Equations For Two-Layer QG Model

The two-layer quasigeostrophic evolution equations are (1)

\[\partial_t\,{q_1} + \mathsf{J}\left(\psi_1\,, q_1\right) + \beta\, {\psi_1}_x = \text{ssd} \,,\]

and (2)

\[\partial_t\,{q_2} + \mathsf{J}\left(\psi_2\,, q_2\right)+ \beta\, {\psi_2}_x = -r_{ek}\nabla^2 \psi_2 + \text{ssd}\,,\]

where the horizontal Jacobian is \(\mathsf{J}\left(A\,, B\right) = A_x B_y - A_y B_x\). Also in (1) and (2) ssd denotes small-scale dissipation (in turbulence regimes, ssd absorbs enstrophy that cascades towards small scales). The linear bottom drag in (2) dissipates large-scale energy.

The potential vorticities are (3)

\[{q_1} = \nabla^2\psi_1 + F_1\left(\psi_2 - \psi_1\right)\,,\]

and (4)

\[{q_2} = \nabla^2\psi_2 + F_2\left(\psi_1 - \psi_2\right)\,,\]

where

\[F_1 \equiv \frac{k_d^2}{1 + \delta}\,, \qquad \text{and} \qquad F_2 \equiv \delta \,F_1\,,\]

with the deformation wavenumber

\[k_d^2 \equiv\frac{f_0^2}{g'}\frac{H_1+H_2}{H_1 H_2}\,,\]

where \(H = H_1 + H_2\) is the total depth at rest.

Forced-dissipative equations

We are interested in flows driven by baroclinic instabilty of a base-state shear \(U_1-U_2\). In this case the evolution equations (1) and (2) become (5)

\[\partial_t\,{q_1} + \mathsf{J}\left(\psi_1\,, q_1\right) + \beta_1\, {\psi_1}_x = \text{ssd} \,,\]

and (6)

\[\partial_t\,{q_2} + \mathsf{J}\left(\psi_2\,, q_2\right)+ \beta_2\, {\psi_2}_x = -r_{ek}\nabla^2 \psi_2 + \text{ssd}\,,\]

where the mean potential vorticity gradients are (9,10)

\[\beta_1 = \beta + F_1\,\left(U_1 - U_2\right)\,, \qquad \text{and} \qquad \beta_2 = \beta - F_2\,\left( U_1 - U_2\right)\,.\]

Equations in Fourier space

We solve the two-layer QG system using a pseudo-spectral doubly-peridioc model. Fourier transforming the evolution equations (5) and (6) gives (7)

\[\partial_t\,{\widehat{q}_1} = - \widehat{\mathsf{J}}\left(\psi_1\,, q_1\right) - \text{i}\,k\, \beta_1\, {\widehat{\psi}_1} + \widehat{\text{ssd}} \,,\]

and

\[\partial_t\,{\widehat{q}_2} = - \widehat{\mathsf{J}}\left(\psi_2\,, q_2\right)- \text{i}\,k\, \beta_2\, {\widehat{\psi}_2} + r_{ek}\,\kappa^2\,\, \widehat{\psi}_2 + \widehat{\text{ssd}}\,,\]

where, in the pseudo-spectral spirit, \(\widehat{\mathsf{J}}\) means the Fourier transform of the Jacobian i.e., we compute the products in physical space, and then transform to Fourier space.

In Fourier space the “inversion relation” (3)-(4) is

\[\begin{split}\underbrace{\begin{bmatrix} -(\kappa^2 + F_1) \qquad \:\:\:\:F_1\\ \:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:F_2 \qquad - (\kappa^2 + F_2)\; \end{bmatrix}}_{\equiv \,\mathsf{M_2}} \begin{bmatrix} \widehat{\psi}_1\\ \widehat{\psi}_2\\ \end{bmatrix} =\begin{bmatrix} \widehat{q}_1\\ \widehat{q}_2\\ \end{bmatrix} \,,\end{split}\]

or equivalently

\[\begin{split}\begin{bmatrix} \widehat{\psi}_1\\ \widehat{\psi}_2\\ \end{bmatrix} =\underbrace{\frac{1}{\text{det}\,\mathrm{M_2}} \begin{bmatrix} -(\kappa^2 + F_2) \qquad \:\:\:\:-F_1\\ \:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:-F_2 \qquad - (\kappa^2 + F_1)\; \end{bmatrix}}_{=\,\mathsf{M_2}^{-1}} \begin{bmatrix} \widehat{q}_1\\ \widehat{q}_2\\ \end{bmatrix} \,,\qquad\end{split}\]

where

\[\qquad \text{det}\,\mathsf{M}_2 = \kappa^2\left(\kappa^2 + F_1 + F_2\right)\,.\]

Marching forward

We use a third-order Adams-Bashford scheme

\[{\widehat{q}_i}^{n+1} = E_f\times\left[{\widehat{q}_i}^{n} + \frac{\Delta t}{12}\left(23\, \widehat{Q}_i^{n} - 16\widehat{Q}_i^{n-1} + 5 \widehat{Q}_i^{n-2}\right)\right]\,,\]

where

\[\widehat{Q}_i^n \equiv - \widehat{\mathsf{J}}\left(\psi_i^n\,, q_i^n\right) - \text{i}\,k\, \beta_i\, {\widehat{\psi}_i^n}, \qquad i = 1,2\,.\]

The AB3 is initialized with a first-order AB (or forward Euler)

\[{\widehat{q}_i}^{1} = E_f\times\left[{\widehat{q}_i}^{0} + \Delta t \widehat{Q}_i^{0}\right]\,,\]

The second step uses a second-order AB scheme

\[{\widehat{q}_i}^{2} = E_f\times\left[{\widehat{q}_i}^{1} + \frac{\Delta t}{2}\left(3\, \widehat{Q}_i^{1} - \widehat{Q}_i^0\right)\right]\,.\]

The small-scale dissipation is achieve by a highly-selective exponential filter

\[\begin{split}E_f =\begin{cases} \text{e}^{-23.6\,\left(\kappa^{\star} - \kappa_c\right)^4}: &\qquad \kappa \ge\kappa_c\\ \,\,\,\,\,\,\,\,\,\,\,1:&\qquad \text{otherwise}\,. \end{cases}\end{split}\]

where the non-dimensional wavenumber is

\[\kappa^{\star} \equiv \sqrt{ (k\,\Delta x)^2 + (l\,\Delta y)^2 }\, ,\]

and \(\kappa_c\) is a (non-dimensional) wavenumber cutoff here taken as \(65\%\) of the Nyquist scale \(\kappa^{\star}_{ny} = \pi\). The parameter \(-23.6\) is obtained from the requirement that the energy at the largest wanumber (\(\kappa^{\star} = \pi\)) be zero whithin machine double precision:

\[\frac{\log 10^{-15}}{(0.35\, \pi)^4} \approx -23.5\,.\]

For experiments with \(|\widehat{q_i}|\ll\mathcal{O}(1)\) one can use a smaller constant.

Diagnostics

The kinetic energy is

\[E = \tfrac{1}{H\,S} \int \tfrac{1}{2} H_1 \, |\boldsymbol{\nabla} \psi_1|^2 + \tfrac{1}{2} H_2 \, |\boldsymbol{\nabla} \psi_2|^2 \, d S\,.\]

The potential enstrophy is

\[Z = \tfrac{1}{H\,S}\int \tfrac{1}{2}H_1 \, q_1^2 + \tfrac{1}{2} H_2 \, q_2^2 \, d S\,.\]

We can use the enstrophy to estimate the eddy turn-over timescale

\[T_e \equiv \frac{2\,\pi}{\sqrt{Z}}\,.\]