notes

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

commit 8ce911ece70e6d388d9cd789443ae126f7eddedd
parent c77d1533d1c61411c2bf6ca72e6c8636357994fa
Author: miksa <milutin@popovic.xyz>
Date:   Sat,  4 Jun 2022 19:10:50 +0200

new chapters

Diffstat:
Mapp_pde/basics_fluids.tex | 390++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++-
Mapp_pde/build/basics_fluids.pdf | 0
2 files changed, 387 insertions(+), 3 deletions(-)

diff --git a/app_pde/basics_fluids.tex b/app_pde/basics_fluids.tex @@ -90,6 +90,7 @@ neither created nor destroyed anywhere in the fluid, leading us to \begin{align} \frac{d}{dt}\left( \int_V \rho(\mathbf{x}, t)\ dV \right) = 0. \end{align} +\textbf{NOT SURE HERE YET!!!!!!!!!!!, CHECK LEIBINZ FORMULA} To get more information we simply ''differentiate under the integral sign``, also known as the Leibniz Rule of Integration, see appendix \ref{appendix:leibniz}, the integral equation representing the rate of change @@ -99,6 +100,7 @@ of mass reads +\int_{\partial V} \rho(\mathbf{x}, t) \mathbf{u}\cdot\mathbf{n}\ dS = 0. \end{align} +\textbf{----------------------} The above equation in \ref{eq:mass balance} is an underlying equation, describing that the rate of change of mass in V is brought about, only by the rate of mass flowing into V across S, and thus the mass does not change. @@ -394,7 +396,7 @@ components Integrating the above expression from bottom to surface, i.e. from $z=b(\mathbf{x}_\perp,t)$ to $z = h (\mathbf{x},t)$ gives \begin{align} - \int_b^h \nabla_\perp \mathbf{u}_\perp\ dz w\bigg|_{z=b}^{z=h} = 0, + \int_b^h \nabla_\perp \mathbf{u}_\perp\ dz + w\bigg|_{z=b}^{z=h} = 0, \end{align} where we insert the conditions on the free surface and on the bottom surface \begin{align} @@ -423,7 +425,7 @@ As a consequence the \textbf{Integrated Mass Condition} is given by b_t}_{=d_t} = 0. \end{align} \subsection{Energy Equation} -To derive the energy equation we start of with Euler's Equation of Motion +To derive the energy equation we start off with Euler's Equation of Motion \begin{align} \mathbf{u} _t + \nabla (\frac{1}{2}\mathbf{u}\mathbf{u}+\frac{P}{\rho}+\Omega) = \mathbf{u}\times @@ -478,7 +480,7 @@ additionally adding $\frac{\partial \Omega}{\partial t} =0$ leads us to \end{align} This is called the \textbf{energy equation} and is a general result for a inviscid and incompressible fluids, which we can apply to study water waves. -We start of with replacing $\nabla = \nabla_\perp + \frac{\partial }{\partial +We start off with replacing $\nabla = \nabla_\perp + \frac{\partial }{\partial z} $ and $\Omega = g z$ and multiplying by $\rho$, then our energy equation in \ref{eq:energy} becomes \begin{align} @@ -551,6 +553,388 @@ can set $P_s=0$, such that $\mathcal{P} =0$ leaving us with the equation We note that the assumption $P_s=0$ is only possible if the coefficient of surface tension is set to 0, which usually is not the case. \section{Dimensional Analysis} +Our derived model of fluid dynamics yields formal connections between +physical quantities. These quantities bear units, e.g. the velocity of fluid +particles $\mathbf{u}$ has the ``SI'' unites of $\frac{m}{s}$, meters per +second. The idea is the make use of these scales and formulate a model, where +the quantities are nondimensionalized, i.e. to get rid of physical units by +scaling each quantity appropriately. The appropriate length scales are that +of the typical water depth $h_0$ and the typical wavelength $\lambda$ of a +surface wave. + +\subsection{Nondimensionalisation} +In summary we use these adaptations +\begin{itemize} + \item $h_0$ for the typical water depth + \item $\lambda$ for the typical wavelength + \item $\frac{\lambda}{\sqrt{g h_0}}$ time scale of wave propagation + \item $\sqrt{g h_0}$ velocity scale of waves in $(x, y)$ + \item $\frac{h_0 \sqrt{g h_0} }{\lambda}$ velocity scale in the $z$ + direction. +\end{itemize} +$(x, z, t)$, then +\begin{align} + u = \psi _z, \qquad w = - \psi_x; +\end{align} +and the scale of $\psi$ must be $h_0\sqrt{g h_0}$. Additionally we write the +boundary condition on the free surface as follows +\begin{align} + h = h_0 + a \eta (\mathbf{x}_\perp, t) = z, +\end{align} +where $a$ is the typical amplitude and $\eta$ nondimensional function. All in +all we have the following scaling for the physical quantities of our context +\begin{align} + &x \rightarrow\ \lambda x, \quad u \rightarrow \sqrt{gh_0} u, \\ + &y \rightarrow\ \lambda y, \quad v \rightarrow \sqrt{gh_0} v, \qquad + t\rightarrow \frac{\lambda}{\sqrt{gh_0}}t,\\ + &z \rightarrow\ h_0 z, \quad w \rightarrow + \frac{h_0\sqrt{gh_0}}{\lambda} w. +\end{align} +with +\begin{align} + h = h_0 + a \eta, \qquad b \rightarrow h_0 b. +\end{align} +The pressure is also rewritten into +\begin{align} + P = P_a + \rho g(h_0 -z) + \rho g h_0 p, +\end{align} +where $P_a$ is the atmospheric pressure, the term $h_0-z$ represent the +hydrostatic pressure distribution, i.e. pressure at depth and the term with the pressure +variable $p$ measures the deviation from the hydrostatic pressure +distribution. Indeed $p\neq 0 $ for wave propagation. Now we can perform a +rescaling of the Euler's Equation of Motion, we introduce the notation +\begin{align} + &t = \frac{\lambda}{\sqrt{gh_0}}\tau,\quad x = \lambda \xi,\quad u = + \sqrt{gh_0} \tilde{u}\\ + &y = \lambda \chi,\quad v = \sqrt{gh_0} \tilde{v}\\ + &z = h_0 \zeta, \quad w = \frac{h_0\sqrt{gh_0} }{\lambda}\tilde{w}. +\end{align} +We start off with the $x$ coordinate, substitute and apply the chain rule +leading us to +\begin{align} + \frac{Du}{Dt} + &= \frac{\partial u}{\partial t} +u \frac{\partial + u}{\partial x} \\ + &= \sqrt{gh_{0}}\frac{\partial \tilde{u}}{\partial \tau} \frac{\partial + \tau}{\partial t} +gh_0 \tilde{u} \frac{\partial \tilde{u}}{\partial \xi} + \frac{\partial \xi}{\partial x} \\ + &= \frac{gh_0}{\lambda} \left( \frac{\partial \tilde{u}}{\partial \tau} + \tilde{u} \frac{\partial \tilde{u}}{\partial \xi} \right), +\end{align} +on the other hand +\begin{align} + \frac{gh_0}{\lambda} \left( \frac{\partial \tilde{u}}{\partial \tau} + \tilde{u} \frac{\partial \tilde{u}}{\partial \xi} \right) + &=-\frac{1}{\rho}\frac{1}{\lambda}\frac{\partial P}{\partial x} \\ + &=-\frac{ g h_0 }{\lambda}\rho \frac{\partial p}{\partial \xi}. +\end{align} +Thereby the rescaling evolves to +\begin{align} + \frac{D \tilde{u}}{D\tau} = -\frac{\partial p}{\partial \xi}. +\end{align} +Because of the same scaling in $y$ we get the same result as in $x$, that is +\begin{align} + \frac{D \tilde{v}}{D\tau} = -\frac{\partial p}{\partial \chi}. +\end{align} +In the $z$ coordinate we have +\begin{align} + \frac{Dw}{Dt} + &= \frac{\partial w}{\partial t} +w \frac{\partial + w}{\partial \zeta} \\ + &= \frac{h_0\sqrt{gh_0}}{\lambda} \frac{\sqrt{gh_0}}{\lambda} + \frac{\partial \tilde{w}}{\partial \tau} + \frac{1}{h_0} + \frac{h_0\sqrt{gh_0} }{\lambda} \frac{h_0\sqrt{gh_0}}{\lambda} + \tilde{w}\frac{\partial \tilde{v}}{\partial \zeta}\\ + &= \frac{h_0^2g}{\lambda}\left( \frac{\partial \tilde{w}}{\partial \tau} + + \tilde{w}\frac{\partial \tilde{w}}{\partial \zeta} \right) . +\end{align} +On the other side we have +\begin{align} + \frac{h_0^2g}{\lambda}\left( \frac{\partial \tilde{w}}{\partial \tau} + + \tilde{w}\frac{\partial \tilde{w}}{\partial \zeta} \right) + &= + -\frac{1}{h_0\rho} \frac{\partial P}{\partial z} +g \\ + &=-\frac{1}{h_0\rho}(-\rho gh_0 \frac{\partial \zeta}{\partial \zeta} + \rho gh_0 + \frac{\partial p}{\partial \zeta} ) + g \\ + &= -g \frac{\partial p}{\partial z}. +\end{align} +In total for the $z$ direction we get +\begin{align} + \underbrace{\left( \frac{h_0}{\lambda} \right)^2}_{=: \delta^2} + \frac{Dw}{Dt} = -\frac{\partial p}{\partial z}, +\end{align} +where $\delta$ is the \textbf{long wavelength} or \textbf{shallowness} +parameter, a very important constant for developing model hierarchies. For +clarity we resubstitute for $x, y, z, t, u, v$ and $w$, and for completeness +the we display the equations again, which are +\begin{align}\label{eq:nondim-motion} + \frac{Du}{Dt} = - \frac{\partial p}{\partial x}&, \quad + \frac{Dv}{Dt} = - \frac{\partial p}{\partial y}, \quad + \delta^2\frac{Dw}{Dt} = - \frac{\partial p}{\partial z}, \\ + &\frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} + +\frac{\partial w}{\partial z} = 0. +\end{align} +We can now turn our attention to the boundary conditions, on both free +surface $z=h$ and the bottom $z=b$ we have $z \Rightarrow h_0 z$ and thereby +\begin{align} + z = 1+ + \underbrace{\frac{a}{h_0}}_{:=\varepsilon}\eta(\mathbf{x}_\perp,t) \quad + \text{and}\quad z= b, +\end{align} +where we arrive at our second very important parameter $\varepsilon$ called +the \textbf{amplitude} parameter. As for the kinematic condition, we +substitute the free surface $z=h = 1+\varepsilon \eta$ and get +\begin{align} + \frac{Dz}{Dt} = \varepsilon\left(\eta_t + (\mathbf{u}_\perp + \nabla_\perp)\eta\right) \qquad \text{on}\;\; z= 1+\varepsilon \eta. +\end{align} +Respectively the bottom condition is not changed +\begin{align} + w = b_t + (\mathbf{u}_\perp \nabla_\perp) b \quad \text{on}\;\; z= b. +\end{align} +The general dynamic condition for $h = h(x, y, t)$ yields a rescaling of the +curvature in terms of +\begin{align} + \frac{1}{R} + &= \frac{(1+h_y^2)h_{x x} + (1+h_x^2)h_yy - 2h_xh_yh_{xy} + }{\left(h_x^2+h_y^2 +1 \right)^{\frac{3}{2}} } \\ + &= -\frac{\varepsilon h_0}{\lambda^2} \frac{( + 1+\varepsilon^2\delta^2\eta_y^2 )\eta_{x x}+ + (1+\varepsilon^2\delta^2\eta_x^2)\eta_{yy} - + 2\varepsilon^2\delta^2\eta_x\eta_y\eta_{xy}}{\left( + 1+\varepsilon^2\delta^2\eta_x^2+\varepsilon^2\delta^2\eta_y^2 + \right)^{\frac{3}{2}} }, +\end{align} +together with the pressure difference +\begin{align} + \Delta P = \rho g h_0(p - \varepsilon \eta) = \frac{\Gamma}{R}, +\end{align} +leaving us ultimately with the dynamic condition +\begin{align} + p-\varepsilon\eta= \varepsilon\left( \frac{\Gamma}{\rho g\lambda^2} + \right) \left(\frac{\lambda^2}{\varepsilon h_0}\frac{1}{R}\right), +\end{align} +where $W_e = \frac{\Gamma}{\rho g h_0^2}$ is the \textbf{Weber number}. This +dimensionless parameter can be considered as a measure of the fluid's inertia +compered to its surface tension, which satisfies the relation +\begin{align} + \delta^2 W_e = \frac{\Gamma}{\rho g \lambda^2}. +\end{align} +\subsection{Scaling of Variables} +Admits a simple observation of the governing equations in the last chapter we +notice that $w$ and $p$ on the free surface $z = 1 + \varepsilon\eta$ are +directly proportional to $\varepsilon$. Hence we want to ''scale this way`` +by introducing the following transformation +\begin{align} + p \rightarrow \varepsilon p, \quad w \rightarrow \varepsilon w, \quad + \mathbf{u}_\perp \rightarrow \varepsilon \mathbf{u}_\perp. +\end{align} +Because of this scaling our material derivative changes slightly to +\begin{align}\label{eq:mod-material} + \frac{D}{Dt} = \frac{\partial }{\partial t} + \varepsilon\left(u + \frac{\partial }{\partial x} + v \frac{\partial }{\partial y} + w + \frac{\partial }{\partial z} \right) +\end{align} +A simple recalculation yields the rescaled, nondimensionalized Euler's +Equation of motion are the same as in equations \ref{eq:nondim-motion} with +the modified material derivative from \ref{eq:mod-material}, and the boundary +conditions are +\begin{align} + p &= \eta - \frac{\delta^2\varepsilon h_0}{\lambda^2} \frac{W_e}{R}\\ + w &= \frac{1}{\varepsilon}\eta_t + (\mathbf{u}_\perp \nabla_\perp)\eta + \quad \text{on}\;\; z = 1+\varepsilon\eta\\ + w &=\frac{1}{\varepsilon}b_t + (\mathbf{u}_\perp \nabla_\perp)b \quad + \text{on}\;\; z=b +\end{align} +\subsection{Model Hierarchies} +As we have derived a model of fluid dynamics, with small parameters +$\varepsilon$ and $\delta$, we can conduct a series of classifications and +perform asymptotic analysis on them. The main hierarchies important in this +review are derived from the following problem classifications +\begin{itemize} + \item $\varepsilon\rightarrow 0$: linearized problem, small amplitude + \item $\delta\rightarrow 0$: shallow Water, long-wave + \item$\delta \rightarrow 0;\; \varepsilon~1$: shallow Water, large + amplitude + \item $\delta\ll 1;\; \varepsilon~\delta$: shallow water, medium + amplitude + \item $\delta\ll 1;\; \varepsilon~\delta^2$: shallow water, small + amplitude + \item $\delta \gg 1;\; \varepsilon\delta\ll 1$: deep water, small + steepness. +\end{itemize} + +\section{The Solitary Wave and The KdV Equation} +The solitary wave is a wave of translation, it is stable and can travel long +distances additionally the speed depends on the size of the wave. An +interesting feature is that two solitary waves do not merge together to form +one solitary wave, rather the small wave is overtaken by a larger one. If a +solitary wave is too big for the depth it splits into two, a big and a small +one. Solitary waves arise in the region $\varepsilon=O(\delta^2)$. + + +\subsection{Solitary Wave} +To describe +a solitary wave we begin with Euler's Equation of Motion, where we assume +there is no surface tension we set $W_e = 0$ and additionally assume +irrotational flow $\mathbf{\omega}=\nabla \times \mathbf{u} = 0$. This means +that there exists a velocity potential $\phi(\mathbf{x},t)$ given +by$\mathbf{u} = \nabla \phi$ satisfying the Laplace equation. In regard of a +solitary wave being a plane wave, we rotate our coordinate system such that +the propagation is in the $x$-direction and a stationary \& fixed bottom +$b=0$. Ultimately leaving us with the following model +\begin{align}\label{eq:soliton} +\begin{drcases} + & \phi_{zz} + \delta \phi_{x x } = 0,\\ + &\text{with the boundary conditions}\\ + &\begin{drcases} + &\phi_z = \delta^2 (\eta_t + \varepsilon \phi_x \eta_x) \\ + &\phi_t + \eta + \frac{1}{2}\varepsilon\left( \frac{1}{\delta^2}\phi^2_z + + \phi_x^2\right) =0 + \end{drcases}\quad \text{on}\;\; z = 1+\varepsilon\eta,\\ + &\text{and}\\ + & \phi_z =0 \quad \text{on}\;\; z = b = 0. +\end{drcases} +\end{align} +Since the model arises $\varepsilon = O(\delta^2)$, for convince we set +$\varepsilon=1$. The fact of the matter is we are seeking a traveling wave +solution, thereby we can go into the coordinate system of the traveling wave, +one in the variable $\xi = x - ct$ for a from left to right traveling wave, +where $c$ is the nondimensional speed of the wave. Our goal is to find the +solution for the velocity potential $\phi(\xi, z)$ and the wave profile +$\eta(\xi)$. The chain rule gives us +\begin{align} + \frac{\partial }{\partial x} &= \frac{\partial \xi}{\partial x} + \frac{\partial }{\partial \xi} = \frac{\partial }{\partial \xi}, \\ + \frac{\partial }{\partial t} &= \frac{\partial \xi}{\partial t} + \frac{\partial }{\partial \xi} = -c\frac{\partial }{\partial \xi}. +\end{align} +Together with the equations in \ref{eq:soliton} we obtain +\begin{align}\label{eq:soliton-xi} + \begin{drcases} + & \phi_{zz} + \delta \phi_{\xi\xi} = 0,\\ + &\text{with the boundary conditions}\\ + &\begin{drcases} + &\phi_z = \delta^2 (\phi_\xi -c)\eta_\xi \\ + &-c\phi_\xi + \eta + \frac{1}{2}\varepsilon\left( \frac{1}{\delta^2}\phi^2_z + + \phi_\xi^2\right) =0 + \end{drcases}\quad \text{on}\;\; z = 1+\eta,\\ + &\text{and}\\ + & \phi_z =0 \quad \text{on}\;\; z = b = 0. + \end{drcases} +\end{align} +\subsubsection{Exponential Decay} +We would like to analyze if the equation in \ref{eq:soliton-xi} gives viable a +solution that decays exponentially, we make the ansatz +\begin{align} + \eta \simeq a e^{-\alpha |\psi|},\quad \phi \simeq \psi(z)e^{-\alpha + |\xi|}, \qquad \mid \xi \mid \rightarrow \infty, +\end{align} +where $\alpha>0$ is the exponent. The equations in \ref{eq:soliton-xi} +transforms to +\begin{align} + \psi'' + \alpha^2 \delta^2\psi = 0. +\end{align} +The above equation is a standard well known ordinary differential equation +reading +\begin{align} + \psi = A \cos(\alpha\delta z), +\end{align} +where $A$ is the integration constant. On the free surface $z\simeq 1$ gives +\begin{align} + &-cA\alpha\sin(\alpha\delta) = ca\alpha,\label{eq:sol1}\\ + &cA\alpha \cos(\alpha\delta) = -a \label{eq:sol2}. +\end{align} +Dividing equation \ref{eq:sol1} with equation \ref{eq:sol2} gives +\begin{align} \label{eq:soliton-dispersion} + c^2 = \frac{\tan\left(\alpha\delta \right) }{\alpha\delta}. +\end{align} +We conclude that the solution for such a wave exists provided that the +dispersion relation on the wave propagation speed holds, thereby all solitary +waves exhibit exponential decay in their tail and satisfy the dispersion +relation in equation \ref{eq:soliton-dispersion}. +\subsubsection{Asymptotic Analysis} +The underlining equations in \ref{eq:soliton} extend from $-\infty$ to +$\infty$, so the length scale is much greater than any finite depth of +water. Therefore the classification $\delta \rightarrow 0$ is appropriate for +a solitary wave, this however goes with the assumption +$\varepsilon\rightarrow 0$ otherwise we cannot make an appropriate expansion. +Let us look at the main equation +\begin{align}\label{eq:sol-laplace} + \phi_{zz} + \delta \phi_{x x} = 0. +\end{align} +For small $\delta$ we conduct the $\delta^2 = O(\varepsilon)$ standard ansatz +in asymptotic analysis +\begin{align} + \phi_{\delta}(x, t, z) \simeq \sum_{n=0}^{\infty} \delta^{2n}\phi_n(x, t, + z). +\end{align} +Substituting $\phi_\delta$ into equation \ref{eq:sol-laplace} we get +\begin{align} + \delta^{2\cdot 0}\left( \phi_{0zz} \right) + \delta^{2\cdot 1}\left( + \phi_{1zz}+\phi_{0 x x} \right) + \delta^{2\cdot 2}\left( \phi_{2zz}+ + \phi_{1 x x} \right) + O(\delta^{2\cdot 3}) = 0. +\end{align} +We start off with $O(\delta^{2\cdot0}) $, which gives us an arbitrary function +$\phi_{0} = \theta(x, t)$. Next we may generalize the results for all +$O(\delta^{2\cdot n})$ in the means of +\begin{align} + \phi_{n+1zz} = -\phi_{nx x}\qquad \forall n\in \mathbb{N} . +\end{align} +Therefore leaving us for $\phi_1$ and $\phi_2$ with +\begin{align} + &\phi_1 = -\frac{1}{2} z^2 \theta_0(x,t) + \theta(x, t),\\ + \Rightarrow& \phi_2 = + \frac{1}{24}z^4\theta_0(x,t)-\frac{1}{2}z^2\theta_1(x,t) + \theta_2(x,t). +\end{align} +The boundary condition on the bottom comes around to be +\begin{align} + \phi_{nz} =0 \quad \text{on}\;\; z=0. +\end{align} +The free surface boundary condition $z= 1+\varepsilon\eta$ n evolves more calculation, we consider +only terms up the order of $\delta^2$, initializing with +\begin{align} + &\phi_z = \delta^2(\eta_t + \varepsilon\phi_x \eta_x)\\ + \Leftrightarrow &\frac{1}{\delta}\phi_z = \eta_t + \varepsilon\phi_x + \eta_x, +\end{align} +substituting $\phi_\delta$ into the above proceeds to be +\begin{align} + \frac{1}{\delta^2}\underbrace{\phi_{0z}}_{=0} + \phi_{1z}+ \delta^2\phi_{zz} + O(\delta^{2\cdot 2}) + &= -z\theta_{x x} + \delta^2\left( \frac{1}{6}z^3\theta_{0 x x x x} - z + \theta_{0x x} \right) + O(\delta^{2\cdot 2})\\ + &=-(1+\varepsilon\eta)\theta_{0 x x} + \delta^2\left( + \frac{1}{6}(1+\varepsilon\eta)^3\theta_{0 x x} - +(1+\varepsilon\eta)\theta_{0 x x} \right) +O(\delta^{2\cdot 2})\\ + &= \eta_t + \varepsilon\eta_x \left( \theta_{0x} + \delta^2(\theta_{1x}-\frac{1}{2}( 1+ \varepsilon\eta)^2 \theta_{0x x x} +\right). +\end{align} +The second condition is +\begin{align} + \phi_t + \eta + \frac{1}{2}\varepsilon \left( \frac{1}{\delta}\phi^2_z + \phi_x^2\right) = 0, +\end{align} +becomes after substitution +\begin{align} + \theta_{0t}+ \delta^2\left( -\frac{1}{2}(1+\varepsilon\eta)^2\theta_{0 x xt} + + \theta_{1t}\right) + \eta + O(\delta^{2\cdot 2}) + &=-\frac{1}{2}\delta^2\varepsilon(-(1+\varepsilon\eta)\theta_{0 x x + })^2\\ + &-\frac{1}{2}\left( \theta_{0 x} + \delta^2\left( \theta_{1x} - + \frac{1}{2}(1+\varepsilon\eta)^2\theta_{0 x x x x} \right) \right) ^2 +\end{align} +The simplest case is $\varepsilon,\delta \rightarrow 0$. + + + + + + + + diff --git a/app_pde/build/basics_fluids.pdf b/app_pde/build/basics_fluids.pdf Binary files differ.