Skip to content

Commit

Permalink
Kelvin and Rossby waves
Browse files Browse the repository at this point in the history
  • Loading branch information
milancurcic committed Oct 21, 2024
1 parent 7f7b631 commit c0fa27b
Show file tree
Hide file tree
Showing 3 changed files with 207 additions and 1 deletion.
Binary file added assets/fig_kelvin_wave.pdf
Binary file not shown.
Binary file added assets/fig_rossby_wave.pdf
Binary file not shown.
208 changes: 207 additions & 1 deletion notes.tex
Original file line number Diff line number Diff line change
Expand Up @@ -2851,7 +2851,8 @@ \subsubsection{The complete equation set}
We now proceed to further simplify this equation set to derive a general
solution for the shallow water equations.

\subsection{General solution}
\subsection{Poincaré waves}
\label{sec:poincare_waves}

As we proceed without our intention to derive a solution to the
equations (\ref{eq:shallow_water_final_scalar_u}-\ref{eq:shallow_water_final_scalar_eta}),
Expand Down Expand Up @@ -3055,6 +3056,119 @@ \subsubsection{Inertial oscillations}
Here, it comes out as a limiting case from the general solution which we
couldn't obtain prior to the shallow water approximations and linearization.

\subsection{Kelvin waves}

A special case of the general solution that is particularly relevant to the
atmospheric and oceanic dynamics is that of a linearized shallow water flow
that is bounded on one side by a solid boundary, such as a coastline.
The resulting solution is a special class of gravity waves called
\textit{Kelvin waves}\index{Kelvin waves}, which propagate as a shallow water
gravity wave along the solid boundary and whose propagation direction, as well
as the perturbation scale in the direction away from the boundary, are governed
by the planetary rotation rate.
Kelvin waves appear in both the atmosphere and the ocean.

To derive the Kelvin waves, we start from the linearized shallow water equations
(where we drop the primes for brevity):

\begin{equation}
\frac{\partial u}{\partial t} - f v = - g \frac{\partial \eta}{\partial x}
\end{equation}

\begin{equation}
\frac{\partial v}{\partial t} + f u = - g \frac{\partial \eta}{\partial y}
\end{equation}

\begin{equation}
\frac{\partial \eta}{\partial t} + H \left( \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} \right) = 0
\end{equation}
Now, suppose that our solid boundary is along the $x$-axis at $y = 0$, which
allows us to neglect the meridional flow ($v=0$):

\begin{equation}
\frac{\partial u}{\partial t} = - g \frac{\partial \eta}{\partial x}
\label{eq:kelvin_u}
\end{equation}

\begin{equation}
f u = - g \frac{\partial \eta}{\partial x}
\label{eq:kelvin_v}
\end{equation}

\begin{equation}
\frac{\partial \eta}{\partial t} + H \frac{\partial u}{\partial x} = 0
\label{eq:kelvin_eta}
\end{equation}
Differentiate Eq. \ref{eq:kelvin_u} with respect to time and Eq. \ref{eq:kelvin_eta}
with respect to $x$, and combine them to get:

\begin{equation}
\frac{\partial^2 u}{\partial t^2} - \sqrt{gH} \frac{\partial^2 u}{\partial x^2} = 0
\end{equation}
which is the standard wave equation, whose solution is a wave that propagates
with the phase speed $c = \sqrt{gH}$.
We will thus assume a wave-like solution for $u$, like we did for the Poincaré
waves in Section \ref{sec:poincare_waves}.
However, since we now have a solid boundary at $y = 0$, we should also assume
that the solution should vary in the $y$ direction (because it must be zero
at the boundary, and non-zero elsewhere).
The general solution for $u$ may be:

\begin{equation}
u = \widehat{u}(y) e^{i(k - c t)}
\label{eq:kelvin_u_sol}
\end{equation}
As for the elevation $\eta$, insert Eq. \ref{eq:kelvin_u_sol} into Eq.
\ref{eq:kelvin_eta} to get:

\begin{equation}
\eta = \sqrt{\frac{g}{H}} \widehat{u}(y) e^{i(k - c t)}
\label{eq:kelvin_eta_sol}
\end{equation}
We still need to solve for $\widehat{u}(y)$, so we look for the equation that
has a derivative with respect to $y$.
So, insert Eqs. \ref{eq:kelvin_u_sol} and \ref{eq:kelvin_eta_sol} into Eq.
\ref{eq:kelvin_v} to get:

\begin{equation}
f \widehat{u}(y) = - \sqrt{\frac{H}{g}} \frac{\partial \widehat{u}(y)}{\partial y}
\end{equation}
which integrates to:

\begin{equation}
\widehat{u}(y) = \widehat{u}_0 e^{-\frac{y}{L_d}}
\label{eq:kelvin_u_sol_y}
\end{equation}
where

\begin{equation}
L_d = \frac{\sqrt{gH}}{f}
\label{eq:rossby_deformation_radius}
\end{equation}
is the \textit{Rossby radius of deformation}\index{deformation!Rossby radius}
\index{Rossby!radius of deformation},
which is the length scale at which planetary rotation becomes important
relative to the effects of gravity (or buoyancy, in stratified flows).
The complete solutions for the shallow water Kelvin waves are then:

\begin{equation}
u = \widehat{u}_0 e^{-\frac{y}{L_d}} e^{i(k - c t)}
\end{equation}

\begin{equation}
\eta = \sqrt{\frac{H}{g}} \widehat{u}_0 e^{-\frac{y}{L_d}} e^{i(k - c t)}
\end{equation}

\begin{figure}[h]
\centering
\includegraphics[width=0.4\textwidth]{assets/fig_kelvin_wave.pdf}
\caption{
Kelvin waves propagating eastward along the equator and decaying rapidly
away to either side.
This is Fig. 4.5 in Vallis (EAOD).
}
\end{figure}

\subsection{Conservative properties}

We now look at some conservative properties of the shallow water equations,
Expand Down Expand Up @@ -3292,6 +3406,98 @@ \subsubsection{Energy}
The total energy of the system $E$ is thus conserved and entirely governed by
the divergence of the energy flux $\mathbf{F}$.

\subsection{Rossby waves}

One emerging pattern from the conservation of potential vorticity arises if
the planetary vorticity $f$ is allowed to vary with latitude.
This is true on a sphere where $f = 2 \Omega \sin(\theta)$, or on a $\beta$-plane
where $f = f_0 + \beta y$.
This pattern is called \textit{Rossby waves}\index{Rossby waves} (also called
\textit{planetary waves}\index{Planetary waves}) and is among the most important
class of motions in both the ocean and the atmosphere.

To derive the solution for Rossby waves, we start from the shallow-water potential
vorticity conservation equation:

\begin{equation}
\frac{d}{dt} \left( \frac{\zeta + f}{h} \right) = 0
\end{equation}
To simplify the derivation, we will assume a flat bottom so that

\begin{equation}
\frac{d(\zeta + f)}{dt} = 0
\end{equation}
Expand the Lagrangian derivative to get:

\begin{equation}
\frac{\partial \zeta}{\partial t} + \left(\mathbf{u} \cdot \nabla \right) \zeta + v \beta = 0
\label{eq:swe_potential_vorticity_beta}
\end{equation}
which is the potential vorticity conservation equation on a $\beta$-plane.

We still have only one equation with two unknowns (relative vorticity $\zeta$
and velocity $\mathbf{u}$), so we somehow need to reduce them to one unknown
variable.
One approach is to introduce a streamfunction $\psi$ such that:

\begin{equation}
(u, v) = \left( - \frac{\partial \psi}{\partial y}, \frac{\partial \psi}{\partial x} \right)
\label{eq:swe_streamfunction}
\end{equation}
We can then express the relative vorticity as:

\begin{equation}
\zeta = \frac{\partial v}{\partial x} - \frac{\partial u}{\partial y} =
\frac{\partial^2 \psi}{\partial x^2} + \frac{\partial^2 \psi}{\partial y^2}
= \nabla^2 \psi
\label{eq:swe_relative_vorticity}
\end{equation}
Insert Eqs. \ref{eq:swe_streamfunction} and \ref{eq:swe_relative_vorticity}
into Eq. \ref{eq:swe_potential_vorticity_beta} to get:

\begin{equation}
\frac{\partial}{\partial t} \nabla^2 \psi + U \frac{\partial}{\partial x} \nabla^2 \psi + \beta \frac{\partial \psi}{\partial x} = 0
\label{eq:swe_streamfunction_beta}
\end{equation}
which is the potential vorticity equation on a $\beta$-plane in terms of the
streamfunction.

As before, assume a wave-like solution but this time for the streamfunction:

\begin{equation}
\psi = \widehat{\psi} e^{i(kx - \omega t)}
\end{equation}
and insert it into Eq. \ref{eq:swe_streamfunction_beta} to get the dispersion
relation for Rossby waves:

\begin{equation}
\omega = U k - \frac{\beta}{k}
\label{eq:swe_rossby_wave_dispersion_1d}
\end{equation}
The phase speed of Rossby waves is:

\begin{equation}
c = \frac{\omega}{k} = U - \frac{\beta}{k^2}
\label{eq:swe_rossby_wave_phase_speed_1d}
\end{equation}

\begin{figure}[h]
\centering
\includegraphics[width=\textwidth]{assets/fig_rossby_wave.pdf}
\caption{
A two-dimensional (x-y) Rossby wave.
An initial disturbance displaces a material line at constant latitude
(the straight horizontal line) to the solid line marked $\eta(t=0)$.
Conservation of potential vorticity, $\zeta + \beta y$, leads to the
production of relative vorticity, $\zeta$, as shown.
The associated velocity field (arrows on the circles) then advects the
fluid parcels, and the material line evolves into the dashed line with
the phase propagating westward.
This is Fig. 6.3 in Vallis (EAOD).
}
\label{fig:swe_rossby_wave}
\end{figure}

\subsection{Exercises}

\begin{enumerate}
Expand Down

0 comments on commit c0fa27b

Please sign in to comment.