Skip to content

Commit

Permalink
Geostrophic balance and Boussinesq
Browse files Browse the repository at this point in the history
  • Loading branch information
milancurcic committed Sep 23, 2024
1 parent 9783dc3 commit 33ebe7f
Show file tree
Hide file tree
Showing 2 changed files with 215 additions and 11 deletions.
Binary file added assets/fig_geostrophic_balance.pdf
Binary file not shown.
226 changes: 215 additions & 11 deletions notes.tex
Original file line number Diff line number Diff line change
Expand Up @@ -1299,7 +1299,7 @@ \subsubsection{Gravity}
This completes the full system of momentum conservation equations in the
Cartesian coordinate system.

\subsection{Hydrostatic approximation}
\subsection{Hydrostatic balance}

Take Eq. \ref{eq:momentum_navier_stokes_scalar_w} and assume that the vertical
acceleration $\frac{dw}{dt}$ is small compared to $g$, and that the spatial
Expand Down Expand Up @@ -1444,6 +1444,7 @@ \subsubsection{In the ocean}
equation of state for seawater is not easily derived.
Instead, we assume that the ocean is a single-component fluid, and we use the
density field $\rho$ as the equation of state.

\begin{equation}
\rho = \rho(T, S, p) \\
= \rho_0 \left[ 1 - \beta_T(T-T_0) + \beta_S(S-S_0) + \beta_p(p-p_0) \right]
Expand Down Expand Up @@ -1502,6 +1503,7 @@ \subsubsection{In the ocean}
numerical ocean circulation models.

\subsection{Nondimensionalization and scaling}
\label{sec:nondimensionalization_and_scaling}

A useful technique to simplify the analysis of the governing equations is to
scale the variables using characteristic values for each of the variables.
Expand Down Expand Up @@ -1607,7 +1609,7 @@ \subsection{Exercises}
(c) a river inflow into the ocean;
(d) a breaking ocean surface wave;
(e) water flowing through a pipe with a diameter of 0.1 m and flow speed of 1 m s$^{-1}$.
Assume $nu = 10^{-5} m^2 s^{-1}$ for air and $nu = 10^{-6} m^2 s^{-1}$ for water.
Assume $\nu = 10^{-5} m^2 s^{-1}$ for air and $\nu = 10^{-6} m^2 s^{-1}$ for water.

\end{enumerate}

Expand Down Expand Up @@ -1993,7 +1995,113 @@ \subsection{f-plane and $\beta$-plane approximations}
where $f_0 = 2\Omega \sin\theta_0$ and $\beta = \partial f/\partial y = (2\Omega\cos\theta_0) / R_E$
(where $R_E$ is the radius of the Earth).

\subsection{Geostrophic approximation}
\subsection{Geostrophic balance}

Now that we have incorporated the effects of rotation into our equations of motion,
let's evaluate the scales of the terms in the horizontal momentum equations.
We will start from Eq. \ref{eq:momentum_navier_stokes_rotating}, use the f-plane
notation for the Coriolis term, ignore the viscous terms, and drop the gravity
term as we're looking at the flow in the horizontal plane:

\begin{equation}
\frac{\partial \mathbf{u}}{\partial t} +
(\mathbf{u} \cdot \nabla) \mathbf{u} +
\mathbf{f} \times \mathbf{u} =
- \frac{1}{\rho} \nabla p
\end{equation}

As we did in Section \ref{sec:nondimensionalization_and_scaling}, let's scale
each term on the left-hand side with their characteristic scales for mesoscale
ocean flow ($L \sim 10^5\ m$, $T \sim 10^6\ s$, $U \sim 10^{-1}\ m/s$):

\begin{itemize}
\item $\frac{\partial \mathbf{u}}{\partial t} \sim \frac{U}{T} \sim 10^{-7}$
\item $(\mathbf{u} \cdot \nabla) \mathbf{u} \sim \frac{U^2}{L} \sim 10^{-7}$
\item $\mathbf{f} \times \mathbf{u} \sim f_0 U \sim 10^{-6}$
\end{itemize}
This means that on these oceanic scales ($L \sim 100\ km$, $T \sim 1\ day$),
the inertial terms are of the same order of magnitude as the Coriolis term.
In other words, rotation here is much more important than the local rate of
change or advection.
Also, whatever the scale of the pressure gradient term is, it is the only
term that can balance the rotation.
Thus, if we can state that the inertial terms can be neglected, we can also
state:

\begin{equation}
\mathbf{f} \times \mathbf{u} \approx - \frac{1}{\rho} \nabla p
\end{equation}
or, in scalar component form:

\begin{equation}
f u \approx - \frac{1}{\rho} \frac{\partial p}{\partial y}
\end{equation}

\begin{equation}
f v \approx \frac{1}{\rho} \frac{\partial p}{\partial x}
\end{equation}

This balance is called the
\textit{geostrophic balance}\index{Geostrophic!balance}\index{Balance!geostrophic},
and it is a key concept in geophysical fluid dynamics.
It states that the flow is governed by the balance between the rotation and the
pressure gradient force.
Although the geostrophic balance is strictly an approximation and it never holds
exactly, large scale oceanic ($L \sim 100\ km$ and larger) and atmospheric
($L \sim 1000\ km$ and larger) flows are often in geostrophic balance.
For the analysis of geophysical flows at such scales, it is then useful to
define the \textit{geostrophic velocity}\index{Geostrophic!velocity} as:

\begin{equation}
u_g = - \frac{1}{\rho f} \frac{\partial p}{\partial y}
\end{equation}

\begin{equation}
v_g = \frac{1}{\rho f} \frac{\partial p}{\partial x}
\end{equation}
Notice that the geostrophic flow is always perpendicular to the pressure gradient,
which means it is parallel to the isobars (lines of constant pressure).
This also means that the isobars are streamlines of the geostrophic flow.
In the northern hemisphere ($f > 0$), the geostrophic flow is cyclonic
(counter-clockwise) around the low-pressure region and anti-cyclonic
(clockwise) around the high-pressure region.
In the southern hemisphere ($f < 0$), it is the opposite.
A nearly geostrophic flow is illustrated in Fig. \ref{fig:geostrophic_flow}.

\begin{figure}[h]
\centering
\includegraphics[width=0.8\textwidth]{assets/fig_geostrophic_balance.pdf}
\caption{
Geostrophic flow with a positive value of the Coriolis parameter $f$.
Flow is parallel to the lines of constant pressure (isobars).
Cyclonic flow is anticlockwise around a low pressure region and
anticyclonic flow is clockwise around a high. If $f$ were negative, as in
the Southern Hemisphere, (anti)cyclonic flow would be (anti)clockwise.
This is Fig. 2.5 in AOFD (Vallis, 2017).
}
\label{fig:geostrophic_flow}
\end{figure}

\subsection{Rossby number}

Recall that we required the inertial terms to be much smaller than the Coriolis
term for the geostrophic approximation to hold.
Like we did earlier with the Reynolds number to quantify how turbulent a flow is,
we can define the \textit{Rossby number}\index{Rossby!number} as:

\begin{equation}
\text{Ro} \equiv
\frac{\text{Advection}}{\text{Rotation}} =
\frac{\left( \mathbf{u} \cdot \nabla \right) \mathbf{u}}{\mathbf{f} \times \mathbf{u}}
\approx \frac{\frac{U^2}{L}}{fU}
\approx \frac{U}{fL}
\end{equation}
Though the Rossby number characterizes the relative importance of rotation in
the flow, notice that the rotation term is in the denominator.
The Rossby number is thus small for flows in which rotation dominates over
advection.
In general, flows with a Rossby number of 0.1 or smaller are considered
approximately geostrophically balanced.

\subsection{Exercises}

Expand Down Expand Up @@ -2054,6 +2162,8 @@ \subsection{The Boussinesq equations}
\frac{d p_0}{d z} = - \rho_0 g
\end{equation}

\subsection{Momentum balance}

Let's first apply the Boussinesq approximation to the momentum balance.
Recall the Navier-Stokes equation with rotation
(Eq. \ref{eq:momentum_navier_stokes_rotating}), while neglecting the viscosity
Expand All @@ -2063,10 +2173,10 @@ \subsection{The Boussinesq equations}
\frac{d \mathbf{u}}{dt} = - \frac{1}{\rho} \nabla p - f \mathbf{k} \times \mathbf{u} + \mathbf{g}
\end{equation}
Apply Eqs. \ref{eq:boussinesq_density}-\ref{eq:boussinesq_pressure} to the
above equation:
above equation to get:

\begin{equation}
\left( \rho_0 + \delta \rho \right) \left( \frac{d \mathbf{u}}{dt} + f \mathbf{k} \times \mathbf{u} \right) =
\left( \rho_0 + \delta \rho \right) \left( \frac{d \mathbf{u}}{dt} + \mathbf{f} \times \mathbf{u} \right) =
- \nabla \left( p_0 + \delta p \right)
+ \left( \rho_0 + \delta \rho \right) \mathbf{g}
\end{equation}
Expand All @@ -2080,23 +2190,95 @@ \subsection{The Boussinesq equations}
on the left-hand side:

\begin{equation}
\rho_0 \left( \frac{d \mathbf{u}}{dt} + f \mathbf{k} \times \mathbf{u} \right) =
\rho_0 \left( \frac{d \mathbf{u}}{dt} + \mathbf{f} \times \mathbf{u} \right) =
- \nabla \delta p + \delta \rho\ \mathbf{g}
\end{equation}

\begin{equation}
\frac{d \mathbf{u}}{dt} + f \mathbf{k} \times \mathbf{u} =
\frac{d \mathbf{u}}{dt} + \mathbf{f} \times \mathbf{u} =
- \frac{1}{\rho_0} \nabla \delta p + \frac{\delta \rho}{\rho_0} \mathbf{g}
\end{equation}
We can express $- g \delta \rho / \rho_0$ as $b$, \textit{buoyancy}\index{buoyancy}:
For convenience of notation, let's now define \textit{buoyancy}\index{buoyancy}
as $b = - g \delta \rho / \rho_0$, and re-write the above to obtain the
Boussinesq momentum equation:

\begin{equation}
\frac{d \mathbf{u}}{dt} + f \mathbf{k} \times \mathbf{u} =
\frac{d \mathbf{u}}{dt} + \mathbf{f} \times \mathbf{u} =
- \frac{1}{\rho_0} \nabla \delta p + b \mathbf{k}
\end{equation}
This equation states that now that we are in a gradually stratified fluid,
the gravity term is scaled by $\delta \rho / \rho_0$ to yield the appropriate
vertical acceleration, and the pressure gradient is due to the relatively
small perturbations in density $\delta \rho$ around the mean density $\rho_0$.

\subsection{Continuity}

As we did for the momentum equation, we'll now apply the Boussinesq approximation
(\textit{i.e.} $\rho = \rho_0 + \delta \rho$, $\delta \rho \ll \rho_0$) to the
continuity equation.
Recall the continuity equation in its complete form:

\begin{equation}
\frac{d \rho}{dt} + \rho \nabla \cdot \mathbf{u} = 0
\end{equation}
Insert Eq. \ref{eq:boussinesq_density} to get:

\begin{equation}
\frac{d\delta \rho}{dt} + \left( \rho_0 + \delta \rho \right) \nabla \cdot \mathbf{u} = 0
\end{equation}
Then, if we can state that that $d\delta \rho / dt \ll \rho_0 \nabla \cdot \mathbf{u}$,
which we will for the Boussinesq approximation, we recover the original
continuity equation for incompressible flows:

\begin{equation}
\nabla \cdot \mathbf{u} = 0
\end{equation}
Note that we do not say that strictly $d \delta \rho / dt = 0$, but rather that
we can neglect it in this equation in favor of the velocity divergence term.
The evolution of $\delta \rho$ is still governed by the evolution of buoyancy,
which in turn is governed by the evolution of the temperature and salinity fields
and the equation of state.
The buoyancy $b = - g \delta \rho / \rho_0$ evolves as:

\begin{equation}
\frac{d b}{dt} = \dot{b}
\end{equation}
and the equation of state can be expressed in terms of buoyancy:

\begin{equation}
b = b(T, S, p)
\end{equation}
which is just another form of Eq. \ref{eq:equation_of_state_ocean}.

Finally the temperature and salinity evolve as before, following
Eqs. \ref{eq:temperature_equation_ocean} and \ref{eq:salinity_equation_ocean},
respectively.

\subsection{Complete system of equations}

The full system of Boussinesq equations for the ocean are then:

\begin{equation}
\frac{d \mathbf{u}}{dt} + \mathbf{f} \times \mathbf{u} =
- \frac{1}{\rho_0} \nabla \delta p + b \mathbf{k}
\end{equation}

\begin{equation}
\nabla \cdot \mathbf{u} = 0
\end{equation}

\begin{equation}
\frac{d T}{dt} = \dot{T}
\end{equation}

\begin{equation}
\frac{d S}{dt} = \dot{S}
\end{equation}

\begin{equation}
b = b(T, S, p)
\end{equation}

%\newpage
%\section{Shallow water systems}

%\newpage
%\section{Boundary layers}
Expand Down Expand Up @@ -2292,6 +2474,28 @@ \section{Quick reference}
\beta = \frac{\partial f}{\partial y} = \frac{2\Omega\cos(\theta_0)}{R_E}
\end{equation}

\textbf{Geostrophic balance:}

\begin{equation}
\mathbf{f} \times \mathbf{u} = - \frac{1}{\rho} \nabla p
\end{equation}

\textbf{Geostrophic velocity:}

\begin{equation}
u_g = - \frac{1}{\rho f} \frac{\partial p}{\partial y}
\end{equation}

\begin{equation}
v_g = \frac{1}{\rho f} \frac{\partial p}{\partial x}
\end{equation}

\textbf{Rossby number:}

\begin{equation}
\text{Ro} \equiv \frac{\left( \mathbf{u} \cdot \nabla \right) \mathbf{u}}{\mathbf{f} \times \mathbf{u}} \approx \frac{U}{fL}
\end{equation}

\printindex

\end{document}

0 comments on commit 33ebe7f

Please sign in to comment.