notes

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

basics_fluids.tex (42231B)


      1 \include{./preamble.tex}
      2 
      3 \usepackage{amsmath}
      4 \numberwithin{equation}{section}
      5 
      6 \begin{document}
      7 
      8 \maketitle
      9 \tableofcontents
     10 
     11 \section{Governing Equations of Fluid Dynamics}
     12 We first start of with a fluid with a density
     13 \begin{align}
     14     \rho(\mathbf{x}, t),
     15 \end{align}
     16 in three dimensional Cartesian coordinates $\mathbf{x} = (x, y, z)$ at time
     17 $t$. For water-wave applications, we should note that we take
     18 $\rho=\text{constant}$, but we will go into this fact later. The fluid moves
     19 in time and space with a velocity field
     20 \begin{align}
     21     \mathbf{u}(\mathbf{x}, t) = (u, v, w).
     22 \end{align}
     23 Additionally it is also described by its pressure
     24 \begin{align}
     25     P(\mathbf{x}, t),
     26 \end{align}
     27 generally depending on time and position. When thinking of e.g. water the
     28 pressure increases the deeper we go, that is with decreasing or increasing $z$
     29 direction (depending how we set up our system $z$ pointing up or down
     30 respectively).
     31 
     32 The general assumption in fluid dynamics is the \textbf{Continuum
     33 Hypothesis}, which assumes continuity of $\textbf{u}, \rho$ and $P$ in
     34 $\mathbf{x}$ and $t$. In other words, we premise that the velocity field,
     35 density and pressure are ''nice enough`` functions of position and time, such
     36 that we can do all the differential operations we desire in the framework of
     37 differential analysis.
     38 \subsection{Mass Conservation}
     39 Our aim is to derive a model of the fluid and its dynamics, with respect to
     40 time and position, in the most general way. This is usually done thinking
     41 of the density of a given fluid, which is a unit mass per unit volume,
     42 intrinsically  an integral representation to derive these equations suggests
     43 by itself.
     44 
     45 Let us now thing of an arbitrary fluid. Within this fluid we define a fixed
     46 volume $V$ relative to a chosen inertial frame and bound it by a surface $S$
     47 within the fluid, such that the fluid motion $\mathbf{u}(\mathbf{x}, t)$ may
     48 cross the surface $S$. The fluid density is given by $\rho(\mathbf{x}, t)$,
     49 thereby the mass of the fluid in the defined Volume $V$ is an integral
     50 expression
     51 \begin{align}
     52     m = \int_V \rho(\mathbf{x}, t) dV.
     53 \end{align}
     54 The figure bellow \ref{fig:volume}, expresses the above described picture.
     55 \begin{figure}[H]
     56     \centering
     57   \begin{tikzpicture}[>=latex,scale=1, xscale=1, opacity=.8]
     58 % second sphere
     59     \begin{scope}[rotate=10, xscale=3, yscale=2, shift={(2.3,-0.2)}]
     60       \coordinate (O) at (0,0);
     61       \shade[ball color=gray!10!] (0,0) coordinate(Hp) circle (1) ;
     62 
     63       \draw[thick] (O) circle (1);
     64       \draw[rotate=5] (O) ellipse (1cm and 0.66cm);
     65       \draw[rotate=90] (O) ellipse (1cm and 0.33cm);
     66 \node[circle, fill=black, inner sep=1pt] at (0.15, 0.25) {} ; \draw[-latex, thick] (0.15, 0.25) -- (1, 1) ;
     67       \node[right] at (1, 1) {$\mathbf{u}(\mathbf{x}, t)$};
     68 
     69       \node[] at (O) {$V$};
     70       \node[] at (0.55, -0.25) {$\rho(\mathbf{x}, t)$};
     71 
     72       \draw[-] (0.76, -0.66) -- (1.2, -0.7);
     73       \node[right] at (1.2, -0.7) {$S$};
     74 
     75       \draw[-latex, thick] (-0.25, -0.65) -- (-1, -1);
     76       \node[left] at (-1, -1) {$\mathbf{n}$};
     77 
     78     \end{scope}
     79 
     80 % axis
     81   \end{tikzpicture}
     82   \caption{Volume bounded by a surface in a fluid with density and momentum,
     83   with a surface normal vector $\mathbf{n}$ \label{fig:volume}}
     84 \end{figure}
     85 
     86 Since we want to figure out the fluid's dynamics, we can consider the rate
     87 of change in the completely arbitrary $V$. The rate of change of mass needs to
     88 disappear, i.e. it is equal to zero since we cannot lose mass. Matter (mass) is
     89 neither created nor destroyed anywhere in the fluid, leading us to
     90 \begin{align}
     91     \frac{d}{dt}\left( \int_V \rho(\mathbf{x}, t)\ dV \right) = 0.
     92 \end{align}
     93 \textbf{NOT SURE HERE YET!!!!!!!!!!!, CHECK LEIBINZ FORMULA}
     94 To get more information we simply ''differentiate under the integral
     95 sign``, also known as the Leibniz Rule of Integration, see appendix
     96 \ref{appendix:leibniz}, the integral equation representing the rate of change
     97 of mass reads
     98 \begin{align}\label{eq:mass balance}
     99     \frac{dm}{dt} = \int_V \frac{\partial \rho(\mathbf{x}, t)}{\partial t}\ dV
    100     +\int_{\partial V} \rho(\mathbf{x}, t) \mathbf{u}\cdot\mathbf{n}\ dS
    101     = 0.
    102 \end{align}
    103 \textbf{----------------------}
    104 The above equation in \ref{eq:mass balance} is an underlying equation, describing that the rate of
    105 change of mass in V is brought about, only by the rate of mass flowing into
    106 V across S, and thus the mass does not change.
    107 
    108 For the second integral in \ref{eq:mass balance} we utilize the Gaussian
    109 integration law to acquire an integral over the volume
    110 \begin{align}
    111     \int_{\partial V} \rho(\mathbf{x}, t) \mathbf{u} \cdot \mathbf{n} \ dS =
    112     \int_V \nabla (\rho \mathbf{u})\ dV.
    113 \end{align}
    114 Thereby we can put everything inside the volume integral
    115 \begin{align}
    116     \frac{d m}{dt} = \int_V \left(\partial_t \rho + \nabla(\rho \mathbf{u}) \right) \ dV = 0.
    117 \end{align}
    118 Everything under the integral sign needs to be zero, thus we obtain
    119 the \textbf{Equation of Mass Conservation} or in the general sense also
    120 called the \textbf{Continuity Equation}
    121 \begin{align}\label{eq:continuity}
    122     \partial_t \rho + \nabla(\rho \mathbf{u}) = 0
    123 \end{align}
    124 
    125 In light of the results of the equation of mass conservation
    126 in \ref{eq:continuity}, an product rule gives
    127 \begin{align}
    128     \partial_t \rho + (\nabla \rho)\mathbf{u} + \rho(\nabla \mathbf{u}),
    129 \end{align}
    130 for notational purposes, we define the \textbf{material/convective derivative}
    131 as follows
    132 \begin{align}
    133     \frac{D}{Dt} = \partial_t + \mathbf{u}\nabla.
    134 \end{align}
    135 With the material derivative the equation of mass conservation reads
    136 \begin{align}
    137     \frac{D\rho}{Dt} + \rho \nabla\mathbf{u} = 0
    138 \end{align}
    139 We may undertake the first case separation, initiating $\rho = \text{cosnt.}$
    140 called \textbf{incompressible flow} causes the material derivative of $\rho$ to
    141 be zero, and thereby
    142 \begin{align}
    143     \frac{D\rho}{Dt} = 0 \quad \Rightarrow \quad \nabla \mathbf{u} = 0,
    144 \end{align}
    145 following that the divergence of the velocity field is zero, in this case
    146 $\mathbf{u}$ is called \textbf{solenoidal}.
    147 \subsection{Euler's Equation of Motion}
    148 Additional consideration we undertake is the assumption of an
    149 \textbf{inviscid} fluid, that is we set viscosity to zero. Otherwise we would
    150 get a viscous contribution under the integral which results in the
    151 Navier-Stokes equation. In this regard we apply Newton's second law to our
    152 fluid in terms of infinitesimal pieces $\delta V$ of the fluid. The
    153 acceleration divides into two terms, a \textbf{body force} given by gravity
    154 of earth in the $z$ coordinate $\mathbf{F} = (0, 0, -g)$ and a
    155 \textbf{local/short-rage force} described by the stress tensor in the fluid.
    156 In the inviscid case we the local force retains the pressure $P$, producing a
    157 normal force, with respect to the surface, acting onto any infinitesimal
    158 element in the fluid. The integral formulation of the force would be
    159 \begin{align}
    160     \int_V \rho \mathbf{F}\ dV - \int_S P\mathbf{n}\ dV.
    161 \end{align}
    162 Now applying the Gaussian rule of integration on the second integral over the
    163 surface, the resulting force in per unit volume is
    164 \begin{align}
    165     \int_V \left(\rho \mathbf{F} - \nabla P\right)\ dV.
    166 \end{align}
    167 The acceleration of the fluid particles is given by $\frac{D\mathbf{u}}{Dt}$,
    168 and thus the total force per unit volume on the other hand is
    169 \begin{align}
    170     \int_V \rho \frac{D\mathbf{u}}{Dt}\ dV =
    171     \int_V \left(\rho \mathbf{F} - \nabla P\right)\ dV.
    172 \end{align}
    173 Newton's Second Law for a fluid in an Volume is essentially saying that the
    174 rate of change of momentum of the fluid in the fixed volume $V$, which is the particle
    175 acceleration is the resulting force acting on V together with the rate of
    176 flow of momentum across the surface $S$ into the volume $V$. Hence we arrive
    177 at the \textbf{Euler's Equation(s) of Motion}
    178 \begin{align}
    179     \frac{D\mathbf{u}}{Dt} = \left(\frac{\partial \mathbf{u}}{\partial t}
    180     (\mathbf{u}\nabla)\mathbf{u}\right)  =
    181     -\frac{1}{\rho}\nabla P + \mathbf{F}.
    182 \end{align}
    183 As a side note we have mentioned that there is another contribution if the
    184 fluid is viscid. Indeed there is a tangential force due to the velocity
    185 gradient, which into introduces the additional term
    186 \begin{align}
    187     \mu \nabla^2 \mathbf{u}, \qquad
    188     \mu = \text{viscosity of the Fluid}.
    189 \end{align}
    190 Thereby the equations become
    191 \begin{align}
    192     \rho\frac{D\mathbf{u}}{Dt}
    193     =  -\nabla P + \rho \mathbf{F} + \mu \nabla^2 \mathbf{u}.
    194 \end{align}
    195 
    196 For now we have separated two simplifications, that define an
    197 \textbf{idealized/perfect fluid}
    198 \begin{enumerate}
    199     \item \textbf{incompressible} $\qquad \mu=0$
    200     \item  \textbf{inviscid} $\quad \rho = \text{const.},\ \nabla \mathbf{u}=
    201         0$
    202 \end{enumerate}
    203 \subsection{Vorticity and irrotational Flow}
    204 The curl of the velocity field $\mathbf{\omega} = \nabla \times \mathbf{u}$
    205 of a fluid (i.e. the vorticity), describes a spinning motion of the fluid
    206 near a position $\mathbf{x}$ at time $t$. The vorticity is an important
    207 property of a fluid, flows or regions of flows where $\mathbf{\omega}=0$ are
    208 \textbf{irrotational}, and thus can be modeled and analyzed following well
    209 known routine methods. Even though real flows are rarely irrotational
    210 anywhere (!), in water wave theory wave problems, from the classical aspect
    211 of vorticity have a minor contribution. Hence we can assume irrotational flow
    212 modeling water waves. To arrive at the vorticity in the equations of motions
    213 derived in the last section we resort to a differential identity derived in appendix
    214 \ref{appendix:diff identity}, which gives for the material derivative
    215 \begin{align}
    216     \frac{D\mathbf{u}}{Dt} = \frac{\partial \mathbf{u}}{\partial t}
    217     \nabla(\frac{1}{2}\mathbf{u}\mathbf{u)}
    218     - \left( \mathbf{u}\times (\nabla \times  \mathbf{u} \right).
    219 \end{align}
    220 Thus the equations of motion become
    221 \begin{align}
    222     \frac{\partial \mathbf{u}}{\partial t} + \nabla\left(
    223     \frac{1}{2}\mathbf{u}\mathbf{u} + \frac{P}{\rho} + \Omega \right)
    224     = \mathbf{u} \times  \mathbf{\omega},
    225 \end{align}
    226 where $\Omega$ is the force potential per
    227 unite mass given by $\mathbf{F} = -\nabla \Omega$.
    228 
    229 At this point we may differentiate between \textbf{stead and unsteady flow}.
    230 For \textbf{Steady Flow} we assume that $\mathbf{u}, P$ and $\Omega$ are time
    231 independent, thus we get
    232 \begin{align}
    233       \nabla\left( \frac{1}{2}\mathbf{u}\mathbf{u} + \frac{P}{\rho} + \Omega
    234       \right)  = \mathbf{u} \times  \mathbf{\omega}.
    235 \end{align}
    236 It is general knowledge that the gradient of a function $\nabla f$ is
    237 perpendicular the level sets of $f(\mathbf{x})$, where $f(\mathbf{x}) =
    238 \text{const.}$. Thus $\mathbf{u} \times  \mathbf{\omega}$ is orthogonal to
    239 the surfaces  where
    240 \begin{align} \label{eq:bernoulli}
    241     \frac{1}{2}\mathbf{u}\mathbf{u} + \frac{P}{\rho} + \Omega =
    242     \text{const.},
    243 \end{align}
    244 The above equation is called \textbf{Bernoulli's Equation}.
    245 
    246 Secondly \textbf{Unsteady Flow} but irrotational (+ incompressible), first of
    247 all gives us the condition for the existence of a velocity potential $\phi$
    248 in the sense
    249 \begin{align}
    250     \mathbf{\omega} = \nabla \times  \mathbf{u} = 0  \quad \Rightarrow \quad
    251     \mathbf{u} = \nabla \phi,
    252 \end{align}
    253 where $\phi$ needs to satisfy the Laplace equation
    254 \begin{align}
    255     \Delta \phi = 0.
    256 \end{align}
    257 According to the Theorem of Schwartz we may exchange $\frac{\partial
    258 }{\partial t}$ and $\nabla$, giving us an expression for the material
    259 derivative
    260 \begin{align}
    261     \nabla\left( \frac{\partial \phi}{\partial t} +\frac{1}{2}
    262     \mathbf{u}\mathbf{u} + \frac{P}{\rho}  + \Omega \right) = 0
    263 \end{align}
    264 Thus the expression differentiated by the $\nabla$ operator is an arbitrary
    265 function $f(\mathbf{x}, t)$, writing
    266 \begin{align}
    267      \frac{\partial \phi}{\partial t} +\frac{1}{2}
    268     \mathbf{u}\mathbf{u} + \frac{P}{\rho}  + \Omega = f(\mathbf{x}, t).
    269 \end{align}
    270 The function $f(\mathbf{x}, t)$ can be removed by gauge transformation of
    271 $\phi \rightarrow \phi + \int f(\mathbf{x}, t)\ dt$, never the less this is
    272 not further discussed and left to the reader in the reference.
    273 \subsection{Boundary Conditions for water waves}
    274 The boundary conditions for water-wave problems vary, generally on the
    275 simplification we undertake. At the surface, called the free surface as in
    276 free from the velocity conditions, we have the atmospheric stress on the
    277 fluid. The stress component would again have a viscid component, this however
    278 is only relevant when modeling surface wind, in this review we model the
    279 fluid as unaffectedly and within reason as inviscid. The atmosphere employs
    280 only a pressure on the surface, this pressure is taken to be the atmospheric
    281 pressure, dependent on time and point in space. Thereby  any surface tension
    282 effects can also include a scenario at a curved surface (e.g. wave), giving
    283 rise to the pressure difference across the surface. A more precise
    284 description would use Thermodynamics to derive boundary conditions coupling
    285 water surface and the air above it, yet the density component of air
    286 compared to that of water makes our ansatz viable. The described conditions
    287 are called the \textbf{dynamic conditions}
    288 
    289 An additional condition revolves around the fluid particles on the moving
    290 surface, called the \textbf{kinematic condition}. This condition bounds
    291 the vertical velocity component on the surface.
    292 
    293 The logical step now is to define boundary conditions on the bod of the
    294 fluid, i.e. the bottom. If the viscid case bottom is impermeable, we a no
    295 slip condition to all fluid particles $\mathbf{u}_\text{bottom}= 0$. If we
    296 assume that the fluid is inviscid then the bottom becomes a surface of the
    297 fluid in the sense that the fluid particles in contact with the bed move in
    298 the surface, we more or less mirror the kinematic condition of the surface.
    299 For many problems the condition is going to vary, in most cases the bottom
    300 will be rigid and fixed not necessarily horizontal. This condition is simply
    301 called the \textbf{bottom condition}.
    302 \subsubsection{Kinematic Condition}
    303 Obtaining the free surface is the primary objective in the theory of modeling
    304 water waves, represented by
    305 \begin{align}
    306     z = h(\mathbf{x}_\perp, t),
    307 \end{align}
    308 where $\mathbf{x}_\perp = (x, y)$ in Cartesian, or $\mathbf{x}_\perp = (r,
    309 \theta)$ in cylindrical coordinates. A surfaces that moves with the fluid,
    310 always contains the same fluid particles, described as
    311 \begin{align}
    312     \frac{D}{Dt}\left(z - h(\mathbf{x}_\perp, t  \right) = 0.
    313 \end{align}
    314 Upon expanding the derivative we get
    315 \begin{align}
    316     \frac{Dz}{Dt} - \frac{Dh}{Dt}
    317     &= \frac{\partial z}{\partial t}+
    318     (\mathbf{u}\nabla)z - \frac{\partial h}{\partial t} -(\mathbf{u}\nabla)\\
    319     &= w - \left(h_t - (\mathbf{u}_\perp \nabla_\perp) h\right) = 0,
    320 \end{align}
    321 where the subscript $\perp$ describes the components with regard to
    322 $\mathbf{x}_\perp$. The \textbf{kinematic condition} reads
    323 \begin{align}
    324     w = h_t - (\mathbf{u}_\perp \nabla_\perp) h \qquad \text{on}\;\;
    325     z=h(\mathbf{u}_\perp, t).
    326 \end{align}
    327 
    328 \subsubsection{Dynamic Condition}
    329 As described in the prescript of this section, the case of an inviscid fluid,
    330 requires that only the pressure $P$ needs to be described on the free surface
    331 $z = h(\mathbf{x}_\perp, t)$. Assuming incompressible, irrotational,
    332 unsteady flow and setting $P=P_a$ for atmospheric pressure and $\Omega =
    333 g\cdot z$ for the force per unit mass potential the equations of motion are
    334 \begin{align}
    335     \frac{\partial \phi}{\partial t} +\frac{1}{2}\mathbf{u}\mathbf{u}
    336     + P_\frac{a}{\rho}+gh = f(t) \qquad \text{on}\;\; on z=h.
    337 \end{align}
    338 Somewhere $\|\mathbf{x}_\perp\| \rightarrow \infty$ the fluid reaches
    339 equilibrium and is thereby stationary, thereby has no motion and the pressure
    340 is $P=P_a$ and the surface is a constant $h = h_0$ $f(t)$ is
    341 \begin{align}
    342     f(t) = \frac{P_a}{\rho}+gh_0.
    343 \end{align}
    344 The simplest description for the \textbf{dynamic condition} may be written as
    345 \begin{align}
    346     \frac{\partial \phi}{\partial t}
    347     +\frac{1}{2}\mathbf{u}\mathbf{u}+g(h-h_0) = 0 \qquad \text{on}\;\; z=h.
    348 \end{align}
    349 
    350 Regarding the pressure difference on a curved surface, we may expand the
    351 dynamic condition by introducing the pressure difference known as the
    352 \textbf{Young-Laplace Equation}
    353 \begin{align}
    354     \Delta P = \frac{\Gamma}{R},
    355 \end{align}
    356 where $\Gamma>0$ is the coefficient of surface tension and $\frac{1}{R}$ is
    357 the curvature representing an implicit function, in our case the implicit
    358 function is $z - h(\mathbf{x}_\perp, t)$ for fixed time. The curvature in
    359 Cartesian coordinates takes the form
    360 \begin{align}
    361     \frac{1}{R} = \frac{(1+h_y^2)h_{x x}+(1+h_y^2)h_{yy} -
    362     2h_xh_yh_{xy}}{\left( h_x^2+h_y^2+1 \right)^{\frac{3}{2}} },
    363 \end{align}
    364 the derivation is precisely described in \ref{appendix:curvature}
    365 
    366 
    367 
    368 \subsubsection{The Bottom Condition}
    369 The representation for the bottom is
    370 \begin{align}
    371    z = b(\mathbf{x}_\perp, t),
    372 \end{align}
    373 where the fluid surface needs to satisfy
    374 \begin{align}
    375     \frac{D}{Dt} \left(z - b(\mathbf{x}_\perp) \right)  = 0.
    376 \end{align}
    377 Hence we arrive at the bottom boundary conditions
    378 \begin{align}
    379     w = b_t + (\mathbf{u}_\perp \nabla_\perp)b \qquad \text{on}\;\; z=b ,
    380 \end{align}
    381 where $b(\mathbf{x}_\perp, t)$ is already known for most water wave
    382 problems. If we consider a stationary bottom then the time derivative
    383 vanishes, leaving us with the following condition
    384 \begin{align}
    385     w = (\mathbf{u}_\perp \nabla_\perp)b \qquad \text{on}\;\; z=b
    386 \end{align}
    387 
    388 
    389 \subsubsection{Integrated Mass Condition}
    390 In this section we want to combine the kinematics of both the free and the
    391 bottom surface with the mass conservation equation on the perpendicular
    392 components
    393 \begin{align}
    394     \nabla \mathbf{u} = \nabla_\perp \mathbf{u}_\perp + w_z = 0 .
    395 \end{align}
    396 Integrating the above expression from bottom to surface, i.e. from
    397 $z=b(\mathbf{x}_\perp,t)$ to $z = h (\mathbf{x},t)$ gives
    398 \begin{align}
    399     \int_b^h \nabla_\perp \mathbf{u}_\perp\ dz + w\bigg|_{z=b}^{z=h} = 0,
    400 \end{align}
    401 where we insert the conditions on the free surface and on the bottom surface
    402 \begin{align}
    403     w &= h_t + (\mathbf{u}_{\perp \text{s}} \nabla_\perp) h \quad
    404     \text{on}\;\; z = h\\
    405     w &= b_t + (\mathbf{u}_{\perp \text{b}} \nabla_\perp) h \quad
    406     \text{on}\;\; z =b,
    407 \end{align}
    408 with the subscript $s$ and $b$ indicating the evaluation of a quantity
    409 on the free surface and the bottom surface respectively. Inserting the
    410 boundary conditions we get
    411 \begin{align}
    412     \int_b^h \nabla_\perp \mathbf{u}_\perp
    413     + h_t + (\mathbf{u}_{\perp \text{s}} \nabla_\perp) h
    414     - b_t - (\mathbf{u}_{\perp \text{b}} \nabla_\perp) b= 0.
    415 \end{align}
    416 To simplify the equation we resort again to the Leibniz Rule of Integration
    417 \begin{align}
    418      \int_b^h \nabla_\perp\mathbf{u}_\perp =
    419     \nabla_\perp \int_b^h \mathbf{u}_\perp\ dz - (\mathbf{u}_{\perp \text{s}}
    420     \nabla_\perp)h - (\mathbf{u}_{\perp \text{b}})b.
    421 \end{align}
    422 As a consequence the \textbf{Integrated Mass Condition} is given by
    423 \begin{align}
    424     \nabla_\perp \int_b^h \mathbf{u}_\perp\ dz  + \underbrace{h_t -
    425     b_t}_{=d_t} = 0.
    426 \end{align}
    427 \subsection{Energy Equation}
    428 To derive the energy equation we start off with Euler's Equation of Motion
    429 \begin{align}
    430   \mathbf{u} _t + \nabla
    431   (\frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}+\Omega) = \mathbf{u}\times
    432   \mathbf{w},
    433 \end{align}
    434 multiplying the equation with $\mathbf{u}$ we get
    435 \begin{align}
    436   &\mathbf{u}\mathbf{u} _t \label{eq:energy1} \\
    437   &+(\mathbf{u}\nabla)(\frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}+\Omega)\label{eq:energy2}\\
    438   &= \mathbf{u}(\mathbf{u}\times
    439   \mathbf{w})\label{eq:energy3}.
    440 \end{align}
    441 The first equation given in \ref{eq:energy1} can we rewritten using inverse
    442 product rule of differentiation
    443 \begin{align}
    444     \mathbf{u}\frac{\partial \mathbf{u}}{\partial t}
    445     &= \frac{\partial
    446     }{\partial t} (\mathbf{u}\mathbf{u}) - \frac{\partial \mathbf{u}}{\partial t}
    447     \mathbf{u} \\
    448     &= \frac{\partial
    449     }{\partial t} (\mathbf{u}\mathbf{u}) - \mathbf{u}\frac{\partial
    450     \mathbf{u}}{\partial t}\\
    451     \Rightarrow\quad & \mathbf{u} \frac{\partial \mathbf{u}}{\partial t}  =
    452     \frac{1}{2}\frac{\partial }{\partial t} (\mathbf{u}\mathbf{u}).
    453 \end{align}
    454 Then we may add
    455 \begin{align}
    456     \left(\frac{1}{2} \mathbf{u}\mathbf{u}+\frac{P}{\rho} +\Omega  \right)
    457     \underbrace{(\nabla u)}_{=0} = 0,
    458 \end{align}
    459 to above not changing anything. Thereby getting
    460 \begin{align}
    461     \frac{\partial }{\partial t} (\frac{1}{2}\mathbf{u}\mathbf{u})
    462     +(\mathbf{u}\nabla \mathbf{u})\left(
    463     \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho} \right)
    464     +\left( \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho} + \Omega \right)
    465     (\nabla \mathbf{u}) = 0.
    466 \end{align}
    467 Applying the product rule we can simplify
    468 \begin{align}
    469     \frac{\partial }{\partial t} \left(\frac{1}{2}\mathbf{u}\mathbf{u}\right)
    470     +\nabla \left(\mathbf{u}\left(\mathbf{u}(
    471     \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}\right) \right) = 0,
    472 \end{align}
    473 additionally adding $\frac{\partial \Omega}{\partial t}  =0$ leads us to
    474 \begin{align}
    475     \underbrace{\frac{\partial }{\partial t} \left(\frac{1}{2}\mathbf{u}\mathbf{u}
    476     +\Omega\right)}_{\text{change of total energy density}}
    477     +\underbrace{\nabla \left(\mathbf{u}\left(\mathbf{u}(
    478     \frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}\right)
    479 \right)}_{\text{energy flow of the velocity field}} = 0.\label{eq:energy}
    480 \end{align}
    481 This is called the \textbf{energy equation} and is a general result for a
    482 inviscid and incompressible fluids, which we can apply to study water waves.
    483 We start off with replacing $\nabla = \nabla_\perp + \frac{\partial }{\partial
    484 z} $ and $\Omega = g z$ and multiplying by $\rho$, then our energy equation
    485 in \ref{eq:energy} becomes
    486 \begin{align}
    487     \frac{\partial }{\partial t}\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho
    488     g z\right)  + \nabla_\perp\left( \mathbf{u}_\perp\left(
    489     \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho gz \right)  \right)
    490     \frac{\partial}{\partial z} \left( w\left(
    491     \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho g z \right)  \right)  = 0.
    492 \end{align}
    493 Integrating from bottom to top, i.e. from bed to free surface gets us to
    494 \begin{align}
    495     &\int_b^h\frac{\partial }{\partial t}\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho
    496     g z\right)\ dz  \label{eq:e-int1}\\
    497     &+ \int_b^h \nabla_\perp\left( \mathbf{u}_\perp\left(
    498     \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho gz \right)  \right)\
    499     dz\label{eq:e-int2}\\
    500     &+ \left(\frac{\partial}{\partial z} \left( w\left(
    501     \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho g z \right)
    502 \right)\right)\Bigg|_b^h \label{eq:e-int3}
    503     = 0.
    504 \end{align}
    505 For equation \ref{eq:e-int1} we use Leibniz Rule of Integration, leaving us
    506 with
    507 \begin{align}
    508     \int_b^h\frac{\partial }{\partial t}\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho
    509     g z\right)\ dz
    510     &= \frac{\partial }{\partial t} \int_b^h
    511     \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho gz \ dz\\
    512     &+ \left( \frac{1}{2}\rho \mathbf{u}_s \mathbf{u}_s  + \rho g h \right)
    513     h_t\\
    514     &- \left( \frac{1}{2}\rho \mathbf{u}_b \mathbf{u}_b  + \rho g b \right)
    515     b_t
    516 \end{align}
    517 For equation \ref{eq:e-int2} we again take note of the Leibniz Rule of
    518 Integration, getting
    519 \begin{align}
    520     \int_b^h \nabla_\perp\left( \mathbf{u}_\perp\left(
    521     \frac{1}{2}\rho\mathbf{u}\mathbf{u}+P+\rho gz \right)  \right)\
    522     dz
    523     &= \nabla_\perp \int_b^h \mathbf{u}_\perp\left(
    524     \frac{1}{2}\rho\mathbf{u}\mathbf{u} + P + \rho g z \right) \ dz\\
    525     &- \left( \frac{1}{2}\rho \mathbf{u}_s\mathbf{u}_s + P + \rho g h \right)
    526     \left( \mathbf{u}_{\perp s} \nabla_\perp \right) h\\
    527     &+\left( \frac{1}{2}\rho \mathbf{u}_b\mathbf{u}_b + P + \rho g b \right)
    528     \left( \mathbf{u}_{\perp b} \nabla_\perp \right) b
    529 \end{align}
    530 Thereby transforming our equation into
    531 \begin{align}
    532     \frac{\partial }{\partial t} \underbrace{\int_b^h \frac{1}{2}\rho
    533     \mathbf{u}\mathbf{u}+\rho g z\ dz}_{=:\mathcal{E}}
    534     + \nabla_\perp&\underbrace{\int_b^h
    535     \mathbf{u}_\perp\left( \frac{1}{2}\rho\mathbf{u}\mathbf{u} + \rho g z
    536 \right)\ dz}_{:=\mathcal{F}}
    537 + \underbrace{P_s h_t - P_b b_t}_{:=\mathcal{P}} = 0\\
    538 \nonumber\\
    539     &\frac{\partial \mathcal{E}}{\partial t}
    540     + \nabla_\perp \mathcal{F} + \mathcal{P} = 0,
    541 \end{align}
    542 where $\mathcal{E}$ represents the energy in the flow per unit horizontal
    543 area, since we are integrating from bed to free surface. Where $\mathcal{F}$
    544 is the horizontal energy flux vector and lastly $\mathcal{P} = P_s h_t -
    545 P_b b_t$ is the net energy input due to the pressure forces doing work on the
    546 upper and lower boundaries, i.e. bottom and free surface of the fluid.
    547 Assuming stationary rigid bottom condition and constant surface pressure, we
    548 can set $P_s=0$, such that $\mathcal{P} =0$ leaving us with the equation
    549 \begin{align}
    550     \frac{\partial \mathcal{E}}{\partial t}
    551     + \nabla_\perp \mathcal{F} = 0.
    552 \end{align}
    553 We note that the assumption $P_s=0$ is only possible if the coefficient of
    554 surface tension is set to 0, which usually is not the case.
    555 \section{Dimensional Analysis}
    556 Our derived model of fluid dynamics yields formal connections between
    557 physical quantities. These quantities bear units, e.g. the velocity of fluid
    558 particles $\mathbf{u}$ has the ``SI'' unites of $\frac{m}{s}$, meters per
    559 second. The idea is the make use of these scales and formulate a model, where
    560 the quantities are nondimensionalized, i.e. to get rid of physical units by
    561 scaling each quantity appropriately. The appropriate length scales are that
    562 of the typical water depth $h_0$ and the typical wavelength $\lambda$ of a
    563 surface wave.
    564 
    565 \subsection{Nondimensionalisation}
    566 In summary we use these adaptations
    567 \begin{itemize}
    568     \item $h_0$ for the typical water depth
    569     \item $\lambda$ for the typical wavelength
    570     \item $\frac{\lambda}{\sqrt{g h_0}}$ time scale of wave propagation
    571     \item $\sqrt{g h_0}$ velocity scale of waves in $(x, y)$
    572     \item $\frac{h_0 \sqrt{g h_0} }{\lambda}$ velocity scale in the $z$
    573         direction.
    574 \end{itemize}
    575 $(x, z, t)$, then
    576 \begin{align}
    577     u = \psi _z, \qquad w = - \psi_x;
    578 \end{align}
    579 and the scale of $\psi$ must be $h_0\sqrt{g h_0}$. Additionally we write the
    580 boundary condition on the free surface as follows
    581 \begin{align}
    582     h  = h_0 + a \eta (\mathbf{x}_\perp, t) = z,
    583 \end{align}
    584 where $a$ is the typical amplitude and $\eta$ nondimensional function. All in
    585 all we have the following scaling for the physical quantities of our context
    586 \begin{align}
    587     &x \rightarrow\ \lambda x, \quad u \rightarrow \sqrt{gh_0} u, \\
    588       &y \rightarrow\ \lambda y, \quad v \rightarrow \sqrt{gh_0} v, \qquad
    589       t\rightarrow \frac{\lambda}{\sqrt{gh_0}}t,\\
    590       &z \rightarrow\ h_0 z, \quad w \rightarrow
    591     \frac{h_0\sqrt{gh_0}}{\lambda} w.
    592 \end{align}
    593 with
    594 \begin{align}
    595     h = h_0 + a \eta, \qquad  b \rightarrow h_0 b.
    596 \end{align}
    597 The pressure is also rewritten into
    598 \begin{align}
    599     P = P_a + \rho g(h_0 -z) + \rho g h_0 p,
    600 \end{align}
    601 where $P_a$ is the atmospheric pressure, the term $h_0-z$ represent the
    602 hydrostatic pressure distribution, i.e. pressure at depth and the term with the pressure
    603 variable $p$  measures the deviation from the hydrostatic pressure
    604 distribution. Indeed $p\neq 0 $ for wave propagation. Now we can perform a
    605 rescaling of the Euler's Equation of Motion, we introduce the notation
    606 \begin{align}
    607     &t = \frac{\lambda}{\sqrt{gh_0}}\tau,\quad x = \lambda \xi,\quad u =
    608     \sqrt{gh_0} \tilde{u}\\
    609     &y = \lambda \chi,\quad v = \sqrt{gh_0} \tilde{v}\\
    610     &z = h_0 \zeta, \quad w = \frac{h_0\sqrt{gh_0} }{\lambda}\tilde{w}.
    611 \end{align}
    612 We start off with the $x$ coordinate, substitute and apply the chain rule
    613 leading us to
    614 \begin{align}
    615     \frac{Du}{Dt}
    616     &= \frac{\partial u}{\partial t} +u \frac{\partial
    617     u}{\partial x} \\
    618     &= \sqrt{gh_{0}}\frac{\partial \tilde{u}}{\partial \tau} \frac{\partial
    619     \tau}{\partial t} +gh_0 \tilde{u} \frac{\partial \tilde{u}}{\partial \xi}
    620     \frac{\partial \xi}{\partial x} \\
    621     &= \frac{gh_0}{\lambda} \left( \frac{\partial \tilde{u}}{\partial \tau}
    622     \tilde{u} \frac{\partial \tilde{u}}{\partial \xi} \right),
    623 \end{align}
    624 on the other hand
    625 \begin{align}
    626     \frac{gh_0}{\lambda} \left( \frac{\partial \tilde{u}}{\partial \tau}
    627     \tilde{u} \frac{\partial \tilde{u}}{\partial \xi} \right)
    628     &=-\frac{1}{\rho}\frac{1}{\lambda}\frac{\partial P}{\partial x} \\
    629     &=-\frac{ g h_0 }{\lambda}\rho \frac{\partial p}{\partial \xi}.
    630 \end{align}
    631 Thereby the rescaling evolves to
    632 \begin{align}
    633     \frac{D \tilde{u}}{D\tau} = -\frac{\partial p}{\partial \xi}.
    634 \end{align}
    635 Because of the same scaling in $y$ we get the same result as in $x$, that is
    636 \begin{align}
    637     \frac{D \tilde{v}}{D\tau} = -\frac{\partial p}{\partial \chi}.
    638 \end{align}
    639 In the $z$ coordinate we have
    640 \begin{align}
    641     \frac{Dw}{Dt}
    642     &= \frac{\partial w}{\partial t} +w \frac{\partial
    643     w}{\partial \zeta} \\
    644     &= \frac{h_0\sqrt{gh_0}}{\lambda} \frac{\sqrt{gh_0}}{\lambda}
    645     \frac{\partial \tilde{w}}{\partial \tau}  + \frac{1}{h_0}
    646     \frac{h_0\sqrt{gh_0} }{\lambda} \frac{h_0\sqrt{gh_0}}{\lambda}
    647     \tilde{w}\frac{\partial \tilde{v}}{\partial \zeta}\\
    648     &= \frac{h_0^2g}{\lambda}\left( \frac{\partial \tilde{w}}{\partial \tau}
    649     + \tilde{w}\frac{\partial \tilde{w}}{\partial \zeta} \right) .
    650 \end{align}
    651 On the other side we have
    652 \begin{align}
    653     \frac{h_0^2g}{\lambda}\left( \frac{\partial \tilde{w}}{\partial \tau}
    654     + \tilde{w}\frac{\partial \tilde{w}}{\partial \zeta} \right)
    655     &=
    656     -\frac{1}{h_0\rho} \frac{\partial P}{\partial z} +g \\
    657     &=-\frac{1}{h_0\rho}(-\rho gh_0 \frac{\partial \zeta}{\partial \zeta}
    658     \rho gh_0
    659     \frac{\partial p}{\partial \zeta} ) + g  \\
    660     &= -g \frac{\partial p}{\partial z}.
    661 \end{align}
    662 In total for the $z$ direction we get
    663 \begin{align}
    664    \underbrace{\left( \frac{h_0}{\lambda} \right)^2}_{=: \delta^2}
    665     \frac{Dw}{Dt} = -\frac{\partial p}{\partial z},
    666 \end{align}
    667 where $\delta$ is the \textbf{long wavelength} or \textbf{shallowness}
    668 parameter, a very important constant for developing model hierarchies. For
    669 clarity we resubstitute for $x, y, z, t, u, v$ and $w$, and for completeness
    670 the we display the equations again, which are
    671 \begin{align}\label{eq:nondim-motion}
    672     \frac{Du}{Dt} = - \frac{\partial p}{\partial x}&, \quad
    673     \frac{Dv}{Dt} = - \frac{\partial p}{\partial y}, \quad
    674     \delta^2\frac{Dw}{Dt} = - \frac{\partial p}{\partial z}, \\
    675     &\frac{\partial u}{\partial x} + \frac{\partial v}{\partial y}
    676     +\frac{\partial w}{\partial z}  = 0.
    677 \end{align}
    678 We can now turn our attention to the boundary conditions, on both free
    679 surface $z=h$ and the bottom $z=b$ we have $z \Rightarrow h_0 z$ and thereby
    680 \begin{align}
    681     z = 1+
    682     \underbrace{\frac{a}{h_0}}_{:=\varepsilon}\eta(\mathbf{x}_\perp,t) \quad
    683     \text{and}\quad z= b,
    684 \end{align}
    685 where we arrive at our second very important parameter $\varepsilon$ called
    686 the \textbf{amplitude} parameter. As for the kinematic condition, we
    687 substitute the free surface $z=h = 1+\varepsilon \eta$ and get
    688 \begin{align}
    689     \frac{Dz}{Dt} = \varepsilon\left(\eta_t + (\mathbf{u}_\perp
    690         \nabla_\perp)\eta\right) \qquad \text{on}\;\; z= 1+\varepsilon \eta.
    691 \end{align}
    692 Respectively the bottom condition is not changed
    693 \begin{align}
    694     w = b_t + (\mathbf{u}_\perp \nabla_\perp) b \quad \text{on}\;\; z= b.
    695 \end{align}
    696 The general dynamic condition for $h = h(x, y, t)$ yields a rescaling of the
    697 curvature in terms of
    698 \begin{align}
    699    \frac{1}{R}
    700    &= \frac{(1+h_y^2)h_{x x} + (1+h_x^2)h_yy - 2h_xh_yh_{xy}
    701    }{\left(h_x^2+h_y^2 +1  \right)^{\frac{3}{2}} } \\
    702    &= -\frac{\varepsilon h_0}{\lambda^2} \frac{(
    703    1+\varepsilon^2\delta^2\eta_y^2 )\eta_{x x}+
    704     (1+\varepsilon^2\delta^2\eta_x^2)\eta_{yy} -
    705     2\varepsilon^2\delta^2\eta_x\eta_y\eta_{xy}}{\left(
    706     1+\varepsilon^2\delta^2\eta_x^2+\varepsilon^2\delta^2\eta_y^2
    707     \right)^{\frac{3}{2}} },
    708 \end{align}
    709 together with the pressure difference
    710 \begin{align}
    711     \Delta P = \rho g h_0(p - \varepsilon \eta) = \frac{\Gamma}{R},
    712 \end{align}
    713 leaving us ultimately with the dynamic condition
    714 \begin{align}
    715     p-\varepsilon\eta= \varepsilon\left( \frac{\Gamma}{\rho g\lambda^2}
    716     \right) \left(\frac{\lambda^2}{\varepsilon h_0}\frac{1}{R}\right),
    717 \end{align}
    718 where $W_e = \frac{\Gamma}{\rho g h_0^2}$ is the \textbf{Weber number}. This
    719 dimensionless parameter can be considered as a measure of the fluid's inertia
    720 compered to its surface tension, which satisfies the relation
    721 \begin{align}
    722     \delta^2 W_e = \frac{\Gamma}{\rho g \lambda^2}.
    723 \end{align}
    724 \subsection{Scaling of Variables}
    725 Admits a simple observation of the governing equations in the last chapter we
    726 notice that $w$ and $p$ on the free surface $z = 1 + \varepsilon\eta$ are
    727 directly proportional to $\varepsilon$. Hence we want to ''scale this way``
    728 by introducing the following transformation
    729 \begin{align}
    730     p \rightarrow \varepsilon p, \quad w \rightarrow \varepsilon w, \quad
    731     \mathbf{u}_\perp \rightarrow \varepsilon \mathbf{u}_\perp.
    732 \end{align}
    733 Because of this scaling our material derivative changes slightly to
    734 \begin{align}\label{eq:mod-material}
    735     \frac{D}{Dt} = \frac{\partial }{\partial t} + \varepsilon\left(u
    736     \frac{\partial }{\partial x}  + v \frac{\partial }{\partial y}  + w
    737     \frac{\partial }{\partial z} \right)
    738 \end{align}
    739 A simple recalculation yields the rescaled, nondimensionalized Euler's
    740 Equation of motion are the same as in equations \ref{eq:nondim-motion} with
    741 the modified material derivative from \ref{eq:mod-material}, and the boundary
    742 conditions are
    743 \begin{align}
    744     p &= \eta - \frac{\delta^2\varepsilon h_0}{\lambda^2} \frac{W_e}{R}\\
    745     w &= \frac{1}{\varepsilon}\eta_t + (\mathbf{u}_\perp \nabla_\perp)\eta
    746     \quad \text{on}\;\; z = 1+\varepsilon\eta\\
    747     w &=\frac{1}{\varepsilon}b_t + (\mathbf{u}_\perp \nabla_\perp)b \quad
    748     \text{on}\;\; z=b
    749 \end{align}
    750 \subsection{Model Hierarchies}
    751 As we have derived a model of fluid dynamics, with small parameters
    752 $\varepsilon$ and $\delta$, we can conduct a series of classifications and
    753 perform asymptotic analysis on them. The main hierarchies important in this
    754 review are derived from the following problem classifications
    755 \begin{itemize}
    756     \item $\varepsilon\rightarrow 0$: linearized problem, small amplitude
    757     \item $\delta\rightarrow 0$: shallow Water, long-wave
    758     \item$\delta \rightarrow 0;\; \varepsilon~1$: shallow Water, large
    759         amplitude
    760     \item $\delta\ll 1;\; \varepsilon~\delta$: shallow water, medium
    761         amplitude
    762     \item $\delta\ll 1;\; \varepsilon~\delta^2$: shallow water, small
    763         amplitude
    764     \item $\delta \gg 1;\; \varepsilon\delta\ll 1$: deep water, small
    765         steepness.
    766 \end{itemize}
    767 
    768 \section{The Solitary Wave and The KdV Equation}
    769 The solitary wave is a wave of translation, it is stable and can travel long
    770 distances additionally the speed depends on the size of the wave. An
    771 interesting feature is that two solitary waves do not merge together to form
    772 one solitary wave, rather the small wave is overtaken by a larger one. If a
    773 solitary wave is too big for the depth it splits into two, a big and a small
    774 one. Solitary waves arise in the region $\varepsilon=O(\delta^2)$.
    775 
    776 
    777 \subsection{Solitary Wave}
    778 To describe
    779 a solitary wave we begin with Euler's Equation of Motion, where we assume
    780 there is no surface tension we set $W_e = 0$ and additionally assume
    781 irrotational flow $\mathbf{\omega}=\nabla \times  \mathbf{u} = 0$. This means
    782 that there exists a velocity potential $\phi(\mathbf{x},t)$ given
    783 by$\mathbf{u} = \nabla \phi$ satisfying the Laplace equation. In regard of a
    784 solitary wave being a plane wave, we rotate our coordinate system such that
    785 the propagation is in the $x$-direction and a stationary \& fixed bottom
    786 $b=0$. Ultimately leaving us with the following model
    787 \begin{align}\label{eq:soliton}
    788 \begin{drcases}
    789    & \phi_{zz} + \delta \phi_{x x }  = 0,\\
    790    &\text{with the boundary conditions}\\
    791    &\begin{drcases}
    792     &\phi_z = \delta^2 (\eta_t + \varepsilon \phi_x \eta_x) \\
    793     &\phi_t + \eta +  \frac{1}{2}\varepsilon\left( \frac{1}{\delta^2}\phi^2_z
    794     + \phi_x^2\right)  =0
    795   \end{drcases}\quad \text{on}\;\; z = 1+\varepsilon\eta,\\
    796    &\text{and}\\
    797    & \phi_z =0 \quad \text{on}\;\; z = b = 0.
    798 \end{drcases}
    799 \end{align}
    800 Since the model arises $\varepsilon = O(\delta^2)$, for convince we set
    801 $\varepsilon=1$. The fact of the matter is we are seeking a traveling wave
    802 solution, thereby we can go into the coordinate system of the traveling wave,
    803 one in the variable $\xi = x - ct$ for a from left to right traveling wave,
    804 where $c$ is the nondimensional speed of the wave. Our goal is to find the
    805 solution for the velocity potential $\phi(\xi, z)$ and the wave profile
    806 $\eta(\xi)$. The chain rule gives us
    807 \begin{align}
    808     \frac{\partial }{\partial x} &= \frac{\partial \xi}{\partial x}
    809     \frac{\partial }{\partial \xi}  = \frac{\partial }{\partial \xi}, \\
    810     \frac{\partial }{\partial t} &= \frac{\partial \xi}{\partial t}
    811     \frac{\partial }{\partial \xi}  = -c\frac{\partial }{\partial \xi}.
    812 \end{align}
    813 Together with the equations in \ref{eq:soliton} we obtain
    814 \begin{align}\label{eq:soliton-xi}
    815     \begin{drcases}
    816    & \phi_{zz} + \delta \phi_{\xi\xi}  = 0,\\
    817    &\text{with the boundary conditions}\\
    818    &\begin{drcases}
    819     &\phi_z = \delta^2 (\phi_\xi -c)\eta_\xi \\
    820     &-c\phi_\xi + \eta +  \frac{1}{2}\varepsilon\left( \frac{1}{\delta^2}\phi^2_z
    821     + \phi_\xi^2\right)  =0
    822   \end{drcases}\quad \text{on}\;\; z = 1+\eta,\\
    823    &\text{and}\\
    824    & \phi_z =0 \quad \text{on}\;\; z = b = 0.
    825     \end{drcases}
    826 \end{align}
    827 \subsubsection{Exponential Decay}
    828 We would like to analyze if the equation in \ref{eq:soliton-xi} gives viable a
    829 solution that decays exponentially, we make the ansatz
    830 \begin{align}
    831     \eta \simeq a e^{-\alpha |\psi|},\quad \phi \simeq \psi(z)e^{-\alpha
    832     |\xi|}, \qquad  \mid \xi \mid \rightarrow \infty,
    833 \end{align}
    834 where $\alpha>0$ is the exponent. The equations in \ref{eq:soliton-xi}
    835 transforms to
    836 \begin{align}
    837     \psi'' + \alpha^2 \delta^2\psi = 0.
    838 \end{align}
    839 The above equation is a standard well known ordinary differential equation
    840 reading
    841 \begin{align}
    842     \psi = A \cos(\alpha\delta z),
    843 \end{align}
    844 where $A$ is the integration constant. On the free surface $z\simeq 1$ gives
    845 \begin{align}
    846     &-cA\alpha\sin(\alpha\delta) = ca\alpha,\label{eq:sol1}\\
    847     &cA\alpha \cos(\alpha\delta) = -a \label{eq:sol2}.
    848 \end{align}
    849 Dividing equation \ref{eq:sol1} with equation \ref{eq:sol2} gives
    850 \begin{align} \label{eq:soliton-dispersion}
    851     c^2 = \frac{\tan\left(\alpha\delta  \right) }{\alpha\delta}.
    852 \end{align}
    853 We conclude that the solution for such a wave exists provided that the
    854 dispersion relation on the wave propagation speed holds, thereby all solitary
    855 waves exhibit exponential decay in their tail and satisfy the dispersion
    856 relation in equation \ref{eq:soliton-dispersion}.
    857 \subsubsection{Asymptotic Analysis}
    858 The underlining equations in \ref{eq:soliton} extend from $-\infty$ to
    859 $\infty$, so the length scale is much greater than any finite depth of
    860 water. Therefore the classification $\delta \rightarrow 0$ is appropriate for
    861 a solitary wave, this however goes with the assumption
    862 $\varepsilon\rightarrow 0$ otherwise we cannot make an appropriate expansion.
    863 Let us look at the main equation
    864 \begin{align}\label{eq:sol-laplace}
    865     \phi_{zz} + \delta \phi_{x x} = 0.
    866 \end{align}
    867 For small $\delta$ we conduct the $\delta^2 = O(\varepsilon)$ standard ansatz
    868 in asymptotic analysis
    869 \begin{align}
    870     \phi_{\delta}(x, t, z) \simeq \sum_{n=0}^{\infty} \delta^{2n}\phi_n(x, t,
    871     z).
    872 \end{align}
    873 Substituting $\phi_\delta$ into equation \ref{eq:sol-laplace} we get
    874 \begin{align}
    875     \delta^{2\cdot 0}\left( \phi_{0zz} \right)  + \delta^{2\cdot 1}\left(
    876     \phi_{1zz}+\phi_{0 x x} \right)  + \delta^{2\cdot 2}\left( \phi_{2zz}+
    877     \phi_{1 x x} \right)  + O(\delta^{2\cdot 3}) = 0.
    878 \end{align}
    879 We start off with $O(\delta^{2\cdot0}) $, which gives us an arbitrary function
    880 $\phi_{0} = \theta(x, t)$. Next we may generalize the results for all
    881 $O(\delta^{2\cdot n})$ in the means of
    882 \begin{align}
    883     \phi_{n+1zz}  = -\phi_{nx x}\qquad \forall n\in \mathbb{N} .
    884 \end{align}
    885 Therefore leaving us for $\phi_1$ and $\phi_2$ with
    886 \begin{align}
    887     &\phi_1 = -\frac{1}{2} z^2 \theta_0(x,t) + \theta(x, t),\\
    888     \Rightarrow& \phi_2 =
    889     \frac{1}{24}z^4\theta_0(x,t)-\frac{1}{2}z^2\theta_1(x,t) + \theta_2(x,t).
    890 \end{align}
    891 The boundary condition on the bottom comes around to be
    892 \begin{align}
    893     \phi_{nz} =0 \quad \text{on}\;\; z=0.
    894 \end{align}
    895 The free surface boundary condition $z= 1+\varepsilon\eta$ n evolves more calculation, we consider
    896 only terms up the order of $\delta^2$, initializing with
    897 \begin{align}
    898     &\phi_z = \delta^2(\eta_t + \varepsilon\phi_x \eta_x)\\
    899     \Leftrightarrow &\frac{1}{\delta}\phi_z = \eta_t + \varepsilon\phi_x
    900     \eta_x,
    901 \end{align}
    902 substituting $\phi_\delta$ into the above proceeds to be
    903 \begin{align}
    904     \frac{1}{\delta^2}\underbrace{\phi_{0z}}_{=0} + \phi_{1z}+ \delta^2\phi_{zz}
    905     O(\delta^{2\cdot 2})
    906     &= -z\theta_{x x} + \delta^2\left( \frac{1}{6}z^3\theta_{0 x x x x} - z
    907     \theta_{0x x} \right) + O(\delta^{2\cdot 2})\\
    908     &=-(1+\varepsilon\eta)\theta_{0 x x} + \delta^2\left(
    909     \frac{1}{6}(1+\varepsilon\eta)^3\theta_{0 x x} -
    910 (1+\varepsilon\eta)\theta_{0 x x} \right) +O(\delta^{2\cdot 2})\\
    911     &= \eta_t + \varepsilon\eta_x \left( \theta_{0x}
    912     \delta^2(\theta_{1x}-\frac{1}{2}( 1+ \varepsilon\eta)^2 \theta_{0x x x}
    913 \right).
    914 \end{align}
    915 The second condition is
    916 \begin{align}
    917     \phi_t + \eta + \frac{1}{2}\varepsilon \left( \frac{1}{\delta}\phi^2_z
    918     \phi_x^2\right)  = 0,
    919 \end{align}
    920 becomes after substitution
    921 \begin{align}
    922     \theta_{0t}+ \delta^2\left( -\frac{1}{2}(1+\varepsilon\eta)^2\theta_{0 x xt}
    923     + \theta_{1t}\right) + \eta + O(\delta^{2\cdot 2})
    924     &=-\frac{1}{2}\delta^2\varepsilon(-(1+\varepsilon\eta)\theta_{0 x x
    925     })^2\\
    926     &-\frac{1}{2}\left( \theta_{0 x} + \delta^2\left( \theta_{1x} -
    927     \frac{1}{2}(1+\varepsilon\eta)^2\theta_{0 x x x x}  \right)  \right) ^2
    928 \end{align}
    929 The simplest case is $\varepsilon,\delta \rightarrow 0$.
    930 
    931 
    932 
    933 
    934 
    935 
    936 
    937 
    938 
    939 
    940 
    941 
    942 \newpage
    943 \include{appendix.tex}
    944 
    945 
    946 
    947 \nocite{johnson_1997}
    948 \nocite{vallis_2017}
    949 \nocite{constantin_tsunami}
    950 \nocite{rupert_2009}
    951 \nocite{mathe-physik}
    952 
    953 \printbibliography
    954 
    955 \end{document}