Finished chapter 5
This commit is contained in:
parent
022cf18b43
commit
80d4bae49f
@ -280,6 +280,7 @@ correlations itself.
|
|||||||
\mynote{introduce citations from PRX/PRA above}
|
\mynote{introduce citations from PRX/PRA above}
|
||||||
|
|
||||||
\section{The Master Equation}
|
\section{The Master Equation}
|
||||||
|
\label{sec:master}
|
||||||
|
|
||||||
A quantum trajectory is stochastic in nature, it depends on the exact
|
A quantum trajectory is stochastic in nature, it depends on the exact
|
||||||
timings of the quantum jumps which are determined randomly. This makes
|
timings of the quantum jumps which are determined randomly. This makes
|
||||||
|
BIN
Chapter5/Figs/comp.pdf
Normal file
BIN
Chapter5/Figs/comp.pdf
Normal file
Binary file not shown.
BIN
Chapter5/Figs/figure3.pdf
Normal file
BIN
Chapter5/Figs/figure3.pdf
Normal file
Binary file not shown.
@ -83,7 +83,9 @@ non-Hermitian Hamiltonian thus extending the notion of quantum Zeno
|
|||||||
dynamics into the realm of non-Hermitian quantum mechanics joining the
|
dynamics into the realm of non-Hermitian quantum mechanics joining the
|
||||||
two paradigms.
|
two paradigms.
|
||||||
|
|
||||||
\section{Large-Scale Dynamics due to Weak Measurement}
|
\section{Quantum Measurement Induced Dynamics}
|
||||||
|
|
||||||
|
\subsection{Large-Scale Dynamics due to Weak Measurement}
|
||||||
|
|
||||||
We start by considering the weak measurement limit when photon
|
We start by considering the weak measurement limit when photon
|
||||||
scattering does not occur frequently compared to the tunnelling rate
|
scattering does not occur frequently compared to the tunnelling rate
|
||||||
@ -542,7 +544,7 @@ described by this model, is that the width does in fact stay roughly
|
|||||||
constant. It is only in the later stages when the oscillations reach
|
constant. It is only in the later stages when the oscillations reach
|
||||||
maximal amplitude that the width becomes visibly reduced.
|
maximal amplitude that the width becomes visibly reduced.
|
||||||
|
|
||||||
\section{Three-Way Competition}
|
\subsection{Three-Way Competition}
|
||||||
|
|
||||||
Now it is time to turn on the inter-atomic interactions,
|
Now it is time to turn on the inter-atomic interactions,
|
||||||
$U/J^\mathrm{cl} \ne 0$. As a result the atomic dynamics will change
|
$U/J^\mathrm{cl} \ne 0$. As a result the atomic dynamics will change
|
||||||
@ -787,8 +789,6 @@ into the strong measurement regime in Fig. \ref{fig:squeezing} as the
|
|||||||
$U$-dependence flattens out with increasing measurement strength as
|
$U$-dependence flattens out with increasing measurement strength as
|
||||||
the $\gamma/U \gg 1$ regime is reached.
|
the $\gamma/U \gg 1$ regime is reached.
|
||||||
|
|
||||||
\section{Quantum Zeno Dynamics}
|
|
||||||
|
|
||||||
\subsection{Emergent Long-Range Correlated Tunnelling}
|
\subsection{Emergent Long-Range Correlated Tunnelling}
|
||||||
|
|
||||||
When $\gamma \rightarrow \infty$ the measurement becomes
|
When $\gamma \rightarrow \infty$ the measurement becomes
|
||||||
@ -1085,10 +1085,716 @@ which as we prove can successfully compete with other short-range
|
|||||||
processes in many-body systems. This opens promising opportunities for
|
processes in many-body systems. This opens promising opportunities for
|
||||||
future research.
|
future research.
|
||||||
|
|
||||||
\subsection{Non-Hermitian Dynamics in the Quantum Zeno Limit}
|
\section{Non-Hermitian Dynamics in the Quantum Zeno Limit}
|
||||||
|
|
||||||
|
In the previous section we provided a rather high-level analysis of
|
||||||
|
the strong measurement limit in our quantum gas model. We showed that
|
||||||
|
global measurement in the strong, but not projective, limit leads to
|
||||||
|
correlated tunnelling events which can be highly delocalised. Multiple
|
||||||
|
examples for different optical geometries and measurement operators
|
||||||
|
demonstrated the incredible felixbility and potential in engineering
|
||||||
|
dynamics for ultracold gases in an optical lattice. We also claimed
|
||||||
|
that the behaviour of the system is described by the Hamiltonian given
|
||||||
|
in Eq. \eqref{eq:hz}. Having developed a physical and intuitive
|
||||||
|
understanding of the dynamics in the quantum Zeno limit we will now
|
||||||
|
provide a more rigorous, low-level and fundamental understanding of
|
||||||
|
the process.
|
||||||
|
|
||||||
|
\subsection{Suppression of Coherences in the Density Matrix}
|
||||||
|
|
||||||
\subsection{Steady-State of the Non-Hermitian Hamiltonian}
|
At this point we deviate from the quantum trajectory approach and we
|
||||||
|
resort to a master equation as introduced in section
|
||||||
|
\ref{sec:master}. We do this, because we have seen that the emergent
|
||||||
|
long-range correlated tunnelling is a feature of all trajectories and
|
||||||
|
mostly depends on the geometry of the measurement. Therefore, a
|
||||||
|
general approach starting from an unconditioned state should be able
|
||||||
|
to reveal these features. However, we will later make use of the fact
|
||||||
|
that we are in possession of a measurement record and obtain a
|
||||||
|
conditioned state. Furthermore, we first consider the most general
|
||||||
|
case of an open system subject to a quantum measurement and only limit
|
||||||
|
ourselves to the quantum gas model later on. This demonstrates that
|
||||||
|
the dynamics we observed in the previous section are a feature of
|
||||||
|
measurement rather than our specific model.
|
||||||
|
|
||||||
\section{Conclusions}
|
As introduced in section \ref{sec:master} we consider a state
|
||||||
|
described by the density matrix $\hat{\rho}$ whose isolated behaviour
|
||||||
|
is described by the Hamiltonian $\H_0$ and when measured the jump
|
||||||
|
operator $\c$ is applied to the state at each detection
|
||||||
|
\cite{MeasurementControl}. The master equation describing its time
|
||||||
|
evolution when we ignore the measurement outcomes is given by
|
||||||
|
\begin{equation}
|
||||||
|
\dot{\hat{\rho}} = -i [ \H_0 , \hat{\rho} ] + \c \hat{\rho} \cd - \frac{1}{2}(
|
||||||
|
\cd\c \hat{\rho} + \hat{\rho} \cd\c ).
|
||||||
|
\end{equation}
|
||||||
|
We also define $\c = \lambda \op$ and $\H_0 = \nu \h$. The exact
|
||||||
|
definition of $\lambda$ and $\nu$ is not so important as long as these
|
||||||
|
coefficients can be considered to be some measure of the relative size
|
||||||
|
of these operators. They would have to be determined on a case-by-case
|
||||||
|
basis, because the operators $\c$ and $\H_0$ may be unbounded. If
|
||||||
|
these operators are bounded, one can simply define them such that
|
||||||
|
$||\op|| \sim O(1)$ and $||\h|| \sim O(1)$. If they are unbounded, one
|
||||||
|
possible approach would be to identify the relevant subspace of which
|
||||||
|
dynamics we are interested in and scale the operators such that the
|
||||||
|
eigenvalues of $\op$ and $\h$ in this subspace are $\sim O(1)$.
|
||||||
|
|
||||||
|
We will once again use projectors $P_m$ which have no effect on states
|
||||||
|
within a degenerate subspace of $\c$ ($\op$) with eigenvalue $c_m$
|
||||||
|
($o_m$), but annihilate everything else. For convenience we will also
|
||||||
|
use the following definition $\hat{\rho}_{mn} = P_m \hat{\rho} P_n$.
|
||||||
|
Note that these are submatrices of the density matrix, which in
|
||||||
|
general are not single matrix elements. Therefore, we can write the
|
||||||
|
master equation that describes this open system as a set of equations
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:master}
|
||||||
|
\dot{\hat{\rho}}_{mn} = -i K P_m \left[ \h \sum_r \hat{\rho}_{rn}
|
||||||
|
- \sum_r \hat{\rho}_{mr} \h \right] P_n + \lambda^2 \left[ o_m
|
||||||
|
o_n^* - \frac{1}{2} \left( |o_m|^2 + |o_n|^2 \right) \right] \hat{\rho}_{mn},
|
||||||
|
\end{equation}
|
||||||
|
where the first term describes coherent evolution whereas the second
|
||||||
|
term causes dissipation.
|
||||||
|
|
||||||
|
First, note that for the density submatrices for which $m = n$,
|
||||||
|
$\hat{\rho}_{mm}$, the dissipative term vanishes. This means that
|
||||||
|
these submatrices are subject to coherent evolution only and do not
|
||||||
|
experience losses and they are thus decoherence free subspaces. It is
|
||||||
|
crucial to note that these submatrices are simply the density matrices
|
||||||
|
of the individual degenerate Zeno subspaces. Interestingly, any state
|
||||||
|
that consists only of these decoherence free subspaces, i.e.~
|
||||||
|
$\hat{\rho} = \sum_m \hat{\rho}_{mm}$, and that commutes with the
|
||||||
|
Hamiltonian, $[\hat{\rho}, \hat{H}_0] = 0$, will be a steady state.
|
||||||
|
This can be seen by substituting this ansatz into
|
||||||
|
Eq. \eqref{eq:master} which yields $\dot{\hat{\rho}}_{mn} = 0$ for all
|
||||||
|
$m$ and $n$. These states can be prepared dissipatively using known
|
||||||
|
techniques \cite{diehl2008}, but it is not required that the state be
|
||||||
|
a dark state of the dissipative operator as is usually the case.
|
||||||
|
|
||||||
|
Second, we consider a large detection rate, $\lambda^2 \gg \nu$, for
|
||||||
|
which the coherences, i.e.~ the density submatrices $\hat{\rho}_{mn}$
|
||||||
|
for which $m \ne n$, will be heavily suppressed by dissipation. We can
|
||||||
|
adiabatically eliminate these cross-terms by setting
|
||||||
|
$\dot{\hat{\rho}}_{mn} = 0$, to get
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:intermediate}
|
||||||
|
\hat{\rho}_{mn} = \frac{\nu}{\lambda^2} \frac{i P_m \left[ \h \sum_r \hat{\rho}_{rn} - \sum_r \hat{\rho}_{mr} \h \right] P_n } {o_m o_n^* - \frac{1}{2} \left( |o_m|^2 + |o_n|^2 \right)}
|
||||||
|
\end{equation}
|
||||||
|
which tells us that they are of order $\nu/\lambda^2 \ll
|
||||||
|
1$. Therefore, the resulting density matrix will be given by
|
||||||
|
$\hat{\rho} \approx \sum_m \hat{\rho}_{mm}$ which consists solely of
|
||||||
|
the individual Zeno subspace density matrices. One can easily recover
|
||||||
|
the projective Zeno limit by considering $\lambda \rightarrow \infty$
|
||||||
|
when all the subspaces completely decouple. This is exactly the
|
||||||
|
$\gamma \rightarrow \infty$ limit discussed in the previous
|
||||||
|
section. However, we have seen that it is crucial we only consider,
|
||||||
|
$\lambda^2 \gg \nu$, but not infinite. If the subspaces do not
|
||||||
|
decouple completely, then transitions within a single subspace can
|
||||||
|
occur via other subspaces in a manner similar to Raman transitions. In
|
||||||
|
Raman transitions population is transferred between two states via a
|
||||||
|
third, virtual, state that remains empty throughout the process. By
|
||||||
|
avoiding the infinitely projective Zeno limit we open the option for
|
||||||
|
such processes to happen in our system where transitions within a
|
||||||
|
single Zeno subspace occur via a second, different, Zeno subspace even
|
||||||
|
though the occupation of the intermediate states will remian
|
||||||
|
negligible at all times.
|
||||||
|
|
||||||
|
A single quantum trajectory results in a pure state as opposed to the
|
||||||
|
density matrix and in general, there are many density matrices that
|
||||||
|
have non-zero and non-negligible $m = n$ submatrices,
|
||||||
|
$\hat{\rho}_{mm}$, even when the coherences are small. They correspond
|
||||||
|
to a mixed states containing many Zeno subspaces and it is not clear
|
||||||
|
what the pure states that make up these density matrices are. However,
|
||||||
|
we note that for a single pure state the density matrix can consist of
|
||||||
|
only a single diagonal submatrix $\hat{\rho}_{mm}$. To understand
|
||||||
|
this, consider the state $| \Phi \rangle$ and take it to span exactly
|
||||||
|
two distinct subspaces $P_a$ and $P_b$ ($a \ne b$). This wavefunction
|
||||||
|
can thus be written as
|
||||||
|
$| \Phi \rangle = P_a | \Phi \rangle + P_b | \Phi \rangle$. The
|
||||||
|
corresponding density matrix is given by
|
||||||
|
\begin{equation}
|
||||||
|
\hat{\rho}_\Psi = P_a | \Phi \rangle \langle \Phi | P_a + P_a | \Phi
|
||||||
|
\rangle \langle \Phi | P_b + P_b | \Phi \rangle \langle \Phi | P_a +
|
||||||
|
P_b | \Phi \rangle \langle \Phi | P_b.
|
||||||
|
\end{equation}
|
||||||
|
If the wavefunction has significant components in both subspaces then
|
||||||
|
in general the density matrix will not have negligible coherences,
|
||||||
|
$\hat{\rho}_{ab} = P_a | \Phi \rangle \langle \Phi | P_b$. A density
|
||||||
|
matrix with just diagonal components must be in either subspace $a$,
|
||||||
|
$| \Phi \rangle = P_a | \Phi \rangle$, or in subspace $b$,
|
||||||
|
$| \Phi \rangle = P_b | \Phi \rangle$. Therefore, a density matrix of
|
||||||
|
the form $\hat{\rho} = \sum_m \hat{\rho}_{mm}$ without any cross-terms
|
||||||
|
between different Zeno subspaces can only be composed of pure states
|
||||||
|
that each lie predominantly within a single subspace. However, because
|
||||||
|
we will not be dealing with the projective limit, the wavefunction
|
||||||
|
will in general not be entirely confined to a single Zeno subspace. We
|
||||||
|
have seen that the coherences are of order $\nu/\lambda^2$. This would
|
||||||
|
require the wavefunction components to satisfy
|
||||||
|
$P_a | \Phi \rangle \approx O(1)$ and
|
||||||
|
$P_b | \Phi \rangle \approx O(\nu/\lambda^2)$ (or vice-versa). This in
|
||||||
|
turn implies that the population of the states outside of the dominant
|
||||||
|
subspace (and thus the submatrix $\hat{\rho}_{bb}$) will be of order
|
||||||
|
$\langle \Phi | P_b^2 | \Phi \rangle \approx
|
||||||
|
O(\nu^2/\lambda^4)$. Therefore, these pure states, even though they
|
||||||
|
span multiple Zeno subspaces, cannot exist in a meaningful coherent
|
||||||
|
superposition in this limit. This means that a density matrix that
|
||||||
|
spans multiple Zeno subspaces has only classical uncertainty about
|
||||||
|
which subspace is currently occupied as opposed to the uncertainty due
|
||||||
|
to a quantum superposition. This is anlogous to the simple qubit
|
||||||
|
example we considered in section \ref{sec:master}.
|
||||||
|
|
||||||
|
\subsection{Quantum Measurement vs. Dissipation}
|
||||||
|
|
||||||
|
This is where quantum measurement deviates from dissipation. If we
|
||||||
|
have access to a measurement record we can infer which Zeno subspace
|
||||||
|
is occupied, because we know that only one of them can be occupied at
|
||||||
|
any time. We have seen that since the density matrix cross-terms are
|
||||||
|
small we know \emph{a priori} that the individual wavefunctions
|
||||||
|
comprising the density matrix mixture will not be coherent
|
||||||
|
superpositions of different Zeno subspaces and thus we only have
|
||||||
|
classical uncertainty which means we can resort to clasical
|
||||||
|
probability methods. Each individual experiment will at any time be
|
||||||
|
predominantly in a single Zeno subspace with small cross-terms and
|
||||||
|
negligible occupations in the other subspaces. With no measurement
|
||||||
|
record our density matrix would be a mixture of all these
|
||||||
|
possibilities. We can try and determine the Zeno subspace around which
|
||||||
|
the state evolves in a single experiment from the number of
|
||||||
|
detections, $m$, in time $t$.
|
||||||
|
|
||||||
|
The detection distribution on time-scales shorter than dissipation (so
|
||||||
|
we can approximate as if we were in a fully Zeno regime) can be
|
||||||
|
obtained by integrating over the detection times \cite{mekhov2009pra}
|
||||||
|
to get
|
||||||
|
\begin{equation}
|
||||||
|
P(m,t) = \sum_n \frac{[|c_n|^2 t]^m} {m!} e^{-|c_n|^2 t} \mathrm{Tr} (\rho_{nn}).
|
||||||
|
\end{equation}
|
||||||
|
For a state that is predominantly in one Zeno subspace, the
|
||||||
|
distribution will be approximately Poissonian (up to
|
||||||
|
$O(\nu^2 / \lambda^4)$, the population of the other
|
||||||
|
subspaces). Therefore, in a single experiment we will measure
|
||||||
|
$m = |c_0|^2t \pm \sqrt{|c_0|^2t}$ detections (note, we have assumed
|
||||||
|
$|c_0|^2 t$ is large enough to approximate the distribution as
|
||||||
|
normal. This is not necessary, we simply use it here to not have to
|
||||||
|
worry about the asymmetry in the deviation around the mean value). The
|
||||||
|
uncertainty does not come from the fact that $\lambda$ is not
|
||||||
|
infinite. The jumps are random events with a Poisson
|
||||||
|
distribution. Therefore, even in the full projective limit we will not
|
||||||
|
observe the same detection trajectory in each experiment even though
|
||||||
|
the system evolves in exactly the same way and remains in a perfectly
|
||||||
|
pure state.
|
||||||
|
|
||||||
|
If the basis of $\c$ is continuous (e.g. free particle position or
|
||||||
|
momentum) then the deviation around the mean will be our upper bound
|
||||||
|
on the deviation of the system from a pure state evolving around a
|
||||||
|
single Zeno subspace. However, continuous systems are beyond the scope
|
||||||
|
of this work and we will confine ourselves to discrete systems. Though
|
||||||
|
it is important to remember that continuous systems can be treated
|
||||||
|
this way, but the error estimate (and thus the mixedness of the state)
|
||||||
|
will be different.
|
||||||
|
|
||||||
|
For a discrete system it is easier to exclude all possibilities except
|
||||||
|
for one. The error in our estimate of $|c_0|^2$ in a single experiment
|
||||||
|
decreases as $1/\sqrt{t}$ and thus it can take a long time to
|
||||||
|
confidently determine $|c_0|^2$ to a sufficient precision this
|
||||||
|
way. However, since we know that it can only take one of the possible
|
||||||
|
values from the set $\{|c_n|^2 \}$ it is much easier to instead
|
||||||
|
exclude all the other values.
|
||||||
|
|
||||||
|
In an experiment we can use Bayes' theorem to infer the state of our
|
||||||
|
system as follows
|
||||||
|
\begin{equation}
|
||||||
|
p(c_n = c_0 | m) = \frac{ p(m | c_n = c_0) p(c_n = c_0) }{ p(m) },
|
||||||
|
\end{equation}
|
||||||
|
where $p(x)$ denotes the probability of the discrete event $x$ and
|
||||||
|
$p(x|y)$ the conditional probability of $x$ given $y$. We know that
|
||||||
|
$p(m | c_n = c_0)$ is simply given by a Poisson distribution with mean
|
||||||
|
$|c_0|^2 t$. $p(m)$ is just a normalising factor and $p(c_n = c_0)$ is
|
||||||
|
our \emph{a priori} knowledge of the state. Therefore, one can get the
|
||||||
|
probability of being in the right Zeno subspace from
|
||||||
|
\begin{align}
|
||||||
|
p(c_n & = c_0 | m) = \frac{ p_0(c_n = c_0) \frac{ \left( |c_0|^2 t
|
||||||
|
\right)^{2m} } {m!} e^{-|c_0|^2 t}} {\sum_n p_0(c_n) \frac{
|
||||||
|
\left( |c_n|^2 t \right)^{2m} } {m!} e^{-|c_n|^2 t}} \nonumber \\
|
||||||
|
& = p_0(c_n = c_0) \left[ \sum_n p_0(c_n) \left( \frac{
|
||||||
|
|c_n|^2 } { |c_0|^2 } \right)^{2m} e^{\left( |c_0|^2 -
|
||||||
|
|c_n|^2 \right) t} \right]^{-1},
|
||||||
|
\end{align}
|
||||||
|
where $p_0$ denotes probabilities at $t = 0$. In a real experiment one
|
||||||
|
could prepare the initial state to be close to the Zeno subspace of
|
||||||
|
interest and thus it would be easier to deduce the state. Furthermore,
|
||||||
|
in the middle of an experiment if we have already established the Zeno
|
||||||
|
subspace this will be reflected in these \emph{a priori} probabilities
|
||||||
|
again making it easier to infer the correct subspace. However, we will
|
||||||
|
consider the worst case scenario which might be useful if we don't
|
||||||
|
know the initial state or if the Zeno subspace changes during the
|
||||||
|
experiment, a uniform $p_0(c_n)$.
|
||||||
|
|
||||||
|
This probability is a rather complicated function as $m$ is a
|
||||||
|
stochastic quantity that also increases with $t$. We want it to be as
|
||||||
|
close to $1$ as possible. In order to devise an appropriate condition
|
||||||
|
for this we note that in the first line all terms in the denominator
|
||||||
|
are Poisson distributions of $m$. Therefore, if the mean values
|
||||||
|
$|c_n|^2 t$ are sufficiently spaced out, only one of the terms in the
|
||||||
|
sum will be significant for a given $m$ and if this happens to be the
|
||||||
|
one that corresponds to $c_0$ we get a probability close to
|
||||||
|
unity. Therefore, we set the condition such that it is highly unlikely
|
||||||
|
that our measured $m$ could be produced by two different distributions
|
||||||
|
\begin{align}
|
||||||
|
\sqrt{|c_0|^2 t} \ll ||c_0|^2 - |c_n|^2| t, \forall n \ne 0 \\
|
||||||
|
\sqrt{|c_n|^2 t} \ll ||c_0|^2 - |c_n|^2| t, \forall n \ne 0
|
||||||
|
\end{align}
|
||||||
|
The left-hand side is the standard deviation of $m$ if the system was
|
||||||
|
in subspace $P_0$ or $P_n$. The right-hand side is the difference in
|
||||||
|
the mean detections between the subspace $n$ and the one we are
|
||||||
|
interested in. The condition becomes more strict if the subspaces
|
||||||
|
become less distinguishable as it becomes harder to confidently
|
||||||
|
determine the correct state. Once again, using $\c = \lambda \hat{o}$
|
||||||
|
where $\hat{o} \sim O(1)$ we get
|
||||||
|
\begin{equation}
|
||||||
|
t \gg \frac{1}{\lambda^2} \frac{|o_{0,n}|^2} {(|o_0|^2 - |o_n|^2|)^2}.
|
||||||
|
\end{equation}
|
||||||
|
Since detections happen on average at an average rate of order
|
||||||
|
$\lambda^2$ we only need to wait for a few detections to satisfy this
|
||||||
|
condition. Therefore, we see that even in the worst case scenario of
|
||||||
|
complete ignorance of the state of the system we can very easily
|
||||||
|
determine the correct subspace. Once it is established for the first
|
||||||
|
time, the \emph{a priori} information can be updated and it will
|
||||||
|
become even easier to monitor the system.
|
||||||
|
|
||||||
|
However, it is important to note that physically once the quantum
|
||||||
|
jumps deviate too much from the mean value the system is more likely
|
||||||
|
to change the Zeno subspace (due to measurement backaction) and the
|
||||||
|
detection rate will visibly change. Therefore, if we observe a
|
||||||
|
consistent detection rate it is extremely unlikely that it can be
|
||||||
|
produced by two different Zeno subspaces so in fact it is even easier
|
||||||
|
to determine the correct state, but the above estimate serves as a
|
||||||
|
good lower bound on the necessary detection time.
|
||||||
|
|
||||||
|
Having derived the necessary conditions to confidently determine which
|
||||||
|
Zeno subspace is being observed in the experiment we can make another
|
||||||
|
approximation thanks to measurement which would be impossible in a
|
||||||
|
purely dissipative open system. If we observe a number of detections
|
||||||
|
consistent with the subspace $P_m = P_0$ we can set
|
||||||
|
$\hat{\rho}_{mn} \approx 0$ for all cases when both $m \ne 0$ and
|
||||||
|
$n \ne 0$ leaving our density matrix in the form
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:approxrho}
|
||||||
|
\hat{\rho} = \hat{\rho}_{00} + \sum_{r\ne0} (\hat{\rho}_{0r} +
|
||||||
|
\hat{\rho}_{r0}).
|
||||||
|
\end{equation}
|
||||||
|
We can do this, because the other states are inconsistent with the
|
||||||
|
measurement record. We know from the previous section that the system
|
||||||
|
must lie predominantly in only one of the Zeno subspaces and when that
|
||||||
|
is the case, $\hat{\rho}_{0r} \approx O(\nu/\lambda^2)$ and for
|
||||||
|
$m \ne 0$ and $n \ne 0$ we have
|
||||||
|
$\hat{\rho}_{mn} \approx O(\nu^2/\lambda^4)$. Therefore, this amounts
|
||||||
|
to keeping first order terms in $\nu/\lambda^2$ in our approximation.
|
||||||
|
|
||||||
|
This is a crucial step as all $\hat{\rho}_{mm}$ matrices are
|
||||||
|
decoherence free subspaces and thus they can all coexist in a mixed
|
||||||
|
state decreasing the purity of the system without
|
||||||
|
measurement. Physically, this means we exclude trajectories in which
|
||||||
|
the Zeno subspace has changed (measurement isn't fully projective). By
|
||||||
|
substituting Eq. \eqref{eq:intermediate} into Eq. \eqref{eq:master} we
|
||||||
|
see that this happens at a rate of $\nu^2 / \lambda^2$. However, since
|
||||||
|
the two measurement outcomes cannot coexist any transition between
|
||||||
|
them happens in discrete transitions (which we know about from the
|
||||||
|
change in the detection rate as each Zeno subspace will correspond to
|
||||||
|
a different rate) and not as continuous coherent evolution. Therefore,
|
||||||
|
we can postselect in a manner similar to Refs. \cite{otterbach2014,
|
||||||
|
lee2014prx, lee2014prl}, but our requirements are significantly more
|
||||||
|
relaxed - we do not require a specific single trajectory, only that it
|
||||||
|
remains within a Zeno subspace. Furthermore, upon reaching a steady
|
||||||
|
state, these transitions become impossible as the coherences
|
||||||
|
vanish. This approximation is analogous to optical Raman transitions
|
||||||
|
where the population of the excited state is neglected. Here, we can
|
||||||
|
make a similar approximation and neglect all but one Zeno subspace
|
||||||
|
thanks to the additional knowledge we gain from knowing the
|
||||||
|
measurement outcomes.
|
||||||
|
|
||||||
|
\subsection{The Non-Hermitian Hamiltonian}
|
||||||
|
|
||||||
|
Rewriting the master equation using $\c = c_0 + \delta \c$, where
|
||||||
|
$c_0$ is the eigenvalue corresponding to the eigenspace defined by the
|
||||||
|
projector $P_0$ which we used to obtain the density matrix in
|
||||||
|
Eq. \eqref{eq:approxrho}, we get
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:finalrho}
|
||||||
|
\dot{\hat{\rho}} = -i \left( \H_\mathrm{eff} \hat{\rho} - \hat{\rho}
|
||||||
|
\H_\mathrm{eff}^\dagger \right) + \delta \c \hat{\rho} \delta \cd,
|
||||||
|
\end{equation}
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:Ham}
|
||||||
|
\H_\mathrm{eff} = \H_0 + i \left( c_0^*\c - \frac{|c_0|^2}{2} - \frac{\cd\c}{2} \right).
|
||||||
|
\end{equation}
|
||||||
|
The first term in Eq. \eqref{eq:finalrho} describes coherent evolution
|
||||||
|
due to the non-Hermitian Hamiltonian $\H_\mathrm{eff}$ and the second
|
||||||
|
term is decoherence due to our ignorance of measurement outcomes. When
|
||||||
|
we substitute our approximation of the density matrix
|
||||||
|
$\hat{\rho} = \hat{\rho}_{00} + \sum_{r\ne0} (\hat{\rho}_{0r} +
|
||||||
|
\hat{\rho}_{r0})$ into Eq. \eqref{eq:finalrho}, the last term
|
||||||
|
vanishes, $\delta \c \hat{\rho} \delta \cd = 0$. This happens, because
|
||||||
|
$\delta \c P_0 \hat{\rho} = \hat{\rho} P_0 \delta \c^\dagger = 0$. The
|
||||||
|
projector annihilates all states except for those with eigenvalue
|
||||||
|
$c_0$ and so the operator $\delta \c = \c - c_0$ will always evaluate
|
||||||
|
to $c_0 - c_0 = 0$. Recall that we defined
|
||||||
|
$\hat{\rho}_{mn} = P_m \hat{\rho} P_n$ which means that every term in
|
||||||
|
our approximate density matrix contains the projector $P_0$. However,
|
||||||
|
it is important to note that this argument does not apply to other
|
||||||
|
second order terms in the master equation, because some terms only
|
||||||
|
have the projector $P_0$ applied from one side,
|
||||||
|
e.g.~$\hat{\rho}_{0m}$. The term $\delta \c \hat{\rho} \delta \cd$
|
||||||
|
applies the fluctuation operator from both sides so it does not matter
|
||||||
|
in this case, but it becomes relevant for terms such as
|
||||||
|
$\delta \cd \delta \c \hat{\rho}$. It is important to note that this
|
||||||
|
term does not automatically vanish, but when the explicit form of our
|
||||||
|
approximate density matrix is inserted, it is in fact zero. Therefore,
|
||||||
|
we can omit this term using the information we gained from
|
||||||
|
measurement, but keep other second order terms, such as
|
||||||
|
$\delta \cd \delta \c \rho$ in the Hamiltonian which are the origin of
|
||||||
|
other second-order dynamics. This could not be the case in a
|
||||||
|
dissipative system.
|
||||||
|
|
||||||
|
Ultimately we find that a system under continuous measurement for
|
||||||
|
which $\lambda^2 \gg \nu$ in the Zeno subspace $P_0$ is described by
|
||||||
|
the deterministic non-Hermitian Hamiltonian $\H_\mathrm{eff}$ in
|
||||||
|
Eq. \eqref{eq:Ham} and thus obeys the following Schr\"{o}dinger
|
||||||
|
equation
|
||||||
|
\begin{equation}
|
||||||
|
i \frac{\mathrm{d} | \Psi \rangle}{\mathrm{d}t} = \left[\H_0 + i \left(
|
||||||
|
c_0^*\c - \frac{|c_0|^2}{2} - \frac{\cd\c}{2} \right) \right] |
|
||||||
|
\Psi \rangle.
|
||||||
|
\end{equation}
|
||||||
|
Of the three terms in the parentheses the first two represent the
|
||||||
|
effects of quantum jumps due to detections (which one can think of as
|
||||||
|
`reference frame' shifts between different degenerate eigenspaces) and
|
||||||
|
the last term is the non-Hermitian decay due to information gain from
|
||||||
|
no detections. It is important to emphasize that even though we
|
||||||
|
obtained a deterministic equation, we have not neglected the
|
||||||
|
stochastic nature of the detection events. The detection trajectory
|
||||||
|
seen in an experiment will have fluctuations around the mean
|
||||||
|
determined by the Zeno subspace, but there simply are many possible
|
||||||
|
measurement records with the same outcome. This is just like the fully
|
||||||
|
projective Zeno limit where the system remains perfectly pure in one
|
||||||
|
of the possible projections, but the detections remain randomly
|
||||||
|
distributed in time.
|
||||||
|
|
||||||
|
One might then be concerned that purity is preserved even though we
|
||||||
|
might be averaging over many trajectories within this Zeno
|
||||||
|
subspace. We have neglected the small terms $\hat{\rho}_{m,n}$
|
||||||
|
($m,n \ne 0$) which are $O(\nu^2/\lambda^4)$ and thus they are not
|
||||||
|
correctly accounted for by our approximation. This means that we have
|
||||||
|
an $O(\nu^2/\lambda^4)$ error in our density matrix. The purity
|
||||||
|
given by
|
||||||
|
\begin{equation}
|
||||||
|
\mathrm{Tr}(\hat{\rho}^2) = \mathrm{Tr}(\hat{\rho}^2_{00} + \sum_{m \ne
|
||||||
|
0} \hat{\rho}_{0m}\hat{\rho}_{m0}) + \mathrm{Tr}(\sum_{m,n\ne0}
|
||||||
|
\hat{\rho}_{mn} \hat{\rho}_{nm})
|
||||||
|
\end{equation}
|
||||||
|
where the second term contains the terms not accounted for by our
|
||||||
|
approximation thus introduces an $O(\nu^4/\lambda^8)$
|
||||||
|
error. Therefore, this discrepancy is negligible in our
|
||||||
|
approximation. The pure state predicted by $\H_\mathrm{eff}$ is only
|
||||||
|
an approximation, albeit a good one, and the real state will be mixed
|
||||||
|
to a small extent. Whilst perfect purity within the Zeno subspace
|
||||||
|
$\hat{\rho}_{00}$ is expected due to the measurement's strong
|
||||||
|
decoupling effect, the nearly perfect purity when transitions outside
|
||||||
|
the Zeno subspace are included is a nontrivial result. Similarly, in
|
||||||
|
Raman transitions the population of the neglected excited state is
|
||||||
|
also non-zero, but negligible. Furthermore, this equation does not
|
||||||
|
actually require the adiabatic elimination used in
|
||||||
|
Eq. \eqref{eq:intermediate} (we only used it to convince ourselves
|
||||||
|
that the coherences are small) and such situations may be considered
|
||||||
|
provided all approximations remain valid. In a similar way the limit
|
||||||
|
of linear optics is derived from the physics of a two-level nonlinear
|
||||||
|
medium, when the population of the upper state is neglected and the
|
||||||
|
adiabatic elimination of coherences is not required.
|
||||||
|
|
||||||
|
\subsection{Non-Hermitian Dynamics in Ultracold Gases}
|
||||||
|
|
||||||
|
We finally return to our quantum gas model inside of a cavity. We
|
||||||
|
start by considering the simplest case of a global multi-site
|
||||||
|
measurement of the form $\hat{D} = \hat{N}_K = \sum_i^K \n_i$, where
|
||||||
|
the sum is over $K$ illuminated sites. The effective Hamiltonian
|
||||||
|
becomes
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:nHH2}
|
||||||
|
\hat{H}_\mathrm{eff} = \hat{H}_0 - i \gamma \left( \delta \hat{N}_K \right)^2,
|
||||||
|
\end{equation}
|
||||||
|
where $ \delta \hat{N}_K = \hat{N}_K - N^0_K$ and $N^0_K$ is the Zeno
|
||||||
|
subspace eigenvalue. It is now obvious that continuous measurement
|
||||||
|
squeezes the fluctuations in the measured quantity, as expected, and
|
||||||
|
that the only competing process is the system's own dynamics.
|
||||||
|
|
||||||
|
In this case, if we adiabatically eliminate the density matrix
|
||||||
|
cross-terms and substitute Eq. \eqref{eq:intermediate} into
|
||||||
|
Eq. \eqref{eq:master} for this system we obtain an effective
|
||||||
|
Hamiltonian within the Zeno subspace defined by $N_K$
|
||||||
|
\begin{equation}
|
||||||
|
\H_\varphi = P_0 \left[ \H_0 - i \frac{J^2}{\gamma}
|
||||||
|
\sum_\varphi \sum_{\substack{\langle i \in \varphi, j \in \varphi^\prime
|
||||||
|
\rangle \\ \langle k \in \varphi^\prime, l \in \varphi
|
||||||
|
\rangle}} b^\dagger_i b_j b^\dagger_k b_l \right] P_0,
|
||||||
|
\end{equation}
|
||||||
|
where $\varphi$ denotes a set of sites belonging to a single mode and
|
||||||
|
$\varphi^\prime$ is the set's complement (e.g. odd and even or
|
||||||
|
illuminated and non-illuminated sites) and $P_0$ is the projector onto
|
||||||
|
the eigenspace with $N_K^0$ atoms in the illuminated area. We focus on
|
||||||
|
the case when the second term is not only significant, but also leads
|
||||||
|
to dynamics within a Zeno subspace that are not allowed by
|
||||||
|
conventional quantum Zeno dynamics accounted for by the first
|
||||||
|
term. The second term represents second-order transitions via other
|
||||||
|
subspaces which act as intermediate states much like virtual states in
|
||||||
|
optical Raman transitions. This is in contrast to the conventional
|
||||||
|
understanding of the Zeno dynamics for infinitely frequent projective
|
||||||
|
measurements (corresponding to $\gamma \rightarrow \infty$) where such
|
||||||
|
processes are forbidden \cite{facchi2008}. Thus, it is the weak
|
||||||
|
quantum measurement that effectively couples the states. Note that
|
||||||
|
this is a special case of the equation in Eq. \eqref{eq:hz} which can
|
||||||
|
be obtained by considering a more general two mode setup.
|
||||||
|
|
||||||
|
\subsection{Small System Example}
|
||||||
|
|
||||||
|
To get clear physical insight, we initially consider three atoms in
|
||||||
|
three sites and choose our measurement operator such that
|
||||||
|
$\hat{D} = \n_2$, i.e.~only the middle site is subject to measurement,
|
||||||
|
and the Zeno subspace defined by $n_2 = 1$. Such an illumination
|
||||||
|
pattern can be achieved with global addressing by crossing two beams
|
||||||
|
and placing the nodes at the odd sites and the antinodes at even
|
||||||
|
sites. This means that $P_0 \H_0 P_0 = 0$. However, the
|
||||||
|
first and third sites are connected via the second term. Diagonalising
|
||||||
|
the Hamiltonian reveals that out of its ten eigenvalues all but three
|
||||||
|
have a significant negative imaginary component of the order $\gamma$
|
||||||
|
which means that the corresponding eigenstates decay on a time scale
|
||||||
|
of a single quantum jump and thus quickly become negligible. The three
|
||||||
|
remaining eigenvectors are dominated by the linear superpositions of
|
||||||
|
the three Fock states $|2,1,0 \rangle$, $|1, 1, 1 \rangle$, and
|
||||||
|
$|0,1,2 \rangle$. Whilst it is not surprising that these components
|
||||||
|
are the only ones that remain as they are the only ones that actually
|
||||||
|
lie in the Zeno subspace $n_2 = 1$, it is impossible to solve the full
|
||||||
|
dynamics by just considering these Fock states alone as they are not
|
||||||
|
coupled to each other in $\hat{H}_0$. The components lying outside of
|
||||||
|
the Zeno subspace have to be included to allow intermediate steps to
|
||||||
|
occur via states that do not belong in this subspace, much like
|
||||||
|
virtual states in optical Raman transitions.
|
||||||
|
|
||||||
|
An approximate solution for $U=0$ can be written for the
|
||||||
|
$\{|2,1,0 \rangle, |1,1,1 \rangle, |0,1,2 \rangle\}$ subspace by
|
||||||
|
multiplying each eigenvector with its corresponding time evolution
|
||||||
|
\begin{equation}
|
||||||
|
| \Psi(t) \rangle \propto \left( \begin{array}{c}
|
||||||
|
z_1 + \sqrt{2} z_2 e^{-6 J^2 t / \gamma} + z_3 e^{-12 J^2 t / \gamma} \\
|
||||||
|
-\sqrt{2} \left(z_1 - z_3 e^{-12 J^2 t / \gamma} \right) \\
|
||||||
|
z_1 - \sqrt{2} z_2 e^{-6 J^2 t / \gamma} + z_3 e^{-12 J^2 t /
|
||||||
|
\gamma} \\
|
||||||
|
\end{array}
|
||||||
|
\right), \nonumber
|
||||||
|
\end{equation}
|
||||||
|
where $z_i$ denote the overlap between the eigenvectors and the
|
||||||
|
initial state, $z_i = \langle v_i | \Psi (0) \rangle$, with
|
||||||
|
$| v_1 \rangle = (1, -\sqrt{2}, 1)/2$,
|
||||||
|
$| v_2 \rangle = (1, 0, -1)/\sqrt{2}$, and
|
||||||
|
$| v_3 \rangle = (1, \sqrt{2}, 1)/2$. The steady state as
|
||||||
|
$t \rightarrow \infty$ is given by
|
||||||
|
$| v_1 \rangle = (1, -\sqrt{2}, 1)/2$. This solution is illustrated in
|
||||||
|
Fig. \ref{fig:comp} which clearly demonstrates dynamics beyond the
|
||||||
|
canonical understanding of quantum Zeno dynamics as tunnelling occurs
|
||||||
|
between states coupled via a different Zeno subspace.
|
||||||
|
|
||||||
|
\begin{figure}[hbtp!]
|
||||||
|
\includegraphics[width=\linewidth]{comp}
|
||||||
|
\caption[Fock State Populations in a Zeno
|
||||||
|
Subspace]{Populations of the Fock states in the Zeno subspace
|
||||||
|
for $\gamma/J = 100$ and initial state $| 2,1,0 \rangle$. It
|
||||||
|
is clear that quantum Zeno dynamics occurs via Raman-like
|
||||||
|
processes even though none of these states are connected in
|
||||||
|
$\hat{H}_0$. The dynamics occurs via virtual intermediate
|
||||||
|
states outside the Zeno subspace. The system also tends to a
|
||||||
|
steady state which minimises tunnelling effectively
|
||||||
|
suppressing fluctuations. The lines are solutions to the
|
||||||
|
non-Hermitian Hamiltonian, and the dots are points from a
|
||||||
|
stochastic trajectory calculation.\label{fig:comp}}
|
||||||
|
\end{figure}
|
||||||
|
|
||||||
|
\subsection{Steady State of non-Hermitian Dynamics}
|
||||||
|
|
||||||
|
A distinctive difference between Bose-Hubbard model ground states and
|
||||||
|
the final steady state,
|
||||||
|
$| \Psi \rangle = [|2,1,0 \rangle - \sqrt{2} |1,1,1\rangle +
|
||||||
|
|0,1,2\rangle]/2$, is that its components are not in phase. Squeezing
|
||||||
|
due to measurement naturally competes with inter-site tunnelling which
|
||||||
|
tends to spread the atoms. However, from Eq. \eqref{eq:nHH2} we see
|
||||||
|
the final state will always be the eigenvector with the smallest
|
||||||
|
fluctuations as it will have an eigenvalue with the largest imaginary
|
||||||
|
component. This naturally corresponds to the state where tunnelling
|
||||||
|
between Zeno subspaces (here between every site) is minimised by
|
||||||
|
destructive matter-wave interference, i.e.~the tunnelling dark state
|
||||||
|
defined by $\hat{T} |\Psi \rangle = 0$, where
|
||||||
|
$\hat{T} = \sum_{\langle i, j \rangle} \bd_i b_j$. This is simply the
|
||||||
|
physical interpretation of the steady states we predicted for
|
||||||
|
Eq. \eqref{eq:master}. Crucially, this state can only be reached if
|
||||||
|
the dynamics aren't fully suppressed by measurement and thus,
|
||||||
|
counter-intuitively, the atomic dynamics cooperate with measurement to
|
||||||
|
suppress itself by destructive interference. Therefore, this effect is
|
||||||
|
beyond the scope of traditional quantum Zeno dynamics and presents a
|
||||||
|
new perspective on the competition between a system's short-range
|
||||||
|
dynamics and global measurement backaction.
|
||||||
|
|
||||||
|
We now consider a one-dimensional lattice with $M$ sites so we extend
|
||||||
|
the measurement to $\hat{D} = \N_\text{even}$ where every even site is
|
||||||
|
illuminated. The wavefunction in a Zeno subspace must be an
|
||||||
|
eigenstate of $\c$ and we combine this with the requirement for it to
|
||||||
|
be in the dark state of the tunnelling operator (eigenstate of $\H_0$
|
||||||
|
for $U = 0$) to derive the steady state. These two conditions in
|
||||||
|
momentum space are
|
||||||
|
\begin{equation}
|
||||||
|
\hat{T} | \Psi \rangle = \sum_{\text{RBZ}} \left[ \bd_k b_k -
|
||||||
|
\bd_{q} b_{q} \right] \cos(ka) |\Psi \rangle = 0, \nonumber
|
||||||
|
\end{equation}
|
||||||
|
\begin{equation}
|
||||||
|
\Delta \N |\Psi \rangle = \sum_{\text{RBZ}} \left[ \bd_k b_{-q} +
|
||||||
|
\bd_{-q} b_k \right] | \Psi \rangle= \Delta N |\Psi \rangle, \nonumber
|
||||||
|
\end{equation}
|
||||||
|
where $b_k = \frac{1}{\sqrt{M}} \sum_j e^{i k j a} b_j$,
|
||||||
|
$\Delta \hat{N} = \hat{D} - N/2$, $q = \pi/a - k$, $a$ is the lattice
|
||||||
|
spacing, $N$ the total atom number, and we perform summations over the
|
||||||
|
reduced Brillouin zone (RBZ), $-\pi/2a < k \le \pi/2a$, as the
|
||||||
|
symmetries of the system are clearer this way. Now we define
|
||||||
|
\begin{equation}
|
||||||
|
\hat{\alpha}_k^\dagger = \bd_k \bd_q - \bd_{-k} \bd_{-q},
|
||||||
|
\end{equation}
|
||||||
|
\begin{equation}
|
||||||
|
\hat{\beta}_\varphi^\dagger = \bd_{\pi/2a} + \varphi \bd_{-\pi/2a},
|
||||||
|
\end{equation}
|
||||||
|
where $\varphi = \Delta N / | \Delta N |$, which create the smallest
|
||||||
|
possible states that satisfy the two equations for $\Delta N = 0$ and
|
||||||
|
$\Delta N \ne 0$ respectively. Therefore, by noting that
|
||||||
|
\begin{align}
|
||||||
|
\left[ \hat{T}, \hat{\alpha}_k^\dagger \right] & = 0, \\
|
||||||
|
\left[ \hat{T}, \hat{\beta}_\varphi^\dagger \right] & = 0, \\
|
||||||
|
\left[ \Delta \N, \hat{\alpha}_k^\dagger \right] & = 0, \\
|
||||||
|
\left[ \Delta \N, \hat{\beta}_\varphi^\dagger \right] & = \varphi
|
||||||
|
\hat{\beta}_\varphi^\dagger,
|
||||||
|
\end{align}
|
||||||
|
we can now write the equation for the $N$-particle steady state
|
||||||
|
\begin{equation}
|
||||||
|
\label{eq:ss}
|
||||||
|
| \Psi \rangle \propto \left[ \prod_{i=1}^{(N - |\Delta N|)/2}
|
||||||
|
\left( \sum_{k = 0}^{\pi/2a} \phi_{i,k} \hat{\alpha}_k^\dagger
|
||||||
|
\right) \right] \left( \hat{\beta}_\varphi^\dagger \right)^{|
|
||||||
|
\Delta N |} | 0 \rangle, \nonumber
|
||||||
|
\end{equation}
|
||||||
|
where $\phi_{i,k}$ are coefficients that depend on the trajectory
|
||||||
|
taken to reach this state and $|0 \rangle$ is the vacuum state defined
|
||||||
|
by $b_k |0 \rangle = 0$. Since this a dark state (an eigenstate of
|
||||||
|
$\H_0$) of the atomic dynamics, this state will remain stationary even
|
||||||
|
with measurement switched-off. Interestingly, this state is very
|
||||||
|
different from the ground states of the Bose-Hubbard Hamiltonian, it
|
||||||
|
is even orthogonal to the superfluid state, and thus it cannot be
|
||||||
|
obtained by cooling or projecting from an initial ground state. The
|
||||||
|
combination of tunnelling with measurement is necessary.
|
||||||
|
|
||||||
|
\begin{figure}[hbtp!]
|
||||||
|
\includegraphics[width=\linewidth]{figure3}
|
||||||
|
\caption[Non-Hermitian Steady State]{A trajectory simulation
|
||||||
|
for eight atoms in eight sites, initially in
|
||||||
|
$|1,1,1,1,1,1,1,1 \rangle$, with periodic boundary
|
||||||
|
conditions and $\gamma/J = 100$. (a), The fluctuations in
|
||||||
|
$\c$ where the stochastic nature of the process is clearly
|
||||||
|
visible on a single trajectory level. However, the general
|
||||||
|
trend is captured by the non-Hermitian Hamiltonian. (b), The
|
||||||
|
local density variance. Whilst the fluctuations in the
|
||||||
|
global measurement operator decrease, the fluctuations in
|
||||||
|
local density increase due to tunnelling via states outside
|
||||||
|
the Zeno subspace. (c), The momentum distribution. The
|
||||||
|
initial Fock state has a flat distribution which with time
|
||||||
|
approaches the steady state distribution of two identical
|
||||||
|
and symmetric distributions centred at $k = \pi/2a$ and
|
||||||
|
$k = -\pi/2a$.\label{fig:steady}}
|
||||||
|
\end{figure}
|
||||||
|
|
||||||
|
In order to prepare the steady state one has to run the experiment and
|
||||||
|
wait until the photocount rate remains constant for a sufficiently
|
||||||
|
long time. Such a trajectory is illustrated in Fig. \ref{fig:steady}
|
||||||
|
and compared to a deterministic trajectory calculated using the
|
||||||
|
non-Hermitian Hamiltonian. It is easy to see from
|
||||||
|
Fig. \ref{fig:steady}(a) how the stochastic fluctuations around the
|
||||||
|
mean value of the observable have no effect on the general behaviour
|
||||||
|
of the system in the strong measurement regime. By discarding these
|
||||||
|
fluctuations we no longer describe a pure state, but we showed how
|
||||||
|
this only leads to a negligible error. Fig. \ref{fig:steady}(b) shows
|
||||||
|
the local density variance in the lattice. Not only does it grow
|
||||||
|
showing evidence of tunnelling between illuminated and non-illuminated
|
||||||
|
sites, but it grows to significant values. This is in contrast to
|
||||||
|
conventional quantum Zeno dynamics where no tunnelling would be
|
||||||
|
allowed at all. Finally, Fig. \ref{fig:steady}(c) shows the momentum
|
||||||
|
distribution of the trajectory. We can clearly see that it deviates
|
||||||
|
significantly from the initial flat distribution of the Fock
|
||||||
|
state. Furthermore, the steady state does not have any atoms in the
|
||||||
|
$k=0$ state and thus is orthogonal to the superfluid state as
|
||||||
|
discussed.
|
||||||
|
|
||||||
|
To obtain a state with a specific value of $\Delta N$ postselection
|
||||||
|
may be necessary, but otherwise it is not needed. The process can be
|
||||||
|
optimised by feedback control since the state is monitored at all
|
||||||
|
times \cite{ivanov2014}. Furthermore, the form of the measurement
|
||||||
|
operator is very flexible and it can easily be engineered by the
|
||||||
|
geometry of the optical setup \cite{elliott2015, mazzucchi2016} which
|
||||||
|
can be used to design a state with desired properties.
|
||||||
|
|
||||||
|
\section{Conclusions}
|
||||||
|
|
||||||
|
In this chapter we have demonstrated that global quantum measurement
|
||||||
|
backaction can efficiently compete with standard local processes in
|
||||||
|
many-body systems. This introduces a completely new energy and time
|
||||||
|
scale into quantum many-body research. This is made possbile by the
|
||||||
|
ability to structure the spatial profile of the measurement on a
|
||||||
|
microscopic scale comparable to the lattice period without the need
|
||||||
|
for single site addressing. The extreme flexibility of the setup
|
||||||
|
considered allowed us to effectively tailor long-range entanglement
|
||||||
|
and correlations present in the system. We showed that the competition
|
||||||
|
between the global backaction and usual atomic dynamics leads to the
|
||||||
|
production of spatially multimode macroscopic superpositions which
|
||||||
|
exhibit large-scale oscillatory dynamics which could be used for
|
||||||
|
quantum information and metrology. We subsequently demonstrated that
|
||||||
|
when on-site atomic interactions are introduced the dynamics become
|
||||||
|
much more complicated with different regimes of behaviour where
|
||||||
|
measurement and interactions can either compete or cooperate. In the
|
||||||
|
strong measurement regime we showed that conventional quantum Zeno
|
||||||
|
dynamics can be realised, but more interestingly, by considering a
|
||||||
|
strong, but not projective, limit of measurement we observe a new type
|
||||||
|
of nonlocal dynamics. It turns out that a global measurement scheme
|
||||||
|
leads to correlations between spatially separated tunnelling events
|
||||||
|
which conserve the Zeno subspace via Raman-like processes which would
|
||||||
|
be forbidden in the canonical fully projective limit. We subsequently
|
||||||
|
presented a rigorous analysis of the underlying process of this new
|
||||||
|
type of quantum Zeno dynamics in which we showed that in this limit
|
||||||
|
quantum trajectories can be described by a deterministic non-Hermitian
|
||||||
|
Hamiltonian. In contrast to previous works, it is independent of the
|
||||||
|
underlying system and there is no need to postselect a particular
|
||||||
|
exotic trajectory \cite{lee2014prx, lee2014prl}. Finally, we have
|
||||||
|
shown that the system will always tend towards the eigenstate of the
|
||||||
|
Hamiltonian with the best squeezing of the observable and the atomic
|
||||||
|
dynamics, which normally tend to spread the distribution, cooperates
|
||||||
|
with measurement to produce a state in which tunnelling is suppressed
|
||||||
|
by destructive matter-wave interference. A dark state of the
|
||||||
|
tunnelling operator will have zero fluctuations and we provided an
|
||||||
|
expression for the steady state which is significantly different from
|
||||||
|
the ground state of the Hamiltonian. This is in contrast to previous
|
||||||
|
works on dissipative state preparation where the steady state had to
|
||||||
|
be a dark state of the measurement operator \cite{diehl2008}.
|
||||||
|
|
||||||
|
Such globally paired tunnelling due to a fundamentally new phenomenon,
|
||||||
|
global quantum measurement backaction, can enrich the physics of
|
||||||
|
long-range correlated systems beyond relatively short-range
|
||||||
|
interactions expected from standard dipole-dipole interactions
|
||||||
|
\cite{sowinski2012, omjyoti2015}. These nonlocal high-order processes
|
||||||
|
entangle regions of the optical lattice that are disconnected by the
|
||||||
|
measurement. Using different detection schemes, we showed how to
|
||||||
|
tailor density-density correlations between distant lattice
|
||||||
|
sites. Quantum optical engineering of nonlocal coupling to
|
||||||
|
environment, combined with quantum measurement, can allow the design
|
||||||
|
of nontrivial system-bath interactions, enabling new links to quantum
|
||||||
|
simulations~\cite{stannigel2013} and thermodynamics~\cite{erez2008}
|
||||||
|
and extend these directions to the field of non-Hermitian quantum
|
||||||
|
mechanics, where quantum optical setups are particularly
|
||||||
|
promising~\cite{lee2014prl}. Importantly, both systems and baths,
|
||||||
|
designed by our method, can be strongly correlated systems with
|
||||||
|
internal long-range entanglement.
|
||||||
|
@ -211,4 +211,7 @@
|
|||||||
\renewcommand{\a}{a} % in case we decide to put hats on
|
\renewcommand{\a}{a} % in case we decide to put hats on
|
||||||
\renewcommand{\c}{\hat{c}}
|
\renewcommand{\c}{\hat{c}}
|
||||||
\newcommand{\cd}{\hat{c}^\dagger}
|
\newcommand{\cd}{\hat{c}^\dagger}
|
||||||
\renewcommand{\b}[1]{\mathbf{#1}}
|
\renewcommand{\b}[1]{\mathbf{#1}}
|
||||||
|
\newcommand{\op}{\hat{o}}
|
||||||
|
\newcommand{\h}{\hat{h}}
|
||||||
|
\newcommand{\N}{\hat{N}}
|
@ -978,3 +978,65 @@ doi = {10.1103/PhysRevA.87.043613},
|
|||||||
year = {2005},
|
year = {2005},
|
||||||
publisher = {American Physical Society},
|
publisher = {American Physical Society},
|
||||||
}
|
}
|
||||||
|
@article{otterbach2014,
|
||||||
|
title = {Dissipative Preparation of Spatial Order in Rydberg-Dressed
|
||||||
|
Bose-Einstein Condensates},
|
||||||
|
author = {Otterbach, Johannes and Lemeshko, Mikhail},
|
||||||
|
journal = {Phys. Rev. Lett.},
|
||||||
|
volume = {113},
|
||||||
|
issue = {7},
|
||||||
|
pages = {070401},
|
||||||
|
numpages = {6},
|
||||||
|
year = {2014},
|
||||||
|
month = {Aug},
|
||||||
|
publisher = {American Physical Society},
|
||||||
|
}
|
||||||
|
@article{lee2014prx,
|
||||||
|
title = {Heralded Magnetism in Non-Hermitian Atomic Systems},
|
||||||
|
author = {Lee, Tony E. and Chan, Ching-Kit},
|
||||||
|
journal = {Phys. Rev. X},
|
||||||
|
volume = {4},
|
||||||
|
issue = {4},
|
||||||
|
pages = {041001},
|
||||||
|
numpages = {13},
|
||||||
|
year = {2014},
|
||||||
|
month = {Oct},
|
||||||
|
publisher = {American Physical Society},
|
||||||
|
}
|
||||||
|
@article{lee2014prl,
|
||||||
|
title = {Entanglement and Spin Squeezing in Non-Hermitian Phase Transitions},
|
||||||
|
author = {Lee, Tony E. and Reiter, Florentin and Moiseyev, Nimrod},
|
||||||
|
journal = {Phys. Rev. Lett.},
|
||||||
|
volume = {113},
|
||||||
|
issue = {25},
|
||||||
|
pages = {250401},
|
||||||
|
numpages = {5},
|
||||||
|
year = {2014},
|
||||||
|
month = {Dec},
|
||||||
|
publisher = {American Physical Society},
|
||||||
|
}
|
||||||
|
@article{stannigel2013,
|
||||||
|
title = {{Constrained Dynamics via the Zeno Effect in Quantum
|
||||||
|
Simulation: Implementing Non-Abelian Lattice Gauge
|
||||||
|
Theories with Cold Atoms}},
|
||||||
|
author = {Stannigel, K. and Hauke, P. and Marcos, D. and Hafezi,
|
||||||
|
M. and Diehl, S. and Dalmonte, M. and Zoller, P.},
|
||||||
|
journal = {Phys. Rev. Lett.},
|
||||||
|
volume = {112},
|
||||||
|
issue = {12},
|
||||||
|
pages = {120406},
|
||||||
|
numpages = {6},
|
||||||
|
year = {2014},
|
||||||
|
month = {Mar},
|
||||||
|
publisher = {American Physical Society},
|
||||||
|
doi = {10.1103/PhysRevLett.112.120406}
|
||||||
|
}
|
||||||
|
@article{erez2008,
|
||||||
|
title={{Repulsively bound atom pairs in an optical lattice}},
|
||||||
|
author={Erez, N. and Gordon, G. and Nest, M. and Kurizki, G.},
|
||||||
|
journal={Nature},
|
||||||
|
volume={452},
|
||||||
|
pages = {724},
|
||||||
|
year={2008},
|
||||||
|
publisher={Nature Publishing Group}
|
||||||
|
}
|
||||||
|
@ -1,7 +1,7 @@
|
|||||||
% ******************************* PhD Thesis Template **************************
|
% ******************************* PhD Thesis Template **************************
|
||||||
% Please have a look at the README.md file for info on how to use the template
|
% Please have a look at the README.md file for info on how to use the template
|
||||||
|
|
||||||
\documentclass[a4paper,12pt,times,numbered,print,chapter]{Classes/PhDThesisPSnPDF}
|
\documentclass[a4paper,12pt,times,numbered,print]{Classes/PhDThesisPSnPDF}
|
||||||
|
|
||||||
% ******************************************************************************
|
% ******************************************************************************
|
||||||
% ******************************* Class Options ********************************
|
% ******************************* Class Options ********************************
|
||||||
|
Reference in New Issue
Block a user