1. Introduction
Steadily translating dipolar vortices, also known as modons, arise as exact solutions in many geophysical fluid systems (Flierl 1987). They can be viewed as a robust pair of opposite-signed vortices whose mutual interaction leads to steady propagation of the entire structure. Vortex dipoles have been used as models for cold-core oceanic rings (Flierl 1979), atmospheric blocks (McWilliams 1980; Butchart et al. 1989), and troughs and localized jets at the tropopause (Van Tuyl and Young 1982; Hakim 2000; Cunningham and Keyser 2000). Applications to tropopause disturbances have been limited by the fact that, except for Berestov (1979), the available dipole solutions all come from barotropic or layered models. For these applications, the surface quasigeostrophic equations (sQG; Juckes 1994; Held et al. 1995) offer a most natural and simple description of the flow, in which potential vorticity is uniform away from the strong gradients at the tropopause, and the tropopause itself is represented as a rigid boundary. Models based on sQG have been applied extensively to the midlatitude tropopause (Rivest et al. 1992; Juckes 1994; Muraki and Hakim 2001; Hakim et al. 2002). Moreover, vortex dipoles provide useful idealizations of localized jets at the tropopause even when isolated, coherent dipoles are not present, as in a baroclinic wave [e.g., Fig. 4 of Plougonven and Snyder (2005), in which the jet and associated diffluent region downstream of the trough resemble the flow in a vortex dipole]. In this note, we derive a three-dimensional, continuously stratified modon for sQG. This modon is the simplest model for a localized jet on the tropopause that retains realistic vertical structure. Since the troposphere is of finite depth and typically exhibits large-scale temperature gradients, we also extend the basic dipole solution to a fluid layer between two rigid boundaries, and to the presence of uniform background vertical shear and horizontal potential temperature gradient.
2. The modon construction and the sQG dipole
Flows in sQG, in contrast, have uniform PV and their dynamics are associated with a potential temperature disturbance on a horizontal boundary. Thus in sQG, the PV advection equation is satisfied identically by imposing uniform interior PV (Q = 0). The dipole motion is then determined solely through the evolution of surface potential temperature (2).
Numerical solutions for ψ̂(σ) given α are simply obtained by matrix inversion after approximating the integral in (11) by a discrete quadrature (such as Simpson’s rule) over a sufficiently large interval (0 ≤ σ ≤ Σ). Radial profiles of the surface streamfunction (solid line) and potential temperature (dashed line) are shown in Fig. 1a. The Fourier–Bessel spectral amplitudes ψ̂(σ) are shown in Fig. 1b (100 of 512 modes shown; Σ = 256). The line indicates an asymptotic slope of −7/2, which is the characteristic (Fourier–Bessel) spectral decay expected when the θ field has a jump in the radial derivative.
The value α ≈ 4.1213 is found by a simple root-finding procedure that enforces the modon boundary (r = R) to be a separating streamline (Ψs = 0). This feature of the solution is apparent in the total streamfunction Ψ(x, y, 0) shown in Fig. 2. As with other modons, the dipole’s propagation speed is simply proportional to its amplitude. A positive value of c makes the upper part of the dipole (Fig. 2) a warm, cyclonic anomaly, which results in a modon that propagates eastward relative to background winds.
For a variety of flow systems, the stability of dipoles has been inferred through time-dependent simulations (Flierl 1987). We have verified that the propagation and spatial structure for the sQG dipole, and the variants that follow, are robust within the sQG dynamics to the addition of noise into the initial condition. The numerical sQG simulation uses a dealiased pseudospectral method, described in Hakim et al. (2002), and includes weak ∇8 hyperdiffusion. Unlike the analytic sQG dipole, the numerical simulations assume periodicity in both horizontal directions.
Figure 3 shows contours of θs at t = 8 from a simulation beginning from an sQG dipole (c = 1; R = 1) with superimposed initial noise in θs. The simulation uses 2562 collocation points, a domain of size L = 8, and initial Gaussian noise that is uncorrelated in space and has variance that is 5% of the dipole’s maximum θs. Even after propagating a distance of nearly 8, or one complete circuit of the periodic domain, the dipole’s structure is little changed from the analytic solution. Longer simulations reveal similar, stable behavior.
3. Dipole in a finite layer
This observation illustrates the general principle that, in sQG, the horizontal scales whose Rossby heights are deeper than the layer behave roughly as those of the barotropic vorticity equation. The decrease of the normalized θsmax, seen in Fig. 4a, corresponds to the fact that, as the layer thins, weaker potential temperature disturbances are required to maintain a fixed velocity perturbation.
4. Dipole in a horizontal temperature gradient
The Ψ–Θ relation (18) is equivalent to beginning as in the general Berestov construction (5) with an αi and αo in (8). To see this, note first that nonzero αo implies that the disturbances Ψs and ψsz as defined in (6) can no longer be vanishing at r → ∞. Rather, it is necessary to introduce the Eady shear as in (16) with αo/R = λ/c. The result can then be rewritten in the form (18). The zero exterior condition of (18) is required mathematically as the Hankel transform can only be applied for disturbances that decay as r → ∞.
5. Conclusions
Vortex dipoles are the simplest idealization of a localized jet at the tropopause, yet existing dipole solutions pertain only to barotropic or layered models or to dipoles with significant anomalies of interior PV. We have developed a vortex-dipole solution for sQG that is more appropriate to tropopause disturbances, in that its dynamics are driven only by potential temperature advection on a single surface (approximating the tropopause) yet its vertical structure extends continuously into the uniform-PV interior. Rather than the algebraic conditions produced by matching across the dipole’s boundary in other dipole solutions, the sQG dipole requires the solution of the three-dimensional Laplace equation and yields an integral equation for the coefficient in the ψ–θ relation within the dipole. The basic sQG modon can also be extended to a troposphere of finite depth with uniform potential temperature at the lower boundary, and to the presence of uniform background vertical shear and horizontal potential temperature gradient. Numerical simulations with the pseudospectral sQG model of Hakim et al. (2002) indicate that the basic sQG dipole, and the variants presented here, are robust to initial background noise. However, we are currently investigating the nature and stability of similar dipole solutions for the primitive equations, in particular for their interaction with inertia–gravity waves. Other modifications not included in this Note include a superimposed monopolar rider and an sQG dipole in the presence of planetary vorticity gradients. The construction technique also likely allows extension to weak shear (Swenson 1986), tropopause deformation (Rivest et al. 1992), and possibly, beyond QG corrections (Muraki et al. 1999).
Acknowledgments
DJM gratefully acknowledges support through NSF DMR-9704724, NSERC RGPIN238928, the Alfred P. Sloan Foundation, and the Visitor Program of the Mesoscale and Microscale Meteorology Division at NCAR. Author CS was partially supported by NSF Grant CMG-0327582. We are indebted to Andreas Dörnbrack for alerting us to the paper of Meleshko and van Heijst. Last, special thanks are given to Greg Hakim and Phil Cunningham for their enthusiastic support of this analytical work.
REFERENCES
Berestov, A. L., 1979: Solitary Rossby waves. Izv. Acad. Sci. USSR, Atmos. Oceanic Phys., 15 , 443–447.
Butchart, N., K. Haines, and J. C. Marshall, 1989: A theoretical and diagnostic study of solitary waves and atmospheric blocking. J. Atmos. Sci., 46 , 2063–2078.
Cunningham, P., and D. Keyser, 2000: Analytical and numerical modelling of jetstreaks: Barotropic dynamics. Quart. J. Roy. Meteor. Soc., 126 , 3187–3217.
Flierl, G. R., 1979: A simple model for warm and cold core rings. J. Geophys. Res., 84 , 57–65.
Flierl, G. R., 1987: Isolated eddy models in geophysics. Annu. Rev. Fluid Mech., 19 , 493–530.
Hakim, G. J., 2000: Climatology of coherent structures on the extratropical tropopause. Mon. Wea. Rev., 128 , 385–406.
Hakim, G. J., C. Snyder, and D. J. Muraki, 2002: A new surface model for cyclone–anticyclone asymmetry. J. Atmos. Sci., 59 , 2405–2420.
Held, I. M., R. T. Pierrehumbert, S. T. Garner, and K. L. Swanson, 1995: Surface quasigeostrophic dynamics. J. Fluid Mech., 282 , 1–20.
Juckes, M., 1994: Quasigeostrophic dynamics of the tropopause. J. Atmos. Sci., 51 , 2756–2768.
Kondo, J., 1991: Integral Equations. Clarendon, 466 pp.
Lamb, H., 1945: Hydrodynamics. Dover, 738 pp.
McWilliams, J. C., 1980: An application of equivalent modons to atmospheric blocking. Dyn. Atmos. Oceans, 5 , 43–66.
Meleshko, V. V., and G. J. F. van Heijst, 1994: On Chaplygin’s investigations of two-dimensional vortex structures in an inviscid fluid. J. Fluid Mech., 272 , 157–182.
Morse, P. M., and H. Feshbach, 1953: Methods of Theoretical Physics. Vol. 1 McGraw-Hill, 997 pp.
Muraki, D. J., and G. J. Hakim, 2001: Balanced asymmetries of waves on the tropopause. J. Atmos. Sci., 58 , 237–252.
Muraki, D. J., C. Snyder, and R. Rotunno, 1999: The next-order corrections to quasigeostrophic theory. J. Atmos. Sci., 56 , 1547–1560.
Plougonven, R., and C. Snyder, 2005: Gravity waves excited by jets: Propagation versus generation. Geophys. Res. Lett., 32 .L18802, doi:10.1029/2005GL023730.
Rivest, C., C. A. Davis, and B. F. Farrell, 1992: Upper-tropospheric synoptic-scale waves. Part I: Maintenance as Eady normal modes. J. Atmos. Sci., 49 , 2108–2119.
Swenson, M., 1986: Equivalent modons in simple shear. J. Atmos. Sci., 43 , 3177–3185.
Van Tuyl, A. H., and J. A. Young, 1982: Numerical simulation of nonlinear jet streak adjustment. Mon. Wea. Rev., 110 , 2038–2054.
Verkley, W. T. M., 1987: Stationary barotropic modons in westerly background flows. J. Atmos. Sci., 44 , 2383–2398.
APPENDIX
Details of the Integral Equation
(a) Radial profiles of disturbance streamfunction (solid line) and potential temperature (dashed line) for the basic dipole. The basic dipole streamfunction is scaled on Rc, and the potential temperature is scaled on c. (b) Log–log scatterplot of Fourier–Bessel amplitudes ψ̂(σ) as numerically computed from (11).
Citation: Journal of the Atmospheric Sciences 64, 8; 10.1175/JAS3958.1
Contours of the full streamfunction for the basic dipole in a frame in which the dipole is steady. The contour intervals are Rc/2. Positive values of c correspond to an easterly incident wind.
Citation: Journal of the Atmospheric Sciences 64, 8; 10.1175/JAS3958.1
Contours of the full surface potential temperature for an sQG simulation of the basic dipole that has been initialized with small-scale noise. Despite the noisy background, the dipole has very nearly returned to its center location after one transversal of the periodic domain. The interval of the dark contours is 0.2 of θsmax, and the light contours are at ±0.02 of θsmax.
Citation: Journal of the Atmospheric Sciences 64, 8; 10.1175/JAS3958.1
Variation of the α parameter (thick solid line) for the two dipole variants: (a) finite layer and (b) vertical shear. The thin-layer (H/R → 0) asymptote is shown (thin dashed line) in (a). Also shown in both panels (gray solid line) is the variation of θsmax normalized onto the basic dipole of section 2 (θsmax ≈ 9.24).
Citation: Journal of the Atmospheric Sciences 64, 8; 10.1175/JAS3958.1