notes

uni notes
git clone git://popovic.xyz/notes.git
Log | Files | Refs

chap1.tex (25268B)


      1 \section{Governing Equations of Fluid Mechanics}
      2 We first start off with a fluid with a density given by
      3 \begin{align}
      4     \rho(\mathbf{x}, t),
      5 \end{align}
      6 in three dimensional Cartesian coordinates $\mathbf{x} = (x, y, z)$ at time
      7 $t$. For water-wave applications, we should note that we take
      8 $\rho=\text{constant}$, but we will go into this fact later. The fluid moves
      9 in time and space with a velocity field
     10 \begin{align}
     11     \mathbf{u}(\mathbf{x}, t) = (u, v, w).
     12 \end{align}
     13 Additionally it is also described by its pressure
     14 \begin{align}
     15     P(\mathbf{x}, t),
     16 \end{align}
     17 generally depending on time and position. When thinking of e.g. water the
     18 pressure increases the deeper we go, that is with decreasing or increasing $z$
     19 direction (depending on how we set up our system $z$ pointing up or down
     20 respectively).
     21 
     22 The general assumption in fluid mechanics is the \textbf{Continuum
     23 Hypothesis}, which assumes continuity of $\textbf{u}, \rho$ and $P$ in
     24 $\mathbf{x}$ and $t$. In other words, we premise that the velocity field,
     25 density and pressure are ''nice enough`` functions of position and time, such
     26 that we can do all the differential operations we desire in the framework of
     27 fluid mechanics.
     28 \subsection{Mass Conservation}
     29 Our aim is to derive a model of the fluid and its mechanics, with respect to
     30 time and position, in the most general way. This is usually done thinking
     31 of the density of a given fluid, which is a unit mass per unit volume,
     32 intrinsically  an integral representation to derive these equations suggests
     33 by itself.
     34 
     35 Let us now think of an arbitrary fluid. Within this fluid we define a fixed
     36 volume $V$ relative to a chosen inertial frame and bound it by a surface $S$
     37 within the fluid, such that the fluid motion $\mathbf{u}(\mathbf{x}, t)$ may
     38 cross the surface $S$. The fluid density is given by $\rho(\mathbf{x}, t)$,
     39 thereby the mass of the fluid in the defined Volume $V$ is an integral
     40 expression
     41 \begin{align}
     42     m = \int_V \rho(\mathbf{x}, t) dV.
     43 \end{align}
     44 The figure bellow \ref{fig:volume}, expresses the above described picture.
     45 \begin{figure}[H]
     46     \centering
     47   \begin{tikzpicture}[>=latex,scale=1, xscale=1, opacity=.8]
     48 % second sphere
     49     \begin{scope}[rotate=10, xscale=3, yscale=2, shift={(2.3,-0.2)}]
     50       \coordinate (O) at (0,0);
     51       \shade[ball color=gray!10!] (0,0) coordinate(Hp) circle (1) ;
     52 
     53       \draw[thick] (O) circle (1);
     54       \draw[rotate=5] (O) ellipse (1cm and 0.66cm);
     55       \draw[rotate=90] (O) ellipse (1cm and 0.33cm);
     56 \node[circle, fill=black, inner sep=1pt] at (0.15, 0.25) {} ; \draw[-latex, thick] (0.15, 0.25) -- (1, 1) ;
     57       \node[right] at (1, 1) {$\mathbf{u}(\mathbf{x}, t)$};
     58 
     59       \node[] at (O) {$V$};
     60       \node[] at (0.55, -0.25) {$\rho(\mathbf{x}, t)$};
     61 
     62       \draw[-] (0.76, -0.66) -- (1.2, -0.7);
     63       \node[right] at (1.2, -0.7) {$S$};
     64 
     65       \draw[-latex, thick] (-0.25, -0.65) -- (-1, -1);
     66       \node[left] at (-1, -1) {$\mathbf{n}$};
     67 
     68     \end{scope}
     69 
     70 % axis
     71   \end{tikzpicture}
     72   \caption{Volume bounded by a surface in a fluid with density and momentum,
     73   with a surface normal vector $\mathbf{n}$ \label{fig:volume}}
     74 \end{figure}
     75 
     76 Since we want to figure out the fluid's mechanics, we can consider the rate
     77 of change in the completely arbitrary volume $V$, by
     78 \begin{align}
     79     \frac{d}{dt}\left( \int_V \rho(\mathbf{x}, t)\ dV \right) = \int_V
     80     \frac{\partial \rho(\mathbf{x}, t)}{\partial t} \ dV
     81 \end{align}
     82 On the other hand we have that density is mass over volume, meaning
     83 \begin{align}
     84     dm = \rho \cdot dV.
     85 \end{align}
     86 The infinitesimal volume has the base area $dS$ with hight $h$, which is the
     87 distance in the direction perpendicular to to the base area, leaving us with
     88 $dV = h\ dS$. By definition $\mathbf{n}$ is perpendicular to $dS$, we have
     89 that $h = l \mathbf{n}$. Where $l$ is the unit of length, or velocity times
     90 time $l = \mathbf{u}\ dt$, since mass is flowing out of the surface we
     91 change the sign of the flow leading us to
     92 \begin{align}
     93    dm  = -\rho dV = -\rho \mathbf{u} \cdot \mathbf{n}\ dt\ dS,
     94 \end{align}
     95 All together we have
     96 \begin{align}
     97     \frac{dm}{dt} = -\int_{\partial V} \rho(\mathbf{x},t)
     98     \mathbf{u}\cdot\mathbf{n}\ dS.
     99 \end{align}
    100 Putting both equations on one side leaves us with the equation
    101 \begin{align}\label{eq:mass balance}
    102      \int_V \frac{\partial \rho(\mathbf{x}, t)}{\partial t}\ dV
    103     +\int_{\partial V} \rho(\mathbf{x}, t) \mathbf{u}\cdot\mathbf{n}\ dS
    104     = 0.
    105 \end{align}
    106 The above equation in \ref{eq:mass balance} is an underlying equation, describing that the rate of
    107 change of mass in V is brought about, only by the rate of mass flowing into
    108 V across S, and thus the mass does not change.
    109 
    110 For the second integral in \ref{eq:mass balance} we utilize the Gaussian
    111 integration law to acquire an integral over the volume
    112 \begin{align}
    113     \int_{\partial V} \rho(\mathbf{x}, t) \mathbf{u} \cdot \mathbf{n} \ dS =
    114     \int_V \nabla (\rho \mathbf{u})\ dV.
    115 \end{align}
    116 Thereby we can put everything inside the volume integral
    117 \begin{align}
    118     \int_V \left(\partial_t \rho + \nabla(\rho \mathbf{u}) \right) \ dV = 0.
    119 \end{align}
    120 Everything under the integral sign needs to be zero, thus we obtain
    121 the \textbf{Equation of Mass Conservation} or in the general sense also
    122 called the \textbf{Continuity Equation}
    123 \begin{align}\label{eq:continuity}
    124     \partial_t \rho + \nabla(\rho \mathbf{u}) = 0
    125 \end{align}
    126 
    127 In light of the results of the equation of mass conservation
    128 in \ref{eq:continuity}, the product rule gives
    129 \begin{align}
    130     \partial_t \rho + (\nabla \rho)\mathbf{u} + \rho(\nabla \mathbf{u}),
    131 \end{align}
    132 for notational purposes, we define the \textbf{material/convective derivative}
    133 as follows
    134 \begin{align}
    135     \frac{D}{Dt} = \frac{\partial }{\partial t}  + \mathbf{u}\nabla.
    136 \end{align}
    137 With the material derivative the equation of mass conservation reads
    138 \begin{align}
    139     \frac{D\rho}{Dt} + \rho \nabla\mathbf{u} = 0
    140 \end{align}
    141 We may undertake the first case separation, initiating $\rho = \text{cosnt.}$
    142 called \textbf{incompressible flow} causes the material derivative of $\rho$ to
    143 be zero, and thereby
    144 \begin{align}
    145     \frac{D\rho}{Dt} = 0 \quad \Rightarrow \quad \nabla \mathbf{u} = 0,
    146 \end{align}
    147 following that the divergence of the velocity field is zero, in this case
    148 $\mathbf{u}$ is called \textbf{solenoidal}.
    149 \subsection{Euler's Equation of Motion}
    150 Additional consideration we undertake is the assumption of an
    151 \textbf{inviscid} fluid, that is we set viscosity to zero. Otherwise we would
    152 get a viscous contribution under the integral which results in the
    153 Navier-Stokes equation. In this regard we apply Newton's second law to our
    154 fluid in terms of infinitesimal pieces $\delta V$ of the fluid. The
    155 acceleration divides into two terms, a \textbf{body force} given by gravity
    156 of earth in the $z$ coordinate $\mathbf{F} = (0, 0, -g)$ and a
    157 \textbf{local/short-rage force} described by the stress tensor in the fluid.
    158 In the inviscid case, the local force retains the pressure $P$, producing a
    159 normal force, with respect to the surface, acting onto the infinitesimal
    160 element in the fluid. Summing over all infinitesimal pieces of the fluid,
    161 gives us an integral formulation of the force
    162 \begin{align}
    163     \int_V \rho \mathbf{F}\ dV - \int_S P\mathbf{n}\ dV.
    164 \end{align}
    165 Now applying the Gaussian rule of integration on the second integral over the
    166 surface, the resulting force in per unit volume is
    167 \begin{align}
    168     \int_V \left(\rho \mathbf{F} - \nabla P\right)\ dV.
    169 \end{align}
    170 The acceleration of the fluid particles is given by $\frac{D\mathbf{u}}{Dt}$,
    171 and thus the total force per unit volume on the other hand is
    172 \begin{align}
    173     \int_V \rho \frac{D\mathbf{u}}{Dt}\ dV =
    174     \int_V \left(\rho \mathbf{F} - \nabla P\right)\ dV.
    175 \end{align}
    176 Newton's Second Law for a fluid in a Volume is essentially saying that the
    177 rate of change of momentum of the fluid in the fixed volume $V$, which is the particle
    178 acceleration is the resulting force acting on V together with the rate of
    179 flow of momentum across the surface $S$ into the volume $V$. Hence we arrive
    180 at the \textbf{Euler's Equation(s) of Motion}
    181 \begin{align}
    182     \frac{D\mathbf{u}}{Dt} = \left(\frac{\partial \mathbf{u}}{\partial t}
    183     +(\mathbf{u}\nabla)\mathbf{u}\right)  =
    184     -\frac{1}{\rho}\nabla P + \mathbf{F}.
    185 \end{align}
    186 As a side note we have mentioned that there is another contribution if the
    187 fluid is viscid. Indeed there is a tangential force due to the velocity
    188 gradient, which into introduces the additional term
    189 \begin{align}
    190     \mu \nabla^2 \mathbf{u}, \qquad
    191     \mu = \text{viscosity of the Fluid}.
    192 \end{align}
    193 Thereby the equations become
    194 \begin{align}
    195     \rho\frac{D\mathbf{u}}{Dt}
    196     =  -\nabla P + \rho \mathbf{F} + \mu \nabla^2 \mathbf{u}.
    197 \end{align}
    198 
    199 For now we have separated two simplifications, that define an
    200 \textbf{idealized/perfect fluid}
    201 \begin{enumerate}
    202     \item \textbf{inviscid} $\qquad \mu=0$
    203     \item  \textbf{incompressible} $\quad \rho = \text{const.},\ \nabla \mathbf{u}=
    204         0$
    205 \end{enumerate}
    206 \subsection{Vorticity and irrotational Flow}
    207 The curl of the velocity field $\mathbf{\omega} = \nabla \times \mathbf{u}$
    208 of a fluid (i.e. the vorticity), describes a ``spinning'' motion of the fluid
    209 near a position $\mathbf{x}$ at time $t$. The vorticity is an important
    210 property of a fluid. Flows or regions of flows where $\mathbf{\omega}=0$ are
    211 \textbf{irrotational}, and thus can be modeled and analyzed following well
    212 known routine methods. Even though real flows are rarely irrotational
    213 anywhere (!), in water wave theory wave problems, from the classical aspect
    214 of vorticity has only minor contributions. Hence we can assume irrotational flow
    215 modeling water waves. To arrive at the vorticity in the equations of motions
    216 derived in the last section we resort to a differential identity derived in appendix
    217 \ref{appendix:diff identity}, rewriting the motion derivative in terms of
    218 \begin{align}
    219     \frac{D\mathbf{u}}{Dt} = \frac{\partial \mathbf{u}}{\partial t}
    220     +\nabla(\frac{1}{2}\mathbf{u}\mathbf{u)}
    221     - \left( \mathbf{u}\times (\nabla \times  \mathbf{u} \right).
    222 \end{align}
    223 Thus the equations of motion become
    224 \begin{align}
    225     \frac{\partial \mathbf{u}}{\partial t} + \nabla\left(
    226     \frac{1}{2}\mathbf{u}\mathbf{u} + \frac{P}{\rho} + \Omega \right)
    227     = \mathbf{u} \times  \mathbf{\omega},
    228 \end{align}
    229 where $\Omega$ is the force potential per
    230 unite mass given by $\mathbf{F} = -\nabla \Omega$.
    231 
    232 At this point we may differentiate between \textbf{steady and unsteady flow}.
    233 For \textbf{Steady Flow} we assume that $\mathbf{u}, P$ and $\Omega$ are time
    234 independent, thus we get
    235 \begin{align}
    236       \nabla\left( \frac{1}{2}\mathbf{u}\mathbf{u} + \frac{P}{\rho} + \Omega
    237       \right)  = \mathbf{u} \times  \mathbf{\omega}.
    238 \end{align}
    239 It is general knowledge that the gradient of a function $\nabla f$ is
    240 perpendicular the level sets of $f(\mathbf{x})$, where $f(\mathbf{x}) =
    241 \text{const.}$. Thus $\mathbf{u} \times  \mathbf{\omega}$ is orthogonal to
    242 the surfaces  where
    243 \begin{align} \label{eq:bernoulli}
    244     \frac{1}{2}\mathbf{u}\mathbf{u} + \frac{P}{\rho} + \Omega =
    245     \text{const.},
    246 \end{align}
    247 The above equation is called \textbf{Bernoulli's Equation}.
    248 
    249 Secondly \textbf{Unsteady Flow} but irrotational (+ incompressible), first of
    250 all gives us the condition for the existence of a velocity potential $\phi$
    251 in the sense
    252 \begin{align}
    253     \mathbf{\omega} = \nabla \times  \mathbf{u} = 0  \quad \Rightarrow \quad
    254     \mathbf{u} = \nabla \phi,
    255 \end{align}
    256 where $\phi$ needs to satisfy the Laplace equation
    257 \begin{align}
    258     \Delta \phi = 0.
    259 \end{align}
    260 According to the Theorem of Schwartz we may exchange $\frac{\partial
    261 }{\partial t}$ and $\nabla$, giving us the expression
    262 \begin{align}
    263     \nabla\left( \frac{\partial \phi}{\partial t} +\frac{1}{2}
    264     \mathbf{u}\mathbf{u} + \frac{P}{\rho}  + \Omega \right) = 0
    265 \end{align}
    266 Thus the expression differentiated by the $\nabla$ operator is an arbitrary
    267 function $f(\mathbf{x}, t)$, writing
    268 \begin{align}
    269      \frac{\partial \phi}{\partial t} +\frac{1}{2}
    270     \mathbf{u}\mathbf{u} + \frac{P}{\rho}  + \Omega = f(\mathbf{x}, t).
    271 \end{align}
    272 The function $f(\mathbf{x}, t)$ can be removed by gauge transformation of
    273 $\phi \rightarrow \phi + \int f(\mathbf{x}, t)\ dt$.
    274 \subsection{Boundary Conditions for water waves}
    275 The boundary conditions for water-wave problems vary, generally on the
    276 simplification we undertake. At the surface, called the free surface as in
    277 free from the velocity conditions, we have the atmospheric stress on the
    278 fluid. The stress component would again have a viscid component, this however
    279 is only relevant when modeling surface wind, in this review we model the
    280 fluid within reason as inviscid. The atmosphere employs
    281 only a pressure on the surface, this pressure is taken to be the atmospheric
    282 pressure, dependent on time and point in space. Thereby  any surface tension
    283 effects can also include a scenario at a curved surface (e.g. wave), giving
    284 rise to the pressure difference across the surface. A more precise
    285 description would use Thermomechanics to derive boundary conditions coupling
    286 water surface and the air above it, yet the density component of air
    287 compared to that of water makes our ansatz viable. The described conditions
    288 are called the \textbf{dynamic conditions}
    289 
    290 An additional condition revolves around the fluid particles on the moving
    291 surface, called the \textbf{kinematic condition}. This condition bounds
    292 the vertical velocity component on the surface.
    293 
    294 The logical step now is to define boundary conditions on the bed of the
    295 fluid. In the viscid case bottom is impermeable, we a no
    296 slip condition to all fluid particles $\mathbf{u}_\text{bottom}= 0$. If we
    297 assume that the fluid is inviscid then the bottom becomes a surface of the
    298 fluid in the sense that the fluid particles in contact with the bed move in
    299 the surface, we more or less mirror the kinematic condition of the surface.
    300 For many problems the condition is going to vary, in most cases the bottom
    301 will be rigid and fixed not necessarily horizontal. This condition is simply
    302 called the \textbf{bottom condition}.
    303 \subsubsection{Kinematic Condition}
    304 Obtaining the free surface is the primary objective in the theory of modeling
    305 water waves, represented by
    306 \begin{align}
    307     z = h(\mathbf{x}_\perp, t),
    308 \end{align}
    309 where $\mathbf{x}_\perp = (x, y)$ in Cartesian, or $\mathbf{x}_\perp = (r,
    310 \theta)$ in cylindrical coordinates. A surfaces that moves with the fluid,
    311 always contains the same fluid particles, described as
    312 \begin{align}
    313     \frac{D}{Dt}\left(z - h(\mathbf{x}_\perp, t ) \right) = 0.
    314 \end{align}
    315 Upon expanding the derivative we get
    316 \begin{align}
    317     \frac{Dz}{Dt} - \frac{Dh}{Dt}
    318     &= \frac{\partial z}{\partial t}+
    319     (\mathbf{u}\nabla)z - \frac{\partial h}{\partial t} -(\mathbf{u}\nabla)\\
    320     &= w - \left(h_t - (\mathbf{u}_\perp \nabla_\perp) h\right) = 0,
    321 \end{align}
    322 where the subscript $\perp$ describes the components with regard to
    323 $\mathbf{x}_\perp$. The \textbf{kinematic condition} reads
    324 \begin{align}
    325     w = h_t - (\mathbf{u}_\perp \nabla_\perp) h \qquad \text{on}\;\;
    326     z=h(\mathbf{u}_\perp, t).
    327 \end{align}
    328 
    329 \subsubsection{Dynamic Condition}
    330 As described in the prescript of this section, the case of an inviscid fluid,
    331 requires that only the pressure $P$ needs to be described on the free surface
    332 $z = h(\mathbf{x}_\perp, t)$. Assuming incompressible, irrotational,
    333 unsteady flow and setting $P=P_a$ for atmospheric pressure and $\Omega =
    334 g\cdot z$ for the force per unit mass potential the equations of motion are
    335 \begin{align}
    336     \frac{\partial \phi}{\partial t} +\frac{1}{2}\mathbf{u}\mathbf{u}
    337     + \frac{P_a}{\rho}+gh = f(t) \qquad \text{on}\;\; z=h.
    338 \end{align}
    339 Somewhere $\|\mathbf{x}_\perp\| \rightarrow \infty$ the fluid reaches
    340 equilibrium and is thereby stationary, meaning it has no motion and the
    341 pressure is $P=P_a$, the surface is a constant $h = h_0$ and therefore
    342 \begin{align}
    343     f(t) = \frac{P_a}{\rho}+gh_0.
    344 \end{align}
    345 The \textbf{dynamic condition} may be written as
    346 \begin{align}
    347     \frac{\partial \phi}{\partial t}
    348     +\frac{1}{2}\mathbf{u}\mathbf{u}+g(h-h_0) = 0 \qquad \text{on}\;\; z=h.
    349 \end{align}
    350 
    351 Regarding the pressure difference on a curved surface, we may expand the
    352 dynamic condition by introducing the pressure difference known as the
    353 \textbf{Young-Laplace Equation}
    354 \begin{align}
    355     \Delta P = \frac{\Gamma}{R},
    356 \end{align}
    357 where $\Gamma>0$ is the coefficient of surface tension and $\frac{1}{R}$ is
    358 the curvature representing an implicit function, in our case the implicit
    359 function is $z - h(\mathbf{x}_\perp, t)=0$ for fixed time. The curvature in
    360 Cartesian coordinates takes the form
    361 \begin{align}
    362     \frac{1}{R} = \frac{(1+h_y^2)h_{x x}+(1+h_y^2)h_{yy} -
    363     2h_xh_yh_{xy}}{\left( h_x^2+h_y^2+1 \right)^{\frac{3}{2}} },
    364 \end{align}
    365 the derivation is precisely described in \ref{appendix:curvature}
    366 
    367 
    368 
    369 \subsubsection{The Bottom Condition}
    370 The representation for the bottom is
    371 \begin{align}
    372    z = b(\mathbf{x}_\perp, t),
    373 \end{align}
    374 where the fluid surface needs to satisfy
    375 \begin{align}
    376     \frac{D}{Dt} \left(z - b(\mathbf{x}_\perp) \right)  = 0.
    377 \end{align}
    378 Hence we arrive at the bottom boundary conditions
    379 \begin{align}
    380     w = b_t + (\mathbf{u}_\perp \nabla_\perp)b \qquad \text{on}\;\; z=b ,
    381 \end{align}
    382 where $b(\mathbf{x}_\perp, t)$ is already known for most water wave
    383 problems. If we consider a stationary bottom then the time derivative
    384 vanishes, leaving us with the following condition
    385 \begin{align}
    386     w = (\mathbf{u}_\perp \nabla_\perp)b \qquad \text{on}\;\; z=b
    387 \end{align}
    388 
    389 
    390 \subsubsection{Integrated Mass Condition}
    391 In this section we want to combine the kinematics of both the free and the
    392 bottom surface with the mass conservation equation on the perpendicular
    393 components
    394 \begin{align}
    395     \nabla \mathbf{u} = \nabla_\perp \mathbf{u}_\perp + w_z = 0 .
    396 \end{align}
    397 Integrating the above expression from bottom to surface, i.e. from
    398 $z=b(\mathbf{x}_\perp,t)$ to $z = h (\mathbf{x},t)$ gives
    399 \begin{align}
    400     \int_b^h \nabla_\perp \mathbf{u}_\perp\ dz + w\bigg|_{z=b}^{z=h} = 0,
    401 \end{align}
    402 where we insert the conditions on the free surface and on the bottom surface
    403 \begin{align}
    404     w &= h_t + (\mathbf{u}_{\perp \text{s}} \nabla_\perp) h \quad
    405     \text{on}\;\; z = h\\
    406     w &= b_t + (\mathbf{u}_{\perp \text{b}} \nabla_\perp) h \quad
    407     \text{on}\;\; z =b,
    408 \end{align}
    409 with the subscript $s$ and $b$ indicating the evaluation of a quantity
    410 on the free surface and the bottom surface respectively. Inserting the
    411 boundary conditions we get
    412 \begin{align}
    413     \int_b^h \nabla_\perp \mathbf{u}_\perp
    414     + h_t + (\mathbf{u}_{\perp \text{s}} \nabla_\perp) h,
    415     - b_t - (\mathbf{u}_{\perp \text{b}} \nabla_\perp) b= 0.
    416 \end{align}
    417 To simplify the equation we resort to the Leibniz Rule of Integration
    418 \ref{appendix:leibniz},
    419 \begin{align}
    420      \int_b^h \nabla_\perp\mathbf{u}_\perp =
    421     \nabla_\perp \int_b^h \mathbf{u}_\perp\ dz - (\mathbf{u}_{\perp \text{s}}
    422     \nabla_\perp)h + (\mathbf{u}_{\perp \text{b}})b.
    423 \end{align}
    424 As a consequence the \textbf{Integrated Mass Condition} is given by
    425 \begin{align}
    426     \nabla_\perp \int_b^h \mathbf{u}_\perp\ dz  + \underbrace{h_t -
    427     b_t}_{=d_t} = 0.
    428 \end{align}
    429 %\subsection{Energy Equation}
    430 %To derive the energy equation we start off with Euler's Equation of Motion
    431 %\begin{align}
    432 %  \mathbf{u} _t + \nabla
    433 %  (\frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}+\Omega) = \mathbf{u}\times
    434 %  \mathbf{w},
    435 %\end{align}
    436 %multiplying the equation with $\mathbf{u}$ we get
    437 %\begin{align}
    438 %  &\mathbf{u}\mathbf{u} _t \label{eq:energy1} \\
    439 %  &+(\mathbf{u}\nabla)(\frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}+\Omega)\label{eq:energy2}\\
    440 %  &= \mathbf{u}(\mathbf{u}\times
    441 %  \mathbf{w})\label{eq:energy3}.
    442 %\end{align}
    443 %The first equation given in \ref{eq:energy1} can we rewritten using inverse
    444 %product rule of differentiation
    445 %\begin{align}
    446 %    \mathbf{u}\frac{\partial \mathbf{u}}{\partial t}
    447 %    &= \frac{\partial
    448 %    }{\partial t} (\mathbf{u}\mathbf{u}) - \frac{\partial \mathbf{u}}{\partial t}
    449 %    \mathbf{u} \\
    450 %    &= \frac{\partial
    451 %    }{\partial t} (\mathbf{u}\mathbf{u}) - \mathbf{u}\frac{\partial
    452 %    \mathbf{u}}{\partial t}\\
    453 %    \Rightarrow\quad & \mathbf{u} \frac{\partial \mathbf{u}}{\partial t}  =
    454 %    \frac{1}{2}\frac{\partial }{\partial t} (\mathbf{u}\mathbf{u}).
    455 %\end{align}
    456 %Then we may add
    457 %\begin{align}
    458 %    \left(\frac{1}{2} \mathbf{u}\mathbf{u}+\frac{P}{\rho} +\Omega  \right)
    459 %    \underbrace{(\nabla u)}_{=0} = 0,
    460 %\end{align}
    461 %to above not changing anything. Thereby getting
    462 %\begin{align}
    463 %    \frac{\partial }{\partial t} \left(\frac{1}{2}\mathbf{u}\mathbf{u}\right)
    464 %    +(\mathbf{u}\nabla \mathbf{u})\left(
    465 %    \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho} \right)
    466 %    +\left( \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho} + \Omega \right)
    467 %    (\nabla \mathbf{u}) = 0.
    468 %\end{align}
    469 %Applying the product rule we can simplify
    470 %\begin{align}
    471 %    \frac{\partial }{\partial t} \left(\frac{1}{2}\mathbf{u}\mathbf{u}\right)
    472 %    +\nabla \left(\mathbf{u}\left(\mathbf{u}(
    473 %    \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}\right) \right) = 0,
    474 %\end{align}
    475 %additionally adding $\frac{\partial \Omega}{\partial t}  =0$ leads us to
    476 %\begin{align}
    477 %    \underbrace{\frac{\partial }{\partial t} \left(\frac{1}{2}\mathbf{u}\mathbf{u}
    478 %    +\Omega\right)}_{\text{change of total energy density}}
    479 %    +\underbrace{\nabla \left(\mathbf{u}\left(\mathbf{u}(
    480 %    \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}\right)
    481 %\right)}_{\text{energy flow of the velocity field}} = 0.\label{eq:energy}
    482 %\end{align}
    483 %This is called the \textbf{energy equation} and is a general result for a
    484 %inviscid and incompressible fluids, which we can apply to study water waves.
    485 %We start off with replacing $\nabla = \nabla_\perp + \frac{\partial }{\partial
    486 %z} $ and $\Omega = g z$ and multiplying by $\rho$, then our energy equation
    487 %in \ref{eq:energy} becomes
    488 %\begin{align}
    489 %    \frac{\partial }{\partial t}\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho
    490 %    g z\right)  + \nabla_\perp\left( \mathbf{u}_\perp\left(
    491 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho gz \right)  \right)
    492 %    \frac{\partial}{\partial z} \left( w\left(
    493 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho g z \right)  \right)  = 0.
    494 %\end{align}
    495 %Integrating from bottom to top, i.e. from bed to free surface gets us to
    496 %\begin{align}
    497 %    &\int_b^h\frac{\partial }{\partial t}\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho
    498 %    g z\right)\ dz  \label{eq:e-int1}\\
    499 %    &+ \int_b^h \nabla_\perp\left( \mathbf{u}_\perp\left(
    500 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho gz \right)  \right)\
    501 %    dz\label{eq:e-int2}\\
    502 %    &+ \left(\frac{\partial}{\partial z} \left( w\left(
    503 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho g z \right)
    504 %\right)\right)\Bigg|_b^h \label{eq:e-int3}
    505 %    = 0.
    506 %\end{align}
    507 %For equation \ref{eq:e-int1} we use Leibniz Rule of Integration, leaving us
    508 %with
    509 %\begin{align}
    510 %    \int_b^h\frac{\partial }{\partial t}\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho
    511 %    g z\right)\ dz
    512 %    &= \frac{\partial }{\partial t} \int_b^h
    513 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho gz \ dz\\
    514 %    &+ \left( \frac{1}{2}\rho \mathbf{u}_s \mathbf{u}_s  + \rho g h \right)
    515 %    h_t\\
    516 %    &- \left( \frac{1}{2}\rho \mathbf{u}_b \mathbf{u}_b  + \rho g b \right)
    517 %    b_t
    518 %\end{align}
    519 %For equation \ref{eq:e-int2} we again take note of the Leibniz Rule of
    520 %Integration, getting
    521 %\begin{align}
    522 %    \int_b^h \nabla_\perp\left( \mathbf{u}_\perp\left(
    523 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho gz \right)  \right)\
    524 %    dz
    525 %    &= \nabla_\perp \int_b^h \mathbf{u}_\perp\left(
    526 %    \frac{1}{2}\rho\mathbf{u}\mathbf{u} + P + \rho g z \right) \ dz\\
    527 %    &- \left( \frac{1}{2}\rho \mathbf{u}_s\mathbf{u}_s + P + \rho g h \right)
    528 %    \left( \mathbf{u}_{\perp s} \nabla_\perp \right) h\\
    529 %    &+\left( \frac{1}{2}\rho \mathbf{u}_b\mathbf{u}_b + P + \rho g b \right)
    530 %    \left( \mathbf{u}_{\perp b} \nabla_\perp \right) b
    531 %\end{align}
    532 %Thereby transforming our equation into
    533 %\begin{align}
    534 %    \frac{\partial }{\partial t} \underbrace{\int_b^h \frac{1}{2}\rho
    535 %    \mathbf{u}\mathbf{u}+\rho g z\ dz}_{=:\mathcal{E}}
    536 %    + \nabla_\perp&\underbrace{\int_b^h
    537 %    \mathbf{u}_\perp\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho g z
    538 %\right)\ dz}_{:=\mathcal{F}}
    539 %+ \underbrace{P_s h_t - P_b b_t}_{:=\mathcal{P}} = 0\\
    540 %\nonumber\\
    541 %    &\frac{\partial \mathcal{E}}{\partial t}
    542 %    + \nabla_\perp \mathcal{F} + \mathcal{P} = 0,
    543 %\end{align}
    544 %where $\mathcal{E}$ represents the energy in the flow per unit horizontal
    545 %area, since we are integrating from bed to free surface. Where $\mathcal{F}$
    546 %is the horizontal energy flux vector and lastly $\mathcal{P} = P_s h_t -
    547 %P_b b_t$ is the net energy input due to the pressure forces doing work on the
    548 %upper and lower boundaries, i.e. bottom and free surface of the fluid.
    549 %Assuming stationary rigid bottom condition and constant surface pressure, we
    550 %can set $P_s=0$, such that $\mathcal{P} =0$ leaving us with the equation
    551 %\begin{align}
    552 %    \frac{\partial \mathcal{E}}{\partial t}
    553 %    + \nabla_\perp \mathcal{F} = 0.
    554 %\end{align}
    555 %We note that the assumption $P_s=0$ is only possible if the coefficient of
    556 %surface tension is set to 0, which usually is not the case.