Finish first draft of introduction and method section.

This commit is contained in:
Janita Willumsen 2023-12-19 14:50:06 +01:00
parent 939aa3ad50
commit f27cca4baf
7 changed files with 688 additions and 106 deletions

View File

@ -2,103 +2,48 @@
\begin{document}
\section{Theoretical background}\label{sec:theory}
Scientists have made use of light in devices leading to...
The use of light have facilitated many discoveries through history. The invention
of the compound microscope, which used lenses to focus light, led to the first
observation of animal cells. And the telescope, which led to a greater understanding
of the universe.
The nature of light has long been a subject of interest and discussion. Around year
1600 the first microscope was invented, which made use of light through lenses.
The observation made using the microscope led to a greater understanding of the
world at a microscopic scale. However, the study of light itself have led to (something).
Through the 1600s, the view shifted from particle to wave.
% Introduction?
The nature of light has long been a subject of discussion, from the 1500s
The nature of light has long been a subject of interest and discussion, from the 1500s
and the invention of microscopes, through the 1700s where both a particle theory
and a wave theory. In the late 1600s, Christiaan Huygens proposed the wave theory of light,
which was challenged by Isaac Newton's particle theory. The particle theory made
was the leading theory in the beginning of 1800s, when Thomas Young demonstrated
the interference of light, through his double-slit experiment, the wave theory found
which was challenged by Isaac Newton's particle theory. The particle theory was
the leading theory in the beginning of 1800s, when Thomas Young demonstrated
the interference of light, through his double-slit experiment. The wave theory found
new hold.
In the 1800s, the study of ideal black bodies done by Gustav R. Kirchhoff, lead to a
better understanding of heat radiation. Wilhelm Wien started working on determining the
spectral energy distribution, and Wien's law. The law did make sense for high frequencies,
however, there were inconsistencies when frequency were lower than a certain value.
however, as there were inconsistencies when the frequency were lower than a threshold value.
Wiens law led to an exponential curve, which disagree with the law of conservation.
Max Planck guessed a result which led to Plancks radiation law, which he later derived
using Boltzmanns statistical interpretation of the second law of thermodunamics.
Plancks findings gave rise to Einsteins quantum hypothesis, and later the wave-particle
Max Planck guessed a solution to the problem, which led to Planck's radiation law.
He later derived the radiation law using Boltzmanns statistical interpretation of
the second law of thermodynamics.
Plancks findings gave rise to the quantum hypothesis, and later Einsteins wave-particle
duality \cite{britannica:1998:planck}.
For small atoms classical mechanics are not able to explain the position of a particle,
For small atoms classical mechanics are not sufficient in describing the position of a particle,
and Heisenberg uncertainty priciple suggest that the particles have a wavelike behavior.
The wave-particle duality was later proposed to apply to particles by Louis de Broglie,
which inspired Erwin Schrödinger who proposed a wave function to describe the quantum
state of a particle, resulting in the wave equation.
% Methods?
% Schrödinger
The Schrödinger equation has a general form
\begin{align}
i \hbar \frac{\partial}{\partial t} | \Psi \rangle &= \hat{H} | \Psi \rangle \ ,
\label{eq:schrodinger_general}
\end{align}
where $i$ is the imaginary number, and $\hbar$ is Plancks constant. Here $| \Psi \rangle$
is the quantum state and $\hat{H}$ is a Hamiltonian operator.
For two-dimensional position space, the quantum state can be expressed using the
time-dependent complex-valued wave function $\Psi (x, y, t)$. The square modulus of the wave function $|\psi|^{2}$, predicts probability of finding the particle at position $(x, y)$ at time t.
% Segue to Born
The modulus of the wave function, is related to the probability density function
\begin{align}
p(x, y \ | \ t) &= |\Psi(x, y, t)|^{2} = \Psi^{*}(x, y, t) \Psi(x, y, t)
\label{eq:born_rule}
\end{align}
using the Born rule, where $^{*}$ denotes the complex conjugated wave function.
When the potential is time-independent, the Schrödinger equation can be expressed as
\begin{align*}
i \hbar \frac{\partial}{\partial t} \Psi (x, y, t) &= - \frac{\hbar^{2}}{2m} \bigg( \frac{\partial^{2}}{\partial x^{2}} + \frac{\partial^{2}}{\partial y^{2}} \bigg) \Psi (x, y, t) \\
& \quad + V(x, y, t) \Psi (x, y, t) \numberthis \ .
\label{eq:schrodinger_special}
\end{align*}
When we scale Schrödinger equation by the dimensionful variables, we are left with
a wave function $u$, potential $v$ and the dimensionless equation
\begin{align}
i \frac{\partial u}{\partial t} &= - \frac{\partial^{2} u}{\partial x^{2}} - \frac{\partial^{2} u}{\partial y^{2}} + v(x, y) u \ .
\label{eq:schrodinger_dimensionless}
\end{align}
% Crank-Nicolson
To evaluate the position of a single particle, we have to consider partial differential
equations (PDE). To solve these numerically, we have to discretize Equation \eqref{eq:schrodinger_dimensionless}.
We use the Crank-Nicolson method (CN), which combines the forward and backward finite
difference method. %
% Include the general $\theta$-method
Using the $\theta$-rule \footnote{For $\theta \in [0, 1]$ we can derive Forward Euler using $\theta = 1$, Backward Euler using $\theta = 0$, and $\theta = 1/2$ gives Crank-Nicolson}
with $\theta = 1/2$, CN can be written as
\begin{align}
\frac{u_{\ivec, \jvec}^{n+1} - u_{\ivec, \jvec}^{n}}{\Delta t} &= \frac{1}{2} \bigg[ F_{\ivec, \jvec}^{n+1} + F_{\ivec, \jvec}^{n} \bigg] \ .
\label{eq:crank_nicolson}
\end{align}
To simplify notation and avoid confusion of indices with the imaginary number $i$,
we have used the notation $\ivec, \jvec$ in subscript to indicate the commonly named indices $i, j$
in x- and y-direction. In addition, the superscript $n, n+1$ indicate position in time.
We use CN to derive the discretized Schrödinger equation
\begin{align*}
& u_{\ivec, \jvec}^{n+1} - \frac{i \Delta t}{2 \Delta x^{2}} \big[ u_{\ivec+1, \jvec}^{n+1} - 2u_{\ivec, \jvec}^{n+1} + u_{\ivec-1, \jvec}^{n+1} \big] \\
& - \frac{i \Delta t}{2 \Delta y^{2}} \big[ u_{\ivec, \jvec+1}^{n+1} - 2u_{\ivec, \jvec}^{n+1} + u_{\ivec, \jvec-1}^{n+1} \big] + \frac{i \Delta t}{2} v_{\ivec, \jvec} u_{\ivec, \jvec}^{n+1} \\
&= u_{\ivec, \jvec}^{n} + \frac{i \Delta t}{2 \Delta x^{2}} \big[ u_{\ivec+1, \jvec}^{n} - 2u_{\ivec, \jvec}^{n} + u_{\ivec-1, \jvec}^{n} \big] \\
& \quad + \frac{i \Delta t}{2 \Delta y^{2}} \big[ u_{\ivec, \jvec+1}^{n} - 2u_{\ivec, \jvec}^{n} + u_{\ivec, \jvec-1}^{n} \big] - \frac{i \Delta t}{2} v_{\ivec, \jvec} u_{\ivec, \jvec}^{n} \numberthis \ .
\label{eq:schrodinger_discretized}
\end{align*}
The full derivation of both Equation \eqref{eq:crank_nicolson} and Equation \eqref{eq:schrodinger_discretized}
can be found in Appendix \ref{ap:crank_nicolson}
First, we Taylor
expand the wave equation $u$ around (position and time).
\section{Notes}\label{sec:notes}
\subsection*{Introduction - draft 2}
In classical mechanics we study the kinematics and dynamics of physical objects,
ignoring its intrinsic properties for simplicity. It allows us to describe the
forces acting on an object as well as the motion of the object. We can describe
a planets orbital movement \cite{britannica:2023:kepler}, calculate the ... necessary
to launch satellites into orbit, or simply figure out where a ball is going to land
when you throw it... However, when want to study an object at a microscopic level,
e.g. a single atom, classical mechanics falls short.
% Double-slit experiment
Thomas Young first performed the double-slit experiment in 1801, to demonstrate
the principle of interference of light \cite{britannica:2023:young}, postulating
light as waves. In the study of blackbodies, scientists were not able to describe
@ -107,6 +52,292 @@ contradicted the principle of conservation of energy \cite{britannica:1998:planc
Max Planck assumed that the radiated energy consist of discrete values, or quanta,
to describe the peak in radiated energy.
% Reference: https://tex.stackexchange.com/questions/469109/how-to-change-arrowheads-to-lie-on-a-plane
\begin{tikzpicture}[scale=1.25,every node/.append style={transform shape}]
\foreach \x in {-1,-0.75,...,0} {
\draw (\x,-1) -- (\x,1);
}
\draw[fill=black!10] (0.5,-2,-1) -- (0.5,-2,1) -- (0.5,2,1) -- (0.5,2,-1) -- (0.5,-2,-1);
\fill (0.5,0,0) circle (0.05);
\foreach \r in {0.25,0.5,...,1.75} {
\draw (0.5,0) ++(-60:\r) arc (-60:60:\r);
}
\draw[fill=black!10] (2,-2,-1) -- (2,-2,1) -- (2,2,1) -- (2,2,-1) -- (2,-2,-1);
%\fill (2,0.5) circle (0.05) (2,-0.5) circle (0.05);
\foreach \r in {0.25,0.5,...,2} {
\draw (2,0.5) ++(-60:\r) arc (-60:60:\r);
\draw (2,-0.5) ++(-60:\r) arc (-60:60:\r);
}
\draw[fill=black!10] (4,-2,-1) -- (4,-2,1) -- (4,2,1) -- (4,2,-1) -- (4,-2,-1);
% LABELLING
\begin{scope}[canvas is yz plane at x=2,rotate=-90]
\node[circle,inner sep=0.5mm,fill,label=above:{S${}_1$}] at (0,0.5){};
\node[circle,inner sep=0.5mm,fill,label=below:{S${}_2$}] at (0,-0.5) {};
\draw[dash pattern=on 1.5pt off 1pt,thin] (-0.5,0.5) -- (0,0.5)
(-0.5,-0.5) -- (0,-0.5);
\pgflowlevelsynccm
\draw[|<->|] (-0.5,0.5) -- (-0.5,-0.5) node[midway,right=-0.1cm] {d};
\end{scope}
\begin{scope}[canvas is yz plane at x=0.5,rotate=-90]
\node[below left=-0.1cm] at (0,0) {S${}_0$};
\end{scope}
\begin{scope}[xshift=4cm,yshift=2cm,rotate=-90,canvas is xy plane at z=0]
\fill[white] (0,0) rectangle (4,4);
\begin{axis}[
width=5.575cm,
xmin=-0.5,
xmax=0.5,
ticks=none
]
\addplot [samples=1000,blue
]
{(cos(deg(5*pi*sin(deg(x)))))^(2)*((sin(deg(4*pi*sin(deg(x)))))/(4*pi*sin(deg(x))))^(2)};
\end{axis}
\end{scope}
\draw[thin,densely dashed,blue] (2,0) -- (6.9,0);
\draw[thin,densely dashed,blue] (2,0) -- +(15:2.5);
\draw[thin,densely dashed,blue] (2,0) -- +(-15:2.5);
\draw[thin,densely dashed,blue] (2,0) -- +(32:2.5);
\draw[thin,densely dashed,blue] (2,0) -- +(-32:2.5);
\end{tikzpicture}
% Author: Izaak Neutelings (June 2020)
% Inspiration:
% https://courses.physics.ucsd.edu/2011/Summer/session1/physics2c/diffraction.pdf
% https://tex.stackexchange.com/questions/201830/periodic-shading-in-tikz
% \begin{tikzpicture}[
% nodal/.style={mylightgreen,dashed,very thin},
% declare function={
% %xnode(\n,\dn,\lam,\f) = sqrt( (\n^2+(\n+\dn)^2)*\lambd^2/2 - (\n^2-(\n+\dn)^2)^2*\lambd^4/(4*\a^2) - \a^2/4 );
% xnode(\n,\dn,\lam,\f) = \lam/\f*sqrt( \n^2*(\f^2-\dn^2)+\n*\dn*(\f^2-\dn^2)+\dn^2*\f^2/2-(\f^4+\dn^4)/4 );
% ynode(\n,\dn,\lam,\a) = (2*\n*\dn+\dn^2)*\lam/(2*\f);
% intensity(\y,\lam,\a,\L) = cos(180*\a*\y/(2*\lam*sqrt(\L*\L+\y*\y)))^2;
% }
% ]
% \def\L{3.8} % distance between walls
% \def\H{5.4} % total wall height
% \def\h{2.8} % plane wave height
% \def\t{0.15} % wall thickness
% \def\a{1.15} % slit distance
% \def\d{0.20} % slit size
% \def\N{21} % number of waves
% \def\lambd{0.20} % wavelength
% \def\R{\N*\lambd} % wave radius
% \def\Nlines{3} % number of nodal lines
% \def\A{1.6} % amplitude
% %\def\r{0.06} % point source radius
% %\def\nmax{10}
% \def\nsamples{100}
% \def\ang{62}
% \begin{scope}
% \clip (-\t/2,-\H/2) rectangle (\L,\H/2);
% %\clip (-\t/2,0.7*\a) -- (0.6*\L,\H/2) -- (\L,\H/2) --
% % (\L,-\H/2) -- (0.6*\L,-\H/2) -- (-\t/2,-0.7*\a) -- cycle;
% % NODAL LINES
% \draw[nodal]
% (0.08*\N*\lambd,0) -- (1.06*\R,0);
% \coordinate (NP0) at (\L,0); % to avoid "Dimension too large error"
% \foreach \dn [evaluate={
% \f=\a/\lambd;
% \nmin=2.5+0.2*\dn; %0.501*(-\dn+\f)
% \nmax=10; %(NP0)
% \c=int(\dn<\f);
% \y=\L/sqrt((\a/(\lambd*\dn))^2-1);
% }] in {1,...,\Nlines}{
% \coordinate (NP+\dn) at (\L,\y); % to avoid "Dimension too large error"
% \coordinate (NP-\dn) at (\L,-\y); % to avoid "Dimension too large error"
% \ifnum\c=1
% \draw[nodal,variable=\n,samples=\nsamples,smooth]
% plot[domain=\nmin:\nmax] ({xnode(\n,\dn,\lambd,\f)},{ynode(\n,\dn,\lambd,\f)})
% -- (NP+\dn);
% \draw[nodal,variable=\n,samples=\nsamples,smooth]
% plot[domain=\nmin:\nmax] ({xnode(\n,\dn,\lambd,\f)},{-ynode(\n,\dn,\lambd,\f)})
% -- (NP-\dn);
% \fi
% }
% % WAVES
% \foreach \i [evaluate={\R=\i*\lambd;}] in {1,...,\N}{
% \ifodd\i
% \draw[myblue,line width=0.8] (0,\a/2)++(\ang:\R) arc (\ang:-\ang:\R);
% \draw[myred,line width=0.8] (0,-\a/2)++(\ang:\R) arc (\ang:-\ang:\R);
% \else
% \draw[myblue!80,line width=0.1] (0,\a/2)++(\ang:\R) arc (\ang:-\ang:\R);
% \draw[myred!80,line width=0.1] (0,-\a/2)++(\ang:\R) arc (\ang:-\ang:\R);
% \fi
% }
% \end{scope}
% % PLANE WAVES
% \foreach \i [evaluate={\x=-\i*\lambd;}] in {0,...,5}{
% \ifodd\i
% \draw[myblue,line width=0.8] (\x,-\h/2) -- (\x,\h/2);
% \else
% \draw[myblue,line width=0.1] (\x,-\h/2) -- (\x,\h/2);
% \fi
% }
% % WALL
% \fill[wall]
% (\t/2,\a/2-\d/2) rectangle (-\t/2,-\a/2+\d/2)
% (\t/2,\a/2+\d/2) rectangle (-\t/2,\H/2)
% (\t/2,-\a/2-\d/2) rectangle (-\t/2,-\H/2)
% (\L,-\H/2) rectangle (\L+\t,\H/2);
% % SHADES
% \begin{scope}[shift={(1.08*\L,0)}]
% \def\yz{\L/sqrt((\a/\lambd)^2-1)} % m = +- 1/2
% \def\yZ{\L/sqrt((\a/\lambd/2)^2-1)} % m = +- 1
% \clip (0,-\H/2) rectangle (1.1*\A,\H/2);
% \fill[white] (0,-\H/2) rectangle++ (\A,\H); % to fill seams
% \foreach \i [evaluate={\n=0.5*\i;\yn=\L/sqrt((\a/(2*\lambd*\n))^2-1);
% }] in {1,...,\Nlines}{
% \ifodd\i % if even
% \fill[myshadow] (0,{-\yn-0.1}) rectangle++ (\A,0.2); % to fill seams
% \fill[myshadow] (0,{ \yn-0.1}) rectangle++ (\A,0.2); % to fill seams
% \fi
% }
% \path[left color=myshadow,right color=myshadow,middle color=white,shading angle={180}]
% (0,{-\yz}) rectangle (\A,{\yz});
% \foreach \i [evaluate={
% \n=0.5*\i;
% \m=0.5*(\i+1);
% \yn=\L/sqrt((\a/(2*\lambd*\n))^2-1);
% \ym=\L/sqrt((\a/(2*\lambd*\m))^2-1);
% \dang=mod(\i,2)*180;
% }] in {1,...,\Nlines}{
% \path[left color=myshadow,right color=white,shading angle={\dang}]
% (0,\yn) rectangle (\A,\ym);
% \path[left color=myshadow,right color=white,shading angle={180+\dang}]
% (0,-\yn) rectangle (\A,-\ym);
% }
% \end{scope}
% % INTENSITY
% \begin{scope}[shift={(1.1*\L+1.1*\A,0)}]
% \draw[->,thick] (-0.08*\A,0) -- (1.3*\A,0) node[right=-2] {$\expval{I}$}; % I axis
% \draw[->,thick] (0,-0.52*\H) -- (0,0.54*\H) node[right] {$y$}; % y axis
% \draw[nodal] (NP0) --++ (0.15*\L+2.1*\A,0); % green nodal lines
% \foreach \i [evaluate={\y=\L/sqrt((\a/(\lambd*\i))^2-1)}] in {1,...,\Nlines}{ % green nodal lines
% \draw[nodal] (NP+\i) --++ ({0.15*\L+1.1*\A+\A*intensity(\y,\lambd,\a,\L)},0);
% \draw[nodal] (NP-\i) --++ ({0.15*\L+1.1*\A+\A*intensity(\y,\lambd,\a,\L)},0);
% }
% \draw[myred,thick,variable=\y,samples=\nsamples,smooth,domain=-\H/2:\H/2]
% plot({\A*intensity(\y,\lambd,\a,\L)},\y);
% \foreach \i [evaluate={ % ticks
% \modd=\i; %int(\i);
% \meven=int(\i-1);
% \y=\L/sqrt((\a/(\lambd*\i))^2-1);
% }] in {1,...,\Nlines}{
% \ifodd\i
% \tick{0,-\y}{180} node[right=0,scale=0.85] {$m=-\frac{\modd}{2}$};
% \tick{0,\y}{180} node[right=0,scale=0.85] {$m=+\frac{\modd}{2}$};
% \else
% \tick{0,-\y}{180} node[right=0,scale=0.85] {$m=-\meven$};
% \tick{0,\y}{180} node[right=0,scale=0.85] {$m=+\meven$};
% \fi
% }
% \end{scope}
% \end{tikzpicture}
% Schrödinger
% -----------
% The wave equation, Schrödinger equation has a general form
% \begin{align}
% i \hbar \frac{\partial}{\partial t} | \Psi \rangle &= \hat{H} | \Psi \rangle \ ,
% \label{eq:schrodinger_general}
% \end{align}
% where $i$ is the imaginary number, and $\hbar$ is Plancks constant. Here $| \Psi \rangle$
% is the quantum state and $\hat{H}$ is a Hamiltonian operator. % which represent the energy of the system
% For two-dimensional position space, the quantum state can be expressed using the
% time-dependent complex-valued wave function $\Psi (x, y, t)$. The square modulus of the wave function $|\psi|^{2}$, predicts probability of finding the particle at position $(x, y)$ at time t.
% % Segue to Born
% The modulus of the wave function, is related to the probability density function
% \begin{align}
% p(x, y \ | \ t) &= |\Psi(x, y, t)|^{2} = \Psi^{*}(x, y, t) \Psi(x, y, t)
% \label{eq:born_rule}
% \end{align}
% using the Born rule, where $^{*}$ denotes the complex conjugated wave function.
% When the potential is time-independent, the Schrödinger equation can be expressed as
% \begin{align*}
% i \hbar \frac{\partial}{\partial t} \Psi (x, y, t) &= - \frac{\hbar^{2}}{2m} \bigg( \frac{\partial^{2}}{\partial x^{2}} + \frac{\partial^{2}}{\partial y^{2}} \bigg) \Psi (x, y, t) \\
% & \quad + V(x, y, t) \Psi (x, y, t) \numberthis \ .
% \label{eq:schrodinger_special}
% \end{align*}
% When we scale Schrödinger equation by the dimensionful variables, we are left with
% a wave function $u$, potential $v$ and the dimensionless equation
% \begin{align}
% i \frac{\partial u}{\partial t} &= - \frac{\partial^{2} u}{\partial x^{2}} - \frac{\partial^{2} u}{\partial y^{2}} + v(x, y) u \ .
% \label{eq:schrodinger_dimensionless}
% \end{align}
% Crank-Nicolson
% --------------
% To evaluate the position of a single particle, we have to consider partial differential
% equations (PDE). To solve these numerically, we have to discretize Equation \eqref{eq:schrodinger_dimensionless}.
% We use the Crank-Nicolson method (CN), which combines the forward and backward finite
% difference method.
% Include the general $\theta$-method
% Using the $\theta$-rule \footnote{For $\theta \in [0, 1]$ we can derive Forward Euler using $\theta = 1$, Backward Euler using $\theta = 0$, and $\theta = 1/2$ gives Crank-Nicolson}
% with $\theta = 1/2$, CN can be written as
% \begin{align}
% \frac{u_{\ivec, \jvec}^{n+1} - u_{\ivec, \jvec}^{n}}{\Delta t} &= \frac{1}{2} \bigg[ F_{\ivec, \jvec}^{n+1} + F_{\ivec, \jvec}^{n} \bigg] \ .
% \label{eq:crank_nicolson}
% \end{align}
% To simplify notation and avoid confusion of indices with the imaginary number $i$,
% we have used the notation $\ivec, \jvec$ in subscript to indicate the commonly named indices $i, j$
% in x- and y-direction. In addition, the superscript $n, n+1$ indicate position in time.
% We use CN to derive the discretized Schrödinger equation
% \begin{align*}
% & u_{\ivec, \jvec}^{n+1} - \frac{i \Delta t}{2 \Delta x^{2}} \big[ u_{\ivec+1, \jvec}^{n+1} - 2u_{\ivec, \jvec}^{n+1} + u_{\ivec-1, \jvec}^{n+1} \big] \\
% & - \frac{i \Delta t}{2 \Delta y^{2}} \big[ u_{\ivec, \jvec+1}^{n+1} - 2u_{\ivec, \jvec}^{n+1} + u_{\ivec, \jvec-1}^{n+1} \big] + \frac{i \Delta t}{2} v_{\ivec, \jvec} u_{\ivec, \jvec}^{n+1} \\
% &= u_{\ivec, \jvec}^{n} + \frac{i \Delta t}{2 \Delta x^{2}} \big[ u_{\ivec+1, \jvec}^{n} - 2u_{\ivec, \jvec}^{n} + u_{\ivec-1, \jvec}^{n} \big] \\
% & \quad + \frac{i \Delta t}{2 \Delta y^{2}} \big[ u_{\ivec, \jvec+1}^{n} - 2u_{\ivec, \jvec}^{n} + u_{\ivec, \jvec-1}^{n} \big] - \frac{i \Delta t}{2} v_{\ivec, \jvec} u_{\ivec, \jvec}^{n} \numberthis \ .
% \label{eq:schrodinger_discretized}
% \end{align*}
% The full derivation of both Equation \eqref{eq:crank_nicolson} and Equation \eqref{eq:schrodinger_discretized}
% can be found in Appendix \ref{ap:crank_nicolson}.
% Implementation
The implementation of CN is simplified usein Dirichlet boundary conditions, which
can be found in Table \ref{tab:dirichlet}. In addition, we use Gaussian wave packets
for the initial wave function
\begin{align}
u(x, y, t=0) = e^{- \frac{(x-x_{c})^{2}}{2 \sigma_{x}^{2}} - \frac{(y-y_{c})^{2}}{2 \sigma_{y}^{2}} + ip_{x}x + ip_{y}y}
\end{align}
\begin{algorithm}[H]
\caption{Crank-Nicolson scheme}
\label{algo:cn_scheme}
\begin{algorithmic}
\Procedure{Crank-Nicolson}{$args$}
\State Insert pseudo code $var \leftarrow \text{ some text}$
\EndProcedure
\end{algorithmic}
\end{algorithm}
\section{Notes}\label{sec:notes}
\subsection*{Introduction - draft 2}
In classical mechanics we study the kinematics and dynamics of physical objects,
ignoring their intrinsic properties for simplicity. It allows us to describe the
forces acting on an object as well as the motion of the object. We can describe
a planets orbital movement \cite{britannica:2023:kepler}, calculate the ... necessary
to launch satellites into orbit, or simply figure out where a ball is going to land
when you throw it... However, when want to study an object at a microscopic level,
e.g. a single atom, classical mechanics falls short.
% Thomas Young first performed the double-slit experiment in 1801, to demonstrate
% the principle of interference of light \cite{britannica:2023:young}, postulating
% light as waves. In the study of blackbodies, scientists were not able to describe
% the radiated intensity of increased frequencies using classical mechanichs, as they
% contradicted the principle of conservation of energy \cite{britannica:1998:planck}.
% Max Planck assumed that the radiated energy consist of discrete values, or quanta,
% to describe the peak in radiated energy.
Light as particles -> waves -> particles/packets
Erwin Schrödinger wanted to find a mathematical description of the wave characteristics
of matter, supporting the wave-particle idea. He postulated a function which varies

View File

@ -16,6 +16,66 @@
urldate = {2023-12-15},
}
@article{sinha:2010:multi_order_inference,
author = {Urbasi Sinha and Christophe Couteau and Thomas Jennewein and Raymond Laflamme and Gregor Weihs},
journal = {Science},
title = {Ruling Out Multi-Order Interference in Quantum Mechanics},
volume = {329},
number = {5990},
pages = {418--421},
year = {2010},
doi = {10.1126/science.1190545},
url = {https://www.science.org/doi/abs/10.1126/science.1190545},
urldate = {2023-12-16}
}
@article{young:1804:double_slit,
author = {Thomas Young},
title = {I. The Bakerian Lecture. Experiments and calculations relative to physical optics},
journal = {Philosophical Transactions of the Royal Society of London},
volume = {94},
pages = {1--16},
year = {1804},
doi = {10.1098/rstl.1804.0001},
url = {https://royalsocietypublishing.org/doi/abs/10.1098/rstl.1804.0001},
urldate = {2023-12-16},
}
@book{benacquista:2018:classical_mechanics,
author = {Matthew J. Benacquista and Joseph D. Romano},
title = {Classical Mechanics},
publisher = {Springer International Publishing : Imprint: Springer},
year = {2018},
edition = {1}
}
@book{griffiths:2018:quantum_mechanics,
author = {David J. Griffiths and Darrell F. Schroeter},
title = {Introduction to quantum mechanics},
publisher = {Cambridge University Press},
year = {2018},
edition = {3}
}
@book{springer:2018:compendium_quantum_physics,
author = {Daniel Greenberger and Klaus Hentschel and Friedel Weinert},
title = {Compendium of Quantum Physics},
publisher = {Springer Berlin, Heidelberg},
year = {2009},
edition = {1}
}
@article{key,
author = {Asher Peres and Daniel R. Terno},
title = {Quantum Information and Relativity Theory},
journal = {Reviews of Modern Physics},
number = {1},
volume = {76},
doi = {10.1103/revmodphys.76.93},
year = {2004},
url = {http://dx.doi.org/10.1103/RevModPhys.76.93}
}
# Schrodinger and Crank-Nicolson
@book{wu:2023:quantum,
author = {Biao Wu and Ying Hu},

View File

@ -20,6 +20,7 @@
%% I recommend downloading TeXMaker, because it includes a large library of the most common packages.
\usepackage{physics,amssymb} % mathematical symbols (physics imports amsmath)
\usepackage[no-test-for-array]{nicematrix}
\usepackage{graphicx} % include graphics such as plots
\graphicspath{.images/}
@ -33,6 +34,29 @@
\usepackage{algorithm}
\usepackage[noend]{algpseudocode}
\usepackage{tikz}
% \usepackage{pgfplots}
% \pgfplotsset{compat=1.18}
% \usetikzlibrary{3d}
\usepackage{xcolor}
% \usepackage{etoolbox} %ifthen
% \usetikzlibrary{calc}
% \usetikzlibrary{arrows,arrows.meta}
% \usetikzlibrary{decorations.markings}
% \usetikzlibrary{angles,quotes}
% \usetikzlibrary{fadings}
% \tikzset{>=latex}
% \colorlet{wall}{blue!30!black}
% \colorlet{myblue}{blue!70!black}
% \colorlet{myred}{red!70!black}
% \colorlet{mydarkred}{red!50!black}
% \colorlet{mylightgreen}{green!60!black!70}
% \colorlet{mygreen}{green!60!black}
% \colorlet{myredgrey}{red!50!black!80}
% \colorlet{myshadow}{blue!30!black!90}
% \tikzstyle{wave}=[myblue,thick]
% \tikzstyle{mydashed}=[black!70,dashed,thin]
% \tikzstyle{mymeas}=[{Latex[length=3,width=2]}-{Latex[length=3,width=2]},thin]
% \tikzstyle{mysmallarr}=[-{Latex[length=3,width=2]}]
% \usetikzlibrary{quantikz}
% defines the color of hyperref objects
% Blending two colors: blue!80!black = 80% blue and 20% black
@ -44,7 +68,9 @@
% Biblio stuff
% \def\biblio{\bibliographystyle{plain}\bibliography{../references/references}}
\newcommand\numberthis{\addtocounter{equation}{1}\tag{\theequation}}
\newcommand{\rvline}{\hspace*{-\arraycolsep}\vline\hspace*{-\arraycolsep}}
% Defines indices i and j to avoid confusion with imaginary i
\newcommand{\ivec}{\hat{\imath}}
@ -67,10 +93,10 @@
\maketitle
% Introduction
% \subfile{sections/introduction}
\subfile{sections/introduction}
% Methods
% \subfile{sections/methods}
\subfile{sections/methods}
% Results
% \subfile{sections/results}
@ -79,7 +105,7 @@
% \subfile{sections/conclusion}
% Notes
\subfile{draft}
% \subfile{draft}
\clearpage
\newpage

View File

@ -3,7 +3,7 @@
\begin{document}
\appendix
\section{The Crank-Nicholson method}\label{ap:crank_nicolson}
The Crank-Nicolson \(CN\) approach considers both the forward difference, an explicit scheme,
The Crank-Nicolson (CN) approach consider both the forward difference, an explicit scheme,
\begin{equation*}
\frac{u_{\ivec, \jvec}^{n+1} - u_{\ivec, \jvec}^{n}}{\Delta t} = F_{\ivec, \jvec}^{n} \ ,
\end{equation*}

View File

@ -2,12 +2,63 @@
\begin{document}
\section{Introduction}\label{sec:introduction}
The nature of light has long been a subject of interest and discussion. %
% Important part of human behavior is observing and understanding our surroundings.
% Many big discoveries have been made through observations, verified by mathematical
% explanations. Classical physics is based on calculation predicting something we
% verify by observation etc. But what happens when we move down to the microscopic
% scale, can we still predict the position of a microscopic ball, also called an atom?
In classical mechanics, we study the kinematics and dynamics of physical objects,
while ignoring their intrinsic properties for simplicity. Newton's second law can be
applied to an object to describe its trajectory. % It allows us to describe the
% forces acting on an object as well as the motion of the object. We can describe
% a planets orbital movement \cite{britannica:2023:kepler}, calculate the ... necessary
% to launch satellites into orbit, or simply figure out where a ball is going to land
% when you throw it... However, when want to study an object at a microscopic level,
% e.g. a single atom, classical mechanics falls short.
Elementary particles such as electrons, does not abide by the laws of classical mechanics.
For several years, scientists did not agree on whether light was a particle or a
wave. Through the study of interference of light, and radiation of ideal blackbodies,
it has been shown that light has both wavelike and particle-like characteristics.
This is known as the wave-particle duality, and was showed by Albert Einstein in
1905. %
% Thomas Young studied the interference of light, and found that light to showed
% wavelike characteristics \cite{young:1804:double_slit}. This did not agree with
% Newtons particle-theory
Erwin Schrödinger wanted to find a mathematical description of the wave characteristics
of matter, supporting the wave-particle idea. He postulated a wave function which varies
with position, where the function squared can be interpreted as the probability
of finding an electron at a given position. This resulted in the Schrödinger equation,
a wave eqution of the energy levels for a hydrogen atom. It also shows how a quantum
state evolves with time \cite[p. 81]{wu:2023:quantum}.
We will simulate the time-dependent Schrödinger equation in two dimensions, to
study the light wave interference in the double-slit experiment. In addition, we
will include variations of walls such as single- and triple-slit. To solve the equation,
we will apply the Crank-Nicolson method in 2+1 dimensions.
% However, according to the Heisenberg uncertainty principle, we can't find dx and/or
% dp = 0. dx = sqrt{Var(x)} "spread in position", dp = hat{\Psi}(p) = sqrt{Var(p)}
% dx \cdot dp \geq \frac{\hbar}{2}
% Fourier transform \Psi(x) \doublearrow hat{\Psi}(p)
% \Psi(x) = \int_{infty}^{infty} (alt sum) hat{\Psi}(p) \cos(px) dp sum of different wave forms
% Light - particle or wave
% - double-slit, blackbodies radiation
% The nature of light.
% Position space
% - classical vs quantum mechanics
% Intuitions of the behavior of physical objects, predictable. Not like the quantum
% when we scale down to the atom, our intuition are not as reliable and prediction
% are not as perfect.
% Instead of finding the path of a ball, we find all the possible paths a ball can take.
% The world is not one-dimensional, and modelling it require partial diff eqs
In Section \ref{sec:methods}, we will present the theoretical background for
this experiment, as well as the algorithms and tools used in the implementation.
Continuing with Section \ref{sec:results}, we will present our results and
discuss our findings. Lastly, we will conclude our findings in Section \ref{sec:conclusion}.
\end{document}
% crank-nicolson method!

View File

@ -2,28 +2,231 @@
\begin{document}
\section{Methods}\label{sec:methods} %
\subsection{The Schrödinger equation}\label{ssec:schrodinger} %
% Add something that takes Planck to Schrödinger
% In classical mechanics, we have Newton laws and conservation of energy. In quantum
% mechanics, we have Schrödinger equation.
The Schrödinger equation has a general form
\begin{align}
i \hbar \frac{\partial}{\partial t} | \Psi \rangle &= \hat{H} | \Psi \rangle \ ,
\label{eq:schrodinger_general}
\end{align}
where $i$ is the imaginary number, and $\hbar$ is Plancks constant. $\hat{H}$ is
a Hamiltonian operator, which represent the energy for the system, and $| \Psi \rangle$
is the quantum state. In two-dimensional position space, the quantum state can
be expressed using the time-dependent complex-valued wave function $\Psi (x, y, t)$.
Using Born rule, the square modulus of the wave function is proportional to the
probability density of finding a particle at position $(x, y)$ at time t. The relation
is given by
\begin{align}
p(x, y \ | \ t) &= |\Psi(x, y, t)|^{2} = \Psi^{*}(x, y, t) \Psi(x, y, t) \ ,
\label{eq:born_rule}
\end{align}
where $\Psi^{*}$ denotes the complex conjugated wave function.
% Add something about kinetic and potential energy, to introduce the potential V
When the potential is time-independent, the Schrödinger equation can be expressed as
\begin{align*}
i \hbar \frac{\partial}{\partial t} \Psi (x, y, t) &= - \frac{\hbar^{2}}{2m} \bigg( \frac{\partial^{2}}{\partial x^{2}} + \frac{\partial^{2}}{\partial y^{2}} \bigg) \Psi (x, y, t) \\
& \quad + V(x, y, t) \Psi (x, y, t) \numberthis \ .
\label{eq:schrodinger_special}
\end{align*}
The partial derivatives (...) gives the kinetic energy, and the potental $V$ is
the external environment. In this experiment we will only consider the case where
the potential is time-independent, resulting in $V = V(x, y)$
When we scale Schrödinger equation by the dimensionful variables, we are left with
a wave function $u$, potential $v$ and the dimensionless equation
\begin{align}
i \frac{\partial u}{\partial t} &= - \frac{\partial^{2} u}{\partial x^{2}} - \frac{\partial^{2} u}{\partial y^{2}} + v(x, y) u \ .
\label{eq:schrodinger_dimensionless}
\end{align} %
This gives us the Born rule
\begin{align}
p(x, y \ | \ t) &= |u(x, y, t)|^{2} = u^{*}(x, y, t) u(x, y, t) \ .
\label{eq:born_rule_scaled}
\end{align}
\subsection{Double-slit experiment}\label{ssec:double_slit} %
In the beginning of the 1800s, the general view was that light consisted of particles.
However, in 1801 Thomas Young demonstrated the principle of interference of light
\cite{britannica:2023:young}, while postulating light as waves rather than particles.
% Thomas Young first performed the double-slit experiment in 1801, to demonstrate
% the principle of interference of light \cite{britannica:2023:young}, postulating
% light as waves. In the study of blackbodies, scientists were not able to describe
% the radiated intensity of increased frequencies using classical mechanichs, as they
% contradicted the principle of conservation of energy \cite{britannica:1998:planck}.
% Max Planck assumed that the radiated energy consist of discrete values, or quanta,
% to describe the peak in radiated energy.
The double-slit experiment
\subsection{Schrödinger equation}\label{ssec:schrodinger} %
\subsection{Crank-Nicolson}\label{ssec:crank_nicolson} %
\subsection{The Crank-Nicolson scheme}\label{ssec:crank_nicolson} %
When we evaluate a particle in position space, we have to consider partial differential
equations (PDE). To solve these numerically, we have to discretize Equation \eqref{eq:schrodinger_dimensionless}.
We use the $\theta$-rule \footnote{Using the $\theta$-rule, we can derive Forward Euler using $\theta = 1$, and Backward Euler using $\theta = 0$},
to combine the forward (explicit) and backward (implicit) finite difference method.
The result is a linear combination of the explicit and implicit scheme, given by
\begin{align}
\frac{u_{\ivec, \jvec}^{n+1} - u_{\ivec, \jvec}^{n}}{\Delta t} &= \theta F_{\ivec, \jvec}^{n+1} + (1 - \theta) F_{\ivec, \jvec}^{n} \ ,
\label{eq:theta_rule}
\end{align} %
where $\theta \in [0, 1]$.
To simplify notation and avoid confusion of indices with the imaginary number $i$,
we have used the notation $\ivec, \jvec$ in subscript to indicate the commonly named indices $i, j$
in x- and y-direction. In addition, the superscript $n, n+1$ indicate position in time.
We get the Crank-Nicolson method (CN) when $\theta = 1/2$ gives Crank-Nicolson
\begin{align}
\frac{u_{\ivec, \jvec}^{n+1} - u_{\ivec, \jvec}^{n}}{\Delta t} &= \frac{1}{2} \bigg[ F_{\ivec, \jvec}^{n+1} + F_{\ivec, \jvec}^{n} \bigg] \ .
\label{eq:crank_nicolson_method}
\end{align} %
Using CN, we derive the discretized Schrödinger equation given by
\begin{align*}
& u_{\ivec, \jvec}^{n+1} - \frac{i \Delta t}{2 \Delta x^{2}} \big[ u_{\ivec+1, \jvec}^{n+1} - 2u_{\ivec, \jvec}^{n+1} + u_{\ivec-1, \jvec}^{n+1} \big] \\
& - \frac{i \Delta t}{2 \Delta y^{2}} \big[ u_{\ivec, \jvec+1}^{n+1} - 2u_{\ivec, \jvec}^{n+1} + u_{\ivec, \jvec-1}^{n+1} \big] + \frac{i \Delta t}{2} v_{\ivec, \jvec} u_{\ivec, \jvec}^{n+1} \\
&= u_{\ivec, \jvec}^{n} + \frac{i \Delta t}{2 \Delta x^{2}} \big[ u_{\ivec+1, \jvec}^{n} - 2u_{\ivec, \jvec}^{n} + u_{\ivec-1, \jvec}^{n} \big] \\
& \quad + \frac{i \Delta t}{2 \Delta y^{2}} \big[ u_{\ivec, \jvec+1}^{n} - 2u_{\ivec, \jvec}^{n} + u_{\ivec, \jvec-1}^{n} \big] - \frac{i \Delta t}{2} v_{\ivec, \jvec} u_{\ivec, \jvec}^{n} \numberthis \ .
\label{eq:schrodinger_discretized}
\end{align*} %
The full derivation of both Equation \eqref{eq:crank_nicolson_method} and Equation \eqref{eq:schrodinger_discretized}
can be found in Appendix \ref{ap:crank_nicolson}.
\subsection{The double-slit experiment}\label{ssec:double_slit} %
Thomas Young first performed the double-slit experiment in 1801 to demonstrate the
principle of interference of light \cite{britannica:2023:young}, while postulating
light as waves rather than particles. The double-slit experiment gives a diffraction
pattern, where constructive interference of light result in bright spots, and destructive
interference result in dark spots.
% Something about Heisenberg uncertainty principle
% Might be better to move theory on double-slit here and have a subsection of light
% property etc. as a first section or in introduction?
\subsection{Implementation}\label{ssec:implementation} %
% Add tables of parameters used, initial conditions, notation etc.
A, B are sparse csc matrix
- theory of csc matrix?
% \begin{equation*}
% \begin{pNiceArray}{ccc|ccc}
% \bullet & \bullet & & \bullet & & & & & \\
% \bullet & \bullet & \bullet & & \bullet & & & & \\
% & \bullet & \bullet & & & \bullet & & &
% \end{pNiceArray}
% \end{equation*}
\begin{equation*}
\begin{bmatrix}
\begin{matrix}
\bullet & \bullet & \phantom{\bullet} \\
\bullet & \bullet & \bullet \\
\phantom{\bullet} & \bullet & \bullet
\end{matrix}
& \rvline &
\begin{matrix}
\bullet & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \bullet & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \bullet
\end{matrix}
& \rvline &
\begin{matrix}
\phantom{\bullet} & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \phantom{\bullet}
\end{matrix} \\
\hline
\begin{matrix}
\bullet & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \bullet & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \bullet
\end{matrix}
& \rvline &
\begin{matrix}
\bullet & \bullet & \phantom{\bullet} \\
\bullet & \bullet & \bullet \\
\phantom{\bullet} & \bullet & \bullet
\end{matrix}
& \rvline &
\begin{matrix}
\bullet & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \bullet & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \bullet
\end{matrix} \\
\hline
\begin{matrix}
\phantom{\bullet} & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \phantom{\bullet}
\end{matrix}
& \rvline &
\begin{matrix}
\bullet & \phantom{\bullet} & \phantom{\bullet} \\
\phantom{\bullet} & \bullet & \phantom{\bullet} \\
\phantom{\bullet} & \phantom{\bullet} & \bullet
\end{matrix}
& \rvline &
\begin{matrix}
\bullet & \bullet & \phantom{\bullet} \\
\bullet & \bullet & \bullet \\
\phantom{\bullet} & \bullet & \bullet
\end{matrix}
\end{bmatrix}
\end{equation*}
Notations:
In addition, we use an equal step size in x- and y-direction, $h$ such that
\begin{align*}
x \in [0, 1] && x \rightarrow x_{\ivec} = \ivec h && \ivec = 0, 1, \dots, M-1 \\
y \in [0, 1] && y \rightarrow y_{\jvec} = \jvec h && \jvec = 0, 1, \dots, M-1 \\
t \in [0, T] && t \rightarrow t_{n} = n \Delta t && n = 0, 1, \dots, N_{t}-1
\end{align*}
And simplify indices such that
\begin{align*}
u(x, y, t) \rightarrow u(\ivec h, \jvec h, n \Delta t) \equiv u_{\ivec, \jvec}^{n} \\
v(x, y) \rightarrow u(\ivec h, \jvec h) \equiv v_{\ivec, \jvec}
\end{align*}
which gives a matrix $U^{n}$ that contains elements $u_{\ivec, \jvec}^{n}$, and
a matrix $V$ that contains elements $v_{\ivec, \jvec}$.
\begin{table}[H]
\centering
\begin{tabular}{l l l} % @{\extracolsep{\fill}}
\hline
Position & Value \\
\hline
$u(x=0, y, t)$ & $0$ \\
$u(x=1, y, t)$ & $0$ \\
$u(x, y=0, t)$ & $0$ \\
$u(x, y=1, t)$ & $0$ \\
\hline
\end{tabular}
\caption{Boundary conditions in the xy-plane, also known as Dirichlet boundary conditions.}
\label{tab:boundary_conditions}
\end{table}
For the general setup of the wall, we used
\begin{table}[H]
\centering
\begin{tabular}{l l} % @{\extracolsep{\fill}}
\hline
Parameter & Value \\
\hline
Wall thickness & $0.02$ \\
Wall position & $0.5$ \\
Separator length & $0.05$ \\
Slit aperture & $0.05$ \\
\hline
\end{tabular}
\caption{Wall setup.}
\label{tab:wall_setup}
\end{table}
\begin{table}[H]
\centering
\begin{tabular}{l l l} % @{\extracolsep{\fill}}
\hline
Simulation & $1$ & $2$ \\
\hline
$h$ & $0.005$ & $0.005$ \\
$\Delta t$ & $2.5 \times 10^{-5}$ & $2.5 \times 10^{-5}$ \\
$T$ & $0.008$ & $0.002$ \\
$x_{c}$ & $0.25$ & $0.25$ \\
$\sigma_{x}$ & $0.05$ & $0.05$ \\
$p_{x}$ & $200$ & $200$ \\
$y_{c}$ & $0.5$ & $0.5$ \\
$\sigma_{y}$ & $0.05$ & $0.20$ \\
$p_{y}$ & $0$ & $0$ \\
$v_{0}$ & $0$ & $1 \times10^{10}$ \\
\hline
\end{tabular}
\caption{Wall setup.}
\label{tab:sim_setup}
\end{table}
\subsection{Tools}\label{ssec:tools} %
The double-slit experiment is implemented in C++. We use the Python library
\verb|matplotlib| \cite{hunter:2007:matplotlib} to produce all the plots, and
\verb|seaborn| \cite{waskom:2021:seaborn} to set the theme in the figures.
\end{document}

View File

@ -2,6 +2,17 @@
\begin{document}
\section{Results}\label{sec:results}
\subsection{Deviation}\label{ssec:deviation}
% Problem 3: Discuss approaches to solve Au^{n+1} = b, dealing with sparse matrix...
% Problem 7: Consequenses of solver choice, in regards to accuracy of probability conserved
% Add plot of deviation for both single- and double-slit
\subsection{Time evolution}\label{ssec:time_evolution}
% Problem 8: Colormap, include plot of both Re and Im for different time steps
% Account for color scale
\subsection{Particle detection}\label{ssec:particle_detection}
% Problem 9: Plot detection probability for single-, double- and triple-slit
\end{document}