Skip to content

Commit

Permalink
WIP linear wave theory
Browse files Browse the repository at this point in the history
  • Loading branch information
milancurcic committed Nov 18, 2024
1 parent 6b6a389 commit 6ce31cd
Showing 1 changed file with 186 additions and 24 deletions.
210 changes: 186 additions & 24 deletions notes.tex
Original file line number Diff line number Diff line change
Expand Up @@ -4309,15 +4309,22 @@ \subsection{Turbulent cascade}
\newpage
\section{Boundary layers}

\noindent
\textbf{
Caution: I use some figures from Pope's Turbulent Flows (Chapter 7).
Some of the mathematical notation, such as the
vertical and cross-stream coordinates, and the averaging operators
(angle brackets and overlines) are reverse from the convention that we use
in the main text.
Be aware of this until I implement my own figures.
}
Boundary layers occur when a fluid flows over some kind of boundary, whether
rigid or free, stationary or moving.
They are both interesting and convenient because they constrain the flow near
the boundary and thus allow simplifications that may lead to analytical solutions.
They are important because they are often the dominant flow structure in geophysical
flows.
For example, a planetary boundary layer separates the atmosphere from the surface
of the Earth.
The surface beneath the planetary boundary layer may be rigid (land or sea ice)
or free (ocean), and its roughness and thermodynamic properties may vary greatly
from place to place.
In this chapter, we start from the simplest boundary layer, a channel flow, and
derive the stress and mean velocity profiles in laminar flows.
Then, we zoom into the vertical structure of the boundary layer in turbulent
flows, and examine different regimes that occur depending on the distance from
the boundary.

\subsection{Channel flow}

Expand Down Expand Up @@ -4465,7 +4472,7 @@ \subsubsection{Governing equations and boundary conditions}
The stress thus decreases linearly from $\tau_w$ at the bottom wall to zero at
the centerline, reaching $-\tau_w$ at the top wall.

\subsubsection{Laminar flow}
\subsection{Laminar flow}

What does the velocity profile look like in the case of laminar flow?
We can drop the Reynolds stress term in Eq. (\ref{eq:channel_tau}) and combine
Expand All @@ -4490,7 +4497,7 @@ \subsubsection{Laminar flow}
laminar flows, \textit{i.e.} for relatively small Reynolds numbers.
Let's see what the profiles may look like in turbulent flows.

\subsubsection{Turbulent flow}
\subsection{Turbulent flow}

\begin{figure}[h]
\centering
Expand Down Expand Up @@ -4655,7 +4662,7 @@ \subsubsection{Turbulent flow}
Let's now examine in more detail each of these regions and see if flow structure
varies significantly between them.

\subsubsection{Velocity structure in various wall regions}
\subsection{Velocity structure in various wall regions}

Now, let's look at the time-mean velocity profiles in the turbulent channel flow,
and in various regions near and away from the wall.
Expand Down Expand Up @@ -4879,11 +4886,24 @@ \section{Surface gravity waves}
and so these waves are often called \textit{gravity waves}\index{Gravity waves},
much like the waves we explored as a solution of the shallow water equations in
Chapter \ref{sec:shallow_water_systems}.

\subsection{Governing equations}

\subsubsection{Velocity potential}

In the first part of this chapter, we derive the solution for the
small-amplitude (also often called linear) waves, which are valid when the
wave amplitude is much smaller than the wavelength and the water depth.
This assumption allows for a relatively straightforward solution of the flow
anywhere below the free wavy surface.
Although simplistic in its approximations, the linear wave theory has been
surprisingly successful in predicting the behavior of the waves even when the
assumptions behind it are clearly violated.
The linear wave theory remains the basis of modern wave prediction models that
are used in operational weather and ocean forecasting.
After deriving the linear wave solutions, we will explore their properties
and derive some second-order quantities with implication on mean ocean
circulation.

\subsection{Small-amplitude wave derivation}

\subsubsection{Governing equations}
\label{sec:governing_equations}
Key assumptions are that the fluid is incompressible
($\nabla \cdot \mathbf{u} = 0$), inviscid ($\nu \nabla^2 \mathbf{u} = 0$),
and irrotational ($\nabla \times \mathbf{u} = 0$).
Expand Down Expand Up @@ -4913,13 +4933,7 @@ \subsubsection{Velocity potential}
Although $\phi$ is allowed to vary in both space and time, the Laplace equation
states that at any given time, $\phi$ anywhere in the interior of the fluid is
determined by its values at the boundary (\textit{i.e.} the boundary conditions).

\subsubsection{The Bernoulli equation}
\label{sec:bernoulli}

Although the Laplace equation governs the spatial distribution of the velocity
potential depending on its values at the boundaries, it does not determine how
it evolves in time.
It does not, however, determine how $\phi$ evolves in time.
To do that, we can integrate the Euler equations of motion (introduced back in
Section \ref{sec:momentum}, see Eq. \ref{eq:momentum_euler}) to obtain a
steady-state relationship between the pressure and the velocity of the fluid.
Expand Down Expand Up @@ -5010,10 +5024,158 @@ \subsubsection{The Bernoulli equation}
\frac{\partial \phi}{\partial t} +
\frac{1}{2} \left(u^2 + w^2\right) +
\frac{p}{\rho} + gz= C(t)
\label{eq:bernoulli}
\end{equation}
The Bernoulli equation will serve as a dynamic free surface boundary condition
as we proceed to derive the solutions for the surface gravity waves.

\subsubsection{Boundary conditions}
\label{sec:boundary_conditions}

Now that we established the governing equations to solve, we need to specify the
boundary conditions to determine the velocity potential in the interior.
We will rely on a total of four boundary conditions:

\begin{enumerate}
\item \textbf{Kinematic free surface boundary condition}:
This boundary condition determines the vertical velocity at the free surface
$\eta(x, t)$ by exploiting the fact that the Lagrangian (material) change of
the vertical position is the vertical velocity itself:
\begin{equation}
w = \frac{dz}{dt}\Big|_{z=\eta} = \frac{\partial \eta}{\partial t} + u \frac{\partial \eta}{\partial x}
\end{equation}
Expressed in terms of the velocity potential, this boundary condition becomes:
\begin{equation}
\frac{\partial \phi}{\partial z} =
\frac{\partial \eta}{\partial t} +
\frac{\partial \phi}{\partial x} \frac{\partial \eta}{\partial x}, \text{ at } z=\eta(x, t)
\end{equation}
\item \textbf{Dynamic free surface boundary condition}:
We leverage the Bernoulli equation (Eq. \ref{eq:bernoulli}) at the free surface
($z = \eta$) and set the surface pressure to be zero:
\begin{equation}
\frac{\partial \phi}{\partial t} + \frac{1}{2} \left(u^2 + w^2\right) + g\eta = C(t), \text{ at } z=\eta(x, t)
\end{equation}
\item \textbf{Bottom boundary condition}:
The bottom is rigid and impermeable, so the vertical velocity is zero at the
bottom:
\begin{equation}
w = 0, \text{ at } z = -h
\end{equation}
where $h$ is the mean depth of the fluid.
\item \textbf{Lateral boundary condition}:
At the lateral boundaries, since we're seeking a wave solution, we know that
the velocity potential must be periodic in the horizontal space as well as
time:
\begin{equation}
\phi(x, t) = \phi(x+L, t)
\end{equation}
\begin{equation}
\phi(x, t) = \phi(x, t+T)
\end{equation}
where $L$ is the wavelength and $T$ is the period.
\end{enumerate}
With these four boundary conditions, we are now equipped to solve for the velocity
potential in the interior of the fluid.

\subsubsection{Solution}

Our key equation to solve is the Laplace equation (Eq. \ref{eq:laplace})
for the velocity potential $\phi$ that varies in the horizontal and vertical
direction $x$ and $z$ respectively, as well as time $t$:

\begin{equation}
\nabla^2 \phi(x, z, t) = 0
\label{eq:laplace2}
\end{equation}
To solve this equation, we will rely on the method of separation of variables,
where we assume that the solution can be written as a product of functions that
depend on each coordinate separately:

\begin{equation}
\phi(x, z, t) = \phi_x(x) \phi_z(z) \phi_t(t)
\end{equation}
We can start from the time-dependent part $\phi_t(t)$ and recall the lateral
boundary condition which states that the velocity potential must be periodic
in time, which is true for sines and cosines (and some combinations of them).
For a sine function of a phase $\varphi$, this is true:

\begin{equation}
\sin(\varphi) = \sin(\varphi + 2\pi)
\end{equation}
And expressing it as a function of time:

\begin{equation}
\sin(\omega t) = \sin(\omega t + 2\pi)
\end{equation}
where $\omega$ is the angular frequency in units of radians per second, so that
the phase $\varphi$ has angle units (radians).
We can then write the velocity potential as:

\begin{equation}
\phi(x, z, t) = \phi_x(x) \phi_z(z) \sin(\omega t)
\label{eq:phi1}
\end{equation}
Insert this into Eq. (\ref{eq:laplace2}) to get:

\begin{equation}
\frac{\partial^2 \phi_x}{\partial x^2} \phi_z \sin(\omega t) +
\phi_x \frac{\partial^2 \phi_z}{\partial z^2} \sin(\omega t) = 0
\end{equation}
Divide by $\phi_x \phi_z \sin(\omega t)$ to get:

\begin{equation}
\frac{1}{\phi_x} \frac{\partial^2 \phi_x}{\partial x^2} +
\frac{1}{\phi_z} \frac{\partial^2 \phi_z}{\partial z^2} = 0
\label{eq:phi2}
\end{equation}
Can we separate this even further?
Recall that $\phi_x$ and $\phi_z$ are functions of $x$ and $z$ respectively.
If, for example, we hold $x$ constant and consider variations in $z$, the
first term would remain constant but the second term would not!
You can arrive to the same conclusion by holding $z$ constant and varying $x$.
This would clearly violate Eq. (\ref{eq:phi2}), and so the only way that
equation can hold is if both $\phi_x$ and $\phi_z$ are equal to the same
constant but with opposite signs:

\begin{equation}
\frac{1}{\phi_x} \frac{\partial^2 \phi_x}{\partial x^2} = -k^2
\label{eq:phi3}
\end{equation}

\begin{equation}
\frac{1}{\phi_z} \frac{\partial^2 \phi_z}{\partial z^2} = k^2
\label{eq:phi4}
\end{equation}
where $k$ is the separation constant.
These can also be written as:

\begin{equation}
\frac{\partial^2 \phi_x}{\partial x^2} + k^2 \phi_x = 0
\label{eq:phi5}
\end{equation}

\begin{equation}
\frac{\partial^2 \phi_z}{\partial z^2} - k^2 \phi_z = 0
\label{eq:phi6}
\end{equation}
For real values of $k$, the solutions to these equations are:

\begin{equation}
\phi_x(x) = A \sin(kx) + B \cos(kx)
\end{equation}

\begin{equation}
\phi_z(z) = C e^{kz} + D e^{-kz}
\end{equation}
where $A$, $B$, $C$, and $D$ are constants that are yet to be determined.
We now write our intermediate solution for the velocity potential as:

\begin{equation}
\phi(x, z, t) = \left[ A \sin(kx) + B \cos(kx) \right] \left[ C e^{kz} + D e^{-kz} \right] \sin(\omega t)
\label{eq:phi_inter}
\end{equation}

\newpage
\appendix

Expand Down

0 comments on commit 6ce31cd

Please sign in to comment.