005. Also plotted are the corresponding PV gradients (thick curves).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Comparison of time-averaged zonal mean velocity profiles (thin lines) of QL runs. (a) The solution from Fig. 10b with β* = 1.0 and μ* = 0.054; (b) the strongly forced solution with β* = 0.1 and μ* = 0.005. Also plotted are the corresponding PV gradients (thick curves).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Comparison of time-averaged zonal mean velocity profiles (thin lines) of QL runs. (a) The solution from Fig. 10b with β* = 1.0 and μ* = 0.054; (b) the strongly forced solution with β* = 0.1 and μ* = 0.005. Also plotted are the corresponding PV gradients (thick curves).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Thus, although a detailed study of QL jet asymmetry is not a focus of the present work, our QL numerical solutions are generally consistent with the equilibrated SSST jets presented in Farrell and Ioannou (2007).
e. Discussion of the eddy–eddy nonlinearity
An important effect of eddy–eddy nonlinearity is the stirring of PV, producing an exponential-in-time reduction in the length scale of vorticity fluctuations. Eddy-driven stirring is removed from the QL system by (9): shearing by U(y, t) is the only scale-reduction mechanism acting on the QL eddy vorticity. The small-scale structure evident in the QL PV gradient in the right panel of Fig. 11 may reflect the relative inefficiency of shearing by U(y, t) at removing vorticity fluctuations.
Further differences in the jet structure evident in Fig. 11 can be explained by meandering of the NL jets, so that the zonal average reduces the sharpness of the NL PV gradient. The spectral signature of the NL jet meanders is a high energy mode at
7. Discussion and conclusions
A contribution of this work is the analytic development of the linearized theory of zonostrophic instability within the context of the second-order cumulant expansion (CE2) of Marston et al. (2008) and the stochastic structural stability theory (SSST) of Farrell and Ioannou (2003, 2007). These statistical formulations are equivalent to the correlation dynamics derived in section 3, and that physical-space formulation, in terms of partial differential equations for the correlation functions Ψ and
In the top panel of Fig. 5 we display the curve of neutral zonostrophic stability in the (β*, μ*)-parameter plane obtained by solution of linearized CE2 dynamics. We have shown that with isotropic forcing zonostrophic instability is not a negative-viscosity instability: the hallmark of a negative-viscosity instability is that at the stability boundary the most unstable wavenumber is zero. The deterministic model of anisotropically forced β-plane zonation analyzed by Manfroi and Young (1999) provides a bona fide example of the negative-viscosity case. Instead, for the isotropically and stochastically forced model analyzed here, the onset of zonostrophic instability is at the nonzero meridional wavenumber shown in the bottom panel of Fig. 5; only at large β* does this wavenumber approach zero. Moreover, in section 5 we showed that with isotropic forcing the CE2 eddy viscosity νe is identically zero.
Comparison of QL and NL numerical solutions indicates that the CE2 linear stability boundary does not provide an accurate estimate of the onset of zonostrophic instability for NL flows. This quantitative failure of CE2 is not surprising: neglect of the eddy–eddy nonlinearity is most plausible in cases where most of the energy is in the zonal mean flow: close to the stability boundary the zonal mean flow is only incipient. An outstanding open problem is improving CE2 to account for the missing physics in the eddy–eddy nonlinearity. Another important problem is obtaining analytic insight into the solution of the CE2 system in the regime where CE2 is likely to be valid, namely in the strongly unstable regime where the drag μ* is much less than the critical drag μc and the fraction of energy in the zonal mean flow is substantial.
Acknowledgments
This work was supported by the National Science Foundation under OCE1057838. We thank Nikos Bakas, Oliver Bühler, George Carnevale, Brian Farrell, Petros Ioannou, Brad Marston, Rick Salmon, and Steve Tobias for discussion of these results.
APPENDIX A
Implementation of the Random Forcing ξ(x, t)



















Comparison of the ring forcing δk → 0 (solid line) and the narrow-band forcing with δk = kf/4 (dashed line).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1

Comparison of the ring forcing δk → 0 (solid line) and the narrow-band forcing with δk = kf/4 (dashed line).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
Comparison of the ring forcing δk → 0 (solid line) and the narrow-band forcing with δk = kf/4 (dashed line).
Citation: Journal of the Atmospheric Sciences 69, 5; 10.1175/JAS-D-11-0200.1
APPENDIX B
Rapid Temporal Decorrelation: Derivation of (25)




















APPENDIX C
Derivation of the Dispersion Relation (45)













































APPENDIX D
The Function Q(χ, n)
In this appendix we summarize some properties of the function Q(χ, n) defined in (C20).








APPENDIX E
The Neutral Curve




a. Approximation of the marginal curve, β ≫ 1 and χ0 ≪ 1










b. Approximation of the marginal curve, β ≪ 1 and χ0 ≫ 1
















c. The small-m expansion




REFERENCES
Bakas, N. A., and P. Ioannou, 2011: Structural stability theory of two-dimensional fluid flow under stochastic forcing. J. Fluid Mech., 682, 332–361.
Carnevale, G. F., and P. C. Martin, 1982: Field theoretical techniques in statistical fluid dynamics: With applications to nonlinear wave dynamics. Geophys. Astrophys. Fluid Dyn., 20, 131–164.
Cox, S. M., and P. C. Matthews, 2005: Exponential time differencing for stiff systems. J. Comput. Phys., 176, 430–455.
Cummins, P. F., and G. Holloway, 2010: Reynolds stress and eddy viscosity in direct numerical simulation of sheared-two-dimensional turbulence. J. Fluid Mech., 657, 394–412.
Danilov, S., and V. M. Gryanik, 2004: Barotropic beta-plane turbulence in a regime with strong zonal jets revisited. J. Atmos. Sci., 61, 2283–2295.
Danilov, S., and D. Gurarie, 2004: Scaling, spectra and zonal jets in beta-plane turbulence. Phys. Fluids, 16, 2592–2603.
DelSole, T., 2001: A theory for the forcing and dissipation in stochastic turbulence models. J. Atmos. Sci., 58, 3762–3775.
Diamond, P. H., S.-I. Itoh, K. Itoh, and T. S. Hahm, 2005: Zonal flows in plasmas—A review. Plasma Phys. Contr. Fusion, 47, R35, doi:10.1088/0741-3335/47/5/R01.
Dritschel, D. G., and M. E. McIntyre, 2010: Multiple jets as PV staircases: The Phillips effect and the resilience of eddy-transport barriers. J. Atmos. Sci., 65, 855–874.
Farrell, B. F., and P. J. Ioannou, 1993a: Stochastic forcing of perturbation variance in unbounded shear and deformation flows. J. Atmos. Sci., 50, 200–211.
Farrell, B. F., and P. J. Ioannou, 1993b: Stochastic forcing of the linearized Navier–Stokes equations. Phys. Fluids A, 5, 2600–2609.
Farrell, B. F., and P. J. Ioannou, 2003: Structural stability of turbulent jets. J. Atmos. Sci., 50, 2101–2118.
Farrell, B. F., and P. J. Ioannou, 2007: Structure and spacing of jets in barotropic turbulence. J. Atmos. Sci., 64, 3652–3665.
Galperin, B., S. Sukoriansky, N. Dikovskaya, P. Read, Y. Yamazaki, and R. Wordsworth, 2006: Anisotropic turbulence and zonal jets in rotating flows with a β-effect. Nonlinear Processes Geophys., 13, 83–98.
Holloway, G., 2010: Eddy stress and shear in 2D flows. J. Turbul., 11, 1–14.
Huang, H. P., and W. A. Robinson, 1998: Two-dimensional turbulence and persistent zonal jets in a global barotropic model. Nonlinear Processes Geophys., 55, 611–632.
Lilly, D. K., 1969: Numerical simulation of two-dimensional turbulence. Phys. Fluids, 12, II-240, doi:10.1063/1.1692444.
Maltrud, M. E., and G. K. Vallis, 1991: Energy spectra and coherent structures in forced two-dimensional and beta-plane turbulence. J. Fluid Mech., 228, 321–342.
Manfroi, A. J., and W. Young, 1999: Slow evolution of zonal jets on the beta plane. J. Atmos. Sci., 56, 784–800.
Marston, J. B., E. Conover, and T. Schneider, 2008: Statistics of an unstable barotropic jet from a cumulant expansion. J. Atmos. Sci., 65, 1955–1966.
Nozawa, T., and S. Yoden, 1997: Formation of zonal-band structure in forced two-dimensional turbulence on a rotating sphere. Phys. Fluids, 9, 2081–2093.
O’Gorman, P. A., and T. Schneider, 2007: Recovery of atmospheric flow statistics in a general circulation model without nonlinear eddy–eddy interactions. Geophys. Res. Lett., 34, L22801, doi:10.1029/2007GL031779.
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 beta-plane. J. Fluid Mech., 69, 417–443.
Rhines, P. B., and W. R. Young, 1982: Homogenization of potential vorticity in planetary gyres. J. Fluid Mech., 122, 347–367.
Salmon, R., 1998: Lectures on Geophysical Fluid Dynamics. Oxford University Press, 379 pp.
Scott, R. K., and L. M. Polvani, 2007: Forced-dissipative shallow-water turbulence on the sphere and the atmospheric circulation of the giant planets. J. Atmos. Sci., 64, 3158–3176.
Shepherd, T. G., 1985: Time development of small disturbances to plane Couette flow. J. Atmos. Sci., 42, 1868–1871.
Showman, A. P., 2007: Numerical simulations of forced shallow-water turbulence: Effects of moist convection on large-scale circulation of Jupiter and Saturn. J. Atmos. Sci., 64, 3132–3157.
Smith, K. S., 2004: A local model for planetary atmospheres forced by small-scale convection. J. Atmos. Sci., 61, 1420–1433.
Tobias, S. M., K. Dagon, and J. Marston, 2011: Astrophysical fluid dynamics via direct statistical simulation. Astrophys. J., 727, 127, doi:10.1088/0004-637X/727/2/127.
Vallis, G. K., and M. E. Maltrud, 1993: Generation of mean flow and jets on a beta plane and over topography. J. Phys. Oceanogr., 23, 1346–1362.
Vasavada, A. R., and A. P. Showman, 2005: Jovian atmospheric dynamics: An update after Galileo and Cassini. Rep. Prog. Phys., 68, 1935–1996.
Whitehead, J. A., 1975: Mean flow generated by circulation on a beta-plane: An analogy with the moving flame experiment. Tellus, 27, 358–363.
Williams, G. P., 1978: Planetary circulations: 1. Barotropic representation of Jovian and terrestrial turbulence. J. Atmos. Sci., 35, 1399–1426.
There is also uniform advection by the zonal flow. But that sweeping is eliminated by the difference U1 − U2 in the correlation equation (31) and is therefore inconsequential to Reynolds stresses.
In certain cases the system may be transitioning between a state with n and n + 1 jets. Following Panetta (1993), we then count
From (3), the exact energy power integral is 〈ψξ + μ|∇ψ|2 + (−1)n−1νn|∇n−1ζ|2〉 = 0, where 〈〉 is both a domain integral and a time average.
If β = 0 then the equations of motion are invariant under y → −y and ψ → −ψ. This symmetry, which induces u → u, is broken in both QL and NL by nonzero β. This explains the characteristic east–west asymmetry of U(y, t) on the β plane.
Farrell and Ioannou (2007) also employ a forcing with a different correlation function than our isotropic choice.