tprak

Theoretical Physics Practical Training
git clone git://popovic.xyz/tprak.git
Log | Files | Refs

main.tex (21449B)


      1 \documentclass[a4paper]{article}
      2 
      3 \usepackage[T1]{fontenc}
      4 \usepackage[utf8]{inputenc}
      5 
      6 \usepackage{mathptmx}
      7 
      8 \usepackage[a4paper, total={6in, 8in}]{geometry}
      9 \usepackage{subcaption}
     10 \usepackage[shortlabels]{enumitem}
     11 \usepackage{amssymb}
     12 \usepackage{amsthm}
     13 \usepackage{mathtools}
     14 \usepackage{braket}
     15 \usepackage{bbm}
     16 \usepackage{graphicx}
     17 \usepackage{float}
     18 \usepackage[colorlinks=true,naturalnames=true,plainpages=false,pdfpagelabels=true]{hyperref}
     19 \usepackage[parfill]{parskip}
     20 \usepackage[backend=biber, sorting=none]{biblatex}
     21 \addbibresource{uni.bib}
     22 \pagestyle{myheadings}
     23 \markright{Popovic, Vogel\hfill Unbiased Fitting \hfill}
     24 
     25 \title{University of Vienna\\ Faculty of Physics\\ \vspace{1.25cm} Theoretical Physics Lab-Course 2021S\\
     26 Quantum Entanglement at High Energies}
     27 \author{Milutin Popovic \& Tim Vogel \vspace{1cm}\\ Supervisor: Prof. Dr. Beatrix C.
     28 Hiesmayr}
     29 \date{May the 9th, 2021}
     30 
     31 \begin{document}
     32 \maketitle
     33 \noindent\rule[0.5ex]{\linewidth}{1pt}
     34 \begin{abstract}
     35 In this article we go through the lecture notes of Prof. Dr. Beatrix C.
     36 Hiesmayr and the exercises given to us in the Theoretical Physics Lab-Course
     37 in the summer semester of  2021. Precisely  we look at entangled systems at high
     38 energies, we learn about the Bells inequality and it's violation within the world
     39 of quantum mechanics.
     40 Furthermore we learn about the density matrix approach in quantum mechanics,
     41 which allows us to look at the unstable particles with different decay rates
     42 e.g. the K-meson.
     43 \end{abstract}
     44 \noindent\rule[0.5ex]{\linewidth}{1pt}
     45 
     46 \tableofcontents
     47 
     48 \section{Bell Theorem and Bell Inequality}
     49 The Enstein-Podolsky-Rosen paradox argues that, because of hidden variables
     50 not considered, quantum mechanics is incomplete. J.S. Bell in his paper \cite{bell},
     51 published in 1964 discovered what is known as the Bell's Theorem. The theorem
     52 states in short that "In certain experiments all local realistic theories are
     53 incompatible with quantum mechanics" \cite{bell}. This is achieved by
     54 establishing a so called Bell inequality which satisfies the local realistic
     55 theories but violates quantum mechanics.
     56 
     57 To demonstrate the results of Bell, we consider a Wiegner-type Bell inequality
     58 for spin-$\frac{1}{2}$ particles with a source that generates a Bell state
     59 \begin{align}\label{eq:wtb}
     60     P(\Uparrow \vec{a}, \Uparrow \vec{b}) \leq
     61     P(\Uparrow \vec{a}, \Uparrow \vec{c}) + P(\Uparrow \vec{c}, \Uparrow \vec{b})
     62 \end{align}
     63 
     64 here we denote the joint probability, of say Alice and Bob to find their
     65 particles in the spin-up state, with respect to their orientation $\vec{a},
     66 \vec{b}$, $P(\Uparrow \vec{a}, \Uparrow \vec{b})$. For the source we consider
     67 an antisymmetric Bell state $|\psi ^-\rangle$ in terms of $\vec{a}$.
     68 \begin{align}
     69     |\psi ^-\rangle = \frac{1}{\sqrt{2}}(|\Uparrow \vec{a}; \Downarrow
     70     \vec{a}\rangle - |\Downarrow \vec{a}; \Uparrow \vec{a}\rangle).
     71 \end{align}
     72 Quantum Mechanics tells us how to calculate the probability
     73 \begin{align}\label{eq:prob}
     74     P^{QM}(\Uparrow \vec{a}, \Uparrow \vec{b}) = \left\|\langle\Uparrow \vec{a}; \Uparrow
     75     \vec{b}|\psi ^-\rangle \right\|^2.
     76 \end{align}
     77 
     78 Furthermore quantum mechanics tells us how we can represent the ket
     79 $|\Uparrow \vec{b}\rangle$ in terms $|\Uparrow \vec{a}\rangle$ and the other
     80 way around.
     81 \begin{align}
     82     |\Uparrow \vec{b}\rangle &= \cos(\frac{\phi_{ab}}{2})|\Uparrow
     83     \vec{a}\rangle + e^{i\theta_{ab}} \sin(\frac{\phi_{ab}}{2})|\Downarrow
     84     \vec{a}\rangle\\
     85     |\Uparrow \vec{a}\rangle &= \sin(\frac{\phi_{ab}}{2})|\Uparrow
     86     \vec{b}\rangle - e^{i\theta_{ab}} \cos(\frac{\phi_{ab}}{2})|\Downarrow
     87     \vec{b}\rangle
     88 \end{align}
     89 
     90 where $\theta_{ab}$ is an unphysical phase in this case and is set to zero.
     91 
     92 With this information we can derive the following results for the
     93 probability in Equation \ref{eq:prob} (Task 1)
     94 \begin{align}
     95     P^{QM}(\Uparrow \vec{a}, \Uparrow \vec{b}) &= \left\|\langle\Uparrow \vec{a}; \Uparrow
     96     \vec{b}|\psi ^-\rangle \right\|^2 \\
     97     &=\frac{1}{2} \left\|\langle\Uparrow \vec{a}; \Uparrow \vec{b}|\Uparrow
     98     \vec{a}; \Downarrow \vec{a}\rangle - \langle \Uparrow \vec{a}; \Uparrow
     99     \vec{b}| \Downarrow \vec{a}; \Uparrow \vec{a}\rangle\right\|^2 \\
    100     &= \frac{1}{2}\left\|(\langle\Uparrow \vec{a}|\Uparrow\vec{a}\rangle)
    101     (\langle\Uparrow \vec{b}| \Downarrow \vec{a}\rangle) - (\underbrace{\langle\Uparrow
    102     \vec{a}| \Downarrow \vec{a}\rangle}_{\bot})(\langle\Uparrow \vec{b}|
    103     \Uparrow\vec{a}\rangle)\right\|^2 \\
    104     &=\frac{1}{2}\left\|\langle\Uparrow\vec{b}|\Downarrow\vec{a}\rangle\right\|^2
    105     =\frac{1}{2}\sin^2(\frac{\phi_{ab}}{2})
    106 \end{align}
    107 
    108 For the remaining probabilities in $|\Uparrow \vec{a}\Downarrow \vec{b}\rangle$,$|\Downarrow
    109 \vec{a}\Uparrow\vec{b}\rangle$ and $|\Downarrow\vec{a}\Downarrow\vec{b} \rangle$
    110 we apply the same procedure to get
    111 \begin{align}
    112     P^{QM}(\Uparrow \vec{a}, \Downarrow \vec{b}) &= \left\|\langle\Uparrow
    113     \vec{a}; \Downarrow \vec{b}|\psi^- \rangle\right\|^2 \\
    114     &=\frac{1}{2}\left\|\langle\Downarrow \vec{b}|\Downarrow
    115     \vec{a}\rangle\right\|^2 = \frac{1}{2}\cos^2(\frac{\phi_{ab}}{2})\\
    116     \\
    117     P^{QM}(\Downarrow \vec{a}, \Downarrow \vec{b}) &= \left\|\langle\Downarrow
    118     \vec{a}; \Downarrow \vec{b}|\psi^- \rangle\right\|^2 \\
    119     &=\frac{1}{2}\left\|\langle\Downarrow \vec{b}|\Uparrow
    120     \vec{a}\rangle\right\|^2 = \frac{1}{2}\sin^2(\frac{\phi_{ab}}{2})
    121 \end{align}
    122 
    123 Thus for the Bell inequality we get
    124 
    125 \begin{align}
    126     \frac{1}{2}sin^2(\frac{\phi_{ab}}{2}) \leq
    127     \frac{1}{2}sin^2(\frac{\phi_{ac}}{2})+
    128     \frac{1}{2}sin^2(\frac{\phi_{cb}}{2})
    129 \end{align}
    130 We can take $\vec{a}$, $\vec{b}$ to be in the same plane in $\mathbb{R}^3$ and
    131 $\vec{c}$ ($\in \mathbb{R}^2$) to be the bisector of these two. This would imply
    132 $\phi_{ac}=\phi_{cb}= \phi$ and $\phi_{ab}=2\phi$, which gives us.
    133 \begin{align}
    134     \sin^2(\phi) &\leq 2\sin^2(\frac{\phi}{2})\\
    135     \Leftrightarrow \cos(\frac{\phi}{2}) &\leq \frac{1}{2}
    136 \end{align}
    137 This is not always the case and is thereby an obvious contradiction.
    138 
    139 Now we will derive the original inequality of John Stuart Bell\cite{bell}.
    140 For this we consider the expectation values. The expectation value $E$ is expressed
    141 in terms of probabilities of all possible outcomes.
    142 \begin{align}
    143     E(\vec{a}, \vec{b}) &= P(\Uparrow \vec{a}, \Uparrow \vec{b}) + P(\Downarrow
    144     \vec{a}, \Downarrow \vec{b}) -P(\Uparrow \vec{a}, \Downarrow \vec{b})
    145     -P(\Downarrow \vec{a}, \Uparrow \vec{b})\\
    146     E(\vec{a}, \vec{c}) &= P(\Uparrow \vec{a}, \Uparrow \vec{c}) + P(\Downarrow
    147     \vec{a}, \Downarrow \vec{c}) -P(\Uparrow \vec{a}, \Downarrow \vec{c})
    148     -P(\Downarrow \vec{a}, \Uparrow \vec{c})\\
    149     E(\vec{c}, \vec{b}) &= P(\Uparrow \vec{c}, \Uparrow \vec{b}) + P(\Downarrow
    150     \vec{c}, \Downarrow \vec{b}) -P(\Uparrow \vec{c}, \Downarrow \vec{b})
    151     -P(\Downarrow \vec{c}, \Uparrow \vec{b})
    152 \end{align}
    153 The sum of spin-up/spin-down and spin-down/spin-up probabilities needs to be
    154 one, with this we get.
    155 \begin{align}
    156     P(\Uparrow \vec{a}, \Uparrow \vec{b}) = E(\vec{a}, \vec{b}) - P(\Downarrow
    157     \vec{a}, \Downarrow \vec{b}) + 1\\
    158     P(\Uparrow \vec{a}, \Uparrow \vec{c}) = E(\vec{a}, \vec{c}) - P(\Downarrow
    159     \vec{a}, \Downarrow \vec{c}) + 1\\
    160     P(\Uparrow \vec{c}, \Uparrow \vec{b}) = E(\vec{c}, \vec{b}) - P(\Downarrow
    161     \vec{c}, \Downarrow \vec{b}) + 1
    162 \end{align}
    163 Plugging this into the inequality in equation \ref{eq:wtb} we get
    164 \begin{align}
    165     E(\vec{a}, \vec{b}) - E(\vec{a}, \vec{c}) &\leq E(\vec{c}, \vec{b}) + 1\\
    166     & + P(\Downarrow \vec{a}, \Downarrow \vec{b}) -P(\Downarrow \vec{a},
    167     \Downarrow \vec{c}) -P(\Downarrow \vec{c}, \Downarrow \vec{b})
    168     \label{eq:rightside}
    169 \end{align}
    170 The term in equation \ref{eq:rightside} satisfies the bell inequality if the
    171 term on the left is strictly positive. Thus we get the original Bell inequality
    172 \begin{align}
    173     |E(\vec{a}, \vec{b}) - E(\vec{a}, \vec{c})| \leq E(\vec{c}, \vec{b}) + 1
    174 \end{align}
    175 \section{Quantum System of K-mesons}
    176 Kaons were discovered in 1947 by Rochester and Butler, where cosmic ray
    177 particles hit a lead plate and produced a neutral particle. The neutral
    178 particle is named kaon, it was noticed through its decay in two charged
    179 pions. $K^0$, the kaon, is the first strange particle.
    180 The neutral kaons are states of quarks and anti-quarks, the ket
    181 $|K^0\rangle = |d\bar{s}\rangle$ denotes the neutral K-meson, and its
    182 antiparticle is $|\bar{K}^0\rangle = |\bar{d}s\rangle$. The state of the kaon
    183 is entangled in the sense that the particle $K^0$ can turn into an antiparticle
    184 $\bar{K}^0$ before decay through strangeness oscillation, a particle-antiparticle oscillation.
    185 
    186 
    187 There are two states, the long lived kaon $K_L$ and the short lived kaon $K_S$
    188 which diagonalize the Hamiltonian,
    189 the difference of their decay rates is about a factor of $600$.
    190 The thing is $K_L$ should decay into three pions, while $K_S$ should decay
    191 into two pions. But in 1964 Cronin and Fitch found out that the long lived
    192 kaon can also decay into two pions, which directly implies a $CP$ symmetry violation.
    193 $CP$ symmetry (charge conjugation parity symmetry) in particle physics states
    194 that physics is the same if instead of looking at an particle we looked at its
    195 antiparticle.
    196 
    197 The short lived kaon $K_S$ and the long lived kaon $K_L$ can be expressed in
    198 terms of $|K^0\rangle$ and $|\bar{K^0}\rangle$ as follows,
    199 
    200 \begin{align}
    201     |K_S\rangle &=  \frac{1}{N}(p|K^0\rangle - q|\bar{K}^0\rangle) \\
    202     |K_L\rangle &=  \frac{1}{N}(p|K^0\rangle + q|\bar{K}^0\rangle)
    203 \end{align}
    204 where $p = 1+\varepsilon$, $q=1-\varepsilon$ and $N^2 = |p|^2 + |q|^2$,
    205 $\varepsilon$ is called the $CP$ violating parameter and can be measured, with
    206 a magnitude of $|\varepsilon| \approx 10^{-3}$ \cite{Bertlmann} . Note
    207 that these two states are NOT orthogonal due to the $CP$ violation.
    208 
    209 
    210 Furthermore the connection to the $CP$ basis is,
    211 \begin{align}
    212     |K_1\rangle &=  \frac{1}{\sqrt{2}}(|K^0\rangle - e^{i\alpha}|\bar{K}^0\rangle) \\
    213     |K_2\rangle &=  \frac{1}{\sqrt{2}}(|K^0\rangle + e^{i\alpha}|\bar{K}^0\rangle) \\
    214 \end{align}
    215 here $\alpha$ is an unphysical phase and is conventionally set to zero.
    216 
    217 
    218 \subsection{CP-Symmetry violation}
    219 The violation in symmetry can be seen considering the Bell inequality
    220 \begin{align}\label{eq:leqk}
    221     P(K_S, \bar{K}^0) \leq P(K_S, K_1) + P(K_1, \bar{K}^0)
    222 \end{align}
    223 where $P(K_S, \bar{K}^0)$ is the probability of finding $K_S$ on the right and
    224 $\bar{K}^0$ on the left at the time of measurement at $t=0$, $K_1$ denotes a
    225 completely unphysical state and cannot be measured. Quantum mechanics gives us
    226 the tools to calculate these probabilities, the calculation gives
    227 \begin{align}
    228     |p|&\leq |q|\\
    229     \Leftrightarrow \delta &:= \frac{|p|^2 - |q|^2}{|p|^2+|q|^2} \leq 0
    230 \end{align}
    231 
    232 Changing the choice of $\bar{K}^0$ and $K^0$ in equation \ref{eq:leqk} we
    233 calculate the probabilities again:
    234 \begin{align}
    235     P^{QM}(K_S, K^0)&=\left\|\langle K_S, K^0|\psi ^-\rangle\right\|
    236     ^2=\frac{|q|^2}{4N^2} \\
    237     P^{QM}(K_S, K_1)&= \left\|\langle K_S, K_1|\psi ^-\rangle\right\|^2=
    238     \frac{1}{4N^2}\left|q-pe^{i\alpha}\right|\\
    239     P^{QM}(K_1, K^0)&=\left\|\langle K_1, K^0|\psi ^-\rangle\right\| ^2=
    240     \frac{1}{4}|e^{i\alpha}|^2
    241 \end{align}
    242 Inserting this into the inequality in equation \ref{eq:leqk} and applying some
    243 basic algebra we get
    244 \begin{align}
    245     |q| &\leq |p|\\
    246     \Leftrightarrow \delta &= \frac{|p|^2 - |q|^2}{|p|^2+|q|^2} \geq 0
    247 \end{align}
    248 implying a strict equality
    249 \begin{align}
    250         \delta = 0.
    251 \end{align}
    252 This contradicts the experimental value of $\delta$ (citation from slides).
    253 \begin{align}
    254     \delta_{exp} = (3.27\pm 0.12)\cdot 10^{-1}.
    255 \end{align}
    256 
    257 We can say that the CP-violation is directly related to the violation of the
    258 Bell Inequality and entanglement.
    259 If CP-symmetry were true the long lived kaon would decay via two paths in equal
    260 amounts, which is not the case. Also
    261 according to the Big-Bang-Theory the amount
    262 of matter and antimatter initially created is equal, but according to
    263 experimental results CP-asymmetry means that there is an imbalance in
    264 matter and that physics differs for particles and antiparticles.
    265 
    266 \subsection{Efficient description of decaying quantum systems}
    267 The neutral kaon system, is usually described by an effective Schrödinger-equation,
    268 which is given by the Lioville von Neumann form as:
    269 \begin{equation}
    270     \frac{d}{dt}\rho=-iH_{eff}\rho+i\rho H^\dagger_{eff}
    271 \end{equation}
    272 where $\rho$ is a 2x2 density matrix and $H_{eff}$ non hermitian. The Hamiltonian can
    273 be decomposed via the Wigner-Weisskopf approximation into: $H_{eff}=M-\frac{i}{2}\Gamma$,
    274 with the 2x2 mass matrix $H$ and the 2x2 decay matrix $\Gamma$ both being positive and hermitian.
    275 What we are now concerned with, is the question, whether $H^\dagger_{eff}$
    276 can also be decomposed and the implications the result gives.
    277 Starting from the decomposed $H_{eff}$, applying the dagger, we get:
    278 \begin{equation}
    279     H^\dagger_{eff}=M^T+\frac{i}{2}\Gamma^T
    280 \end{equation}
    281 which leads to:
    282 \begin{equation}
    283     H^\dagger_{eff}=\left( \begin{array}{cc}
    284         M_0+\frac{i}{2}\Gamma_0 & (M_{12})^*+\frac{i}{2}(\Gamma_{12})^* \\
    285         (M_{12})+\frac{i}{2}(\Gamma_{12})^* & M_0+\frac{i}{2}\Gamma_0
    286     \end{array}\right)
    287 \end{equation}
    288 This final matrix can now be brought into the form:
    289 \begin{equation}
    290     \begin{pmatrix}
    291         A^* & B^*r \\
    292         \frac{B^*}{r} & A^*
    293     \end{pmatrix}
    294 \end{equation}
    295 with $A,B and r$ being complex numbers.
    296 \newline
    297 We now compute the Eigenvalues of this matrix, giving:
    298 \begin{equation}
    299     (A^*-\lambda)^2-(B^*)^2 \rightarrow \lambda=A^*\pm B^*
    300 \end{equation}
    301 With this, the Eigenvectors take the form:
    302 \begin{equation}
    303     v_1=\begin{pmatrix}
    304         r^* \\
    305         1
    306     \end{pmatrix}
    307 \end{equation}
    308 and
    309 \begin{equation}
    310     v_2=\begin{pmatrix}
    311         -r^* \\
    312         1
    313     \end{pmatrix}
    314 \end{equation}
    315 We now define the matrices:
    316 \begin{equation}
    317     R^{-1}=
    318   \begin{pmatrix}
    319         r^* & -r^* \\
    320         1 & 1
    321     \end{pmatrix}
    322 \end{equation}
    323 and
    324 \begin{equation}
    325     R= \frac{1}{2r}
    326     \begin{pmatrix}
    327         1 & r^* \\
    328         -1 & r^*
    329     \end{pmatrix}
    330 \end{equation}
    331 and with
    332 \begin{equation}
    333     RH^\dagger_{eff}R^{-1}=
    334     \begin{pmatrix}
    335         A^*+B^* & 0 \\
    336         0 & A^*-B^*
    337     \end{pmatrix}
    338 \end{equation}
    339 we finally find values for $\ket{K_S}$ and $\ket{K_L}$ which explicitly are:
    340 \begin{equation}
    341     \ket{K_S}=\frac{1}{\sqrt{1+|r|^2}}(-r^*\ket{K^0}+\ket{\bar{K^0}})
    342 \end{equation}
    343 and
    344 \begin{equation}
    345     \ket{K_L}=\frac{1}{1+|r|^2}(r^*\ket{K^0}-\ket{\bar{K^0}}
    346 \end{equation}
    347 We now want to calculate the overlap of these eigenvectors. We start with the
    348 overlap in $H_{eff}$, which is given by:
    349 \begin{equation}
    350     \braket{K_S|K_L}=\frac{1-|r|^2}{1+|r|^2}
    351 \end{equation}
    352 We now define $\varepsilon=\frac{1-r}{1+r}\rightarrow r=\frac{1-\varepsilon}{1+\varepsilon}$, which leads to:
    353 \begin{equation}
    354     \braket{K_S|K_L}=\frac{|\varepsilon+1|^2-|\varepsilon-1|^2}{|\varepsilon+1|^2+|\varepsilon-1|^2}
    355 \end{equation}
    356 And with only considering the real part of Epsilon, we finally obtain:
    357 \begin{equation}
    358     \braket{K_S|K_L}=\frac{2Re(\varepsilon}{|\varepsilon|^2+1})
    359 \end{equation}
    360 We now do the same for the overlap in $H^\dagger_{eff}$, and thereby obtain the result:
    361 \begin{equation}
    362     \braket{K_S|K_L}=\frac{1-|r|^2}{1+|r|^2}=\frac{2Re(\varepsilon)}{1+|\varepsilon|^2}
    363 \end{equation}
    364 This means, that in both cases we obtain a CP-violation.
    365 \subsection{Charge asymmetry}
    366 Finally, we look at the following charge asymmetry term, given by:
    367 \begin{equation}
    368     \delta(t)=\frac{P(K^0,t;|K^0|)-P(\bar{K^0},t;|K^0|)}{P(K^0,t;|K^0|)+P(\bar{K^0},t;|K^0|)}
    369 \end{equation}
    370 Fist we define the following ket:
    371 \begin{equation}
    372     \ket{K^0(t)}=\frac{\sqrt{1+|\varepsilon}|^2}{\sqrt{2}(1+\varepsilon)}(\exp{(-i\lambda_st)}\ket{K_S}+\exp{(-i\lambda_Lt)}\ket{K_L}
    373 \end{equation}
    374 with: $\lambda_{S/L}=m{_S/L}-\frac{i}{2}\Gamma_{S/L}$. We now compute the probabilities $P(\bar{K^0},t;|K^0|)$ and $P(K^0,t;|K^0|)$, and obtain:
    375 \begin{equation}
    376     P(K^0,t;|K^0|)=\frac{1}{2|1+\varepsilon|^2}|e^{-i\lambda_S t}+
    377     \varepsilon e^{-i\lambda_L t}|^2
    378 \end{equation}
    379 and
    380 \begin{equation}
    381     P(\bar{K^0},t;|K^0|)=\frac{1}{2|1-\varepsilon|^2}|
    382     e^{-i\lambda_S t}-\varepsilon e^{-i\lambda_L t}|^2
    383 \end{equation}
    384 Then:
    385 \begin{equation}
    386     \delta(t)=\frac{\frac{1}{|1+\varepsilon|^2}|
    387     e^{-i\lambda_S t}+\varepsilon
    388     e^{-i\lambda_L t}|^2-\frac{1}{|1-\varepsilon|^2}|
    389     e^{-i\lambda_S t}-\varepsilon
    390     e^{-i\lambda_L t}|^2}{\frac{1}{|1+\varepsilon|^2}|
    391     e^{-i\lambda_S t}+\varepsilon
    392     e^{-i\lambda_L t}|^2+\frac{1}{|1-\varepsilon|^2}|
    393     e^{-i\lambda_S t}-\varepsilon
    394     e^{-i\lambda_L t}|^2}
    395 \end{equation}
    396 and with the limit of $t\rightarrow 0$ we obtain the result:
    397 \begin{equation}
    398     \delta(t)=\frac{|1-\varepsilon|^2-|1+\varepsilon|^2}{|1+\varepsilon|^2+|1-\varepsilon|^2}
    399 \end{equation}
    400 And by expansion of the leading order in $Re(\varepsilon)$, finally:
    401 \begin{equation}
    402     \frac{|1-\varepsilon|^2-|1+\varepsilon|^2}{|1+\varepsilon|^2+|1-\varepsilon|^2}=2Re(\varepsilon)+\mathcal{O}(\varepsilon^3)
    403 \end{equation}
    404 \subsection{Density Matrix Approach Time Evolution}
    405 In this section we describe an open quantum system with unstable particles
    406 (e.g. K-mesons) with the Lindbad-Gorini-Kossakowsky-Sudarhasanan master equation,
    407 an density matrix approach, by enlarging the Hilbertspace\cite{bgh}. With this
    408 larger Hilbertspace $\textbf{H}_{tot} = \textbf{H}_s \oplus \textbf{H}_f$ we take
    409 into consideration both the "surviving"($\textbf{H}_s$) and the "decaying" or
    410 "final" ($\textbf{H}_f$)
    411 states and thus get a positive time evolution described by a non-hermitian
    412 Hamiltonian $H_{eff}$ and a dissipator $\mathcal{D}$ of the Lindbad operator
    413 $L$. The time evolution of the density matrix $\varrho \in \mathbf{H}_{tot}$ is given by the master
    414 equation in the Lindbad form
    415 \begin{align}\label{eq:master}
    416     \frac{d\varrho}{dt} &= -[H, \varrho] - \mathcal{D}[\varrho]\\
    417     \text{with}\;\;\;\;    \mathcal{D}[\varrho] &= \frac{1}{2} \sum_{j=0} (L^{\dagger}_j L_j \varrho + \varrho
    418     L^{\dagger}_j L_j - L_j \varrho L^{\dagger}_j)
    419 \end{align}
    420 where the density matrix $\varrho$ is a $4x4$ matrix with components
    421 $\varrho_{ij}$ ($i,j = s,f$) which are $2x2$ matrices, with the property
    422 $\varrho^\dagger_{sf} = \varrho_{fs}$
    423 \begin{align}
    424     \varrho =
    425     \begin{pmatrix}
    426         \varrho_{ss} &  \varrho_{sf} \\
    427         \varrho_{fs} &  \varrho_{ff}
    428     \end{pmatrix}.
    429 \end{align}
    430 The Hamiltonian $H$ is an extension of the effective Hamiltonian $H_{eff}$ on
    431 the total Hilbertspace $\textbf{H}_{tot}$
    432 \begin{align}
    433     H =
    434     \begin{pmatrix}
    435         H_{eff} &  0 \\
    436         0 &  0
    437     \end{pmatrix}.
    438 \end{align}
    439 Furthermore the Lindbad generator $L_0$ is defined with
    440 $B:\textbf{H}_s \rightarrow \textbf{H}_f$, where $B^\dagger B = \Gamma$, decay
    441 matrix $\Gamma$ from the effective Hamiltonian $H_{eff}$,
    442 \begin{align}
    443     L_0 =
    444     \begin{pmatrix}
    445         0 &  0 \\
    446         B &  0
    447     \end{pmatrix} \;\;\;\;
    448     L_j =
    449     \begin{pmatrix}
    450         A_j &  0 \\
    451         0 &  0
    452     \end{pmatrix} \;\;\;\;\;  (\text{with}\; j > 0).
    453 \end{align}
    454 
    455 Rewriting the master equation in \ref{eq:master} we get the following
    456 differential equations for the density matrix components
    457 \begin{align}
    458     \dot{\varrho}_{ss} &= -i[H_{eff},\varrho{ss}] - \frac{1}{2}\{B^\dagger
    459     B,\varrho_{ss} \} - \tilde{D}[\varrho_{ss}],\\
    460     \dot{\varrho}_{sf} &= -iH_{eff}\varrho_{sf} - \frac{1}{2} B^\dagger B \varrho_{sf}
    461     -\frac{1}{2}\sum_j A_j^\dagger A_j \varrho_{sf},\\
    462     \dot{\varrho}_{ff} &=B\varrho_{ss}B^\dagger .
    463 \end{align}
    464 with $\tilde{D}[\varrho_{ss}] = \frac{1}{2} \sum_{j=0} (A^{\dagger}_j A_j
    465 \varrho_{ss} + \varrho_{ss}
    466     A^{\dagger}_j A_j - A_j \varrho_{ss} A^{\dagger}_j)$.\newline
    467 
    468 Now we solve these equations for the case without decoherence, meaning the
    469 Lindbad Operators operators $A_j$ disappear and we can rewrite the equations
    470 for $\varrho_{ss}$ above in
    471 \begin{align}
    472     \dot{\varrho_{ss}} &= -[H_{eff}, \varrho_{ss}] - \frac{1}{2} \{\Gamma,
    473     \varrho_{ss}\}=\\
    474     &=-i((M-\frac{i}{2}\Gamma)\varrho_{ss} - \varrho_{ss}(M-\frac{i}{2}\Gamma))
    475     -\frac{1}{2}(\Gamma \varrho_{ss} + \varrho_{ss}\Gamma)\\
    476     &= -i\underbrace{[M, \varrho_{ss}]}_{=0} - \varrho_{ss} \Gamma\\
    477     &= -\varrho_{ss}\Gamma \\
    478     \Rightarrow \;\;\; \varrho_{ss} &= \varrho_{ss}(0) e^{-\Gamma t}.
    479 \end{align}
    480 For $\varrho_{sf}$ we get
    481 \begin{align}
    482     \dot{\varrho}_{sf} &= -i H_{eff} \varrho_{sf} - \frac{1}{2} \Gamma
    483     \varrho_{sf} =\\
    484     &= -iM\varrho_{sf}\\
    485     \Rightarrow \;\;\; \varrho_{sf} &= \varrho_{sf}(0) e^{-iM t}.
    486 \end{align}
    487 And for $\varrho_{ff}$
    488 \begin{align}
    489     \dot{\varrho}_{ff} &= B\varrho_{ss}B^\dagger \\
    490     \Rightarrow \;\;\; \varrho_{ff}&= B\int \varrho_{ss}dt B^\dagger \\
    491     &= -B\varrho_{ss}(0) \Gamma^{-1} e^{-\Gamma t} B^\dagger + \varrho_{ff}(0)
    492 \end{align}
    493 
    494 In reality the decay rates of particles differ e.g. $K_S$ and $K_L$, the
    495 density matrix allows such things to be taken care of by mathematically
    496 extending the Hilbertspace and including the Lindbad operator. We could also
    497 consider a particle with three different decay rates, though the Hamiltonian
    498 would be a nine dimensional. In this regard we might say that the master
    499 equation \ref{eq:master} is a more general Schrödinger equation,
    500 because it not only describes pure quantum states but
    501 also mixed states.
    502 
    503 \nocite{carla}
    504 \nocite{bgh}
    505 \nocite{mexico}
    506 \printbibliography
    507 \end{document}