1: \documentclass[11pt,a4paper,oneside,british,headsepline,idxtotoc, bibtotoc,liststotoc]{scrbook}
2: \usepackage[T1]{fontenc}
3: \usepackage[latin1]{inputenc}
4: \usepackage{amsmath}
5: \usepackage{makeidx}
6: \makeindex
7: \usepackage{graphicx}
8: \usepackage{setspace}
9: \onehalfspacing
10: \usepackage{amssymb}
11:
12: \makeatletter
13:
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
15: \newcommand{\noun}[1]{\textsc{#1}}
16: %% Because html converters don't know tabularnewline
17: \providecommand{\tabularnewline}{\\}
18:
19: \input{db.std}
20:
21: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
22: \typearea[20mm]{12}
23: %\usepackage{hyperref}
24: \usepackage{cite}
25: \providecommand{\openone}{\leavevmode\hbox{\small1\kern-3.8pt\normalsize1}}
26: \setcapindent{0cm}
27: %\raggedbottom
28:
29: \usepackage{babel}
30: \makeatother
31: \begin{document}
32: \begin{center}\thispagestyle{plain}\par\end{center}
33:
34: \vspace{-4cm}
35:
36:
37: \hspace{-4.63cm}\includegraphics[width=21cm]{ucllogo}\vspace{2cm}
38:
39:
40: \begin{center}\textsf{\textbf{\LARGE Quantum State Transfer with Spin
41: Chains}} \par\end{center}
42:
43: \begin{center}\vspace{1cm}
44: \par\end{center}
45:
46: \begin{center}\textsf{\textbf{\LARGE Daniel Klaus Burgarth}}\vspace{1cm}
47: \par\end{center}
48:
49: \begin{center}A thesis submitted to the University of London\\
50: for the degree of Doctor of Philosophy\par\end{center}
51:
52: \begin{center}\vspace{1cm}
53: \par\end{center}
54:
55: \begin{center}Department of Physics and Astronomy\\
56: University College London\vspace{1cm}
57: \par\end{center}
58:
59: \begin{center}December 2006\par\end{center}
60:
61:
62: \chapter*{Declaration}
63:
64: I, Daniel Klaus Burgarth, confirm that the work presented in this
65: thesis is my own. Where information has been derived from other sources,
66: I confirm that this has been indicated in the thesis.
67:
68:
69: \chapter*{Abstract}
70:
71: In the last few decades the idea came up that by making use of the
72: superposition principle from Quantum Mechanics, one can process information
73: in a new and much faster way. Hence a new field of information technology,
74: QIT (Quantum Information Technology), has emerged. From a physics
75: point of view it is important to find ways of implementing these new
76: methods in real systems. One of the most basic tasks required for
77: QIT is the ability to connect different components of a Quantum Computer
78: by \emph{quantum wires} that obey the superposition principle. Since
79: superpositions can be very sensitive to noise this turns out to be
80: already quite difficult. Recently, it was suggested to use chains
81: of permanently coupled spin-1/2 particles (\emph{quantum chains})
82: for this purpose. They have the advantage that no external control
83: along the wire is required during the transport of information, which
84: makes it possible to isolate the wire from sources of noise. The purpose
85: of this thesis is to develop and investigate advanced schemes for
86: using quantum chains as wires. We first give an introduction to basic
87: quantum state transfer and review existing advanced schemes by other
88: authors. We then introduce two new methods which were created as a
89: part of this thesis. First, we show how the fidelity of transfer can
90: be made perfect by performing measurements at the receiving end of
91: the chain. Then we introduce a scheme which is based on performing
92: unitary operations at the end of the chain. We generalise both methods
93: and discuss them from the more fundamental point of view of mixing
94: properties of a quantum channel. Finally, we study the effects of
95: a non-Markovian environment on quantum state transfer.
96:
97:
98: \chapter*{Acknowledgements}
99:
100: Most of all, I would like to thank my supervisor Sougato Bose for
101: much inspiration and advice. I am very grateful for many inspiring
102: and fruitful discussions and collaborations with Vittorio Giovannetti,
103: and with Floor Paauw, Christoph Bruder, Jason Twamley, Andreas Buchleitner
104: and Vladimir Korepin. Furthermore I would like to thank all my teachers
105: and those who have guided and motivated me along my journey through
106: physics, including Heinz-Peter Breuer, Francesco Petruccione, Lewis
107: Ryder, John Strange, Werner Riegler, Carsten Schuldt and Rolf Bussmann.
108: I acknowledge financial support by the UK Engineering and Physical
109: Sciences Research Council through the grant GR/S62796/01. Finally
110: I would like to thank my parents for their loving support.
111:
112:
113: \chapter*{Notation}
114:
115: %\addcontentsline{toc}{chapter}{Notation}
116:
117: \begin{quotation}
118: X,Y,Z\hfill{}Pauli matrices
119:
120: $X_{n},Y_{n},Z_{n}$\hfill{}Pauli matrices acting on the Hilbert-space
121: of qubit $n$
122:
123: $|0\rangle,|1\rangle$ \hfill{}Single qubit state in the canonical
124: basis
125:
126: $|\boldsymbol{0}\rangle$\hfill{}Quantum chain in the product state
127: $|0\rangle\otimes\cdots\otimes|0\rangle$
128:
129: $|\boldsymbol{n}\rangle$\hfill{}''Single excitation'' state $X_{n}|\boldsymbol{0}\rangle$
130:
131: $\mbox{Tr}_{X}$\hfill{}Partial trace over subsystem $X$
132:
133: $||\ldots||$\hfill{}Euclidean vector norm
134:
135: $||\ldots||_{1}$\hfill{}Trace norm
136:
137: $||\ldots||_{2}$\hfill{}Euclidean matrix norm
138: \end{quotation}
139: \begin{flushleft}We also use the following graphical representation:\vspace{5mm}
140: \includegraphics[width=0.663\columnwidth]{definitions}\par\end{flushleft}
141:
142: \tableofcontents{}
143:
144:
145: \chapter{Introduction}
146:
147: The Hilbert space that contains the states of quantum mechanical objects
148: is huge, scaling exponentially with the number of particles described.
149: In 1982, Richard Feynman suggested to make use of this as a resource
150: for \emph{simulating} quantum mechanics in a \emph{quantum computer,}
151: i.e. a device where the physical interaction could be {}``programmed''
152: to yield a specific Hamiltonian. This has led to the new fields of
153: Quantum Computation and Quantum Information. A quantum computer can
154: solve questions one could never imagine to solve using an ordinary
155: computer. For example, it can factorise large numbers into primes
156: efficiently, a task of greatest importance for cryptography. It may
157: thus be a surprise that more than twenty years after the initial ideas,
158: these devices still haven't been built or only in ridiculously small
159: size. The largest quantum computer so far can only solve problems
160: that any child could solve within seconds. A closer look reveals that
161: the main problem in the realisation of quantum computers is the {}``programming'',
162: i.e. the design of a specific (time-dependent) Hamiltonian, usually
163: described as a set of discrete unitary gates. This turns out to be
164: extremely difficult because we need to connect microscopic objects
165: (those behaving quantum mechanically) with macroscopic devices that
166: \emph{control} the microscopic behaviour. Even if one manages to find
167: a link between the micro- and the macroscopic world, such as laser
168: pulses and electric or magnetic fields, then the connection introduces
169: not only control but also noise (dissipation and decoherence) to the
170: microscopic system, and its quantum behaviour is diminished.
171:
172: The vision of this thesis is to develop theoretical methods narrowing
173: the gap between what is imagined theoretically and what can be done
174: experimentally. As a method we consider chains (or more general graphs)
175: of \emph{permanently coupled} quantum systems. This idea has been
176: originally put forward by S. Bose for the specific task of quantum
177: communication~\cite{Bose2003}. Due to the permanent coupling, these
178: devices can in principle be built in such a way that they don't require
179: external control to perform their tasks, just like a mechanical clockwork.
180: This also overcomes the problem of decoherence as they can be separated
181: from any source of noise. Unfortunately, most schemes that have been
182: developed so far still require external control, though much less
183: than an {}``ordinary'' quantum computer. Furthermore, internal dispersion
184: in these devices is leading to a decrease of their fidelity. A third
185: problem is, that for building these devices the permanent couplings
186: still need to be realised, although only once, and experimental constraints
187: such as resolution and errors need to be considered. We are thus left
188: with the following questions: which is the best way to perform quantum
189: state transfer using a permanently coupled graph? How much control
190: do we need, and how difficult will it be to implement the couplings?
191: How do errors and noise affect the scheme? All these points are highly
192: related and it cannot be expected to find an \emph{absolute,} i.e.
193: system independent answer. The purpose of this research is to develop
194: advanced schemes for the transfer of quantum information, to improve
195: and generalise existing ideas, to relate them to each other and to
196: investigate their stability and efficiency.
197:
198:
199: \section{Quantum Computation and Quantum Information\label{sec:Quantum-Computation}}
200:
201: In this Section we review some of the basic concepts of Quantum Computation.
202: We will be very brief and only focus on those aspects that we require
203: later on in the thesis. A more detailed introduction can be found
204: in~\cite{NIELSEN}.
205:
206: In information science, an algorithm is a list of instructions that
207: a computer performs on a given input to achieve a specific task. For
208: instance, a \emph{factoring algorithm} has an arbitrary integer as
209: its input, and gives its prime factors as an output. A \emph{quantum
210: factoring algorithm} can be thought of in a similar way, i.e. it has
211: an integer as input, and its prime factors as an output. \emph{In-between}
212: however it encodes information in a quantum mechanical system. Due
213: to the superposition principle, the information of a quantum system
214: cannot be represented as \emph{bits.} The valid generalisation of
215: the bit to the quantum case is called \emph{qubit.} The possible states
216: of a qubit are written as\begin{equation}
217: \alpha|0\rangle+\beta|1\rangle,\end{equation}
218: where $\alpha,\beta$ are normalised complex coefficients, and $|0\rangle$
219: and $|1\rangle$ are vectors of a two-dimensional complex vector space.
220: Peter W. Shor has shown in a famous paper~\cite{Shor1997} that the
221: \emph{detour} of representing the intermediate part of a factoring
222: algorithm in a quantum system (as well as using quantum gates, see
223: below) can be very beneficial: it runs \emph{much} faster. This is
224: important, because many cryptographic methods rely on factoring algorithms
225: being slow. \index{Shor's algorithm}Shor's algorithm is definitely
226: not the only reason why it would be very nice to have a \emph{quantum
227: computer\index{quantum computer},} i.e. a machine that represents
228: information in a quantum way and can perform instructions on it, and
229: many more details can be found in the textbook mentioned above.
230:
231: Algorithms on a computer can be represented as list of logical operations
232: on bits. Likewise, a (standard) quantum algorithm can be represented
233: as a list of \emph{quantum logical operations}, or \emph{\index{quantum gates}quantum
234: gates,} acting on qubits. The most general quantum algorithm is given
235: by an arbitrary unitary operator. A \emph{universal set of gates}
236: is a set such that any quantum algorithm (i.e. unitary operator) can
237: be decomposed into a sequence of gates belonging to this set. In the
238: \emph{standard model} of quantum computation, one assumes that such
239: a set is available on the machine~\cite{DiVincenzo2000}. Also the
240: ability to perform measurements is assumed. We refer to this as the
241: \emph{full control} case.
242:
243: From a information theoretic point of view, qubits are not only \emph{useful}
244: objects to perform algorithms with, but also very interesting from
245: a fundamental point of view. To give a (too simple) analogy consider
246: the following. If you read the word ''chocolate'', you can associate
247: a positive/negative or neutral feeling of whether you would like to
248: eat some chocolate now. However, what was the state of your mind concerning
249: chocolate \emph{before you read} the word? \emph{}Unless you were
250: already craving for chocolate\index{chocolate} beforehand, or you
251: have just eaten a lot, your mind was probably \emph{undecided}. Moreover,
252: it would have been very difficult - if not impossible - to describe
253: to someone in plain language which opinion you had about the chocolate
254: before you read the word.
255:
256: In a similar manner, the quantum information contained in a single
257: \emph{arbitrary and unknown} qubit\index{arbitrary and unknown qubit}
258: cannot be described by classical information. When it is measured,
259: it behaves like a normal bit in the sense that the outcome is only
260: $0$ or $1,$ but when it is not measured, it behaves in some way
261: as if it was undecided between $0$ and 1. Of course one has to be
262: very careful with these analogies. But for the purpose of this thesis
263: it is important to stress that quantum information cannot be transported
264: by any classical methods~\cite{WERNER}. This is why it is so important
265: and also so difficult to develop new wires, dubbed \emph{quantum wires},
266: that are capable of doing this.
267:
268:
269: \section{Quantum state transfer along short distances\label{sec:Quantum-state-transfer}}
270:
271: In theory, additional devices for the transfer of unknown quantum
272: states are not required for building a quantum computer, unless it
273: is being used for typical quantum communication purposes, such as
274: secret key distribution \cite{DiVincenzo2000}. This is because the
275: \emph{universal set of gates} on the quantum computer can be used
276: to transfer quantum states by applying sequences of two-qubit swap
277: gates (Fig.~\ref{fig:swapping0}). %
278: \begin{figure}[htbp]
279: \begin{centering}\includegraphics[width=0.7\columnwidth]{swapping}\par\end{centering}
280:
281:
282: \caption{\label{fig:swapping0}In areas of universal control, quantum states
283: can easily be transferred by sequences of unitary swap gates $S_{j,k}$
284: between nearest neighbours.}
285: \end{figure}
286:
287:
288: However in practice it is crucial to minimise the required number
289: of quantum gates, as each gate typically introduces \emph{errors}.
290: In this light it appears costly to perform $N-1$ swap gates between
291: nearest neighbours to just move a qubit state over a distance of $N$
292: sites. For example, Shor's algorithm on $N$ qubits can be implemented
293: by only $\log N$ quantum gating operations~\cite{Cleve2000} if
294: long distant qubit gates are available. These long distant gates could
295: consist of local gates followed by a quantum state transfer. If however
296: the quantum state transfer is implemented as a sequence of local gates,
297: then the number of operations blows up to the order of $N$ gates.
298: The quantum state transfer can even be thought of as the source of
299: the \emph{power of quantum computation}, as any quantum circuit with
300: $\log N$ gates and \emph{local gates only} can be efficiently simulated
301: on a classical computer~\cite{Josza2006,Short2006}.
302:
303: A second reason to consider devices for quantum state transfer is
304: related to \emph{scalability}\index{scalability}. While small quantum
305: computers have already been built~\cite{Haffner2005}, it is very
306: difficult to build large arrays of fully controllable qubits. A \emph{black
307: box}\index{black box} that transports unknown quantum states could
308: be used to build larger quantum computers out of small components
309: by connecting them. Likewise, quantum state transfer can be used to
310: connect \emph{different} components of a quantum computer, such as
311: the processor and the memory (see also Fig.~\ref{fig:quantumcomputer}).
312: On larger distances, flying qubits such as photons, ballistic electrons
313: and guided atoms/ions are considered for this purpose~\cite{Skinner2003,Kielpinski2002}.
314: However, converting back and forth between stationary qubits and mobile
315: carriers of quantum information and interfacing between different
316: physical implementations of qubits is very difficult and worthwhile
317: only for short communication distances. This is the typical situation
318: one has to face in solid state systems, where quantum information
319: is usually contained in the states of \emph{fixed objects} such as
320: quantum dots or Josephson junctions. In this case permanently coupled
321: \emph{quantum chains} have recently been proposed as prototypes of
322: reliable quantum communication lines~\cite{Bose2003,LLOYD}. A quantum
323: chain\index{quantum chain} (also referred to as \emph{\index{spin chain}spin
324: chain}) is a one-dimensional array of qubits which are coupled by
325: some Hamiltonian (cf. Fig.~\ref{fig:qchain}). These couplings can
326: transfer states \emph{without external classical control.} In many
327: cases, such permanent couplings are easy to build in solid state devices
328: (in fact a lot of effort usually goes into \emph{suppressing} them).
329: The qubits can be of the \emph{same type} as the other qubits in the
330: device, so no interfacing is required.%
331: \begin{figure}[htbp]
332: \begin{centering}\includegraphics[width=0.7\columnwidth]{quantumcomputer}\par\end{centering}
333:
334:
335: \caption{\label{fig:quantumcomputer}Schematic layout of a quantum computer.
336: The solid arrows represent the flow of quantum information, and the
337: dashed arrows the flow of classical information.}
338: \end{figure}
339: %
340: \begin{figure}[htbp]
341: \begin{centering}\includegraphics[width=0.7\columnwidth]{qchain}\par\end{centering}
342:
343:
344: \caption{\label{fig:qchain}Permanently coupled quantum chains can transfer
345: quantum states without control along the line. Note that the ends
346: still need to be controllable to initialise and read out quantum states.}
347: \end{figure}
348:
349:
350: Another related motivation to consider quantum chains is that they
351: can simplify the \emph{layout} of quantum devices on wafers. A typical
352: chip can contain millions of qubits, and the fabrication of many qubits
353: is in principle no more difficult than the fabrication of a single
354: one. In the last couple of years, remarkable progress was made in
355: experiments with quantum dots~\cite{KOPPENS,HANSON} and super-conducting
356: qubits~\cite{YAMA,CHIO}. It should however be emphasised that for
357: initialisation, control and readout, those qubits have to be connected
358: to the macroscopic world (see Fig.~\ref{fig:quantumcomputer}). For
359: example, in a typical flux qubit gate, microwave pulses are applied
360: onto specific qubits of the sample. This requires many (classical)
361: wires on the chip, which is thus a \emph{compound} of quantum and
362: classical components. The macroscopic size of the classical control
363: is likely to be the bottleneck of the scalability as a whole. In this
364: situation, quantum chains are useful in order to keep some distance
365: between the controlled quantum parts. A possible layout for such a
366: quantum computer is shown in Fig.~\ref{fig:connect2}. It is built
367: out of blocks of qubits, some of which are dedicated to communication
368: and therefore connected to another block through a quantum chain.
369: Within each block, arbitrary unitary operations can be performed in
370: a fast and reliable way (they may be decomposed into single and two-qubit
371: operations). Such blocks do not currently exist, but they are the
372: focus of much work in solid state quantum computer architecture. The
373: distance between the blocks is determined by the length of the quantum
374: chains between them. It should be large enough to allow for classical
375: control wiring of each block, but short enough so that the time-scale
376: of the quantum chain communication is well below the time-scale of
377: decoherence in the system.%
378: \begin{figure}[htbp]
379: \begin{centering}\includegraphics[width=0.9\columnwidth]{connect5}\par\end{centering}
380:
381:
382: \caption{\label{fig:connect2}Small blocks (grey) of qubits (white circles)
383: connected by quantum chains. Each block consists of (say) 13 qubits,
384: 4 of which are connected to outgoing quantum chains (the thick black
385: lines denote their nearest-neighbour couplings). The blocks are connected
386: to the macroscopic world through classical wires (thin black lines
387: with black circles at their ends) through which arbitrary unitary
388: operations can be triggered on the block qubits. The quantum chains
389: require no external control.}
390: \end{figure}
391:
392:
393: Finally, an important reason to study quantum state transfer in quantum
394: chains stems from a more fundamental point of view. Such systems in
395: principle allow tests of Bell-inequalities and non-locality in solid-state
396: experiments well before the realisation of a quantum computer. Although
397: quantum transport is quite an established field, the quantum information
398: point of view offers many new perspectives. Here, one looks at the
399: transport of information rather than excitations, and at entanglement~\cite{Lakshminarayan2006,Subrahmanyam2004,Eisert2004,Palma2004}
400: rather than correlation functions. It has recently been shown that
401: this sheds new light on well-known physical phenomena such as quantum
402: phase transitions~\cite{Plenio,Plenio2006,Verstraete,Verstraete2004},
403: quantum chaos~\cite{Monteiro2006,Twamley2005,Santos2004,Boness2006}
404: and localisation~\cite{Winter,Apollaro2006}. Furthermore, quantum
405: information takes on a more \emph{active} attitude. The correlations
406: of the system are not just calculated, but one also looks at how they
407: may be \emph{changed}.
408:
409:
410: \section{Implementations and experiments\index{experiments}\label{sec:Implementations-and-experiments}}
411:
412: As we have seen above, the main advantage of state transfer with quantum
413: chains is that the qubits can be of the same type as those used for
414: the quantum computation. Therefore, most systems that are thought
415: of as possible realisations of a quantum computer can also be used
416: to build quantum chains. Of course there has to be some coupling between
417: the qubits. This is typically easy to achieve in solid state systems,
418: such as Josephson junctions with charge qubits~\cite{Bruder2005a,Tsomokos2006},
419: flux qubits~\cite{Bruder,Bruder2005}\index{flux qubits} (see also
420: Fig.~\ref{fig:A-quantum-chain}) or quantum dots dots using the electrons~\cite{Loss1998,Greentree2004}
421: or excitons~\cite{DAmico,Lovett}. Other systems where quantum chain
422: Hamiltonians can at least be \emph{simulated} are NMR qubits~\cite{Suter2006,Jones,Zhang2005}
423: and optical lattices~\cite{Garcia-Ripoll2003}. Such a simulation
424: is particularly useful in the latter case, where local control is
425: extremely difficult. Finally, qubits in cavities~\cite{Falci2005,Paternostro2005}
426: and coupled arrays of cavities were considered \cite{Bose,Plenioa}.%
427: \begin{figure}[tbh]
428: \begin{centering}\includegraphics[width=0.65\paperwidth]{delft}\par\end{centering}
429:
430:
431: \caption{\label{fig:A-quantum-chain}A quantum chain consisting of $N=20$
432: flux qubits~\cite{Bruder2005} (picture and experiment by Floor Paauw,
433: TU Delft). The chain is connected to four larger SQUIDS for readout
434: and gating.}
435: \end{figure}
436:
437:
438: For the more fundamental questions, such as studies of entanglement
439: transfer, non-locality and coherent transport, the quantum chains
440: could also be realised by systems which are not typically thought
441: of as qubits, but which are \emph{natural spin chains.} These can
442: be molecular systems~\cite{EXCITON} or quasi-1D solid state materials~\cite{Motoyama1996,Gambardella2002}.
443:
444:
445: \section{Basic communication protocol\label{sec:Basic-transport-protocol}}
446:
447: We now review the most basic transport protocol for quantum state
448: transfer, initially suggested in~\cite{Bose2003}. For the sake of
449: simplicity, we concentrate on the linear chain setting, though more
450: general graphs of qubits can be considered in the same way. The protocol
451: consists of the following steps:
452:
453: \begin{enumerate}
454: \item Initialise the quantum chain in the ground state \begin{equation}
455: |G\rangle.\end{equation}
456:
457: \item Put an arbitrary and unknown qubit with (possibly mixed) state $\rho$
458: at the sending end of the chain \begin{equation}
459: \rho\otimes\mbox{Tr}_{1}\left\{ |G\rangle\langle G|\right\} .\end{equation}
460:
461: \item Let the system evolve under its Hamiltonian $H$ for a time $t$\begin{equation}
462: \exp\left\{ -iHt\right\} \rho\otimes\mbox{Tr}_{1}\left\{ |G\rangle\langle G|\right\} \exp\left\{ iHt\right\} .\end{equation}
463:
464: \item Pick up the quantum state at the end of the chain\begin{equation}
465: \sigma\equiv\mbox{Tr}_{1,\ldots,N-1}\left[\exp\left\{ -iHt\right\} \rho\otimes\mbox{Tr}_{1}\left\{ |G\rangle\langle G|\right\} \exp\left\{ iHt\right\} \right].\end{equation}
466:
467: \end{enumerate}
468: Some practical aspects how to realise these steps are discussed in
469: the next section. For the moment, we will concentrate on the \emph{quality}
470: of quantum state transfer given that the above steps can be performed.
471: From a quantum information perspective, the above equations describe
472: a \emph{quantum channel}\index{quantum channel}~\cite{WERNER} $\tau$
473: that maps input states $\rho$ at one end of the chain to output states
474: $\tau(\rho)=\sigma$ on the other end. A very simple measure of the
475: quality of such a quantum channel is the \emph{fidelity}~\cite{Uhlmann1976,Jozsa1994,NIELSEN}\emph{\index{fidelity}}
476: \begin{equation}
477: F(\rho,\sigma)\equiv\left(\mbox{Tr}\sqrt{\rho^{1/2}\sigma\rho^{1/2}}\right)^{2}.\end{equation}
478: More advanced measures of the quality of transfer will be discussed
479: in Chapter \ref{cha:Multi-rail-and-Capacity}. Note also that some
480: authors define the fidelity without taking the square of the trace.
481: It is a real-valued, symmetric function with range between $0$ and
482: $1,$ assuming unity if and only if $\rho=\sigma.$ Since the transported
483: state that is an unknown result of some quantum computation, we are
484: interested in the \emph{minimal fidelity}\index{minimal fidelity}\begin{equation}
485: F_{0}\equiv\mbox{min}_{\rho}F(\rho,\tau(\rho)).\label{eq:minfidelity}\end{equation}
486: We remark that some authors also assume an equal distribution of input
487: states and compute the \emph{average fidelity}~\cite{Bose2003}.
488: Using the strong concavity of the fidelity~\cite{NIELSEN} and the
489: linearity of $\tau$ we find that the minimum must be assumed on pure
490: input states,\begin{equation}
491: F_{0}=\mbox{min}_{\psi}\langle\psi|\tau(\psi)|\psi\rangle.\label{eq:minfidelity2}\end{equation}
492: In the present context, $F_{0}=F_{0}(H,t)$ is a function of of the
493: Hamiltonian $H$ of the quantum chain (through the specific role of
494: the ground state in the protocol and through the time evolution),
495: and of the time interval $t$ that the system is evolving in the third
496: step of the protocol.
497:
498:
499: \subsection{Initialisation and end-gates\label{sec:Initialisation}}
500:
501: There are two strong assumptions in the protocol from the last section.
502: The first one is that the chain can be initialised in the ground state
503: $|G\rangle.$ How can that be achieved if there is no local control
504: along the chain? The answer appears to be quite easy: one just applies
505: a strong global magnetic field and strong cooling (such as laser cooling
506: or dilution refrigeration) and lets the system reach its ground state
507: by relaxation. The cooling needs to be done for the remaining parts
508: of the quantum computer anyway, so no extra devices are required.
509: However there is a problem with the time-scale of the relaxation.
510: If the system is brought to the ground state by cooling, it must be
511: coupled to some environment. But during the quantum computation, one
512: clearly does not want such an environment. This is usually solved
513: by having the time-scale of the computation much smaller (say microseconds)
514: than the time-scale of the cooling (say seconds or minutes). But if
515: the quantum chain should be used multiple times during one computation,
516: then how is it reset between each usage? This is important to avoid
517: memory effects~\cite{Werner2005}, and there are two solutions to
518: this problem. Either the protocol is such that at the end the chain
519: is automatically in the ground state. Such a protocol usually corresponds
520: to \emph{perfect state transfer.} The other way is to use the control
521: at the ends of the chain to bring it back to the ground state. A simple
522: \emph{\index{cooling protocol}cooling protocol} is given by the following:
523: one measures the state of the last qubit of the chain. If it is in
524: $|0\rangle,$ then one just lets the chain evolve again and repeats.
525: If however it is found to be in $|1\rangle,$ one applies the Pauli
526: operator $X$ to flip it before evolving and repeating. It will become
527: clear later on in the thesis that such a protocol typically converges
528: exponentially fast to the ground state of the chain.
529:
530: The second assumption in the last section is that the sender and receiver
531: are capable of swapping in and out the state much quicker than the
532: time-scale of the interaction of the chain. Alternatively, it is assumed
533: that they can switch on and off the interaction between the chain
534: and their memory in such time-scale. It has recently been shown~\cite{Bruder}
535: that this is not a fundamental problem, and that finite switching
536: times can even slightly improve the fidelity if they are carefully
537: included in the protocol. But this requires to solve the full time-dependent
538: Schrödinger equation, and introduces further parameters to the model
539: (i.e. the raise and fall time of the couplings). For the sake of simplicity,
540: we will therefore assume that the end gates are much faster then the
541: time evolution of the chain (see also Section~\ref{sec:Practical-Considerations}).
542:
543:
544: \subsection{Symmetries}
545:
546: The dimensionality of the Hilbert space $\mathcal{H}$ of a quantum
547: chain of $N$ qubits is $2^{N}.$ This makes it quite hopeless in
548: general to determine the minimal fidelity Eq.~(\ref{eq:minfidelity2})
549: for long quantum chains. Most investigations on quantum state transfer
550: with quantum chains up to date are therefore concentrating on Hamiltonians
551: with additional symmetries. With few exceptions \cite{Bruder2005,Plenio,Plenio2006,Kay}
552: Hamiltonians that conserve the number of excitations are considered.
553: In this case the Hilbert space is a direct sum of subspaces invariant
554: under the time evolution,\begin{equation}
555: \mathcal{H}=\bigoplus_{\ell=0}^{N}\mathcal{H}_{\ell},\end{equation}
556: with $\mbox{dim}\mathcal{H}_{\ell}=\binom{N}{\ell},$ and where $\ell$
557: is the number of excitations. These Hamiltonians are \emph{much} easier
558: to handle both analytically and numerically, and it is also easier
559: to get an intuition of the dynamics. Furthermore, they occur quite
560: naturally as a coupling between qubits in the relevant systems. We
561: stress though that there is \emph{no fundamental} reason to restrict
562: quantum chain communication to this case.
563:
564:
565: \subsection{Transfer functions\index{transfer functions}}
566:
567: The space $\mathcal{H}_{0}$ only contains the state $|\boldsymbol{0}\rangle$
568: which is thus always an eigenstate of $H.$ We will assume here that
569: it is also the ground state,\begin{equation}
570: |G\rangle=|\boldsymbol{0}\rangle.\end{equation}
571: This can be achieved by applying a strong global magnetic field (or
572: equivalent) to the system. The space $\mathcal{H}_{1}$ is spanned
573: by the vectors $\left\{ |\boldsymbol{k}\rangle,k=1,\ldots,N\right\} $
574: having exactly one excitation. The above protocol becomes:
575:
576: \begin{enumerate}
577: \item Initialise the quantum chain in the ground state \begin{equation}
578: |\boldsymbol{0}\rangle\end{equation}
579:
580: \item Put an arbitrary and unknown qubit in the pure state $|\psi\rangle=\alpha|0\rangle+\beta|1\rangle$
581: at the sending end of the chain \begin{equation}
582: \alpha|\boldsymbol{0}\rangle+\beta|\boldsymbol{1}\rangle\end{equation}
583:
584: \item Let the system evolve for a time $t$\begin{equation}
585: \alpha|\boldsymbol{0}\rangle+\beta\exp\left\{ -iHt\right\} |\boldsymbol{1}\rangle\end{equation}
586:
587: \item Pick up the quantum state at the end of the chain (see~\cite{Bose2003})\begin{equation}
588: \tau(\psi)=(1-p(t))|0\rangle\langle0|+p(t)|\psi\rangle\langle\psi|,\end{equation}
589: with the minimal fidelity given by\begin{align}
590: F_{0} & =\mbox{min}_{\psi}\langle\psi|\tau(\psi)|\psi\rangle\label{eq:minfidelity3}\\
591: & =p(t)+(1-p(t))\mbox{min}_{\psi}\left|\langle0|\psi\rangle\right|^{2}=p(t).\end{align}
592:
593: \end{enumerate}
594: The function $p(t)$ is the transition probability from the state
595: $|\boldsymbol{1}\rangle$ to $|\boldsymbol{N}\rangle$ given by\begin{equation}
596: p(t)=\left|\langle\boldsymbol{N}|\exp\left\{ -iHt\right\} |\boldsymbol{1}\rangle\right|^{2}.\end{equation}
597: We see that in the context of quantum state transfer, a \emph{single}
598: parameter suffices to characterise the properties of an excitation
599: conserving chain. The averaged fidelity~\cite{Bose2003} is also
600: easily computed as\begin{equation}
601: \bar{F}=\frac{\sqrt{p(t)}}{3}+\frac{p(t)}{6}+\frac{1}{2}.\label{eq:averagedfid}\end{equation}
602: Even more complex measures of transfer such as the quantum capacity
603: only depend on $p(t)$~\cite{Giovannetti2005}. It is also a physically
604: intuitive quantity, namely a particular matrix element of the time
605: evolution operator, \begin{align}
606: f_{n,m}(t) & \equiv\langle\boldsymbol{n}|\exp\left\{ -iHt\right\} |\boldsymbol{m}\rangle\label{eq:spintrans}\\
607: & =\sum_{k}e^{-iE_{k}t}\langle\boldsymbol{n}|E_{k}\rangle\langle E_{k}|\boldsymbol{m}\rangle,\end{align}
608: where $|E_{k}\rangle$ and $E_{k}$ are the eigenstates and energy
609: levels of the Hamiltonian in $\mathcal{H}_{1}.$
610:
611:
612: \subsection{Heisenberg Hamiltonian}
613:
614: The Hamiltonian chosen in~\cite{Bose2003} is a \index{Heisenberg Hamiltonian}Heisenberg
615: Hamiltonian\begin{equation}
616: H=-\frac{J}{2}\sum_{n=1}^{N-1}\left(X_{n}X_{n+1}+Y_{n}Y_{n+1}+Z_{n}Z_{n+1}\right)-B\sum_{n=1}^{N}Z_{n}+c,\label{eq:heisenberg}\end{equation}
617: with a constant term \begin{equation}
618: c=\frac{J(N-1)}{2}+NB\end{equation}
619: added to set the ground state energy to $0.$ For $J>0$ it fulfils
620: all the assumptions discussed above, namely its ground state is given
621: by $|\boldsymbol{0}\rangle$ and it conserves the number of excitations
622: in the chain. The Heisenberg interaction is very common and serves
623: here as a typical and analytically solvable model for quantum state
624: transfer.
625:
626: In the first excitation subspace $\mathcal{H}_{1}$, the Heisenberg
627: Hamiltonian Eq.~(\ref{eq:heisenberg}) is expressed in the basis
628: $\left\{ |\boldsymbol{n}\rangle\right\} $ as \begin{equation}
629: \left(\begin{array}{cccccc}
630: 1 & -1\\
631: -1 & 2 & -1\\
632: & -1 & 2 & \ddots\\
633: & & \ddots & \ddots & -1\\
634: & & & -1 & 2 & -1\\
635: & & & & -1 & 1\end{array}\right).\label{eq:matrix}\end{equation}
636: A more general study of such \emph{tridiagonal} matrices can be found
637: in a series of articles on coherent dynamics~\cite{Eberly1977,Bialynicka-Birula1977,Cook1979,Shore1979}.
638: Some interesting analytically solvable models have also been identified~\cite{Ekert2004,Bialynicka-Birula1977,Cook1979}
639: (we shall come back to that point later).
640:
641: For the present case, the eigenstates of Eq~(\ref{eq:matrix}) are~\cite{Bose2003}
642: \begin{equation}
643: |E_{k}\rangle=\sqrt{\frac{1+\delta_{k0}}{N}}\sum_{n=1}^{N}\cos\left[\frac{\pi k}{2N}(2n-1)\right]|\boldsymbol{n}\rangle\quad(k=0,\ldots,N-1),\end{equation}
644: with the corresponding energies given by\begin{equation}
645: E_{k}=2B+2J\left[1-\cos\frac{\pi k}{N}\right].\label{eq:eigenfrequencies}\end{equation}
646: The parameter $B$ has no relevance for the fidelity but determines
647: the stability of the ground state (the energy of the first excited
648: state is given by $2B).$ The minimal fidelity\index{minimal fidelity}
649: for a Heisenberg chain is given by
650:
651: \begin{equation}
652: \boxed{p(t)=N^{-2}\left|1+\sum_{k=1}^{N-1}\exp\left\{ -2iJt(1-\cos\frac{\pi k}{N})\right\} (-1)^{k}\left(1+\cos\frac{\pi k}{N}\right)\right|^{2}.}\label{eq:resultfid}\end{equation}
653: As an example, Fig~\ref{fig:fidelity} shows $p(t)$ for $N=50$.
654: %
655: \begin{figure}[htbp]
656: \begin{centering}\includegraphics[width=0.5\paperwidth]{fidelity}\par\end{centering}
657:
658:
659: \caption{\label{fig:fidelity}Minimal fidelity $p(t)$ for a Heisenberg chain
660: of length $N=50.$}
661: \end{figure}
662:
663:
664:
665: \subsection{Dynamic and Dispersion\label{sub:Dynamics-and-Dispersion}}
666:
667: Already in~\cite{Bose2003} has been realised that the fidelity for
668: quantum state transfer along spin chains will in general not be perfect.
669: The reason for the imperfect transfer is the \emph{dispersion}\index{dispersion}~\cite{Linden2004}
670: of the information along the chain. Initially the quantum information
671: is localised at the sender, but as it travels through the chain it
672: also spreads (see Fig.~\ref{fig:snapshots} and Fig.~\ref{fig:dispersion}).
673: This is not limited to the Heisenberg coupling considered here, but
674: a very common quantum effect. Due to the dispersion, the probability
675: amplitude peak that reaches Bob is typically small, and becomes even
676: smaller as the chains get longer. %
677: \begin{figure}[htbp]
678: \begin{centering}\includegraphics[width=0.55\paperwidth]{50}\par\end{centering}
679:
680:
681: \caption{\label{fig:snapshots}Snapshots of the time evolution of a Heisenberg
682: chain with $N=50.$ Shown is the distribution $|f_{n,1}(t)|^{2}$
683: of the wave-function in space at different times if initially localised
684: at the first qubit.}
685: \end{figure}
686: %
687: \begin{figure}[htbp]
688: \begin{centering}\includegraphics[width=0.55\paperwidth]{dispersion}\par\end{centering}
689:
690:
691: \caption{\label{fig:dispersion}Mean and variance of the state $|\boldsymbol{1}\rangle$
692: as a function of time. Shown is the case $N=50$ with the y-axis giving
693: the value \emph{relative} to the mean $N/2+1$ and variance $(N^{2}-1)/12$
694: of an equal distribution $\frac{1}{\sqrt{N}}\sum|\boldsymbol{n}\rangle.$}
695: \end{figure}
696:
697:
698: The fidelity given Eq.~(\ref{eq:resultfid}) is shown in Fig.~\ref{fig:fidelity}.
699: We can see that a wave of quantum information is travelling across
700: the chain. It reaches the other end at a time of approximately \begin{equation}
701: t_{\mbox{peak}}\approx\frac{N}{2J}\end{equation}
702: As a rough estimate of the scaling of the fidelity with respect to
703: the chain length around this peak we can use~\cite{Bose2003,Abramowitz1972}
704: (see also Fig.~\ref{fig:airyapprox}) \begin{equation}
705: |f_{N,1}(t)|^{2}\approx|2J_{N}(\frac{2t}{J})|^{2}\approx|\left(\frac{16}{N}\right)^{1/3}\mbox{ai}[\left(\frac{2}{N}\right)^{1/3}(N-\frac{2t}{J})]|^{2},\label{eq:approxf}\end{equation}
706: where $J_{N}(x)$ is a Bessel function of first kind and $ai(x)$
707: is the Airy function. The airy function $ai(x)$ has a maximum of
708: $0.54$ at $x=-1.02.$ Hence we have \begin{equation}
709: p(t_{\mbox{peak}})=|f_{N,1}(\frac{N}{2J})|^{2}\approx1.82N^{-2/3}.\label{eq:estimate}\end{equation}
710: It is however possible to find times where the fidelity of the chain
711: is much higher. The reason for this is that the wave-packet is reflected
712: at the ends of the chain and starts interfering with itself (Fig~\ref{fig:fidelity}).
713: As the time goes on, the probability distribution becomes more and
714: more random. Sometimes high peaks at the receiving end occur. From
715: a theoretical point of view, it is interesting to determine the \emph{maximal
716: peak}\index{maximal peak} occurring, i.e. \begin{equation}
717: p_{M}(T)\equiv\max_{0<t<T}p(t).\label{eq:maximal}\end{equation}
718: As we can see in Fig.~\ref{fig:longtime} there is quite a potential
719: to improve from the estimate Eq.~(\ref{eq:estimate}).%
720: \begin{figure}[htbp]
721: \begin{centering}\includegraphics[width=0.55\paperwidth]{airyapprox}\par\end{centering}
722:
723:
724: \caption{\label{fig:airyapprox}Approximation of the transfer amplitude for
725: $N=50$ around the first maximum by Bessel and Airy functions \cite{Bose2003,Abramowitz1972}. }
726: \end{figure}
727: %
728: \begin{figure}[htbp]
729: \begin{centering}\includegraphics[width=0.55\paperwidth]{longtimes}\par\end{centering}
730:
731:
732: \caption{\label{fig:longtime}$p_{M}(T)$ as a function of $T$ for different
733: chain lengths. The solid curve is given by $1.82{(2T)}^{-2/3}$ and
734: corresponds to the first peak of the probability amplitude (Eq.~\ref{eq:estimate})}
735: \end{figure}
736:
737:
738: We will now show a perhaps surprising connection of the function $p_{M}(T)$
739: to number theory. Some speculations on the dependence of the fidelity
740: on the chain length being divisible by $3$ were already made in~\cite{Bose2003},
741: but not rigorously studied. As it turns out, for chains with \emph{prime
742: number length} the maximum of the fidelity is actually converging
743: to unity (see Fig.~\ref{fig:longtime}). To show this, we first prove
744: the following
745:
746: \begin{lemma}
747: \label{lem:primindependence}Let $N$ be an odd prime. Then the set\begin{equation}
748: \left\{ \cos\frac{k\pi}{N}\quad(k=0,1,\ldots,,\frac{N-1}{2})\right\} \end{equation}
749: is linear independent over the rationals $\mathbb{Q}$.
750: \end{lemma}
751: \begin{proof}
752: Assume that \begin{equation}
753: \sum_{k=0}^{\frac{N-1}{2}}\lambda_{k}\cos\frac{k\pi}{N}=0\end{equation}
754: with $\lambda_{k}\in\mathbb{Q}.$ It follows that \begin{equation}
755: \sum_{k=0}^{\frac{N-1}{2}}\lambda_{k}\left\{ \exp\frac{ik\pi}{N}+\exp\frac{-ik\pi}{N}\right\} =0\end{equation}
756: and hence\begin{equation}
757: \sum_{k=0}^{\frac{N-1}{2}}\lambda_{k}\exp\frac{ik\pi}{N}-\sum_{k=0}^{\frac{N-1}{2}}\lambda_{k}\exp\frac{i(N-k)\pi}{N}=0.\end{equation}
758: Changing indexes on the second sum we get\begin{equation}
759: \sum_{k=0}^{\frac{N-1}{2}}\lambda_{k}\exp\frac{ik\pi}{N}-\sum_{k=\frac{N+1}{2}}^{N}\lambda_{N-k}\exp\frac{ik\pi}{N}=0.\end{equation}
760: and finally \begin{equation}
761: \sum_{k=0}^{N-1}\tilde{\lambda}_{k}\exp\frac{ik\pi}{N}=0,\label{eq:exp}\end{equation}
762: where\begin{align}
763: \tilde{\lambda}_{0} & =2\lambda_{0}\\
764: \tilde{\lambda}_{k} & =\lambda_{k}\quad(k=1,\ldots,\frac{N-1}{2})\\
765: \tilde{\lambda}_{k} & ={-\lambda}_{N-k}\quad(k=\frac{N+1}{2},\ldots,N-1).\end{align}
766: Since $N$ is prime, the roots of unity in Eq.~(\ref{eq:exp}) are
767: all primitive and therefore linearly independent over $\mathbb{Q}$~\cite[Theorem 3.1, p. 313]{Lang1984}.
768: Hence $\lambda_{k}=0$ for all $k.$
769: \end{proof}
770: \begin{framedtheorem}
771: [Half recurrence]Let $N$ be an odd prime. For a Heisenberg chain
772: of length $N$ we have\begin{equation}
773: \lim_{T\rightarrow\infty}p_{M}(T)=\lim_{T\rightarrow\infty}\left[\max_{0<t<T}p(t)\right]=1.\end{equation}
774:
775: \end{framedtheorem}
776: \begin{proof}
777: The eigenfrequencies of the Hamiltonian in the first excitation sector
778: $\mathcal{H}_{1}$ are given by \begin{equation}
779: E_{k}=2B+2J\left[1-\cos\frac{\pi k}{N}\right]\quad(k=0,1,\ldots,,N-1).\label{eq:eigenfrequencies2}\end{equation}
780: Using Kronecker's theorem~\cite{Hemmer1958} and Lemma~\ref{lem:primindependence},
781: the equalities \begin{equation}
782: \exp\left\{ itE_{k}\right\} =\left(-1\right)^{k}e^{2(B+J)t}\quad(k=0,1,\ldots,,\frac{N-1}{2})\label{eq:set}\end{equation}
783: can be fulfilled \emph{arbitrarily well} by choosing an appropriate
784: $t.$ Since\begin{equation}
785: \cos\frac{k\pi}{N}=-\cos\frac{(N-k)\pi}{N},\end{equation}
786: the equalities~(\ref{eq:set}) are then also fulfilled arbitrarily
787: well for $k=0,\ldots,N-1.$ This is known as as sufficient condition
788: for perfect state transfer in mirror symmetric chains~\cite{Bose2005},
789: where the eigenstates can be chosen such that they are alternately
790: symmetric and antisymmetric. Roughly speaking, Eq.~(\ref{eq:set})
791: introduces the correct phases (a sign change for the antisymmetric
792: eigenstates) to move the state $|1\rangle$ to $|N\rangle$ and hence
793: the theorem.
794: \end{proof}
795: \begin{remark}
796: The time-scale for finding high valued peaks is however \emph{exponential}
797: in the chain length~\cite{Hemmer1958}. Therefore the above theorem
798: has little practical use. For non-prime chain lengths, the eigenfrequencies
799: are not sufficiently independent to guarantee a perfect state transfer,
800: with the algebraic dimensionality of the roots of unity for non-prime
801: $N$ given by the Euler totient function $\phi(N)$~\cite[Theorem 3.1, p. 313]{Lang1984}.
802: We also remark that due to its asymptotic character, the above result
803: is not contradicting~\cite{Landahl2005}, where it was shown that
804: chains longer than $N\ge4$ never have perfect fidelity.
805: \end{remark}
806: Having proved that there are many chains that can in principle perform
807: arbitrarily well, it is important to find a cut-off time for the optimisation
808: Eq.~(\ref{eq:maximal}). Faster transfer than linear in $N$ using
809: local Hamiltonians is impossible due to the Lieb-Robinson bound~\cite{Bravyi2006,Nachtergaele2006},
810: which is a ''speed limit'' in non-relativistic quantum mechanics
811: giving rise to a well defined group velocity. Transport faster than
812: this group velocity is exponentially suppressed. Going back to the
813: motivation of quantum state transfer, a natural comparison~\cite{DAmico}
814: for the time-scale of quantum state transfer is given by the time
815: it would take to perform a sequence of swap gates\index{swap gates}
816: (cf. Fig~\ref{fig:swapping0}) that are realised by a pairwise switchable
817: coupling Hamiltonian \begin{equation}
818: \frac{J}{2}(X_{n}X_{n+1}+Y_{n}Y_{n+1}).\end{equation}
819: This time is linear in the chain length: \begin{equation}
820: t_{\mbox{swap}}=\frac{(N-1)\pi}{2J}.\end{equation}
821: Ideally one could say that the time for quantum state transfer should
822: not take much longer than this. However one may argue that there is
823: a trade-off between quick transfer on one hand, and minimising control
824: on the other hand. A second cut-off time may be given by the \emph{decoherence
825: time} of the specific implementation. But short decoherence times
826: could always be counteracted by increasing the chain coupling $J.$
827: A more general and implementation independent limit is given by the
828: requirement that the \emph{peak width\index{peak width}} ${\Delta t}_{\mbox{peak}}$
829: should not be too small with respect to the total time. Otherwise
830: it is difficult to pick up the state at the correct time. For the
831: first peak, we can estimate the width by using the full width at half
832: height of the airy function. From Eq.~(\ref{eq:approxf}) we get
833: an absolute peak width of $\Delta t_{\mbox{peak}}\approx0.72N^{1/3}/J$
834: and a relative width of\begin{equation}
835: \frac{\Delta t_{\mbox{peak}}}{t_{\mbox{peak}}}\approx1.44N^{-2/3}.\end{equation}
836: This is already quite demanding from an experimental perspective and
837: we conclude that the transfer time should not be chosen much longer
838: than those of the first peak.
839:
840:
841: \subsection{How high should $p(t)$ be?\label{sub:How-high-should}}
842:
843: We have not discussed yet what the actual value of $p(t)$ should
844: be to make such a spin chain useful as a device for quantum state
845: transfer. $p(t)=0$ corresponds to no state transfer, $p(t)=1$ to
846: a perfect state transfer. But what are the relevant scales for intermediate
847: $p(t)$? In practice, the quantum transfer will suffer from additional
848: external noise (Chapter~\ref{cha:Problems-and-Practical}) and also
849: the quantum computer itself is likely to be very noisy. From this
850: point of view, requiring $p(t)=1$ seems a bit too demanding.
851:
852: From a theoretical perspective, it is interesting that for any $p(t)>0,$
853: one can already do things which are impossible using classical channels,
854: namely entanglement transfer\index{entanglement transfer} and distillation\index{distillation}~\cite{NIELSEN}.
855: The entanglement of formation\index{entanglement of formation} between
856: the sender (\emph{Alice}) and the receiver (\emph{Bob}) is simply
857: given by $\sqrt{p(t)}$~\cite{Bose2003}. This entanglement can be
858: partially distilled~\cite{Horodecki1997} into singlets, which could
859: be used for state transfer using teleportation~\cite{NIELSEN}. It
860: is however not known \emph{how much}, i.e. at which rate, entanglement
861: can be distilled (we will develop lower bounds for the entanglement
862: of distillation\index{entanglement of distillation} in Section~\ref{sec:con}
863: and Section~\ref{s:sec3}). Also, entanglement distillation is a
864: quite complex procedure that requires local unitary operations and
865: measurements, additional classical communication, and multiple chain
866: usages; and few explicit protocols are known. This is likely to preponderate
867: the benefits of using a quantum chain.
868:
869: When the chain is used without encoding and further operation, the
870: averaged fidelity Eq.~(\ref{eq:averagedfid}) becomes better than
871: the classical%
872: \footnote{By ''classical fidelity'', we mean the fidelity that can be achieved
873: by optimising the following protocol: Alice performs measurements
874: on her state and sends Bob the outcome through a classical communication
875: line. Bob then tries to rebuild the state that Alice had before the
876: measurement based on the information she sent. For qubits, the classical
877: fidelity is given by $2/3$\cite{Horodecki1999}.%
878: } averaged fidelity\index{classical averaged fidelity}~\cite{Bose2003}
879: when $p(t)>3-2\sqrt{2}.$ Following the conclusion from the last subsection
880: that the first peak is the most relevant one, this would mean that
881: only chains with length until $N=33$ perform better than the classical
882: fidelity.
883:
884: Finally, the \emph{quantum capacity}\index{quantum capacity}~\cite{Giovannetti2005,SHOR}
885: of the channel becomes non-zero only when $p(t)>1/2,$ corresponding
886: to chain lengths up to $N=6.$ Roughly speaking, it is a measure of
887: the number of perfectly transmitted qubits per chain usage that can
888: be achieved asymptotically using encoding and decoding operations
889: on multiple channel usages. The quantum capacity considered here is
890: not assumed to be assisted by a classical communication, and the threshold
891: of $p(t)>0.5$ to have a non-zero quantum capacity is a result of
892: the non-cloning theorem~\cite{NIELSEN}. This is not contradicting
893: the fact that entanglement distillation is possible for \emph{any}
894: $p(t)>0,$ as the entanglement distillation protocols require additional
895: classical communication.
896:
897: All the above points are summarised in Fig.~\ref{fig:capacity}.
898: We can see that only very short chains reach reasonable values (say
899: $>0.6$) for the minimal fidelity. %
900: \begin{figure}[htbp]
901: \begin{centering}\includegraphics[width=0.6\paperwidth]{capacity}\par\end{centering}
902:
903:
904: \caption{\label{fig:capacity}Quantum capacity, entanglement of formation
905: (EOF), a lower bound for the entanglement of distillation (EOD) and
906: the averaged fidelity as a function of $p(t).$ We also show the corresponding
907: chain length which reaches this value as a first peak and the classical
908: threshold $3-2\sqrt{2}$. The explicit expression for the quantum
909: capacity plotted here is given in~\cite{Giovannetti2005}, and the
910: lower bound of the entanglement of distillation will be derived in
911: Section~\ref{s:sec3}.}
912: \end{figure}
913:
914:
915:
916: \section{Advanced communication protocols\label{sec:Advanced-transfer-protocols}}
917:
918: We have seen in the last section that without much further effort,
919: i.e. entanglement distillation, unmodulated Heisenberg chains are
920: useful only when they are very short. Shortly after the initial proposal~\cite{Bose2003}
921: it has been shown that there are ways to achieve even \emph{perfect
922: state transfe}r with arbitrarily long chains. These advanced proposals
923: can roughly be grouped into four categories, which we will now briefly
924: describe.
925:
926:
927: \subsection{Engineered Hamiltonians\label{sub:Engineered-Hamiltonians}}
928:
929: The Heisenberg model chosen by Bose features many typical aspects
930: of coherent transport, i.e. the wave-like behaviour, the dispersion,
931: and the almost-periodicity of the fidelity. These features do not
932: depend so much on the specific choices of the parameters of the chain,
933: such as the couplings strengths. There are however \emph{specific
934: couplings} for quantum chains that show a quite different time evolution,
935: and it was suggested in~\cite{Landahl2004} and independently in~\cite{Lambropoulos2004a}
936: to use these to achieve a \emph{perfect} state transfer: \begin{equation}
937: H=-J\sum_{n=1}^{N-1}\sqrt{n(N-n)}\left(X_{n}X_{n+1}+Y_{n}Y_{n+1}\right)\label{eq:engineered}\end{equation}
938: These values for engineered couplings\index{engineered couplings}
939: also appear in a different context in~\cite{Cook1979,Peres1985}.
940: The time evolution under the Hamiltonian~(\ref{eq:engineered}) features
941: an additional \emph{mirror symmetry:} the wave-packet disperses initially,
942: but the dispersion is reversed after its centre has passed the middle
943: of the chain (Fig.~\ref{fig:snapshotseng}).%
944: \begin{figure}[htbp]
945: \begin{centering}\includegraphics[width=0.6\paperwidth]{50_eng}\par\end{centering}
946:
947:
948: \caption{\label{fig:snapshotseng}Snapshots of the time evolution of a quantum
949: chain with engineered couplings~(\ref{eq:engineered}) for $N=50.$
950: Shown is the distribution of the wave-function in space at different
951: times if initially localised at the first qubit (compare Fig.~\ref{fig:snapshots}).}
952: \end{figure}
953: This approach has been extended by various authors~\cite{Bose2005,Yung2006,Sun2006a,Sun2005c,Sun2005a,Lambropoulos2006,Lambropoulos2004,Ericsson2005,Kay,Kay2006,Kay2006a,Stolze2005,Jing-Fu2006,Landahl2005,Ekert2004,Eisert2004},
954: and many other choices of parameters for perfect or near perfect state
955: transfer in various settings were found~\cite{Ekert2004,Stolze2005,Kay2006}.
956:
957:
958: \subsection{Weakly coupled sender and receiver\label{sub:Weakly-coupled-sender}}
959:
960: A different approach of tuning the Hamiltonian was suggested in~\cite{Semiao2005}.
961: There, only the first and the last couplings $j$ of the chain are
962: engineered to be \emph{much weaker} than the remaining couplings $J$
963: of the chain, which can be quite arbitrary. The fidelity can be made
964: arbitrarily high by making the edge coupling strengths smaller. It
965: was shown~\cite{Bednarska,Bednarska2005} that to achieve a fidelity
966: of $1-\delta$ in a chain of odd length, it takes approximately a
967: time of
968:
969: \begin{equation}
970: 2N\pi/\sqrt{\delta}\end{equation}
971: and the coupling ratio has to be approximately $j/J\approx\sqrt{\delta/N}.$
972: Some specific types of quantum chains which show high fidelity for
973: similar reasons were also investigated~\cite{Sun2005,Bose2006,Bose2006a,Pasquale2006}.
974:
975:
976: \subsection{Encoding\label{sub:Encoding}}
977:
978: We have seen in Subsec.~\ref{sub:How-high-should} that if $p(t)<1/2,$
979: the fidelity cannot be improved by using any encoding/decoding strategy
980: (because the quantum capacity is zero). However it is possible to
981: \emph{change the protocol} described in Sec.~\ref{sec:Basic-transport-protocol}
982: slightly such that the fidelity is much higher\emph{.} This can be
983: thought of as a \emph{''}hardware encoding\emph{''}, and was suggested
984: first in~\cite{Linden2004}. There, it was assumed that the chain
985: consists of three sections: one part of length $\approx2N^{1/3}$
986: controlled by the sending party, one ''free'' part of length $N$
987: and one part of length $\approx2.8N^{1/3}$ controlled by the receiving
988: party. The sender encodes the qubit not only in a single qubit of
989: the chain, but in a \emph{Gaussian-modulated superposition} of his
990: qubits. These Gaussian packets are known to have minimal dispersion.
991: Likewise, the receiver performs a decoding operation on all qubits
992: he controls. Near-perfect fidelity can be reached.
993:
994:
995: \subsection{Time-dependent control\label{sub:Time-dependent-control}}
996:
997: Finally, a number of authors found ways of improving the fidelity
998: by time-dependent control of some parameters of the Hamiltonian. In~\cite{Haselgrove2005}
999: it is shown that if the end couplings can be controlled as arbitrary
1000: (in general complex valued) smooth functions of time the encoding
1001: scheme~\cite{Linden2004} could be \emph{simulated} without the requirement
1002: of additional operations and qubits. Another possibility to achieve
1003: perfect state transfer is to have an Ising interaction with additionally
1004: pulsed global rotations~\cite{Fitzsimons2006,Jones,Raussendorf2005}.
1005: Further related methods of manipulating the transfer by global fields
1006: were reported in \cite{Sun2006,Maruyama,Monteiro2006,Korepin2005,Yang2006,Boness2006}.
1007:
1008:
1009: \section{Motivation and outline of this work}
1010:
1011: While the advanced transfer protocols have shown that in principle
1012: high fidelity can be achieved with arbitrarily long chains, they have
1013: all come at a cost. Engineering each coupling of the Hamiltonian puts
1014: extra demands on the experimental realisation, which is often already
1015: at its very limits just to ensure the \emph{coherence} of the system.
1016: Furthermore, the more a scheme relies on particular properties of
1017: the Hamiltonian, the more it will be affected by imperfections in
1018: its implementation~\cite{Fazio2005,Jing-Fu2006}. For example, simulating
1019: an engineered chain of length $N=50$ with a (relative) disorder of
1020: $5\%$, we get a fidelity peak of $0.95\pm0.02.$ For a disorder of
1021: $10\%$ we get $0.85\pm0.05.$ The weakly coupled system is very stable
1022: for off-site disorder~\cite{Semiao2005}, but suffers strongly from
1023: on-site disorder (i.e. magnetic fields in $z-$direction) at the ends
1024: of the chain. For example, for a chain of $N=50$ with edge couplings
1025: $j=0.01$ and the remaining couplings being $J=1,$ we find that already
1026: a magnetic field of the order of $0.00001$ lowers the fidelity to
1027: $0.87\pm0.12.$ For fields of the order of $0.00005$ we find $0.45\pm0.32.$
1028: This is because these fluctuations must be small with respect to the
1029: \emph{small} coupling, so there is a double scaling. Also, the time-scale
1030: of the transfer is longer than in other schemes (note though that
1031: this may sometimes even be useful for having enough time to pick up
1032: the received state). On the other hand, encoding and time-dependent
1033: control require additional resources and gating operations. It is
1034: not possible to judge independently of the realisation which of the
1035: above schemes is the ''most practical'' one. We summarise the different
1036: aspects that are important in the following five criteria for quantum
1037: state transfer\index{criteria for quantum state transfer}:
1038:
1039: \begin{enumerate}
1040: \item \emph{High efficiency:} How does the fidelity depend on the length
1041: of the chain? Which rate~\cite{Fazio,Kay2006,Yung2006} can be achieved?
1042: \item \emph{Minimal control}: How many operations are required to achieve
1043: a certain fidelity? \emph{Where}%
1044: \footnote{For example, gates at the ends of the chain are always needed for
1045: write-in and read-out, and thus ''cheaper'' than gates along the
1046: chain. Global control along the whole chain is often easier than local
1047: control.%
1048: } is control required?
1049: \item \emph{Minimal resources:} What additional resources are required?
1050: \item \emph{Minimal design}: How general is the coupling type%
1051: \footnote{Often the coupling type is already fixed by the experiment%
1052: }? What values of the coupling strengths are allowed?
1053: \item \emph{Robustness}: How is the fidelity affected by static disorder,
1054: by time-dependent disorder, by gate and timing errors, and by external
1055: noise such as decoherence and dissipation?
1056: \end{enumerate}
1057: At the start of this research, only the engineering and encoding schemes
1058: were available. The engineering schemes are strong in the points 2
1059: and 3, but quite weak in the points 4 and 5. The encoding scheme on
1060: the other hand has its weakness in points 2 and 3. It was hence desirable
1061: to develop more balanced schemes. Since most experiments in Quantum
1062: Information are extremely sensitive and at the cutting edge of their
1063: parameters (i.e. requiring extremely low temperatures, well tuned
1064: lasers, and so forth, to maintain their quantum behaviour), we particularly
1065: wanted to find schemes which are strong in the points 4 and 5. Also,
1066: from a more fundamental point of view, we were interested in seeing
1067: how much information on the state of a quantum chain could be obtained
1068: by the receiver in principle, and how the receiver might even be able
1069: to \emph{prepare} states on the whole chain.
1070:
1071: The main achievements of this thesis are two schemes for the transfer
1072: of quantum information using measurements (Chapter~\ref{cha:Dual-Rail}
1073: and~\ref{cha:Multi-rail-and-Capacity}) or unitary operations (Chapter~\ref{cha:Full-read-and}
1074: and~\ref{cha:Single-memory}) at the receiving end of the chain.
1075: Since both schemes use convergence properties of quantum operations,
1076: it seemed natural to investigate these properties in a more abstract
1077: way (Chapter~\ref{cha:Ergodicity-and-mixing}). There, we found a
1078: new way of characterising mixing maps, which has applications beyond
1079: quantum state transfer, and may well be relevant for other fields
1080: such as chaos theory or statistical physics. Finally, in Chapter~\ref{cha:Problems-and-Practical}
1081: we discuss problems quantum state transfer in the presence of external
1082: noise. The results in Chapters~\ref{cha:Multi-rail-and-Capacity}-\ref{cha:Single-memory}
1083: were developed in collaboration with Vittorio Giovannetti from Scuola
1084: Normale, Pisa. Much of the material discussed in this thesis has been
1085: published or submitted for publication~\cite{Bose2006b,DUALRAIL,RANDOMRAIL,Giovannetti2006,Giovannetti,Bose2006c,Burgarth,MULTIRAIL,MEMORYSWAP,Bose2005e}.
1086:
1087:
1088: \chapter{Dual Rail encoding\label{cha:Dual-Rail}}
1089:
1090:
1091: \section{Introduction}
1092:
1093: The role of measurement in quantum information theory has become more
1094: active recently. Measurements are not only useful to obtain information
1095: about some state or for preparation, but also, instead of gates, for
1096: quantum computation \cite{Raussendorf2001}. In the context of quantum
1097: state transfer, it seems first that measurements would spoil the coherence
1098: and destroy the state. The first indication that measurements can
1099: actually be used to transfer quantum information along anti-ferromagnetic
1100: chains was given in~\cite{Verstraete2004}. However there the measurements
1101: had to be performed along the whole chain. This may in some cases
1102: be easier than to perform swap gates, but still requires high local
1103: accessibility. We take a ''hybrid'' approach here: along the chain,
1104: we let the system evolve coherently, but at the receiving end, we
1105: try to \emph{help} the transfer by measuring. The main disadvantage
1106: of the encoding used in the protocols above is that once the information
1107: dispersed, there is no way of finding out where it is without destroying
1108: it. A dual rail encoding~\cite{Chuang1996} as used in quantum optics
1109: on the other hand allows us to perform parity type measurements that
1110: do \emph{not} spoil the coherence of the state that is sent. The outcome
1111: of the measurement tells us if the state has arrived at the end (corresponding
1112: to a perfect state transfer) or not. We call this \emph{conclusively
1113: perfect state transfer}.\index{conclusively perfect state transfer.}
1114: Moreover, by performing repetitive measurements, the probability of
1115: success can be made arbitrarily close to unity. As an example of such
1116: an \emph{\index{amplitude delaying channel}amplitude delaying channel},
1117: we show how two parallel Heisenberg spin chains can be used as quantum
1118: wires. Perfect state transfer with a probability of failure lower
1119: than $P$ in a Heisenberg chain of $N$ qubits can be achieved in
1120: a time-scale of the order of $0.33J^{-1}N^{1.7}|\ln P|$. We demonstrate
1121: that our scheme is more robust to decoherence and non-optimal timing
1122: than any scheme using single spin chains.
1123:
1124: We then generalise the dual rail encoding to disordered quantum chains.
1125: The scheme performs well for both spatially correlated and uncorrelated
1126: fluctuations if they are relatively weak (say 5\%). Furthermore, we
1127: show that given a quite arbitrary pair of quantum chains, one can
1128: check whether it is capable of perfect transfer by only local operations
1129: at the ends of the chains, and the system in the middle being a \emph{black
1130: box}. We argue that unless some specific symmetries are present in
1131: the system, it \emph{will} be capable of perfect transfer when used
1132: with dual rail encoding. Therefore our scheme puts minimal demand
1133: not only on the control of the chains when using them, but also on
1134: the design when building them.
1135:
1136: This Chapter is organised as follows. In Section \ref{sec:con}, we
1137: suggest a scheme for quantum communication using two parallel spin
1138: chains of the most natural type (namely those with constant couplings).
1139: We require modest encodings (or gates) and measurements only at the
1140: ends of the chains. The state transfer is \emph{conclusive}, which
1141: means that it is possible to tell by the outcome of a quantum measurement,
1142: without destroying the state, if the transfer took place or not. If
1143: it did, then the transfer was \emph{perfect}. The transmission time
1144: for conclusive transfer is not longer than for single spin chains.
1145: In Section \ref{sec:Arbit}, we demonstrate that our scheme offers
1146: even more: if the transfer was not successful, then we can wait for
1147: some time and just repeat the measurement, without having to resend
1148: the state. By performing sufficiently many measurements, the probability
1149: for perfect transfer approaches unity. Hence the transfer is \emph{arbitrarily
1150: perfect}. We will show in Section \ref{sec:Estimati} that the time
1151: needed to transfer a state with a given probability scales in a reasonable
1152: way with the length of the chain. In Section \ref{sec:Decohere} we
1153: show that encoding to parallel chains and the conclusiveness also
1154: makes our protocol more robust to decoherence (a hitherto unaddressed
1155: issue in the field of quantum communication through spin chains).
1156: In the last part of this chapter, we show how this scheme can be generalised
1157: to disordered chains (Sections~\ref{sec:Disordered-chains}-\ref{sec:Numerical-Examples})
1158: and even coupled chains (Section~\ref{sec:Coupled-chains}).
1159:
1160:
1161: \section{Scheme for conclusive transfer\label{sec:con}}
1162:
1163: We intend to propose our scheme in a system-independent way with occasional
1164: references to systems where conditions required by our scheme are
1165: achieved. We assume that our system consists of two identical uncoupled
1166: spin-$1/2$-chains $(1)$ and $(2)$ of length $N$, described by
1167: the Hamiltonian\begin{equation}
1168: H=H^{(1)}\otimes I^{(2)}+I^{(1)}\otimes H^{(2)}-E_{g}I^{(1)}\otimes I^{(2)}.\end{equation}
1169: %
1170: \begin{figure}[htbp]
1171: \begin{centering}\includegraphics[width=0.65\paperwidth]{channels_order}\par\end{centering}
1172:
1173:
1174: \caption{\label{fig:channelord}Two quantum chains interconnecting $A$ and
1175: $B$. Control of the systems is only possible at the two qubits of
1176: either end.}
1177: \end{figure}
1178: The term identical states that $H^{(1)}$ and $H^{(2)}$ are the
1179: same apart from the label of the Hilbert space they act on. The requirement
1180: of parallel chains instead of just one is not a real problem, since
1181: in many experimental realisations of spin chains, it is much easier
1182: to produce a whole bunch of parallel uncoupled~\cite{Motoyama1996,Gambardella2002}
1183: chains than just a single one.
1184:
1185: We assume that the ground state of each chain is $\left|\boldsymbol{0}\right\rangle _{i}$,
1186: i.e. a ferromagnetic ground state, with $H^{(i)}\left|\boldsymbol{0}\right\rangle _{i}=E_{g}\left|\boldsymbol{0}\right\rangle _{i},$
1187: and that the subspace consisting of the single spin excitations $\left|\boldsymbol{n}\right\rangle _{i}$
1188: is invariant under $H^{(i)}.$ Let us assume that the state that Alice
1189: wants to send is at the first qubit of the first chain, i.e.\begin{equation}
1190: \left|\boldsymbol{\psi}_{A}\right\rangle _{1}\equiv\alpha\left|\boldsymbol{0}\right\rangle _{1}+\beta\left|\boldsymbol{1}\right\rangle _{1},\end{equation}
1191: and that the second chain is in the ground state $|\boldsymbol{0}\rangle_{2}.$
1192: The aim of our protocol is to transfer quantum information from the
1193: $1$st ({}``Alice'') to the $N$th ({}``Bob'') qubit of the first
1194: chain:\begin{equation}
1195: \left|\boldsymbol{\psi}_{A}\right\rangle _{1}\rightarrow\left|\boldsymbol{\psi}_{B}\right\rangle _{1}\equiv\alpha\left|\boldsymbol{0}\right\rangle _{1}+\beta\left|\boldsymbol{N}\right\rangle _{1}.\end{equation}
1196: The first step (see also Fig.~\ref{fig:summary}) is to encode the
1197: input qubit in a \emph{dual rail}\index{dual rail}~\cite{Chuang1996}
1198: by applying a NOT gate on the first qubit of system $(2)$ controlled
1199: by the first qubit of system $(1)$ being zero, resulting in a superposition
1200: of excitations in both systems, \begin{equation}
1201: \left|\boldsymbol{s}(0)\right\rangle =\alpha\left|\boldsymbol{0,1}\right\rangle +\beta\left|\boldsymbol{1,0}\right\rangle ,\label{eq:superposition}\end{equation}
1202: where we have introduced the short notation $|\boldsymbol{n,m}\rangle\equiv|\boldsymbol{n}\rangle_{1}\otimes|\boldsymbol{m}\rangle_{2}.$
1203: This is assumed to take place in a much shorter time-scale than the
1204: system dynamics. Even though a 2-qubit gate in solid state systems
1205: is difficult, such a gate for charge qubits has been reported~\cite{YAMA}.
1206: For the same qubits, Josephson arrays have been proposed as single
1207: spin chains for quantum communication~\cite{Bruder2005a}. For this
1208: system, both requisites of our scheme are thus available. In fact,
1209: the demand that Alice and Bob can do measurements and apply gates
1210: to their local qubits (i.e. the ends of the chains) will be naturally
1211: fulfilled in practice since we are suggesting a scheme to transfer
1212: information between quantum computers (as described in Section~\ref{sec:Quantum-state-transfer}).%
1213: \begin{figure*}[tbh]
1214: \begin{centering}\includegraphics[width=0.9\textwidth]{par4}\par\end{centering}
1215:
1216:
1217: \caption{\label{fig:summary}Quantum circuit representation of conclusive
1218: and arbitrarily perfect state transfer. The first gate at Alice's
1219: qubits represents a NOT gate applied to the second qubit controlled
1220: by the first qubit being zero. The qubit $\left|\boldsymbol{\psi}_{A}\right\rangle _{1}$
1221: on the left hand side represents an arbitrary input state at Alice's
1222: site, and the qubit $\left|\boldsymbol{\psi}_{B}\right\rangle _{1}$
1223: represents the same state, successfully transferred to Bob's site.
1224: The $t_{\ell}$-gate represents the unitary evolution of the spin
1225: chains for a time interval of $t_{\ell}$.}
1226: \end{figure*}
1227:
1228:
1229: Under the system Hamiltonian, the excitation in Eq. (\ref{eq:superposition})
1230: will travel along the two systems. The state after the time $t_{1}$
1231: can be written as \begin{equation}
1232: \left|\boldsymbol{\phi}(t_{1})\right\rangle =\sum_{n=1}^{N}f_{n,1}(t_{1})\left|\boldsymbol{s}(n)\right\rangle ,\end{equation}
1233: where $\left|\boldsymbol{s}(n)\right\rangle =\alpha\left|\boldsymbol{0,n}\right\rangle +\beta\left|\boldsymbol{n,0}\right\rangle $
1234: and the complex amplitudes $f_{n,1}(t_{1})$ are given by Eq.~(\ref{eq:spintrans}).
1235: We can \emph{decode} the qubit by applying a CNOT gate at Bob's site.
1236: Assuming that this happens on a time-scale much shorter than the evolution
1237: of the chain, the resulting state is given by \begin{equation}
1238: \sum_{n=1}^{N-1}f_{n,1}(t_{1})\left|\boldsymbol{s}(n)\right\rangle +f{}_{N,1}(t_{1})\left|\boldsymbol{\psi}_{B}\right\rangle _{1}\otimes\left|\boldsymbol{N}\right\rangle _{2}.\end{equation}
1239: Bob can now perform a measurement on his qubit of system $(2).$
1240: If the outcome of this measurement is $1$, he can conclude that the
1241: state $\left|\boldsymbol{\psi}\right\rangle _{1}^{(1)}$ has been
1242: successfully transferred to him. This happens with the probability
1243: $\left|f_{N,1}(t_{1})\right|^{2}.$ If the outcome is $0$, the system
1244: is in the state \begin{equation}
1245: \frac{1}{\sqrt{P(1)}}\sum_{n=1}^{N-1}f_{n,1}(t_{1})\left|\boldsymbol{s}(n)\right\rangle ,\label{eq:firstmeas}\end{equation}
1246: where $P(1)=1-\left|f_{N,1}(t_{1})\right|^{2}$ is the probability
1247: of \emph{failure} for the first measurement. If the protocol stopped
1248: here, and Bob would just assume his state as the transferred one,
1249: the channel could be described as an \emph{amplitude damping channel}~\cite{Giovannetti2005},
1250: with exactly the same fidelity as the single chain scheme discussed
1251: in~\cite{Bose2003}. Note that here the encoding is symmetric with
1252: respect to $\alpha$ and $\beta,$ so the minimal fidelity is the
1253: same as the averaged one.
1254:
1255: But success probability is more valuable than fidelity: Bob has gained
1256: knowledge about his state, and may reject it and ask Alice to retransmit
1257: (this is known as a \emph{\index{quantum erasure channel}quantum
1258: erasure channel}~\cite{Bennett1997}). Of course in general the state
1259: that Alice sends is the unknown result of some quantum computation
1260: and cannot be sent again easily. This can be overcome in the following
1261: way: Alice sends one e-bit on the dual rail first. If Bob measures
1262: a success, he tells Alice, and they both start to teleport the unknown
1263: state. If he measures a failure, they reset the chains and start again.
1264: Since the joint probability of failure converges exponentially fast
1265: to zero this is quite efficient. In fact the conclusive transfer of
1266: entanglement is possible even on a \emph{single chain} by using the
1267: same chain again instead of a second one \cite{Wan}. This can be
1268: seen as a very simple \index{entanglement distillation}entanglement
1269: distillation procedure, achieving a rate of $|f_{N,1}(t)|^{2}/2.$
1270: However the chain needs to be reset between each transmission (see
1271: Section~\ref{sec:Initialisation} for problems related to this),
1272: and Alice and Bob require classical communication. We will show in
1273: the next section, that the reuse of the chain(s) is not necessary,
1274: as arbitrarily perfect state transfer can already achieved in the
1275: first transmission.
1276:
1277:
1278: \section{Arbitrarily perfect state transfer\label{sec:Arbit}}
1279:
1280: Because Bob's measurement has not revealed anything about the input
1281: state (the success probability is independent of the input state),
1282: the information is still residing in the chain. By letting the state
1283: (\ref{eq:firstmeas}) evolve for another time $t_{2}$ and applying
1284: the CNOT gate again, Bob has another chance of receiving the input
1285: state. The state before performing the second measurement is easily
1286: seen to be \begin{equation}
1287: \frac{1}{\sqrt{P(1)}}\sum_{n=1}^{N}\left\{ f_{n,1}(t_{2}+t_{1})-f_{n,N}(t_{2})f_{N,1}(t_{1})\right\} \left|\boldsymbol{s}(n)\right\rangle .\label{eq:second_meas}\end{equation}
1288: Hence the probability to receive the qubit at Bobs site at the second
1289: measurement is\begin{equation}
1290: \frac{1}{P(1)}\left|f_{N,1}(t_{2}+t_{1})-f_{N,N}(t_{2})f_{N,1}(t_{1})\right|^{2}.\end{equation}
1291: If the transfer was still unsuccessful, this strategy can be repeated
1292: over and over. Each time Bob has a probability of failed state transfer
1293: that can be obtained from the generalisation of Eq. (\ref{eq:second_meas})
1294: to an arbitrary number of iterations. The joint probability that Bob
1295: fails to receive the state all the time is just the product of these
1296: probabilities. We denote the joint probability of failure for having
1297: done $l$ unsuccessful measurements as $P(\ell)$. This probability
1298: depends on the time intervals $t_{\ell}$ between the $(\ell-1)$th
1299: and $\ell$th measurement, and we are interested in the case where
1300: the $t_{\ell}$ are chosen such that the transfer is fast. It is possible
1301: to write a simple algorithm that computes $P(\ell)$ for any transition
1302: amplitude $f_{r,s}(t).$ Figure \ref{fig:heisenberg} shows some results
1303: for the Heisenberg Hamiltonian given by Eq.~(\ref{eq:heisenberg}).%
1304: \begin{figure}[tbh]
1305: \begin{centering}\includegraphics[width=0.8\columnwidth]{optimisation2}\par\end{centering}
1306:
1307:
1308: \caption{\label{fig:heisenberg}Semilogarithmic plot of the joint probability
1309: of failure $P(\ell)$ as a function of the number of measurements
1310: $\ell$. Shown are Heisenberg spin-$1/2$-chains with different lengths
1311: $N$. The times between measurements $t_{\ell}$ have been optimised
1312: numerically. }
1313: \end{figure}
1314:
1315:
1316: An interesting question is whether the joint probability of failure
1317: can be made arbitrarily small with a large number of measurements.
1318: In fact, the times $t_{\ell}$ can be chosen such that the transfer
1319: becomes arbitrarily perfect. We will prove this in the next Chapter,
1320: where a generalisation of the dual rail scheme and a much wider class
1321: of Hamiltonians is considered. In the limit of large number of measurements,
1322: the spin channel will not damp the initial amplitude, but only \emph{delay}
1323: it.
1324:
1325:
1326: \section{Estimation of the \index{time-scale}time-scale the transfer\label{sec:Estimati}}
1327:
1328: The achievable fidelity is an important, but not the only criterion
1329: of a state transfer protocol. In this Section, we give an heuristic
1330: approach to estimate the time that it needs to achieve a certain fidelity
1331: in a Heisenberg spin chain. The comparison with numeric examples is
1332: confirming this approach.
1333:
1334: Let us first describe the dynamic of the chain in a very qualitative
1335: way. Once Alice has initialised the system, an excitation wave packet
1336: will travel along the chain. As shown in Subsection~\ref{sub:Dynamics-and-Dispersion},
1337: it will reach Bob at a time of the order of \begin{equation}
1338: t_{\mbox{peak}}\approx\frac{N}{2J},\end{equation}
1339: with an amplitude of \begin{equation}
1340: \left|f_{N,1}(t_{\mbox{peak}})\right|^{2}\approx1.82N^{-2/3}.\label{eq:peak}\end{equation}
1341: It is then reflected and travels back and forth along the chain.
1342: Since the wave packet is also dispersing, it starts interfering with
1343: its tail, and after a couple of reflections the dynamic is becoming
1344: quite randomly. This effect becomes even stronger due to Bobs measurements,
1345: which change the dynamics by projecting away parts of the wave packet.
1346: We now assume that $2t_{\mbox{peak}}$ (the time it takes for a wave
1347: packet to travel twice along the chain) remains a good estimate of
1348: the time-scale in which significant probability amplitude peaks at
1349: Bobs site occur, and that Eq.~(\ref{eq:peak}) remains a good estimate
1350: of the amplitude of these peaks%
1351: \footnote{This is not a strong assumption. If the excitation was fully randomly
1352: distributed, the probability would scale as $N^{-1}.$ By searching
1353: for good arrival times, this can be slightly increased to $N^{-2/3}.$%
1354: }. Therefore, the joint probability of failure is expected to scale
1355: as\begin{equation}
1356: P(\ell)\approx\left(1-1.82N^{-2/3}\right)^{\ell}\label{eq:probscale}\end{equation}
1357: in a time of the order of \begin{equation}
1358: t(\ell)\approx2t_{max}\ell=J^{-1}N\ell.\label{eq:timescale}\end{equation}
1359: If we combine Eq. (\ref{eq:probscale}) and (\ref{eq:timescale})
1360: and solve for the time $t(P)$ needed to reach a certain probability
1361: of failure $P$, we get for $N\gg1$ \begin{equation}
1362: t(P)\approx0.55J^{-1}N^{5/3}\left|\ln P\right|.\end{equation}
1363: We compare this rough estimate with exact numerical results in Fig.
1364: \ref{cap:Timefit}. The best fit for the range shown in the figure
1365: is given by\begin{equation}
1366: \boxed{t(P)=0.33J^{-1}N^{5/3}\left|\ln P\right|.}\label{eq:fit3}\end{equation}
1367: We can conclude that the transmission time for arbitrarily perfect
1368: transfer is scaling not much worse with the length $N$ of the chains
1369: than the single spin chain schemes. Despite of the logarithmic dependence
1370: on $P,$ the time it takes to achieve high fidelity is still reasonable.
1371: For example, a system with $N=100$ and $J=20K*k_{B}$ will take approximately
1372: $1.3ns$ to achieve a fidelity of 99\%. In many systems, decoherence
1373: is completely negligible within this time-scale. For example, some
1374: Josephson junction systems~\cite{Vion2002} have a decoherence time
1375: of $T_{\phi}\approx500ns$, while trapped ions have even larger decoherence
1376: times.%
1377: \begin{figure}[tbh]
1378: \includegraphics[width=1\columnwidth]{3d}
1379:
1380:
1381: \caption{\label{cap:Timefit}Time $t$ needed to transfer a state with a given
1382: joint probability of failure $P$ across a chain of length $N$. The
1383: points denote exact numerical data, and the fit is given by Eq. (\ref{eq:fit3}).}
1384: \end{figure}
1385:
1386:
1387:
1388: \section{Decoherence and imperfections\label{sec:Decohere}}
1389:
1390: If the coupling between the spins $J$ is very small, or the chains
1391: are very long, the transmission time may no longer be negligible with
1392: respect to the decoherence time. It is interesting to note that the
1393: dual rail encoding then offers some significant general advantages
1394: over single chain schemes. Since we are suggesting a system-independent
1395: scheme, we will not study the effects of specific environments on
1396: our protocol, but just qualitatively point out its general advantages.
1397:
1398: At least theoretically, it is always possible to cool the system down
1399: or to apply a strong magnetic field so that the environment is not
1400: causing further excitations. For example in flux qubit systems, the
1401: system is cooled to $\approx25mK$ to ensure that the energy splitting$\Delta\gg k_{B}T$
1402: \cite{Chiorescu2003}. Then, there are two remaining types of quantum
1403: noise that will occur: phase noise\index{phase noise} and amplitude
1404: damping\index{amplitude damping}. Phase noise is a serious problem
1405: and arises here \emph{only} when an environment can distinguish between
1406: spin flips on the first chain and spin flips on the second chain.
1407: It is therefore important that the environment cannot resolve their
1408: difference. In this case, the environment will only couple with the
1409: total $z$-component\begin{equation}
1410: Z_{n}^{(1)}+Z_{n}^{(2)}\end{equation}
1411: of the spins of both chains at each position $n$. This has been
1412: discussed for spin-boson models in~\cite{Palma1996,Hwang2000} but
1413: also holds for spin environments as long as the chains are close enough.
1414: The qubit is encoded in a decoherence-free subspace\index{decoherence-free subspace}~\cite{Beige2000}
1415: and the scheme is fully robust to phase noise. Even though this may
1416: not be true for all implementations of dual rail encoding, it is worthwhile
1417: noticing it because such an opportunity does not exist \emph{at all}
1418: for single chain schemes, where the coherence between two states with
1419: different total z-component of the spin has to be preserved. Having
1420: shown one way of avoiding phase noise, at least in some systems, we
1421: now proceed to amplitude damping.
1422:
1423: The evolution of the system in presence of amplitude damping of a
1424: rate $\Gamma$ can be easily derived using a quantum-jump approach\index{quantum-jump approach}~\cite{Plenio1998}.
1425: This is based on a quantum master equation approach, which is valid
1426: in the Born-Markov approximation~\cite{OPENQUANTUM} (i.e. it holds
1427: for weakly coupled environments without memory effects). Similarly
1428: to phase noise, it is necessary that the environment acts symmetrically
1429: on the chains. The dynamics is then given by an effective non-Hermitian
1430: Hamiltonian\begin{equation}
1431: H_{eff}=H+i\Gamma\sum_{n}\left(Z_{n}^{(1)}+Z_{n}^{(2)}+2\right)/2\end{equation}
1432: if no jump occurs. If a jump occurs, the system is back in the ground
1433: state $|\boldsymbol{0}\rangle$. The state of the system before the
1434: first measurement conditioned on no jump is given by\begin{equation}
1435: e^{-\Gamma t}\sum_{n=1}^{N}f_{n,1}(t)\left|\boldsymbol{s}(n)\right\rangle ,\label{eq:jump1}\end{equation}
1436: and this happens with the probability of $e^{-2\Gamma t}$ (the norm
1437: of the above state). If a jump occurs, the system will be in the ground
1438: state \begin{equation}
1439: \sqrt{1-e^{-2\Gamma t}}\left|\boldsymbol{0,0}\right\rangle .\label{eq:jump2}\end{equation}
1440: The density matrix at the time $t$ is given by a mixture of (\ref{eq:jump1})
1441: and (\ref{eq:jump2}). In case of (\ref{eq:jump2}), the quantum information
1442: is completely lost and Bob will always measure an unsuccessful state
1443: transfer. If Bob however measures a success, it is clear that no jump
1444: has occurred and he has the perfectly transferred state. Therefore
1445: the protocol \emph{remains conclusive}, but the success probability
1446: is lowered by $e^{-2\Gamma t}.$ This result is still valid for multiple
1447: measurements, which leave the state (\ref{eq:jump2}) unaltered. The
1448: probability of a successful transfer at each particular measurement
1449: $\ell$ will decrease by $e^{-2\Gamma t(\ell)}$, where $t(\ell)$
1450: is the time at which the measurement takes place. After a certain
1451: number of measurements, the \emph{joint} probability of failure will
1452: no longer decrease. Thus the transfer will no longer be \emph{arbitrarily}
1453: perfect, but can still reach a very high fidelity. Some numerical
1454: examples of the minimal joint probability of failure that can be achieved,
1455: \begin{eqnarray}
1456: \lim_{l\rightarrow\infty}P(\ell) & \approx & \prod_{\ell=1}^{\infty}\left(1-1.35N^{-2/3}e^{-\frac{2\Gamma N}{J}\ell}\right)\label{eq:plimit}\end{eqnarray}
1457: are given in Fig. \ref{fig:losses}. For $J/\Gamma=50K\: ns$ nearly
1458: perfect transfer is still possible for chains up to a length of $N\approx40$.
1459:
1460: Even if the amplitude damping is not symmetric, its effect is weaker
1461: than in single spin schemes. This is because it can be split in a
1462: symmetric and asymmetric part. The symmetric part can be overcome
1463: with the above strategies. For example, if the amplitude damping on
1464: the chains is $\Gamma_{1}$ and $\Gamma_{2}$ with $\Gamma_{1}>\Gamma_{2},$
1465: the state (\ref{eq:jump1}) will be\begin{eqnarray}
1466: & & \sum_{n=1}^{N}f_{n,1}(t)\left\{ \alpha e^{-\Gamma_{2}t}\left|\boldsymbol{0,n}\right\rangle +\beta e^{-\Gamma_{1}t}\left|\boldsymbol{n,0}\right\rangle \right\} \\
1467: & \approx & e^{-\Gamma_{2}t}\sum_{n=1}^{N}f_{n,1}(t)\left|\boldsymbol{s}(n)\right\rangle \label{eq:nodeviation}\end{eqnarray}
1468: provided that $t\ll\left(\Gamma_{1}-\Gamma_{2}\right)^{-1}.$ Using
1469: a chain of length $N=20$ with $J=20K*k_{B}$ and $\Gamma_{1}^{-1}=4ns$,
1470: $\Gamma_{2}^{-1}=4.2ns$ we would have to fulfil $t\ll164ns$. We
1471: could perform approximately $10$ measurements (cf. Eq. (\ref{eq:timescale}))
1472: without deviating too much from the state (\ref{eq:nodeviation}).
1473: In this time, we can use our protocol in the normal way. The resulting
1474: success probability given by the finite version of Eq. (\ref{eq:plimit})
1475: would be $75$\%. A similar reasoning is valid for phase noise, where
1476: the environment can be split into common and separate parts. If the
1477: chains are close, the common part will dominate and the separate parts
1478: can be neglected for short times.%
1479: \begin{figure}[tbh]
1480: \begin{centering}\includegraphics[width=0.8\columnwidth]{converge}\par\end{centering}
1481:
1482:
1483: \caption{\label{fig:losses}The minimal joint probability of failure $P(\ell)$
1484: for chains with length $N$ in the presence of amplitude damping.
1485: The parameter $J/\Gamma$ of the curves is the coupling of the chain
1486: (in Kelvin) divided by the decay rate ($ns^{-1}$).}
1487: \end{figure}
1488:
1489:
1490:
1491: \section{Disordered chains\label{sec:Disordered-chains}}
1492:
1493: The main requirement for perfect transfer with dual rail encoding
1494: in the above is that two \emph{identical} quantum chains have to be
1495: designed. While this is not so much a theoretical problem, for possible
1496: experimental realizations of the scheme~\cite{Bruder2005a} the question
1497: arises naturally how to cope with slight asymmetries of the channels.
1498: We are now going to demonstrate that in many cases, perfect state
1499: transfer with dual rail encoding is possible for quantum chains with
1500: differing Hamiltonians.
1501:
1502: By doing so, we also offer a solution to another and perhaps more
1503: \emph{general} problem: if one implements \emph{any} of the schemes
1504: for quantum state transfer, the Hamiltonians will always be different
1505: from the theoretical ones by some random perturbation. This will lead
1506: to a decrease of fidelity in particular where specific energy levels
1507: were assumed (see~\cite{Fazio2005,Jing-Fu2006} for an analysis of
1508: fluctuations affecting the engineered chains described in Subsection~\ref{sub:Engineered-Hamiltonians}).
1509: This problem can be avoided using the scheme described below. In general,
1510: disorder can lead to a Anderson localisation\index{Anderson localisation}~\cite{Anderson1958,Winter,Apollaro2006}
1511: of the eigenstates (and therefore to low fidelity transport of quantum
1512: information). In this section however this is not relevant, as we
1513: consider only short chains $(N<100)$ and small disorder ($\approx10\%$
1514: of the coupling strength), and the localisation length is much longer
1515: then the length of the chain. We will show numerically that the dual
1516: rail scheme can still achieve arbitrarily perfect transfer for a uniformly
1517: coupled Heisenberg Hamiltonian with disordered coupling strengths
1518: (both for the case of spatially correlated and uncorrelated disorder).
1519: Moreover, for any two quantum chains, we show that Bob and Alice can
1520: check whether their system is capable of dual rail transfer without
1521: directly measuring their Hamiltonians or local properties of the system
1522: along the chains but by only measuring \emph{their} part of the system.
1523:
1524:
1525: \section{Conclusive transfer in the presence of disorder}
1526:
1527: We consider two uncoupled quantum chains $(1)$ and $(2)$, as shown
1528: in Fig. \ref{fig:channel}. The chains are described by the two Hamiltonians
1529: $H^{(1)}$ and $H^{(2)}$ with total Hamiltonian given by\begin{equation}
1530: H=H^{(1)}\otimes I^{(2)}+I^{(1)}\otimes H^{(2)},\end{equation}
1531: and the time evolution operator factorising as\begin{eqnarray}
1532: U(t) & = & \exp\left(-iH^{(1)}t\right)\otimes\exp\left(-iH^{(2)}t\right).\end{eqnarray}
1533: For the moment, we assume that both chains have equal length $N$,
1534: but it will become clear in Section \ref{sec:Tomography} that this
1535: is not a requirement of our scheme. All other assumptions remain as
1536: in the first part of the chapter.%
1537: \begin{figure}[htbp]
1538: \begin{centering}\includegraphics[width=0.5\paperwidth]{channels}\par\end{centering}
1539:
1540:
1541: \caption{\label{fig:channel}Two \emph{disordered} quantum chains interconnecting
1542: $A$ and $B$. Control of the systems is only possible at the two
1543: qubits of either end.}
1544: \end{figure}
1545:
1546:
1547: Initially, Alice encodes the state as\begin{equation}
1548: \alpha\left|\boldsymbol{0,1}\right\rangle +\beta\left|\boldsymbol{1,0}\right\rangle .\end{equation}
1549: This is a superposition of an excitation in the first qubit of the
1550: first chain and an excitation in the first qubit of the second chain.
1551: The state will evolve into\begin{equation}
1552: \sum_{n=1}^{N}\left\{ \alpha g_{n,1}(t)\left|\boldsymbol{0,n}\right\rangle +\beta f_{n,1}(t)\left|\boldsymbol{n,0}\right\rangle \right\} ,\label{eq:evolved}\end{equation}
1553: with \begin{eqnarray}
1554: f_{n,1}(t) & \equiv & \left\langle \boldsymbol{n,0}\left|U(t)\right|\boldsymbol{1,0}\right\rangle \\
1555: g_{n,1}(t) & \equiv & \left\langle \boldsymbol{0,n}\left|U(t)\right|\boldsymbol{0,1}\right\rangle .\end{eqnarray}
1556: In Section~\ref{sec:con}, these functions were identical. For differing
1557: chains this is no longer the case. We may, however, find a time $t_{1}$
1558: such that the modulus of their amplitudes at the last spins are the
1559: same (see Fig. \ref{fig:coincidence}),\begin{equation}
1560: g_{N,1}(t_{1})=e^{i\phi_{1}}f_{N,1}(t_{1}).\label{eq:req1}\end{equation}
1561: %
1562: \begin{figure}[htbp]
1563: \begin{centering}\includegraphics[width=0.8\columnwidth]{ft_dual}\par\end{centering}
1564:
1565:
1566: \caption{\label{fig:coincidence}The absolute values of the transition amplitudes
1567: $f_{N,1}(t)$ and $g_{N,1}(t)$ for two Heisenberg chains of length
1568: $N=10$. The couplings strengths of both chains were chosen randomly
1569: from the interval $\left[0.8J,1.2J\right].$ The circles show times
1570: where Bob can perform measurements without gaining information on
1571: $\alpha$ and $\beta.$}
1572: \end{figure}
1573:
1574:
1575: At this time, the state (\ref{eq:evolved}) can be written as\begin{eqnarray}
1576: \sum_{n=1}^{N-1}\left\{ \alpha g_{n,1}(t_{1})\left|\boldsymbol{0,n}\right\rangle +\beta f_{n,1}(t_{1})\left|\boldsymbol{n,0}\right\rangle \right\} +\nonumber \\
1577: f_{N,1}(t_{1})\left\{ e^{i\phi_{1}}\alpha\left|\boldsymbol{0,N}\right\rangle +\beta\left|\boldsymbol{N,0}\right\rangle \right\} .\end{eqnarray}
1578: Bob decodes the state by applying a CNOT gate on his two qubits,
1579: with the first qubit as the control bit. The state thereafter is\begin{eqnarray}
1580: \sum_{n=1}^{N-1}\left\{ \alpha g_{n,1}(t_{1})\left|\boldsymbol{0,n}\right\rangle +\beta f_{n,1}(t_{1})\left|\boldsymbol{n,0}\right\rangle \right\} +\nonumber \\
1581: f_{N,1}(t_{1})\left\{ e^{i\phi_{1}}\alpha\left|\boldsymbol{0}\right\rangle ^{(1)}+\beta\left|\boldsymbol{N}\right\rangle ^{(1)}\right\} \otimes\left|\boldsymbol{N}\right\rangle ^{(2)}.\end{eqnarray}
1582: Bob then measures his second qubit. Depending on the outcome of this
1583: measurement, the systems will either be in the state\begin{equation}
1584: \frac{1}{\sqrt{p_{1}}}\sum_{n=1}^{N-1}\left\{ \alpha g_{n,1}(t_{1})\left|\boldsymbol{0,n}\right\rangle +\beta f_{n,1}(t_{1})\left|\boldsymbol{n,0}\right\rangle \right\} \label{eq:failure}\end{equation}
1585: or in\begin{equation}
1586: \left\{ e^{i\phi_{1}}\alpha\left|\boldsymbol{0}\right\rangle ^{(1)}+\beta\left|\boldsymbol{N}\right\rangle ^{(1)}\right\} \otimes\left|\boldsymbol{N}\right\rangle ^{(2)},\label{eq:success}\end{equation}
1587: where $p_{1}=1-\left|f_{N,1}(t_{1})\right|^{2}=1-\left|g_{N,1}(t_{1})\right|^{2}$
1588: is the probability that Bob has \emph{not} received the state. The
1589: state (\ref{eq:success}) corresponds to the correctly transferred
1590: state with a \emph{known} phase error (which can be corrected by Bob
1591: using a simple phase gate). If Bob finds the system in the state (\ref{eq:failure}),
1592: the transfer has been unsuccessful, but the information is still in
1593: the chain. We thus see that conclusive transfer is still possible
1594: with randomly coupled chains as long as the requirement (\ref{eq:req1})
1595: is met. This requirement will be further discussed and generalised
1596: in the next section.
1597:
1598:
1599: \section{Arbitrarily perfect transfer in the presence of disorder\label{sec:Arbitrarily-perfect-transferl}}
1600:
1601: If the transfer was unsuccessful, the state (\ref{eq:failure}) will
1602: evolve further, offering Bob further opportunities to receive Alice's
1603: message. For identical quantum chains, leads to a success for any
1604: reasonable Hamiltonian (Section~\ref{sec:Quantum-chains-with}).
1605: For differing chains, this is not necessarily the case, because measurements
1606: are only allowed at times where the probability amplitude at the end
1607: of the chains is equal, and there may be systems where this is never
1608: the case. In this section, we will develop a criterion that generalises
1609: Eq. (\ref{eq:req1}) and allows to check numerically whether a given
1610: system is capable of arbitrarily perfect state transfer.
1611:
1612: The quantity of interest for conclusive state transfer is the joint
1613: probability $P(\ell)$ that after having checked $l$ times, Bob still
1614: has not received the proper state at his end of the chains. Optimally,
1615: this should approach zero if $\ell$ tends to infinity. In order to
1616: derive an expression for $P(\ell),$ let us assume that the transfer
1617: has been unsuccessful for $\ell-1$ times with time intervals $t_{\ell}$
1618: between the the $\ell$th and the $(\ell-1)$th measurement, and calculate
1619: the probability of failure at the $\ell$th measurement. In a similar
1620: manner, we assume that all the $\ell-1$ measurements have met the
1621: requirement of conclusive transfer (that is, Bob's measurements are
1622: unbiased with respect to $\alpha$ and $\beta$) and derive the requirement
1623: for the $\ell$th measurement.
1624:
1625: To calculate the probability of failure for the $\ell$th measurement,
1626: we need to take into account that Bob's measurements disturb the unitary
1627: dynamics of the chain. If the state before a measurement with the
1628: outcome {}``failure'' is $\left|\psi\right\rangle ,$ the state
1629: after the measurement will be\begin{equation}
1630: \frac{1}{\sqrt{p_{\ell}}}Q\left|\psi\right\rangle ,\end{equation}
1631: where $Q$ is the projector\begin{equation}
1632: Q=I-\left|\boldsymbol{0,N}\right\rangle \left\langle \boldsymbol{0,N}\right|-\left|\boldsymbol{N,0}\right\rangle \left\langle \boldsymbol{N,0}\right|,\end{equation}
1633: and $p_{\ell}$ is the probability of failure at the $l$th measurement.
1634: The dynamics of the chain is alternating between unitary and projective,
1635: such that the state before the $\ell$th measurement is given by\begin{equation}
1636: \frac{1}{\sqrt{P(\ell-1)}}\prod_{k=1}^{\ell}\left\{ U(t_{k})Q\right\} \left\{ \alpha\left|\boldsymbol{1,0}\right\rangle +\beta\left|\boldsymbol{0,1}\right\rangle \right\} ,\label{eq:product}\end{equation}
1637: where \begin{equation}
1638: P(\ell-1)=\prod_{\ell=1}^{\ell-1}p_{k}.\label{eq:joint_equal_prod}\end{equation}
1639: Note that the operators in (\ref{eq:product}) do not commute and
1640: that the time ordering of the product (the index $k$ increases from
1641: right to left) is important. The probability that there is an excitation
1642: at the $N$th site of either chain is given by\begin{equation}
1643: \frac{1}{P(\ell-1)}\left\{ \left|\alpha\right|^{2}\left|F(\ell)\right|^{2}+\left|\beta\right|^{2}\left|G(\ell)\right|^{2}\right\} ,\end{equation}
1644: with \begin{equation}
1645: F(\ell)\equiv\left\langle \boldsymbol{N,0}\right|\prod_{k=1}^{\ell}\left\{ U(t_{k})Q\right\} \left|\boldsymbol{1,0}\right\rangle ,\label{eq:onlything}\end{equation}
1646: and \begin{equation}
1647: G(\ell)\equiv\left\langle \boldsymbol{0,N}\right|\prod_{k=1}^{\ell}\left\{ U(t_{k})Q\right\} \left|\boldsymbol{0,1}\right\rangle .\label{eq:onlything2}\end{equation}
1648: Bob's measurements are therefore unbiased with respect to $\alpha$
1649: and $\beta$ if and only if\begin{equation}
1650: \left|F(\ell)\right|=\left|G(\ell)\right|\quad\forall\ell.\label{eq:condition0}\end{equation}
1651: In this case, the state can still be transferred conclusively (up
1652: to a known phase). The probability of failure at the $\ell$th measurement
1653: is given by\begin{equation}
1654: p_{\ell}=1-\frac{\left|F(\ell)\right|^{2}}{P(\ell-1)}.\label{eq:prob_fail_single}\end{equation}
1655: It is easy (but not very enlightening) to show~\cite{RANDOMRAIL}
1656: that the condition (\ref{eq:condition0}) is equivalent to\begin{equation}
1657: \left\Vert \prod_{k=1}^{\ell}\left\{ U(t_{k})Q\right\} \left|\boldsymbol{1,0}\right\rangle \right\Vert =\left\Vert \prod_{k=1}^{\ell}\left\{ U(t_{k})Q\right\} \left|\boldsymbol{0,1}\right\rangle \right\Vert \quad\forall\ell,\label{eq:finalcondition}\end{equation}
1658: and that the joint probability of failure - if at each measurement
1659: the above condition is fulfilled - is simply given by\begin{equation}
1660: P(\ell)=\left\Vert \prod_{k=1}^{\ell+1}\left\{ U(t_{k})Q\right\} \left|\boldsymbol{1,0}\right\rangle \right\Vert ^{2}.\label{eq:finalprob}\end{equation}
1661: It may look as if Eq. (\ref{eq:finalcondition}) was a complicated
1662: multi-time condition for the measuring times $t_{\ell}$, that becomes
1663: increasingly difficult to fulfil with a growing number of measurements.
1664: This is not the case. If proper measuring times have been found for
1665: the first $\ell-1$ measurements, a trivial time $t_{\ell}$ that
1666: fulfils Eq. (\ref{eq:finalcondition}) is $t_{\ell}=0.$ In this case,
1667: Bob measures immediately after the $(\ell-1)$th measurement and the
1668: probability amplitudes on his ends of the chains will be equal - and
1669: zero (a useless measurement). But since the left and right hand side
1670: of Eq. (\ref{eq:finalcondition}) when seen as functions of $t_{\ell}$
1671: are both almost-periodic functions with initial value zero, it is
1672: likely that they intersect many times, unless the system has some
1673: specific symmetry or the systems are completely different. Note that
1674: we do not claim at this point that any pair of chains will be capable
1675: of arbitrary perfect transfer. We will discuss in the next system
1676: how one can check this for a given system by performing some simple
1677: experimental tests.
1678:
1679:
1680: \section{Tomography\index{tomography}\label{sec:Tomography}}
1681:
1682: Suppose someone gives you two different experimentally designed spin
1683: chains. It may seem from the above that knowledge of the full Hamiltonian
1684: of both chains is necessary to check how well the system can be used
1685: for state transfer. This would be a very difficult task, because we
1686: would need access to all the spins along the channel to measure all
1687: the parameters of the Hamiltonian. In fact by expanding the projectors
1688: in Eq. (\ref{eq:finalcondition}) one can easily see that the only
1689: matrix elements of the evolution operator which are relevant for conclusive
1690: transfer are\begin{eqnarray}
1691: f_{N,1}(t) & = & \left\langle \boldsymbol{N,0}\right|U(t)\left|\boldsymbol{1,0}\right\rangle \label{eq:t1}\\
1692: f_{N,N}(t) & = & \left\langle \boldsymbol{N,0}\right|U(t)\left|\boldsymbol{N,0}\right\rangle \label{eq:t2}\\
1693: g_{N,1}(t) & = & \left\langle \boldsymbol{0,N}\right|U(t)\left|\boldsymbol{0,1}\right\rangle \label{eq:t3}\\
1694: g_{N,N}(t) & = & \left\langle \boldsymbol{0,N}\right|U(t)\left|\boldsymbol{0,N}\right\rangle .\label{eq:t4}\end{eqnarray}
1695: Physically, this means that the only relevant properties of the system
1696: are the transition amplitudes to \emph{arrive} at Bob's ends and to
1697: \emph{stay} there. The modulus of $f_{N,1}(t)$ and $f_{N,N}(t)$
1698: can be measured by initialising the system in the states $\left|\boldsymbol{1,0}\right\rangle $
1699: and $\left|\boldsymbol{N,0}\right\rangle $ and then performing a
1700: reduced density matrix tomography at Bob's site at different times
1701: $t$, and the complex phase of these functions is obtained by initialising
1702: the system in $\left(\left|\boldsymbol{0,0}\right\rangle +\left|\boldsymbol{1,0}\right\rangle \right)/\sqrt{2}$
1703: and $\left(\left|\boldsymbol{0,0}\right\rangle +\left|\boldsymbol{N,0}\right\rangle \right)/\sqrt{2}$
1704: instead. In the same way, $g_{N,1}(t)$ and $g_{N,N}(t)$ are obtained.
1705: All this can be done in the spirit of \emph{minimal control} at the
1706: sending and receiving ends of the chain only, and needs to be done
1707: only once. It is interesting to note that the dynamics in the middle
1708: part of the chain is not relevant at all. It is a \emph{\index{black box}black
1709: box} (see Fig. \ref{cap:blackbox}) that may involve even completely
1710: different interactions, number of spins, etc., as long as the total
1711: number of excitations is conserved. %
1712: \begin{figure}[tbh]
1713: \begin{centering}\includegraphics[width=0.8\columnwidth]{blackbox}\par\end{centering}
1714:
1715:
1716: \caption{\label{cap:blackbox}The relevant properties for conclusive transfer
1717: can be determined by measuring the response of the two systems at
1718: their ends only.}
1719: \end{figure}
1720: Once the transition amplitudes {[}Equations (\ref{eq:t1})-(\ref{eq:t4})]
1721: are known, one can search numerically for optimised measurement times
1722: $t_{\ell}$ using Eq. (\ref{eq:finalprob}) and the condition from
1723: Eq. (\ref{eq:finalcondition}).
1724:
1725: One weakness of the scheme described here is that the times at which
1726: Bob measures have to be very precise, because otherwise the measurements
1727: will not be unbiased with respect to $\alpha$ and $\beta.$ This
1728: demand can be relaxed by measuring at times where not only the probability
1729: amplitudes are similar, but also their \emph{slope} (see Fig. \ref{fig:coincidence}).
1730: The computation of these optimal timings for a given system may be
1731: complicated, but they only need to be done once.
1732:
1733:
1734: \section{Numerical Examples\label{sec:Numerical-Examples}}
1735:
1736: In this section, we show some numerical examples for two chains with
1737: Heisenberg couplings $J$ which are fluctuating. The Hamiltonians
1738: of the chains $i=1,2$ are given by\begin{eqnarray}
1739: H^{(i)} & = & \sum_{n=1}^{N-1}J(1+\delta_{n}^{(i)})\left(X_{n}^{(i)}X_{n+1}^{(i)}+Y_{n}^{(i)}Y_{n+1}^{(i)}+Z_{n}^{(i)}Z_{n+1}^{(i)}\right),\end{eqnarray}
1740: where $\delta_{n}^{(i)}$ are uniformly distributed random numbers
1741: from the interval $\left[-\Delta,\Delta\right].$ We have considered
1742: two different cases: in the first case, the $\delta_{n}^{(i)}$ are
1743: completely uncorrelated (i.e. independent for both chains and all
1744: sites along the chain). In the second case, we have taken into account
1745: a spacial correlation of the signs of the $\delta_{n}^{(i)}$ along
1746: each of the chains, while still keeping the two chains uncorrelated.
1747: For both cases, we find that arbitrarily perfect transfer remains
1748: possible except for some very rare realisations of the $\delta_{n}^{(i)}.$
1749:
1750: Because measurements must only be taken at times which fulfil the
1751: condition (\ref{eq:finalcondition}), and these times usually do not
1752: coincide with the optimal probability of finding an excitation at
1753: the ends of the chains, it is clear that the probability of failure
1754: at each measurement will in average be higher than for chains without
1755: fluctuations. Therefore, more measurements have to be performed in
1756: order to achieve the same probability of success. The price for noisy
1757: couplings is thus a longer transmission time and a higher number of
1758: gating operations at the receiving end of the chains. Some averaged
1759: values are given in Table \ref{cap:The-total-time}%
1760: \begin{table}[tbh]
1761: \begin{centering}\begin{tabular}{|c|c|c|c|c|c|}
1762: \hline
1763: &
1764: $\Delta=0$&
1765: $\Delta=0.01$&
1766: $\Delta=0.03$&
1767: $\Delta=0.05$&
1768: $\Delta=0.1$\tabularnewline
1769: \hline
1770: \hline
1771: $t$$\left[\frac{1}{J}\right]$&
1772: $377$&
1773: $524\pm27$&
1774: $694\pm32$&
1775: $775\pm40$&
1776: $1106\pm248$\tabularnewline
1777: \hline
1778: $M$&
1779: $28$&
1780: $43\pm3$&
1781: $58\pm3$&
1782: $65\pm4$&
1783: $110\pm25$\tabularnewline
1784: \hline
1785: \end{tabular}\par\end{centering}
1786:
1787:
1788: \caption{\label{cap:The-total-time}The total time $t$ and the number of
1789: measurements $M$ needed to achieve a probability of success of $99$\%
1790: for different fluctuation strengths $\Delta$ (uncorrelated case).
1791: Given is the statistical mean and the standard deviation. The length
1792: of the chain is $N=20$ and the number of random samples is $10.$
1793: For strong fluctuations $\Delta=0.1$, we also found particular samples
1794: where the success probability could not be achieved within the time
1795: range searched by the algorithm.}
1796: \end{table}
1797: for the Heisenberg chain with uncorrelated coupling fluctuations.
1798:
1799: For the case where the signs of the $\delta_{n}^{(i)}$ are correlated,
1800: we have used the same model as in~\cite{Fazio2005}, introducing
1801: the parameter $c$ such that\begin{equation}
1802: \delta_{n}^{(i)}\delta_{n-1}^{(i)}>0\qquad\textrm{with propability }c,\label{eq:corr1}\end{equation}
1803: and \begin{equation}
1804: \delta_{n}^{(i)}\delta_{n-1}^{(i)}<0\qquad\textrm{with propability }1-c.\label{eq:corr2}\end{equation}
1805: For $c=1$ ($c=0)$ this corresponds to the case where the signs
1806: of the couplings are completely correlated (anti-correlated). For
1807: $c=0.5$ one recovers the case of uncorrelated couplings. We can see
1808: from the numerical results in Table \ref{cap:correlated} that arbitrarily
1809: perfect transfer is possible for the whole range of $c.$ %
1810: \begin{table}[tbh]
1811: \begin{centering}\begin{tabular}{|c|c|c|c|c|c|c|}
1812: \hline
1813: &
1814: $c=0$&
1815: $c=0.1$&
1816: $c=0.3$&
1817: $c=0.7$&
1818: $c=0.9$&
1819: $c=1$\tabularnewline
1820: \hline
1821: \hline
1822: $t$$\left[\frac{1}{J}\right]$&
1823: $666\pm20$&
1824: $725\pm32$&
1825: $755\pm41$&
1826: $797\pm35$&
1827: $882\pm83$&
1828: $714\pm41$\tabularnewline
1829: \hline
1830: $M$&
1831: $256\pm2$&
1832: $62\pm3$&
1833: $65\pm4$&
1834: $67\pm4$&
1835: $77\pm7$&
1836: $60\pm4$\tabularnewline
1837: \hline
1838: \end{tabular}\par\end{centering}
1839:
1840:
1841: \caption{\label{cap:correlated}The total time $t$ and the number of measurements
1842: $M$ needed to achieve a probability of success of $99$\% for different
1843: correlations $c$ between the couplings {[}see Eq. (\ref{eq:corr1})
1844: and Eq. (\ref{eq:corr2})]. Given is the statistical mean and the
1845: standard deviation for a fluctuation strength of $\Delta=0.05$. The
1846: length of the chain is $N=20$ and the number of random samples is
1847: $20.$ }
1848: \end{table}
1849:
1850:
1851: For $\Delta=0$, we know from Section~\ref{sec:Estimati} that the
1852: time to transfer a state with probability of failure $P$ scales as\begin{equation}
1853: t(P)=0.33J^{-1}N^{1.6}\left|\ln P\right|.\label{eq:fit}\end{equation}
1854: If we want to obtain a similar formula in the presence of noise,
1855: we can perform a fit to the exact numerical data. For uncorrelated
1856: fluctuations of $\Delta=0.05,$ this is shown in Fig. \ref{cap:fitfig}.
1857: The best fit is given by\begin{equation}
1858: t(P)=0.2J^{-1}N^{1.9}\left|\ln P\right|.\label{eq:fit2}\end{equation}
1859: We conclude that weak fluctuations (say up to $5$\%) in the coupling
1860: strengths do not deteriorate the performance of our scheme much for
1861: the chain lengths considered. Both the transmission time and the number
1862: of measurements raise, but still in a reasonable way {[}cf. Table~\ref{cap:The-total-time}
1863: and Fig.~\ref{cap:fitfig}]. For larger fluctuations, the scheme
1864: is still applicable in principle, but the amount of junk (i.e. chains
1865: not capable of arbitrary perfect transfer) may get too large.%
1866: \begin{figure}[tbh]
1867: \begin{centering}\includegraphics[width=1\columnwidth]{3d_005}\par\end{centering}
1868:
1869:
1870: \caption{\label{cap:fitfig}Time $t$ needed to transfer a state with a given
1871: joint probability of failure $P$ across a chain of length $N$ with
1872: uncorrelated fluctuations of $\Delta=0.05.$ The points denote numerical
1873: data averaged over $100$ realisations, and the fit is given by Eq.
1874: (\ref{eq:fit2}). This figure should be compared with Fig. \ref{cap:Timefit}
1875: where $\Delta=0.$ }
1876: \end{figure}
1877:
1878:
1879: Note that we have considered the case where the fluctuations $\delta_{n}^{i}$
1880: are constant in time. This is a reasonable assumption if the dynamic
1881: fluctuations (e.g. those arising from thermal noise) can be neglected
1882: with respect to the constant fluctuations (e.g. those arising from
1883: manufacturing errors). If the fluctuations were varying with time,
1884: the tomography measurements in Sec. \ref{sec:Tomography} would involve
1885: a time-average, and Bob would not measure exactly at the correct times.
1886: The transferred state (\ref{eq:success}) would then be affected by
1887: both phase and amplitude noise.
1888:
1889:
1890: \section{Coupled chains\index{coupled chains}\label{sec:Coupled-chains}}
1891:
1892: Let us look at the condition for conclusive transfer in the more general
1893: scenario indicated by Fig.~ \ref{fig:A-black-box}: Alice and Bob
1894: have a black box\index{black box} acting as an amplitude damping
1895: channel in the following way. It has two inputs and two outputs. If
1896: Alice puts in state in the dual rail, \begin{equation}
1897: |\psi\rangle=\alpha|01\rangle+\beta|10\rangle,\label{eq:ampin}\end{equation}
1898: where $\alpha$ and $\beta$ are \emph{arbitrary and unknown} normalised
1899: amplitudes, then the output at Bob is given by\begin{equation}
1900: p|\phi\rangle\langle\phi|+\left(1-p\right)|00\rangle\langle00|,\label{eq:ampout}\end{equation}
1901: with a normalised ''success'' state\begin{equation}
1902: |\phi\rangle=\frac{1}{\sqrt{p}}\left[\alpha f|01\rangle+\beta g|10\rangle+\alpha\tilde{f}|10\rangle+\beta\tilde{g}|01\rangle\right].\end{equation}
1903: This black box describes the behaviour of an arbitrarily coupled qubit
1904: system that conserves the number of excitations and that is initialised
1905: in the all zero state, including parallel uncoupled chains, and coupled
1906: chains.%
1907: \begin{figure}[tbh]
1908: \begin{centering}\includegraphics[width=8cm]{bb2}\par\end{centering}
1909:
1910:
1911: \caption{\label{fig:A-black-box}Most general setting for conclusive transfer:
1912: A \emph{black box} with two inputs and two outputs, acting as an amplitude
1913: damping channel defined by Eqs.~(\ref{eq:ampin}) and (\ref{eq:ampout}) }
1914: \end{figure}
1915:
1916:
1917: From the normalisation of $|\phi\rangle$ it follows that\begin{equation}
1918: p=p(\alpha,\beta)=\left|\alpha f+\beta\tilde{g}\right|^{2}+\left|\beta g+\alpha\tilde{f}\right|^{2}.\end{equation}
1919: We are interested in conclusive transfer: by measuring the observable
1920: $|00\rangle\langle00|$ the Bob can project the output onto either
1921: the failure state $|00\rangle$ or $|\phi\rangle.$ This is clearly
1922: possible, but the question is if the output $|\phi\rangle$ and the
1923: input $|\psi\rangle$ are related by a \emph{unitary} operation.
1924:
1925: If Bob is able to recover the full information that Alice sent, then
1926: $p(\alpha,\beta)$ must be \emph{independent of $\alpha$ and} $\beta$
1927: (otherwise, some information on these amplitudes could be obtained
1928: by the measurement already, which contradicts the non-cloning theorem~\cite{NIELSEN}).
1929: This implies that $p(1,0)=p(0,1),$ i.e.\begin{equation}
1930: \left|f\right|^{2}+\left|\tilde{f}\right|^{2}=\left|\tilde{g}\right|^{2}+\left|g\right|^{2}.\label{eq:numma1}\end{equation}
1931: Because \begin{eqnarray}
1932: p\left(\frac{1}{\sqrt{2}},\frac{1}{\sqrt{2}}\right) & = & \frac{1}{2}\left|f+\tilde{g}\right|^{2}+\frac{1}{2}\left|g+\tilde{f}\right|^{2}\\
1933: & = & p(1,0)+\mbox{Re}\left\{ f^{*}\tilde{g}+g\tilde{f}^{*}\right\} \end{eqnarray}
1934: it also implies that \begin{equation}
1935: \mbox{Re}\left\{ f^{*}\tilde{g}+g\tilde{f}^{*}\right\} =0.\end{equation}
1936: Using the same trick for $p\left(\frac{1}{\sqrt{2}},\frac{i}{\sqrt{2}}\right)$
1937: we get that $\mbox{Im}\left\{ f^{*}\tilde{g}+g\tilde{f}^{*}\right\} =0$
1938: and therefore\begin{equation}
1939: f^{*}\tilde{g}+g\tilde{f}^{*}=0.\label{eq:numma2}\end{equation}
1940: If we write $|\psi\rangle=U|\phi\rangle$ we get\begin{equation}
1941: U=\frac{1}{\sqrt{p}}\left(\begin{array}{cc}
1942: f & \tilde{f}\\
1943: \tilde{g} & g\end{array}\right),\end{equation}
1944: which is a unitary operator if Eq.~(\ref{eq:numma1}) and (\ref{eq:numma2})
1945: hold. We thus come to the conclusion that conclusive transfer with
1946: the black box defined above is possible if and only if the probability
1947: $p$ is independent of $\alpha$ and $\beta.$ It is interesting to
1948: note that a vertical mirror symmetry of the system does not guarantee
1949: this. A counterexample is sketched in Fig.~\ref{cap:counter}: clearly
1950: the initial ({}``dark'') state $|01\rangle-|10\rangle$ does not
1951: evolve, whereas $|01\rangle+|10\rangle$ \emph{does.} Hence the probability
1952: must depend on $\alpha$ and $\beta.$ A trivial case where conclusive
1953: transfer works is given by two uncoupled chains, at times where $|f|^{2}=|g|^{2}.$
1954: This was discussed in Sect.~ \ref{sec:Arbitrarily-perfect-transferl}.
1955: A non-trivial example is given by the coupled system sketched in Fig.~\ref{cap:coupled}.
1956: This can be seen by splitting the Hamiltonian in a horizontal and
1957: vertical component,\begin{equation}
1958: H=H_{v}+H_{z}.\end{equation}
1959: By applying $H_{v}H_{z}$ and $H_{z}H_{v}$ on single-excitation
1960: states it is easily checked that they commute in the first excitation
1961: sector (this is not longer true in higher sectors). Since the probability
1962: is independent of $\alpha$ and $\beta$ in the uncoupled case it
1963: must also be true in the coupled case (a rotation in the subspace
1964: $\left\{ |01\rangle,|10\rangle\right\} $ does not harm).%
1965: \begin{figure}[tbh]
1966: \begin{centering}\includegraphics[width=0.6\columnwidth]{counter}\par\end{centering}
1967:
1968:
1969: \caption{\label{cap:counter}A simple counterexample for a vertically symmetric
1970: system where dual rail encoding is not possible. The black lines represent
1971: exchange couplings.}
1972: \end{figure}
1973: %
1974: \begin{figure}[tbh]
1975: \begin{centering}\includegraphics[width=0.6\columnwidth]{coupled}\par\end{centering}
1976:
1977:
1978: \caption{\label{cap:coupled}An example for a vertically symmetric system
1979: where dual rail encoding is possible. The black lines represent exchange
1980: couplings \emph{of equal strength}.}
1981: \end{figure}
1982:
1983:
1984: A final remark - as Alice and Bob alway only deal with the states
1985: $\left\{ |00\rangle,|10\rangle,|01\rangle\right\} $ it is obvious
1986: that the encoding used in this chapter is really living on \emph{qutrits.}
1987: In some sense it would be more natural to consider permanently coupled
1988: systems of qutrits\index{qutrits}, such as SU(3) chains~\cite{Sutherland1975,DUALRAIL,Bose2005a,Bayat2006}.
1989: The first level of the qutrit $|0\rangle$ is then used as a marker
1990: for ''no information here'', whereas the information is encoded
1991: in the states $|1\rangle$ and $|2\rangle.$ One would have to ensure
1992: that there is no transition between $|0\rangle$ and $|1\rangle,|2\rangle,$
1993: and that the system is initialised in the all zero state.
1994:
1995:
1996: \section{Conclusion\label{sec:Conclusions}}
1997:
1998: In conclusion, we have presented a simple scheme for conclusive and
1999: arbitrarily perfect quantum state transfer. To achieve this, two parallel
2000: spin chains (individually amplitude damping channels) have been used
2001: as one \emph{amplitude delaying channel}. We have shown that our scheme
2002: is more robust to decoherence and imperfect timing than the single
2003: chain schemes. We have also shown that the scheme is applicable to
2004: disordered and coupled chains. The scheme can be used as a way of
2005: improving any of the other schemes from the introduction. For instance,
2006: one may try to engineer the couplings to have a very high probability
2007: of success already at the first measurement, and use further measurements
2008: to compensate the errors of implementing the correct values for the
2009: couplings. We remark that the dual rail protocol is unrelated to error
2010: filtration~\cite{Gisin2005} where parallel channels are used for
2011: filtering out environmental effects on flying qubits, whereas the
2012: purpose of the dual rail protocol is to ensure the \emph{arrival}
2013: of the qubit. Indeed one could combine both protocols to send a qubit
2014: on say four rails to ensure the arrival \emph{and} filter errors.
2015: Finally, we note that in some recent work~\cite{Ericsson2005} it
2016: was shown that our encoding can be used to perform quantum gates while
2017: the state is transferred, and that it can increase the convergence
2018: speed if one performs measurements at intermediate positions~\cite{Bose2005e,Vaucher2005}.
2019:
2020:
2021: \chapter{Multi Rail encoding\label{cha:Multi-rail-and-Capacity}}
2022:
2023:
2024: \section{Introduction}
2025:
2026: In quantum information theory the rate $R$ of transferred qubits
2027: per channel is an important efficiency parameter~\cite{SHOR}. Therefore
2028: one question that naturally arises is whether or not there is any
2029: special meaning in the 1/2 value of $R$ achieved in the dual rail
2030: protocol of the last chapter. We will show now that this is not the
2031: case, because there is a way of bringing $R$ arbitrarily close to
2032: $1$ by considering multi rail encodings. Furthermore, in Section~\ref{sec:Arbit}
2033: it was still left open for which Hamiltonians the probability of success
2034: can be made arbitrarily close to $1.$ Here, we give a sufficient
2035: and easily attainable condition for achieving this goal.
2036:
2037: This chapter is organised as follows: the model and the notation are
2038: introduced in Sec.~\ref{s:sec1}. The efficiency and the fidelity
2039: of the protocol are discussed in Sec.~\ref{s:sec2} and in Sec.~\ref{s:sec3},
2040: respectively. Finally in Sec.~\ref{s:sec4} we prove a theorem which
2041: provides us with a sufficient condition for achieving efficient and
2042: perfect state transfer in quantum chains.
2043:
2044: %
2045: \begin{figure}[tbh]
2046: \begin{centering}\includegraphics[width=0.9\columnwidth]{m_fig1}\par\end{centering}
2047:
2048:
2049: \caption{\label{f:fig1} Schematic of the system: Alice and Bob operate $M$
2050: chains, each containing $N$ spins. The spins belonging to the same
2051: chain interact through the Hamiltonian $H$ which accounts for the
2052: transmission of the signal in the system. Spins of different chains
2053: do not interact. Alice encodes the information in the first spins
2054: of the chains by applying unitary transformations to her qubits. Bob
2055: recovers the message in the last spins of the chains by performing
2056: joint measurements.}
2057: \end{figure}
2058:
2059:
2060:
2061: \section{The model\label{s:sec1}}
2062:
2063: Assume that the two communicating parties operate on $M$ independent
2064: (i.e. non interacting) copies of the chain. This is quite a common
2065: attitude in quantum information theory~\cite{SHOR} where successive
2066: uses of a memoryless channel are formally described by introducing
2067: many parallel copies of the channel (see~\cite{Giovannetti2005}
2068: for a discussion on the possibility of applying this formal description
2069: to quantum chain models). Moreover for the case at hand the assumption
2070: of Alice and Bob dealing with {}``real'' parallel chains seems reasonable
2071: also from a practical point of view~\cite{Motoyama1996,Gambardella2002}.
2072: The idea is to use these copies to improve the overall fidelity of
2073: the communication. As usual, we assume Alice and Bob to control respectively
2074: the first and last qubit of each chain (see Fig.~\ref{f:fig1}).
2075: By preparing any superposition of her spins Alice can in principle
2076: transfer up to $M$ logical qubits. However, in order to improve the
2077: communication fidelity the two parties will find it more convenient
2078: to redundantly encode only a small number (say $Q(M)\leqslant M$)
2079: of logical qubits in the $M$ spins. By adopting these strategies
2080: Alice and Bob are effectively sacrificing the efficiency $R(M)=Q(M)/M$
2081: of their communication line in order to increase its fidelity. This
2082: is typical of any communication scheme and it is analogous to what
2083: happens in quantum error correction theory, where a single logical
2084: qubit is stored in many physical qubits. In the last chapter we have
2085: seen that for $M=2$ it is possible to achieve perfect state transfer
2086: of a single logical qubit with an efficiency equal to $1/2$. Here
2087: we will generalise such result by proving that there exist an optimal
2088: encoding-decoding strategy which asymptotically allows to achieve
2089: perfect state transfer \emph{and} optimal efficiency\index{efficiency},
2090: i.e. \begin{eqnarray}
2091: \lim_{M\rightarrow\infty}R(M)=1\;.\label{efficiency}\end{eqnarray}
2092: Our strategy requires Alice to prepare superpositions of the $M$
2093: chains where $\sim M/2$ of them have a single excitation in the first
2094: location while the remaining are in $|{\boldsymbol{0}}\rangle$. Since
2095: in the limit $M>>1$ the number of qubit transmitted is $\log\binom{M}{M/2}\approx M$,
2096: this architecture guarantees optimal efficiency~(\ref{efficiency}).
2097: On the other hand, our protocol requires Bob to perform collective
2098: measurements on his spins to determine if all the $\sim M/2$ excitations
2099: Alice is transmitting arrived at his location. We will prove that
2100: by repeating these detections many times, Bob is able to recover the
2101: messages with asymptotically perfect fidelity.
2102:
2103: Before beginning the analysis let us introduce some notation. The
2104: following definitions \emph{look} more complicated than they really
2105: \emph{are}; unfortunately we need them to carefully define the states
2106: that Alice uses for encoding the information. In order to distinguish
2107: the $M$ different chains we introduce the label $m=1,\cdots,M$:
2108: in this formalism $|\boldsymbol{n}\rangle_{m}$ represents the state
2109: of $m$-th chain with a single excitation in the $n$-th spin. In
2110: the following we will be interested in those configurations of the
2111: whole system where $K$ chains have a single excitation while the
2112: remaining $M-K$ are in $|\boldsymbol{0}\rangle$, as in the case\begin{equation}
2113: |\boldsymbol{1}\rangle_{1}\otimes|\boldsymbol{1}\rangle_{2}\cdots\otimes|\boldsymbol{1}\rangle_{K}\otimes|\boldsymbol{0}\rangle_{K+1}\cdots\otimes|\boldsymbol{0}\rangle_{M}\label{eq:example}\end{equation}
2114: where for instance the first $K$ chains have an excitation in the
2115: first chain location. Another more general example is given in Fig.
2116: \ref{cap:Example-of-our}. The complete characterisation of these
2117: vectors is obtained by specifying \emph{i)} \emph{which} chains possess
2118: a single excitation and \emph{ii)} \emph{where} these excitations
2119: are located horizontally along the chains. In answering to the point
2120: \emph{i)} we introduce the $K$-element subsets $S_{\ell}$, composed
2121: by the labels of those chains that contain an excitation. Each of
2122: these subsets $S_{\ell}$ corresponds to a subspace of the Hilbert
2123: space $\mathcal{H}(S_{\ell})$ with a dimension $N^{K}.$ The total
2124: number of such subsets is equal to the binomial coefficient $\binom{M}{K}$,
2125: which counts the number of possibilities in which $K$ objects (excitations)
2126: can be distributed among $M$ parties (parallel chains). In particular
2127: for any $\ell=1,\cdots,\binom{M}{K}$ the $\ell$-th subset $S_{\ell}$
2128: will be specified by assigning its $K$ elements, i.e. $S_{\ell}\equiv\{ m_{1}^{(\ell)},\cdots,m_{K}^{(\ell)}\}$
2129: with $m_{j}^{(\ell)}\in\{1,\cdots,M\}$ and $m_{j}^{(\ell)}<m_{j+1}^{(\ell)}$
2130: for all $j=1,\cdots,K$. To characterise the location of the excitations,
2131: point \emph{ii)}, we will introduce instead the $K$-dimensional vectors
2132: $\vec{n}\equiv(n_{1},\cdots,n_{K})$ where $n_{j}\in\{1,\cdots,N\}$.
2133: We can then define \begin{eqnarray}
2134: |\boldsymbol{\vec{n}};\ell\rangle\!\rangle\equiv\bigotimes_{j=1}^{K}|\boldsymbol{n_{j}}\rangle_{m_{j}^{(\ell)}}\;\bigotimes_{m^{\prime}\in{\overline{S}_{\ell}}}|\boldsymbol{0}\rangle_{m^{\prime}}\;,\label{nvec}\end{eqnarray}
2135: where $\overline{S}_{\ell}$ is the complementary of $S_{\ell}$
2136: to the whole set of chains.%
2137: \begin{figure}[tbh]
2138: \begin{centering}\includegraphics[width=0.6\columnwidth]{m_fig2}\par\end{centering}
2139:
2140:
2141: \caption{\label{cap:Example-of-our}Example of our notation for $M=5$ chains
2142: of length $N=6$ with $K=2$ excitations. The state above, given by
2143: $|\boldsymbol{0}\rangle_{1}\otimes|\boldsymbol{3}\rangle_{2}\otimes|\boldsymbol{0}\rangle_{3}\otimes|\boldsymbol{1}\rangle_{4}\otimes|\boldsymbol{0}\rangle_{5},$
2144: has excitations in the chains $m_{1}=2$ and $m_{2}=4$ at the horizontal
2145: position $n_{1}=3$ and $n_{2}=1$. It is in the Hilbert space $\mathcal{H}(S_{6})$
2146: corresponding to the subset $S_{6}=\{2,4\}$ (assuming that the sets
2147: $S_{\ell}$ are ordered in a canonical way, i.e. $S_{1}=\{1,2\},$
2148: $S_{2}=\{1,3\}$ and so on) and will be written as $|(3,1);6\rangle\!\rangle.$
2149: There are $\binom{5}{2}=10$ different sets $S_{\ell}$ and the number
2150: of qubits one can transfer using these states is $\log_{2}10\approx3.$
2151: The efficiency is thus given by $R\approx3/5$ which is already bigger
2152: than in the dual rail scheme.}
2153: \end{figure}
2154:
2155:
2156: The state (\ref{nvec}) represents a configuration where the $j$-th
2157: chain of the subset $S_{\ell}$ is in $|\boldsymbol{n_{j}}\rangle$
2158: while the chains that do not belong to $S_{\ell}$ are in $|\boldsymbol{0}\rangle$
2159: (see Fig. \ref{cap:Example-of-our} for an explicit example). The
2160: kets $|\boldsymbol{\vec{n}};\ell\rangle\!\rangle$ are a natural generalisation
2161: of the states $|\boldsymbol{n}\rangle_{1}\otimes|\boldsymbol{0}\rangle_{2}$
2162: and $|\boldsymbol{0}\rangle_{1}\otimes|\boldsymbol{n}\rangle_{2}$
2163: which were used for the dual rail encoding. They are useful for our
2164: purposes because they are mutually orthogonal, i.e. \begin{eqnarray}
2165: \!\langle\!\langle\boldsymbol{\vec{n}};\ell|\boldsymbol{\vec{n}^{\prime}};\ell^{\prime}\rangle\!\rangle=\delta_{\ell\ell^{\prime}}\;\delta_{\vec{n}\vec{n}^{\prime}}\;,\label{ortho}\end{eqnarray}
2166: and their time evolution under the Hamiltonian does not depend on
2167: $\ell.$ Among the vectors~(\ref{nvec}) those where all the $K$
2168: excitations are located at the beginning of the $S_{\ell}$ chains
2169: play an important role in our analysis. Here $\vec{n}=\vec{1}\equiv(1,\cdots,1)$
2170: and we can write \begin{eqnarray}
2171: |\boldsymbol{\vec{1}};\ell\rangle\!\rangle\equiv\bigotimes_{m\in S_{\ell}}|\boldsymbol{1}\rangle_{m}\;\bigotimes_{m^{\prime}\in{\overline{S}_{\ell}}}|\boldsymbol{0}\rangle_{m^{\prime}}\;.\label{in}\end{eqnarray}
2172: According to Eq.~(\ref{ortho}), for $\ell=1,\cdots,\binom{M}{K}$
2173: these states form orthonormal set of $\binom{M}{K}$ elements. Analogously
2174: by choosing $\vec{n}=\vec{N}\equiv(N,\cdots,N)$ we obtain the orthonormal
2175: set of $\binom{M}{K}$ vectors \begin{equation}
2176: |\boldsymbol{\vec{N}};\ell\rangle\!\rangle\equiv\bigotimes_{m\in S_{\ell}}|\boldsymbol{N}\rangle_{m}\;\bigotimes_{m^{\prime}\in{\overline{S}_{\ell}}}|\boldsymbol{0}\rangle_{m^{\prime}},\end{equation}
2177: where all the $K$ excitations are located at the end of the chains.
2178:
2179:
2180: \section{Efficient encoding}
2181:
2182: \label{s:sec2}
2183:
2184: If all the $M$ chains of the system are originally in $|\boldsymbol{0}\rangle$,
2185: the vectors~(\ref{in}) can be prepared by Alice by locally operating
2186: on her spins. Moreover since these vectors span a $\binom{M}{K}$
2187: dimensional subspace, Alice can encode in the chain $Q(M,K)=\log_{2}\binom{M}{K}$
2188: qubits of logical information by preparing the superpositions, \begin{eqnarray}
2189: |\Phi\rangle\!\rangle=\sum_{\ell}A_{\ell}\;|\boldsymbol{\vec{1}};\ell\rangle\!\rangle\;,\label{logicstate}\end{eqnarray}
2190: with $A_{\ell}$ complex coefficients. The efficiency of such encoding
2191: is hence $R(M,K)=\frac{\log_{2}\binom{M}{K}}{M}$ which maximised
2192: with respect to $K$ gives, \begin{eqnarray}
2193: R(M) & = & \frac{1}{M}\left\{ \begin{array}{ll}
2194: {\log_{2}\binom{M}{M/2}} & \;\mbox{for $M$ even}\\
2195: {\log_{2}\binom{M}{(M-1)/2}} & \;\mbox{for $M$ odd}\;.\end{array}\right.\label{efficiencymax}\end{eqnarray}
2196: The Stirling approximation can then be used to prove that this encoding
2197: is asymptotically efficient~(\ref{efficiency}) in the limit of large
2198: $M$, e.g. \begin{eqnarray}
2199: \log_{2}\binom{M}{M/2} & \approx & \log_{2}\frac{M^{M}}{(M/2)^{M}}=M.\end{eqnarray}
2200: Note that already for $M=5$ the encoding is more efficient (cf.
2201: Fig. \ref{cap:Example-of-our}) than in the dual rail encoding. In
2202: the remaining of the chapter we show that the encoding~(\ref{logicstate})
2203: provides perfect state transfer by allowing Bob to perform joint measurements
2204: at his end of the chains.
2205:
2206:
2207: \section{Perfect transfer\label{s:sec3}}
2208:
2209: Since the $M$ chains do not interact with each other and possess
2210: the same free Hamiltonian $H,$ the unitary evolution of the whole
2211: system is described by $U(t)\equiv\otimes_{m}u_{m}(t)$, with $u_{m}(t)$
2212: being the operator acting on the $m$-th chain. The time evolved of
2213: the input $|\boldsymbol{\vec{1}};\ell\rangle\!\rangle$ of Eq.~(\ref{in})
2214: is thus equal to \begin{eqnarray}
2215: & & U(t)|\boldsymbol{\vec{1}};\ell\rangle\!\rangle=\sum_{\vec{n}}F[\vec{n},\vec{1};t]\;|\boldsymbol{\vec{n}};\ell\rangle\!\rangle\;,\label{time}\end{eqnarray}
2216: where the sum is performed for all $n_{j}=1,\cdots,N$ and \begin{eqnarray}
2217: F[{\vec{n},\vec{n^{\prime}}};t]\equiv f_{n_{1},n_{1}^{\prime}}(t)\cdots f_{n_{K},n_{K}^{\prime}}(t)\;,\label{capitolF}\end{eqnarray}
2218: is a quantity which does \emph{not} depend on $\ell$. In Eq.~(\ref{time})
2219: the term ${\vec{n}}=\vec{N}$ corresponds to having all the $K$ excitations
2220: in the last locations of the chains. We can thus write \begin{eqnarray}
2221: U(t)|\boldsymbol{\vec{1}};\ell\rangle\!\rangle=\gamma_{1}(t)|\boldsymbol{\vec{N}};\ell\rangle\!\rangle+\sqrt{1-|\gamma_{1}(t)|^{2}}\;|\boldsymbol{\xi}(t);\ell\rangle\!\rangle\;,\label{time1}\end{eqnarray}
2222: where \begin{eqnarray}
2223: \gamma_{1}(t) & \equiv & \langle\!\langle\boldsymbol{\vec{N}};\ell|U(t)|\boldsymbol{\vec{1}};\ell\rangle\!\rangle=F[\vec{N},\vec{1};t]\label{gamma1}\end{eqnarray}
2224: is the probability amplitude that all the $K$ excitation of $|\boldsymbol{\vec{1}};\ell\rangle\!\rangle$
2225: arrive at the end of the chains, and \begin{eqnarray}
2226: |\boldsymbol{\xi}(t);\ell\rangle\!\rangle\equiv\sum_{\vec{n}\neq\vec{N}}F_{1}[\vec{n},\vec{1};t]\;|\boldsymbol{\vec{n}};\ell\rangle\!\rangle\;,\label{error}\end{eqnarray}
2227: with \begin{equation}
2228: F_{1}[\vec{n},\vec{1};t]\equiv\frac{F[\vec{n},\vec{1};t]}{\sqrt{1-|\gamma_{1}(t)|^{2}}},\label{eq:13b}\end{equation}
2229: is a superposition of terms where the number of excitations arrived
2230: to the end of the communication line is strictly less then $K$. It
2231: is worth noticing that Eq.~(\ref{ortho}) yields the following relations,
2232: \begin{eqnarray}
2233: \langle\!\langle\boldsymbol{\vec{N}};\ell|\boldsymbol{\xi}(t);\ell^{\prime}\rangle\!\rangle=0,\quad\!\langle\!\langle\boldsymbol{\xi}(t);\ell|\boldsymbol{\xi}(t);{\ell^{\prime}}\rangle\!\rangle=\delta_{\ell\ell^{\prime}}\;,\label{ortho1}\end{eqnarray}
2234: which shows that $\left\{ ||\boldsymbol{\xi}(t);\ell\rangle\!\rangle\right\} $
2235: is an orthonormal set of vectors which spans a subspace orthogonal
2236: to the states $|\boldsymbol{\vec{N}};\ell\rangle\!\rangle.$ The time
2237: evolution of the input state~(\ref{logicstate}) follows by linearity
2238: from Eq.~(\ref{time1}), i.e. \begin{eqnarray}
2239: |\Phi(t)\rangle\!\rangle=\gamma_{1}(t)\;|\Psi\rangle\!\rangle+\sqrt{1-|\gamma_{1}(t)|^{2}}\;|\overline{\Psi}(t)\rangle\!\rangle\;,\label{logicstateout}\end{eqnarray}
2240: with \begin{eqnarray}
2241: |\overline{\Psi}(t)\rangle\!\rangle & \equiv & \sum_{\ell}A_{\ell}\;|\boldsymbol{\xi}(t);\ell\rangle\!\rangle\;,\nonumber \\
2242: |\Psi\rangle\!\rangle & \equiv & \sum_{\ell}A_{\ell}\;|\boldsymbol{\vec{N}};\ell\rangle\!\rangle\;.\label{ok}\end{eqnarray}
2243: The vectors $|\Psi\rangle\!\rangle$ and $|\overline{\Psi}(t)\rangle\!\rangle$
2244: are unitary transformations of the input message~(\ref{logicstate})
2245: where the orthonormal set $\{|\boldsymbol{\vec{1}};\ell\rangle\!\rangle\}$
2246: has been rotated into $\{|\boldsymbol{\vec{N}};\ell\rangle\!\rangle\}$
2247: and $\{|\boldsymbol{\xi}(t);\ell\rangle\!\rangle\}$ respectively.
2248: Moreover $|\Psi\rangle\!\rangle$ is the configuration we need to
2249: have for perfect state transfer at the end of the chain. In fact it
2250: is obtained from the input message~(\ref{logicstate}) by replacing
2251: the components $|\boldsymbol{1}\rangle$ (excitation in the first
2252: spin) with $|\boldsymbol{N}\rangle$ (excitation in the last spin).
2253: From Eq.~(\ref{ortho1}) we know that $|\Psi\rangle\!\rangle$ and
2254: $|\overline{\Psi}(t)\rangle\!\rangle$ are orthogonal. This property
2255: helps Bob to recover the message $|\Psi\rangle\!\rangle$ from $|\Phi(t)\rangle\!\rangle$:
2256: he only needs to perform a collective measurement on the $M$ spins
2257: he is controlling to establish if there are $K$ or less excitations
2258: in those locations. The above is clearly a projective measurement
2259: that can be performed without destroying the quantum coherence associated
2260: with the coefficients $A_{\ell}$. Formally this can described by
2261: introducing the observable \begin{eqnarray}
2262: \Theta\equiv1-\sum_{\ell}|\boldsymbol{\vec{N}};\ell\rangle\!\rangle\langle\!\langle\boldsymbol{\vec{N}};\ell|\;.\label{observable}\end{eqnarray}
2263: A single measurement of $\Theta$ on $|\Phi(t_{1})\rangle\!\rangle$
2264: yields the outcome $0$ with probability $p_{1}\equiv|\gamma_{1}(t_{1})|^{2}$,
2265: and the outcome $+1$ with probability $1-p_{1}$. In the first case
2266: the system will be projected in $|\Psi\rangle\!\rangle$ and Bob will
2267: get the message. In the second case instead the state of the system
2268: will become $|\overline{\Psi}(t_{1})\rangle\!\rangle$. Already at
2269: this stage the two communicating parties have a success probability
2270: equal to $p_{1}$. Moreover, as in the dual rail protocol, the channels
2271: have been transformed into a quantum erasure channel~\cite{Bennett1997}
2272: where the receiver knows if the transfer was successful. Just like
2273: the dual rail encoding, this encoding can be used as a simple entanglement
2274: purification method in quantum chain transfer (see end of Section~\ref{sec:con}).
2275: The rate of entanglement that can be distilled is given by\begin{equation}
2276: R(M)\left|F[\vec{N},\vec{1};t]\right|^{2}=R(M)p(t)^{\left\lfloor M/2\right\rfloor },\end{equation}
2277: where we used Eq.~(\ref{capitolF}) and $p(t)\equiv\left|f_{N,1}(t)\right|^{2}.$
2278: As we can see, increasing $M$ on one hand increases $R(M),$ but
2279: on the other hand decreases the factor $p(t)^{\left\lfloor M/2\right\rfloor }.$
2280: Its maximum with respect to $M$ gives us a lower bound of the entanglement
2281: of distillation\index{entanglement of distillation} for a single
2282: spin chain, as shown in Fig.~\ref{fig:capacity}. We can also see
2283: that it becomes worth encoding on more than \emph{three} chains for
2284: conclusive transfer only when $p(t)>0.8.$
2285:
2286: Consider now what happens when Bob fails to get the right answer from
2287: the measurement. The state on which the chains is projected is explicitly
2288: given by \begin{eqnarray}
2289: |\overline{\Psi}(t_{1})\rangle\!\rangle & = & \sum_{\vec{n}\neq\vec{N}}F_{1}[\vec{n},\vec{1};t_{1}]\sum_{\ell}A_{\ell}|\boldsymbol{\vec{n}};\ell\rangle\!\rangle\;.\label{explicit}\end{eqnarray}
2290: Let us now consider the evolution of this state for another time
2291: interval $t_{2}$. By repeating the same analysis given above we obtain
2292: an expression similar to (\ref{logicstateout}), i.e. \begin{eqnarray}
2293: |\Phi(t_{2},t_{1})\rangle\!\rangle & = & \gamma_{2}\;|\Psi\rangle\!\rangle+\sqrt{1-|\gamma_{2}|^{2}}\;|\overline{\Psi}(t_{2},t_{1})\rangle\!\rangle\;,\label{logicstateout2}\end{eqnarray}
2294: where now the probability amplitude of getting all excitation in
2295: the $N$-th locations is described by \begin{equation}
2296: \gamma_{2}\equiv\sum_{\vec{n}\neq\vec{N}}F[\vec{N},\vec{n};t_{2}]\; F_{1}[\vec{n},\vec{1};t_{1}].\end{equation}
2297: In this case $|\overline{\Psi}(t)\rangle\!\rangle$ is replaced by
2298: \begin{eqnarray}
2299: |\overline{\Psi}(t_{2},t_{1})\rangle\!\rangle & = & \sum_{\ell}A_{\ell}\;|\boldsymbol{\xi}(t_{2},t_{1});\ell\rangle\!\rangle\;,\label{ko2}\end{eqnarray}
2300: with \begin{equation}
2301: |\boldsymbol{\xi}(t_{2},t_{1});\ell\rangle\!\rangle=\sum_{\vec{n}\neq\vec{N}}F_{2}[\vec{n},\vec{1};t_{2},t_{1}]|\boldsymbol{\vec{n}};\ell\rangle\!\rangle,\end{equation}
2302: and $F_{2}$ defined as in Eq.~(\ref{effeq}) (see below). In other
2303: words, the state $|\Phi(t_{2},t_{1})\rangle\!\rangle$ can be obtained
2304: from Eq.~(\ref{logicstateout}) by replacing $\gamma_{1}$ and $F_{1}$
2305: with $\gamma_{2}$ and $F_{2}$. Bob can hence try to use the same
2306: strategy he used at time $t_{1}$: i.e. he will check whether or not
2307: his $M$ qubits contain $K$ excitations. With (conditional) probability
2308: $p_{2}\equiv|\gamma_{2}|^{2}$ he will get a positive answer and his
2309: quantum register will be projected in the state $|\Psi\rangle\!\rangle$
2310: of Eq.~(\ref{ok}). Otherwise he will let the system evolve for another
2311: time interval $t_{3}$ and repeat the protocol. By reiterating the
2312: above analysis it is possible to give a recursive expression for the
2313: conditional probability of success $p_{q}\equiv|\gamma_{q}|^{2}$
2314: after $q-1$ successive unsuccessful steps. The quantity $\gamma_{q}$
2315: is the analogue of $\gamma_{2}$ and $\gamma_{1}$ of Eqs.~(\ref{gamma1})
2316: and (\ref{logicstateout2}). It is given by \begin{eqnarray}
2317: \gamma_{q}\equiv\sum_{\vec{n}\neq\vec{N}}F[\vec{N},\vec{n};t_{q}]\; F_{q-1}[\vec{n},\vec{1},t_{q-1},\cdots,t_{1}]\;,\label{gammaq}\end{eqnarray}
2318: where \begin{eqnarray}
2319: \lefteqn{F_{q-1}[\vec{n},\vec{1};t_{q-1},\cdots,t_{1}]}\label{effeq}\\
2320: & \equiv & \sum_{\vec{n}^{\prime}\neq\vec{N}}\frac{F[\vec{N},\vec{n}^{\prime};t_{q-1}]}{\sqrt{1-|\gamma_{q-1}|^{2}}}F_{q-2}[\vec{n}^{\prime},\vec{1};t_{q-2},\cdots,t_{1}]\nonumber \end{eqnarray}
2321: and $F_{1}[\vec{n},\vec{1},t]$ is given by Eq. (\ref{eq:13b}).
2322: In these equations $t_{q},\cdots,t_{1}$ are the \emph{time-intervals}
2323: that occurred between the various protocol steps. Analogously the
2324: conditional probability of failure at the step $q$ is equal to $1-p_{q}$.
2325: The probability of having $j-1$ failures and a success at the step
2326: $j$-th can thus be expressed as \begin{eqnarray}
2327: \pi(j) & = & p_{j}(1-p_{j-1})(1-p_{j-2})\cdots(1-p_{1})\;,\label{proba}\end{eqnarray}
2328: while the total probability of success after $q$ steps is obtained
2329: by the sum of $\pi(j)$ for all $j=1,\cdots,q$, i.e. \begin{eqnarray}
2330: P_{q} & = & \sum_{j=1}^{q}\pi(j)\;.\label{probtot}\end{eqnarray}
2331: Since $p_{j}\geqslant0$, Eq.~(\ref{probtot}) is a monotonic function
2332: of $q$. As a matter of fact in the next section we prove that under
2333: a very general hypothesis on the system Hamiltonian, the probability
2334: of success $P_{q}$ converges to $1$ in the limit of $q\rightarrow\infty$.
2335: This means that by repeating many times the collective measure described
2336: by $\Theta$ Bob is guaranteed to get, sooner or later, the answer
2337: $0$ and hence the message Alice sent to him. In other words our protocol
2338: allows perfect state transfer in the limit of repetitive collective
2339: measures. Notice that the above analysis applies for all classes of
2340: subsets $S_{\ell}$. The only difference between different choices
2341: of $K$ is in the velocity of the convergence of $P_{q}\rightarrow1$.
2342: In any case, by choosing $K\sim M/2$ Alice and Bob can achieve perfect
2343: fidelity \emph{and} optimal efficiency.
2344:
2345:
2346: \section{Convergence theorem\label{s:sec4}}
2347:
2348: \begin{framedtheorem}
2349: [Arbitrarly perfect transfer]\label{fra:If-there-exists}If there
2350: is no eigenvector $|e_{m}\rangle$ of the quantum chain Hamiltonian
2351: $H$ which is orthogonal to $|\boldsymbol{N}\rangle$, then there
2352: is a choice of the times intervals $t_{q},t_{q-1},\cdots,t_{1}$ such
2353: that the fidelity converges to $1$ as $q\rightarrow\infty.$
2354: \end{framedtheorem}
2355: Before proving this Theorem, let us give an intuitive reasoning for
2356: the convergence. The unitary evolution can be thought of of a \emph{rotation}
2357: in some abstract space, while the measurement corresponds to a \emph{projection.}
2358: The dynamics of the system is then represented by alternating rotations
2359: and projections. In general this will decrease the norm of each vector
2360: to null, unless the rotation axis is \emph{the same} as the projection
2361: axis.
2362:
2363: \begin{proof}
2364: The state of the system at a time interval of $t_{q}$ after the $(q-1)$-th
2365: failure can be expressed in compact form as follows \begin{eqnarray}
2366: |\Phi(t_{q},\cdots,t_{1})\rangle\!\rangle & = & \frac{U(t_{q})\Theta U(t_{q-1})\Theta\cdots U(t_{1})\Theta|\Phi\rangle\!\rangle}{\sqrt{(1-p_{q-1})\cdots(1-p_{1})}}\end{eqnarray}
2367: with $U(t)$ the unitary time evolution generated by the system Hamiltonian,
2368: and with $\Theta$ the projection defined in Eq.~(\ref{observable}).
2369: One can verify for instance that for $q=2$, the above equation coincides
2370: with Eq.~(\ref{logicstateout2}). {[}For $q=1$ this is just (\ref{logicstateout})
2371: evaluated at time $t_{1}$]. By definition the conditional probability
2372: of success at step $q$-th is equal to \begin{equation}
2373: p_{q}\equiv|\langle\!\langle\Psi|\Phi(t_{q},\cdots,t_{1})\rangle\!\rangle|^{2}.\end{equation}
2374: Therefore, Eq.~(\ref{proba}) yields \begin{eqnarray}
2375: \pi(q) & = & |\langle\!\langle\Psi|U(t_{q})\Theta U(t_{q-1})\Theta\cdots U(t_{1})\Theta|\Phi\rangle\!\rangle|^{2}\label{proba0}\\
2376: & = & |\langle\!\langle\boldsymbol{\vec{N}};\ell|U(t_{q})\Theta U(t_{q-1})\Theta\cdots U(t_{1})\Theta|\boldsymbol{\vec{1}};\ell\rangle\!\rangle|^{2}\;,\nonumber \end{eqnarray}
2377: where the second identity stems from the fact that, according to
2378: Eq.~(\ref{ortho}), $U(t)\Theta$ preserves the orthogonality relation
2379: among states $|\boldsymbol{\vec{n}};\ell\rangle\!\rangle$ with distinct
2380: values of $\ell.$ In analogy to the cases of Eqs.~(\ref{capitolF})
2381: and (\ref{gamma1}), the second identity of~(\ref{proba0}) establishes
2382: that $\pi(q)$ can be computed by considering the transfer of the
2383: input $|\boldsymbol{\vec{1}};\ell\rangle\!\rangle$ for \emph{arbitrary}
2384: $\ell$. The expression (\ref{proba0}) can be further simplified
2385: by noticing that for a given $\ell$ the chains of the subset $\overline{S}_{\ell}$
2386: contribute with a unitary factor to $\pi(q)$ and can be thus neglected
2387: (according to~(\ref{in}) they are prepared in $|\boldsymbol{0}\rangle$
2388: and do not evolve under $U(t)\Theta$). Identify $|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}$
2389: and $|\boldsymbol{\vec{N}}\rangle\!\rangle_{\ell}$ with the components
2390: of $|\boldsymbol{\vec{1}};\ell\rangle\!\rangle$ and $|\boldsymbol{\vec{N}};\ell\rangle\!\rangle$
2391: relative to the chains belonging to the subset $S_{\ell}$. In this
2392: notation we can rewrite Eq.~(\ref{proba0}) as \begin{eqnarray}
2393: \pi(q) & = & |_{\ell}\!\langle\!\langle\boldsymbol{\vec{N}}|U_{\ell}(t_{q})\Theta_{\ell}\;\cdots U_{\ell}(t_{1})\Theta_{\ell}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}|^{2}\;,\label{proba1}\end{eqnarray}
2394: where $\Theta_{\ell}=1-|\boldsymbol{\vec{N}}\rangle\!\rangle_{\ell}\langle\!\langle\boldsymbol{\vec{N}}|$
2395: and $U_{\ell}(t)$ is the unitary operator $\otimes_{m\in S_{\ell}}u_{m}(t)$
2396: which describes the time evolution of the chains of $S_{\ell}$. To
2397: prove that there exist suitable choices of $t_{\ell}$ such that the
2398: series~(\ref{probtot}) converges to $1$ it is sufficient to consider
2399: the case $t_{\ell}=t>0$ for all $j=1,\cdots,q$: this is equivalent
2400: to selecting decoding protocols with constant measuring intervals.
2401: By introducing the operator $T_{\ell}\equiv U_{\ell}(t)\Theta_{\ell}$,
2402: Eq.~(\ref{proba1}) becomes thus\begin{eqnarray}
2403: & & \pi(q)=|_{\ell}\!\langle\!\langle\boldsymbol{\vec{N}}|\;(T_{\ell})^{q}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}|^{2}\label{proba2}\\
2404: & & =_{\ell}\!\!\langle\!\langle\boldsymbol{\vec{1}}|(T_{\ell}^{\dag})^{q}|\boldsymbol{\vec{N}}\rangle\!\rangle_{\!\ell}\!\langle\!\langle\boldsymbol{\vec{N}}|\;(T_{\ell})^{q}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}=w(q)-w(q+1)\;,\nonumber \end{eqnarray}
2405: where \begin{equation}
2406: w(j)\equiv_{\ell}\!\langle\!\langle\boldsymbol{\vec{1}}|(T_{\ell}^{\dag})^{j}\;(T_{\ell})^{j}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}=\Vert(T_{\ell})^{j}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}\Vert^{2}\;,\end{equation}
2407: is the norm of the vector $(T_{\ell})^{j}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}$.
2408: Substituting Eq.~(\ref{proba2}) in Eq.~(\ref{probtot}) yields
2409: \begin{eqnarray}
2410: P_{q} & = & \sum_{j=1}^{q}\left[w(j)-w(j+1)\right]=1-w(q+1)\label{probtot1}\end{eqnarray}
2411: where the property $w(1)={_{\ell}\langle\!\langle}\boldsymbol{\vec{1}}|\Theta_{\ell}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}=1$
2412: was employed. Proving the thesis is hence equivalent to prove that
2413: for $q\rightarrow\infty$ the succession $w(q)$ nullifies. This last
2414: relation can be studied using properties of power bounded matrices~\cite{Schott1996}.
2415: In fact, by introducing the norm of the operator $(T_{\ell})^{q}$
2416: we have, \begin{eqnarray}
2417: w(q)=\Vert(T_{\ell})^{q}|\boldsymbol{\vec{1}}\rangle\!\rangle_{\ell}\Vert^{2}\leqslant\Vert(T_{\ell})^{q}\Vert^{2}\leqslant c\left(\frac{1+\rho(T_{\ell})}{2}\right)^{2q}\label{thesis}\end{eqnarray}
2418: where $c$ is a positive constant which does not depend on $q$ (if
2419: $S$ is the similarity transformation that puts $T_{\ell}$ into the
2420: Jordan canonical form, i.e. $J=S^{-1}T_{\ell}S,$ then $c$ is given
2421: explicitly by $c=\| S\|\:\| S^{-1}\|$) and where $\rho(T_{\ell})$
2422: is the spectral radius\index{spectral radius} of $T_{\ell}$, i.e.
2423: the eigenvalue of $T_{\ell}$ with maximum absolute value (N.B. even
2424: when $T_{\ell}$ is not diagonalisable this is a well defined quantity).
2425: Equation~(\ref{thesis}) shows that $\rho(T_{\ell})<1$ is a sufficient
2426: condition for $w(q)\rightarrow0$. In our case we note that, given
2427: any normalised eigenvector $|\lambda\rangle\!\rangle_{\ell}$ of $T_{\ell}$
2428: with eigenvalue $\lambda$ we have \begin{eqnarray}
2429: |\lambda|=\Vert T_{\ell}|\lambda\rangle\!\rangle_{\ell}\Vert=\Vert\Theta_{\ell}|\lambda\rangle\!\rangle_{\ell}\Vert\leqslant1\;,\label{thesis1}\end{eqnarray}
2430: where the inequality follows from the fact that $\Theta_{\ell}$
2431: is a projector. Notice that in Eq.~(\ref{thesis1}) the identity
2432: holds only if $|\lambda\rangle\!\rangle$ is also an eigenvector of
2433: $\Theta_{\ell}$ with eigenvalue $+1$, i.e. only if $|\lambda\rangle\!\rangle_{\ell}$
2434: is orthogonal to $|\boldsymbol{\vec{N}}\rangle\!\rangle_{\ell}$.
2435: By definition $|\lambda\rangle\!\rangle_{\ell}$ is eigenvector $T_{\ell}=U_{\ell}(t)\Theta_{\ell}$:
2436: therefore the only possibility to have the equality in Eq.~(\ref{thesis1})
2437: is that \emph{i)} $|\lambda\rangle\!\rangle_{\ell}$ is an eigenvector
2438: of $U_{\ell}(t)$ (i.e. an eigenvector of the Hamiltonian%
2439: \footnote{Notice that strictly speaking the eigenvectors of the Hamiltonian
2440: are not the same as those of the time evolution operators. The latter
2441: still can have evolution times at which additional degeneracy can
2442: increase the set of eigenstates. A trivial example is given for $t=0$
2443: where \emph{all} states become eigenstates. But it is always possible
2444: to find times $t$ at which the eigenstates of $U(t)$ coincide with
2445: those of $H$. %
2446: } $H_{\ell}^{\mbox{\small{tot}}}$ of the chain subset $S_{\ell}$)
2447: and \emph{ii)} it is orthogonal to $|\boldsymbol{\vec{N}}\rangle\!\rangle_{\ell}$.
2448: By negating the above statement we get a sufficient condition for
2449: the thesis. Namely, if all the eigenvectors $|\vec{E}\rangle\!\rangle_{\ell}$
2450: of $H_{\ell}^{\mbox{\small{tot}}}$ are not orthogonal to $|\boldsymbol{\vec{N}}\rangle\!\rangle_{\ell}$
2451: than the absolute values of the eigenvalues $\lambda$ of $T_{\ell}$
2452: are strictly smaller than $1$ which implies $\rho(T_{\ell})<1$ and
2453: hence the thesis. Since the $S_{\ell}$ channels are identical and
2454: do not interact, the eigenvectors $|\vec{E}\rangle\!\rangle_{\ell}\equiv\bigotimes_{m\in S_{\ell}}|e_{m}\rangle_{m}$
2455: are tensor product of eigenvectors $|e_{m}\rangle$ of the single
2456: chain Hamiltonians $H$. Therefore the sufficient condition becomes
2457: \begin{eqnarray}
2458: _{\ell}\langle\!\langle\vec{E}|\boldsymbol{\vec{N}}\rangle\!\rangle_{\ell}=\prod_{m\in S_{\ell}}{_{m}\!\langle\boldsymbol{N}}|e_{m}\rangle_{m}\neq0\;,\label{last}\end{eqnarray}
2459: which can be satisfied only if ${\langle\boldsymbol{N}}|e_{m}\rangle\neq0$
2460: for all eigenvectors $|e_{m}\rangle$ of the single chain Hamiltonian
2461: $H$.
2462: \end{proof}
2463: \begin{remark}
2464: While we have proven here that for equal time intervals the probability
2465: of success is converging to unity, in practice one may use \emph{optimal}
2466: measuring time intervals $t_{i}$ for a faster transfer (see also
2467: Section~\ref{sec:Estimati}). We also point out that timing errors
2468: may delay the transfer, but will not decrease its fidelity.
2469: \end{remark}
2470:
2471: \section{Quantum chains with nearest-neighbour interactions\label{sec:Quantum-chains-with}}
2472:
2473: It is worth noticing that Eq. (\ref{last}) is a very weak condition,
2474: because eigenstates of Hamiltonians are typically entangled. For instance,
2475: it holds for open chains with nearest neighbour-interactions:
2476:
2477: \begin{framedtheorem}
2478: [Multi rail protocol]\label{thm:Let2}Let $H$ be the Hamiltonian
2479: of an open nearest-neighbour quantum chain that conserves the number
2480: of excitations. If there is a time $t$ such that $f_{1,N}(t)\neq0$
2481: (i.e. the Hamiltonian is capable of transport between Alice and Bob)
2482: then the state transfer can be made arbitrarily perfect by using the
2483: \index{multi rail}multi rail protocol.
2484: \end{framedtheorem}
2485: \begin{proof}
2486: We show by contradiction that the criterion of Theorem~\ref{fra:If-there-exists}
2487: is fulfilled. Assume there exists a normalised eigenvector $\left|e\right\rangle $
2488: of the single chain Hamiltonian $H$ such that \begin{equation}
2489: \langle\boldsymbol{N}|e\rangle=0.\end{equation}
2490: Because $\left|e\right\rangle $ is an eigenstate, we can conclude
2491: that also \begin{equation}
2492: \left\langle e\left|H\right|\boldsymbol{N}\right\rangle =0.\label{eq:h_null}\end{equation}
2493: If we act with the Hamiltonian on the ket in Eq. (\ref{eq:h_null})
2494: we may get some term proportional to $\langle e|\boldsymbol{N}\rangle$
2495: (corresponding to an Ising-like interaction) and some part proportional
2496: to $\langle e|\boldsymbol{N-1}\rangle$ (corresponding to a hopping
2497: term; if this term did not exist, then clearly $f_{1,N}(t)=0$ for
2498: all times). We can thus conclude that \begin{equation}
2499: \langle e|\boldsymbol{N-1}\rangle=0.\label{eq:steptwo}\end{equation}
2500: Note that for a closed chain, e.g. a ring, this need not be the case,
2501: because then also a term proportional to $\langle e|\boldsymbol{N+1}\rangle=\langle e|\boldsymbol{1}\rangle$
2502: would occur. If we insert the Hamiltonian into Eq. (\ref{eq:steptwo})
2503: again, we can use the same reasoning to see that \begin{equation}
2504: \langle e|\boldsymbol{N-2}\rangle=\cdots=\langle e|\boldsymbol{1}\rangle=0\end{equation}
2505: and hence $\left|e\right\rangle =0,$ which is a contradiction to
2506: $\left|e\right\rangle $ being normalised.
2507: \end{proof}
2508:
2509: \section{Comparison with Dual Rail}
2510:
2511: As we have seen above, the Multi Rail protocol allows us in principle
2512: to reach in principle a rate arbitrarily close to one. However for
2513: a fair comparison with the Dual Rail protocol, we should also take
2514: into account the time-scale of the transfer. For the conclusive transfer
2515: of entanglement, we have seen in Section~\ref{s:sec3} that only
2516: for chains which have a success probability higher than $p(t)=0.8$
2517: it is worth encoding on more than three rails. The reason is that
2518: if the probability of success for a single excitation is $p,$ then
2519: the probability of success for $\left\lfloor M/2\right\rfloor $ excitations
2520: on on $M$ parallel chains is lowered to $p^{\left\lfloor M/2\right\rfloor }.$
2521: The protocol for three rails is always more efficient than on two,
2522: as still only one excitation is being used, but three complex amplitudes
2523: can be transferred per usage.
2524:
2525: For arbitrarily perfect transfer, the situation is slightly more complicated
2526: as the optimal choice of $M$ also depends on the joint probability
2527: of failure that one plans to achieve. Let us assume that at each step
2528: of the protocol, the success probability on a single chain is $p.$
2529: Then the number of steps to achieve a given probability of failure
2530: $P$ using $M$ chains is given by\begin{equation}
2531: \ell(P,M)=\max\left\{ \frac{\ln P}{\ln(1-p^{\left\lfloor M/2\right\rfloor })},1\right\} .\end{equation}
2532: If we assume that the total time-scale of the transfer is proportional
2533: to the number of steps, then the number of qubits that can be transferred
2534: per time interval is given by\begin{equation}
2535: v(P,M)\propto R(M)/\ell(P,M).\label{eq:rate}\end{equation}
2536: Optimising this rate with respect to $M$ we find three different
2537: regimes of the joint probability of failure (see Fig.~\ref{f:rates}).
2538: If one is happy with a large $P,$ then the Multi Rail protocol becomes
2539: superior to the Dual Rail for medium $p.$ For intermediate $P,$
2540: the threshold is comparable to the threshold of $p=0.8$ for conclusive
2541: transfer of entanglement. Finally for very low $P$ the Multi Rail
2542: only becomes useful for $p$ very close to one. In all three cases
2543: the threshold is higher than the $p(t)$ that can usually achieved
2544: with unmodulated Heisenberg chains. We can thus conclude that the
2545: Multi Rail protocol only becomes useful for chains which already have
2546: a very good performance.%
2547: \begin{figure}[tbh]
2548: \begin{centering}\includegraphics[width=0.9\columnwidth]{rates}\par\end{centering}
2549:
2550:
2551: \caption{\label{f:rates} Optimal rates (maximisation of Eq.~(\ref{eq:rate}
2552: with respect to $M$) for the Multi Rail protocol. Shown are three
2553: curves corresponding to different values of the joint probability
2554: of failure $P$ one plans to achieve.}
2555: \end{figure}
2556:
2557:
2558:
2559: \section{Conclusion}
2560:
2561: We thus conclude that any nearest-neighbour Hamiltonian that can transfer
2562: quantum information with nonzero fidelity (including the Heisenberg
2563: chains analysed above) is capable of efficient \emph{and} perfect
2564: transfer when used in the context of parallel chains. Hamiltonians
2565: with non-nearest neighbour interactions~\cite{Bose2006,Kay2006}
2566: can also be used as long as the criterion of Theorem~\ref{fra:If-there-exists}
2567: is fulfilled.
2568:
2569:
2570: \chapter{Ergodicity and mixing\label{cha:Ergodicity-and-mixing}}
2571:
2572:
2573: \section{Introduction}
2574:
2575: We have seen above that by applying measurements at the end of parallel
2576: chains, the state of the chain is converging to the ground state,
2577: and the quantum information is transferred to the receiver. Indeed,
2578: repetitive application of the same transformation is the key ingredient
2579: of many controls techniques. Beside quantum state transfer, they have
2580: been exploited to inhibit the decoherence of a system by frequently
2581: perturbing its dynamical evolution~\cite{Viola1998,VIOLA2,VIOLA3,VITALI,SIMON}
2582: (\emph{Bang-Bang control}) or to improve the fidelity of quantum gates~\cite{FRANSON}
2583: by means of frequent measurements (\emph{quantum Zeno-effect}~\cite{PERES}).
2584: Recently analogous strategies have also been proposed in the context
2585: of state preparation~\cite{KUMMERER,WELLENS,HOMOGENIZATION1,HOMOGENIZATION2,TERHAL,YUASA1,YUASA2}.
2586: In Refs.~\cite{HOMOGENIZATION1,HOMOGENIZATION2} for instance, a
2587: \emph{homogenisation} protocol was presented which allows one to transform
2588: any input state of a qubit into a some pre-fixed target state by repetitively
2589: coupling it with an external bath. A similar \emph{thermalisation}
2590: protocol was discussed in Ref.~\cite{TERHAL} to study the efficiency
2591: of simulating classical equilibration processes on a quantum computer.
2592: In Refs.~\cite{YUASA1,YUASA2} repetitive interactions with an externally
2593: monitored environment were exploited instead to implement \emph{purification}
2594: schemes which would allow one to extract pure state components from
2595: arbitrary mixed inputs. %
2596: \begin{figure}[tbh]
2597: \begin{centering}\includegraphics[width=0.75\textwidth]{illus}\par\end{centering}
2598:
2599:
2600: \caption{\label{F1}Schematic examples of the orbits of a ergodic and a mixing
2601: map.}
2602: \end{figure}
2603:
2604:
2605: The common trait of the proposals~\cite{KUMMERER,WELLENS,HOMOGENIZATION1,HOMOGENIZATION2,TERHAL,YUASA1,YUASA2}
2606: and the dual and multi rail protocols is the requirement that repeated
2607: applications of a properly chosen quantum operation $\tau$ converges
2608: to a fixed density matrix $x_{*}$ independently from the input state
2609: $x$ of the system, i.e. \begin{eqnarray}
2610: \tau^{n}(x)\equiv\underbrace{\tau\circ\tau\circ\cdots\circ\tau}_{n}\;(x)\Big|_{n\rightarrow\infty}\longrightarrow\;\; x_{*}\;,\label{mixing0}\end{eqnarray}
2611: with {}``$\circ$'' representing the composition of maps. Following
2612: the notation of Refs.~\cite{RAGINSKY,RICHTER} we call Eq.~(\ref{mixing0})
2613: the \emph{mixing} property of $\tau$. It is related with another
2614: important property of maps, namely \emph{ergodicity} (see Fig. \ref{F1}).
2615: The latter requires the existence of a unique input state $x_{0}$
2616: which is left invariant under a single application of the map%
2617: \footnote{Definition~(\ref{ergodic0}) may sound unusual for readers who are
2618: familiar with a definition of ergodicity from statistical mechanics,
2619: where a map is ergodic if its invariant sets have measure $0$ or
2620: $1.$ The notion of ergodicity used here is completely different,
2621: and was introduced in \cite{RAGINSKY,Raginsky2002,STRICTCONTRATIONS}.
2622: The set $\mathcal{X}$ one should have in mind here is not a measurable
2623: space, but the compact convex set of quantum states. A perhaps more
2624: intuitive definition of ergodicity based on the time average of observables
2625: is given by Lemma~\ref{thm:ergodic}).%
2626: }, i.e. \begin{eqnarray}
2627: \tau(x)=x\qquad\Longleftrightarrow\qquad x=x_{0}\;.\label{ergodic0}\end{eqnarray}
2628: Ergodicity and the mixing property are of high interest not only in
2629: the context of the above quantum information schemes. They also occur
2630: on a more fundamental level in statistical mechanics~\cite{STREATER}
2631: and open quantum systems~\cite{OPENQUANTUM,ALICKI}, where one would
2632: like to study irreversibility and relaxation to thermal equilibrium.
2633:
2634: In the case of quantum transformations one can show that mixing maps
2635: with convergence point $x_{*}$ are also ergodic with fixed point
2636: $x_{0}=x_{*}$. The opposite implication however is not generally
2637: true since there are examples of ergodic quantum maps which are not
2638: mixing (see the following). Sufficient conditions for mixing have
2639: been discussed both in the specific case of quantum channel~\cite{TERHAL,RAGINSKY,STRICTCONTRATIONS}
2640: and in the more abstract case of maps operating on topological spaces~\cite{STREATER}.
2641: In particular the Lyapunov direct method~\cite{STREATER} allows
2642: one to prove that an ergodic map $\tau$ is mixing if there exists
2643: a continuous functional $S$ which, for all points but the fixed one,
2644: is strictly increasing under $\tau$. Here we strengthen this criterion
2645: by weakening the requirement on $S$: our \emph{generalised} Lyapunov
2646: functions are requested only to have limiting values $S(\tau^{n}(x))|_{n\rightarrow\infty}$
2647: which differ from $S(x)$ for all $x\neq x_{0}$. It turns out that
2648: the existence of such $S$ is not just a \emph{sufficient} condition
2649: but also a \emph{necessary} condition for mixing. Exploiting this
2650: fact one can easily generalise a previous result on \emph{strictly
2651: contractive} maps~\cite{RAGINSKY} by showing that maps which are
2652: \emph{asymptotic deformations} (see Definition \ref{asymptdef}) are
2653: mixing. This has, unlike contractivity, the advantage of being a property
2654: independent of the choice of metric (see however~\cite{RICHTER}
2655: for methods of finding {}``tight'' norms). In some cases, the generalised
2656: Lyapunov method permits also to derive an optimal mixing condition
2657: for quantum channels based on the quantum relative entropy. Finally
2658: a slightly modified version of our approach which employs \emph{multi-central}
2659: Lyapunov functions yields a characterisation of (not necessarily mixing)
2660: maps which in the limit of infinitely many applications move all points
2661: toward a proper \emph{subset} (rather than a single point) of the
2662: input space.
2663:
2664: The introduction of a generalised Lyapunov method seems to be sound
2665: not only from a mathematical point of view, but also from a physical
2666: point of view. In effect, it often happens that the informations available
2667: on the dynamics of a system are only those related on its asymptotic
2668: behaviour (e.g. its thermalisation process), its finite time evolution
2669: being instead difficult to characterise. Since our method is explicitly
2670: constructed to exploit asymptotic features of the mapping, it provides
2671: a more effective way to probe the mixing property of the process.
2672:
2673: Presenting our results we will not restrict ourself to the case of
2674: quantum operations. Instead, following~\cite{STREATER} we will derive
2675: them in the more general context of continuous maps operating on topological
2676: spaces~\cite{TOPOLOGYBOOK}. This approach makes our results stronger
2677: by allowing us to invoke only those hypotheses which, to our knowledge,
2678: are strictly necessary for the derivation. It is important to stress
2679: however that, as a particular instance, all the Theorems and Lemmas
2680: presented in this chapter hold for any linear, completely positive,
2681: trace preserving map (i.e. quantum channels) operating on a compact
2682: subset of normed vectors (i.e. the space of the density matrices of
2683: a finite dimensional quantum system). Therefore readers who are not
2684: familiar with topological spaces can simply interpret our derivations
2685: as if they were just obtained for quantum channels acting on a finite
2686: dimensional quantum system.
2687:
2688: This chapter is organised as follows. In Sec.~\ref{GLT} the generalised
2689: Lyapunov method along with some minor results is presented in the
2690: context of topological spaces. Then quantum channels are analysed
2691: in Sec.~\ref{QC} providing a comprehensive summary of the necessary
2692: and sufficient conditions for the mixing property of these maps. Conclusions
2693: and remarks form the end of the chapter in Sec.~\ref{CONCLUSION}.
2694:
2695:
2696: \section{Topological background}
2697:
2698: Let us first introduce some basic topological background required
2699: for this chapter. A more detailed introduction is given in~\cite{TOPOLOGYBOOK}.
2700: Topological spaces are a very elegant way of defining compactness,
2701: convergence and continuity without requiring more than the following
2702: structure:
2703:
2704: \begin{definition}
2705: A \emph{topological space\index{topological space}} is a pair $(\mathcal{X},\mathcal{O})$
2706: of a set $\mathcal{X}$ and a set $\mathcal{O}$ of subsets of $\mathcal{X}$
2707: (called \emph{open} sets) such that
2708: \begin{enumerate}
2709: \item $\mathcal{X}$ and $\emptyset$ are open
2710: \item Arbitrary unions of open sets are open
2711: \item Intersections of two open sets are open
2712: \end{enumerate}
2713: \end{definition}
2714:
2715: \begin{example}
2716: If $\mathcal{X}$ is an arbitrary set, and $\mathcal{O}=\{\mathcal{X},\emptyset\}$, then $(\mathcal{X},\mathcal{O})$ is a topological space. $\mathcal{O}$ is called the \emph{trivial topology}.
2717: \end{example}
2718:
2719: \begin{definition}
2720: A topological space $\mathcal{X}$ is \emph{compact} if any open cover (i.e. a set of open sets such that $\mathcal{X}$ is contained in their union) contains a finite sub-cover.
2721: \end{definition}
2722: \begin{definition}
2723: A sequence $x_n \in \mathcal{X}$ is \emph{convergent} with limit $x_*$ if each open neighbourhood $O(x_*)$ (i.e. a set such that $x_* \in O(x_*) \in \mathcal{O}$ contains all but finitely many points of the sequence.
2724: \end{definition}
2725: \begin{definition}
2726: A map on a topological space is \emph{continous} if the preimage of any open set is open.
2727: \end{definition}
2728: This is already all we require to make useful statements about ergodicity and mixing. However, there are some subtleties which we need to take care of:
2729: \begin{definition}
2730: A topological space is \emph{sequentially compact} if every sequence has a convergent subsequence.
2731: \end{definition}
2732: Sequentially compactness is in general not related to compactness! Another subtlety is that with the above definition, a sequence can converge to many different points. For example, in the trivial topology, \emph{any} sequence converges to \emph{any} point. This motivates
2733: \begin{definition}
2734: A topological space is \emph{Hausdorff} if any two distinct points can by separated by open neighbourhoods.
2735: \end{definition}
2736: A limit of a sequence in a Hausdorff space is unique. All these problems disappear in metrical spaces:
2737: \begin{definition}
2738: A \emph{metric space} is a pair $(\mathcal{X},d)$ of a set $\mathcal{X}$ and a function $d:\mathcal{X}\times\mathcal{X}\rightarrow\mathbb{R}$ such that
2739: \begin{enumerate}
2740: \item $d(x,y)\ge 0$ and $d(x,y)=0 \Leftrightarrow x=y$
2741: \item $d(x,y)=d(y,x)$
2742: \item $d(x,z)\le d(x,y)+d(y,z)$
2743: \end{enumerate}
2744: \end{definition}
2745: A metric space becomes a topological space with the canonical topology
2746: \begin{definition}
2747: A subset $O$ of a metric space $\mathcal{X}$ is \emph{open} if $\forall x\in O$ there is an $\epsilon >0$ such that $\{y \in \mathcal{X} | d(x,y) \le \epsilon\} \subset O.$
2748: \end{definition}In a metric space with the canonical topology, compactness and sequentially
2749: compactness become equivalent. Furthermore, it is automatically Hausdorff
2750: (see Fig.~\ref{fig:topologic}).%
2751: \begin{figure}[htbp]
2752: \begin{centering}\includegraphics[width=0.5\textwidth]{plot1}\par\end{centering}
2753:
2754:
2755: \caption{\label{fig:topologic} Relations between topological spaces~\cite{TOPOLOGYBOOK}.
2756: The space of density matrices on which quantum channels are defined,
2757: is a compact and convex subset of a normed vectors space (the space
2758: of linear operators of the system) which, in the above graphical representation
2759: fits within the set of compact metric spaces.}
2760: \end{figure}
2761:
2762:
2763:
2764: \section{Generalised Lyapunov Theorem\label{GLT}}
2765:
2766:
2767: \subsection{Topological spaces}
2768:
2769: In this section we introduce the notation and derive our main result
2770: (the Generalised Lyapunov Theorem).
2771:
2772: \begin{definition}
2773: Let $\mathcal{X}$ be a topological space and let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2774: be a map. The sequence $x_{n}\equiv\tau^{n}(x)$, where $\tau^{n}$
2775: is a short-hand notation for the $n-$fold composition of $\tau,$
2776: is called the \emph{orbit} of $x.$ An element $x_{*}\in\mathcal{X}$
2777: is called a \emph{fixed point} of $\tau$ if and only if \begin{eqnarray}
2778: \tau(x_{*})=x_{*}\;.\label{defmixing}\end{eqnarray}
2779: $\tau$ is called \emph{ergodic}\index{ergodic} if and only if it
2780: has exactly one \index{fix-point}fixed point. $\tau$ is called \emph{mixing}\index{mixing}
2781: if and only if there exists a \emph{convergence} point $x_{*}\in\mathcal{X}$
2782: such that any orbit converges to it, i.e. \begin{eqnarray}
2783: \lim_{n\rightarrow\infty}x_{n}=x_{*}\quad\forall x\in\mathcal{X}\;.\label{defergo10}\end{eqnarray}
2784:
2785: \end{definition}
2786: A direct connection between ergodicity and mixing can be established
2787: as follows.
2788:
2789: \begin{lemma}
2790: \label{lem:hausdorff} Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2791: be a continuous mixing map on a topological Hausdorff space $\mathcal{X}.$
2792: Then $\tau$ is ergodic.
2793: \end{lemma}
2794: \begin{proof}
2795: Let $x_{*}$ be the convergence point of $\tau$ and let $x\in\mathcal{X}$
2796: arbitrary. Since $\tau$ is continuous we can perform the limit in
2797: the argument of $\tau,$ i.e.\begin{equation}
2798: \tau(x_{*})=\tau\left(\lim_{n\rightarrow\infty}\tau^{n}(x)\right)=\lim_{n\rightarrow\infty}\tau^{n+1}(x)=x_{*},\end{equation}
2799: which shows that $x_{*}$ is a fixed point of $\tau$. To prove that
2800: it is unique assume by contradiction that $\tau$ possesses a second
2801: fixed point $y_{*}\neq x_{*}$. Then $\lim_{n\rightarrow\infty}\tau^{n}(y_{*})=y_{*}\neq x_{*}$,
2802: so $\tau$ could not be mixing (since the limit is unique in a Hausdorff
2803: space -- see Fig.~\ref{fig:topologic}). Hence $\tau$ is ergodic.
2804: \end{proof}
2805: \begin{remarknn}
2806: The converse is not true in general, i.e. not every ergodic map is
2807: mixing (not even in Hausdorff topological spaces). A simple counterexample
2808: is given by $\tau:[-1,1]\rightarrow[-1,1]$ with $\tau(x)\equiv-x$
2809: and the usual topology of $\mathbb{R}$, which is ergodic with fixed
2810: point $0,$ but not mixing since for $x\neq0$, $\tau^{n}(x)=(-1)^{n}x$
2811: is alternating between two points. A similar counterexample will be
2812: discussed in the quantum channel section (see Example~\ref{exm:ergodic}).
2813: \end{remarknn}
2814: A well known criterion for mixing is the existence of a \emph{Lyapunov
2815: function}~\cite{STREATER}.
2816:
2817: \begin{definition}
2818: Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$ be a map on a topological
2819: space $\mathcal{X}.$ A continuous map $S:\mathcal{X}\rightarrow\mathbb{R}$
2820: is called a \emph{(strict) Lyapunov function for $\tau$ around $x_{*}\in\mathcal{X}$}
2821: if and only if\begin{equation}
2822: S\left(\tau(x)\right)>S(x)\quad\forall x\neq x_{*}.\end{equation}
2823:
2824: \end{definition}
2825: \begin{remarknn}
2826: At this point is is neither assumed that $x_{*}$ \emph{is} a fixed
2827: point, nor that $\tau$ is ergodic. Both follows from the theorem
2828: below.
2829: \end{remarknn}
2830: \begin{theorem}
2831: [Lyapunov function]Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2832: be a continuous map on a sequentially compact topological space $\mathcal{X}$.
2833: Let $S:\mathcal{X}\rightarrow\mathbb{R}$ be a Lyapunov function for
2834: $\tau$ around $x_{*}.$ Then $\tau$ is mixing with the fixed point
2835: $x_{*}$.
2836: \end{theorem}
2837: The proof of this theorem is given in~\cite{STREATER}. We will not
2838: reproduce it here, because we will provide a general theorem that
2839: includes this as a special case. In fact, we will show that the requirement
2840: of the strict monotonicity can be \emph{much} weakened, which motivates
2841: the following definition.
2842:
2843: \begin{definition}
2844: Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$ be a map on a topological
2845: space $\mathcal{X}.$ A continuous map $S:\mathcal{X}\rightarrow\mathbb{R}$
2846: is called a \emph{generalised Lyapunov function\index{generalised Lyapunov function}
2847: for $\tau$ around $x_{*}\in\mathcal{X}$} if and only if the sequence
2848: $S\left(\tau^{n}(x)\right)$ is point-wise convergent%
2849: \footnote{Point-wise convergence in this context means that for any fixed $x$
2850: the sequence $S_{n}\equiv S\left(\tau^{n}(x)\right)$ is convergent.%
2851: } for any $x\in\mathcal{X}$ and $S$ fulfils \begin{equation}
2852: S_{*}(x)\equiv\lim_{n\rightarrow\infty}S\left(\tau^{n}(x)\right)\neq S(x)\quad\forall x\neq x_{*}.\label{eq:weakinequality}\end{equation}
2853:
2854: \end{definition}
2855: In general it may be difficult to prove the point-wise convergence.
2856: However if $S$ is monotonic under the action of $\tau$ and the space
2857: is compact, the situation becomes considerably simpler. This is summarised
2858: in the following Lemma.
2859:
2860: \begin{lemma}
2861: \label{lem:monotonlyapov}Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2862: be map on a compact topological space. A continuous map $S:\mathcal{X}\rightarrow\mathbb{R}$
2863: which fulfils\begin{equation}
2864: S\left(\tau(x)\right)\geqslant S(x)\quad\forall x\in\mathcal{X},\label{eq:monotonic}\end{equation}
2865: and\begin{equation}
2866: S_{*}(x)\equiv\lim_{n\rightarrow\infty}S\left(\tau^{n}(x)\right)>S(x)\quad\forall x\neq x_{*}.\end{equation}
2867: for some fixed $x_{*}\in\mathcal{X}$ is a generalised Lyapunov function
2868: for $\tau$ around $x_{*}$.
2869: \end{lemma}
2870: \begin{proof}
2871: It only remains to show the (point-wise) convergence of $S\left(\tau^{n}(x)\right)$.
2872: Since $S$ is a continuous function on a compact space, it is bounded.
2873: By Eq. (\ref{eq:monotonic}) the sequence is monotonic. Any bounded
2874: monotonic sequence converges.
2875: \end{proof}
2876: \begin{corollary}
2877: Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$ be a map on a compact
2878: topological space. A continuous map $S:\mathcal{X}\rightarrow\mathbb{R}$
2879: which fulfils\begin{equation}
2880: S\left(\tau(x)\right)\geqslant S(x)\quad\forall x\in\mathcal{X},\end{equation}
2881: and\begin{equation}
2882: S\left(\tau^{N}(x)\right)>S(x)\quad\forall x\neq x_{*},\end{equation}
2883: for some fixed $N\in\mathbb{N}$ and for some $x_{*}\in\mathcal{X}$
2884: is a generalised Lyapunov function for $\tau$ around $x_{*}$.
2885: \end{corollary}
2886: \begin{remarknn}
2887: This implies that a strict Lyapunov function is a generalised Lyapunov
2888: function (with $N=1$).
2889: \end{remarknn}
2890: We can now state the main result of this section:
2891:
2892: \begin{framedtheorem}
2893: [Generalized Lyapunov function]\label{thm:weaklyapov} Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2894: be a continuous map on a sequentially compact topological space $\mathcal{X}.$
2895: Let $S:\mathcal{X}\rightarrow\mathbb{R}$ be a generalised Lyapunov
2896: function for $\tau$ around $x_{*}.$ Then $\tau$ is mixing with
2897: fixed point $x_{*}$.
2898: \end{framedtheorem}
2899: \begin{proof}
2900: Consider the orbit $x_{n}\equiv\tau^{n}(x)$ of a given $x\in\mathcal{X}.$
2901: Because $\mathcal{X}$ is sequentially compact, the sequence $x_{n}$
2902: has a convergent subsequence (see Fig.~\ref{fig:topologic}), i.e.
2903: $\lim_{k\rightarrow\infty}x_{n_{k}}\equiv\tilde{x}$. Let us assume
2904: that $\tilde{x}\neq x_{*}$ and show that this leads to a contradiction.
2905: By Eq. (\ref{eq:weakinequality}) we know that there exists a finite
2906: $N\in\mathbb{N}$ such that \begin{equation}
2907: S\left(\tau^{N}(\tilde{x})\right)\neq S(\tilde{x}).\label{eq:contra}\end{equation}
2908: Since $\tau^{N}$ is continuous we can perform the limit in the argument,
2909: i.e. \begin{equation}
2910: \lim_{k\rightarrow\infty}\tau^{N}\left(x_{n_{k}}\right)=\tau^{N}(\tilde{x}).\end{equation}
2911: Likewise, by continuity of $S$ we have \begin{equation}
2912: \lim_{k\rightarrow\infty}S\left(x_{n_{k}}\right)=S(\tilde{x}),\label{eq:ss1}\end{equation}
2913: and on the other hand\begin{equation}
2914: \lim_{k\rightarrow\infty}S\left(x_{N+n_{k}}\right)=\lim_{k\rightarrow\infty}S\left(\tau^{N}\left(x_{n_{k}}\right)\right)=S(\tau^{N}\tilde{x}),\label{eq:ss2}\end{equation}
2915: where the second equality stems from the continuity of the map $S$
2916: and $\tau^{N}$. Because $S$ is a generalised Lyapunov function,
2917: the sequence $S\left(x_{n}\right)$ is convergent. Therefore the subsequences
2918: (\ref{eq:ss1}) and (\ref{eq:ss2}) must have the same limit. We conclude
2919: that $S(\tau^{N}\tilde{x})=S(\tilde{x})$ which contradicts Eq. (\ref{eq:contra}).
2920: Hence $\tilde{x}=x_{*}.$ Since we have shown that any convergent
2921: subsequence of $\tau^{n}(x)$ converges to the same limit $x_{*}$,
2922: it follows by Lemma~\ref{lem:subsequences} that $\tau^{n}(x)$ is
2923: converging to $x_{*}.$ Since that holds for arbitrary $x$, it follows
2924: that $\tau$ is mixing.
2925: \end{proof}
2926: \begin{lemma}
2927: \label{lem:subsequences} Let $x_{n}$ be a sequence in a sequentially
2928: compact topological space $\mathcal{X}$ such that any convergent
2929: subsequence converges to $x_{*}.$ Then the sequence converges to
2930: $x_{*}.$
2931: \end{lemma}
2932: \begin{proof}
2933: We prove by contradiction: assume that the sequence does not converge
2934: to $x_{*}.$ Then there exists an open neighbourhood $O(x_{*})$ of
2935: $x_{*}$ such that for all $k\in\mathbb{N},$ there is a $n_{k}$
2936: such that $x_{n_{k}}\notin O(x_{*}).$ Thus the subsequence $x_{n_{k}}$
2937: is in the closed space $\mathcal{X}\backslash O(x_{*}),$ which is
2938: again sequentially compact. $x_{n_{k}}$ has a convergent subsequence
2939: with a limit in $\mathcal{X}\backslash O(x_{*}),$ in particular this
2940: limit is not equal to $x_{*}.$
2941: \end{proof}
2942: There is an even more general way of defining Lyapunov functions which
2943: we state here for completeness. It requires the concept of the quotient
2944: topology~\cite{TOPOLOGYBOOK}.
2945:
2946: \begin{definition}
2947: Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$ be a map on a topological
2948: space $\mathcal{X}.$ A continuous map $S:\mathcal{X}\rightarrow\mathbb{R}$
2949: is called a \emph{multi-central Lyapunov function for $\tau$ around
2950: $\mathcal{F}\subseteq\mathcal{X}$} if and only if the sequence $S\left(\tau^{n}(x)\right)$
2951: is point-wise convergent for any $x\in\mathcal{X}$ and if $S$ and
2952: $\tau$ fulfil the following three conditions: $S$ is constant on
2953: $\mathcal{F}$, $\tau(\mathcal{F})\subseteq\mathcal{F}$, and \begin{equation}
2954: S_{*}(x)\equiv\lim_{n\rightarrow\infty}S\left(\tau^{n}(x)\right)\neq S(x)\quad\forall x\notin\mathcal{F}.\end{equation}
2955:
2956: \end{definition}
2957: For these functions we cannot hope that the orbit is mixing. We can
2958: however show that the orbit is {}``converging'' to the set $\mathcal{F}$
2959: in the following sense:
2960:
2961: \begin{theorem}
2962: [Multi-central Lyapunov function]Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2963: be a continuous map on a sequentially compact topological space $\mathcal{X}.$
2964: Let $S:\mathcal{X}\rightarrow\mathbb{R}$ be a multi-central Lyapunov
2965: function for $\tau$ around $\mathcal{F}.$ Let $\varphi:\mathcal{X}\rightarrow\mathcal{X}/\mathcal{F}$
2966: be the continuous mapping into the quotient space (i.e. $\varphi(x)=[x]$
2967: for $x\in\mathcal{X}\backslash\mathcal{F}$ and $\varphi(x)=[\mathcal{F}]$
2968: for $x\in\mathcal{F})$. Then $\tilde{\tau}:\mathcal{X}/\mathcal{F}\rightarrow\mathcal{X}/\mathcal{F}$
2969: given by $\tilde{\tau}([x])=\varphi\left(\tau\left(\varphi^{-1}([x])\right)\right)$
2970: is mixing with fixed point $[\mathcal{F}]$.
2971: \end{theorem}
2972: \begin{proof}
2973: First note that $\tilde{\tau}$ is well defined because $\varphi$
2974: is invertible on $\mathcal{X}/\mathcal{F}\backslash[\mathcal{F}]$
2975: and $\tau(\mathcal{F})\subseteq\mathcal{F},$ so that $\tilde{\tau}([\mathcal{F}])=[\mathcal{F}]$.
2976: Since $\mathcal{X}$ is sequentially compact, the quotient space $\mathcal{X}/\mathcal{F}$
2977: is also sequentially compact. Note that for $O$ open, $\tilde{\tau}^{-1}(O)=\varphi\left(\tau^{-1}\left(\varphi^{-1}\left(O\right)\right)\right)$
2978: is the image of $\varphi$ of an open set in $\mathcal{X}$ and therefore
2979: (by definition of the quotient topology) open in $\mathcal{X}/\mathcal{F}.$
2980: Hence $\tilde{\tau}$ is continuous. The function $\tilde{S}([x]):\mathcal{X}/\mathcal{F}\rightarrow\mathcal{X}/\mathcal{F}$
2981: given by $\tilde{S}([x])=S(\varphi^{-1}([x]))$ is continuous and
2982: easily seen to be a generalised Lyapunov function around $[\mathcal{F}].$
2983: By Theorem \ref{thm:weaklyapov} it follows that $\tilde{\tau}$ is
2984: mixing.
2985: \end{proof}
2986:
2987: \subsection{Metric spaces}
2988:
2989: We now show that for the particular class of compact topological sets
2990: which posses a metric, the existence of a generalised Lyapunov function
2991: is also a necessary condition for mixing.
2992:
2993: \begin{theorem}
2994: [Lyapunov criterion]\label{thm:glyap} Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
2995: be a continuous map on a compact metric space $\mathcal{X}.$ Then
2996: $\tau$ is mixing with fixed point $x_{*}$ if and only if a generalised
2997: Lyapunov function around $x_{*}$ exists.
2998: \end{theorem}
2999: \begin{proof}
3000: Firstly, in metric spaces compactness and sequential compactness are
3001: equivalent, so the requirements of Theorem \ref{thm:weaklyapov} are
3002: met. Secondly, for any mixing map $\tau$ with fixed point $x_{*},$
3003: a generalised Lyapunov function around $x_{*}$ is given by $S(x)\equiv d(x_{*},x)$.
3004: In fact, it is continuous because of the continuity of the metric
3005: and satisfies \begin{equation}
3006: \lim_{n\rightarrow\infty}S\left(\tau^{n}(x)\right)=d(x_{*},x_{*})=0\leqslant d(x_{*},x)=S(x),\end{equation}
3007: where the equality holds if and only $x=x_{*}.$ We call $d(x_{*},x)$
3008: the \emph{trivial generalised Lyapunov function}.
3009: \end{proof}
3010: \begin{remark}
3011: In the above Theorem we have not used all the properties of the metric.
3012: In fact a continuous \emph{semi-metric} (i.e. without the triangle
3013: inequality) would suffice.
3014: \end{remark}
3015: The trivial Lyapunov function requires knowledge of the fixed point
3016: of the map. There is another way of characterising mixing maps as
3017: those which bring elements closer to \emph{each other} (rather than
3018: closer to the fixed point).
3019:
3020: \begin{definition}
3021: A map $\tau:\mathcal{X}\rightarrow\mathcal{X}$ is on a metric space
3022: is called a \emph{non-expansive map}\index{non-expansive map} if
3023: and only if \begin{equation}
3024: d(\tau(x),\tau(y))\leqslant d(x,y)\quad\forall x,y\in\mathcal{X},\end{equation}
3025: a \emph{weak contraction}\index{weak contraction} if and only if
3026: \begin{equation}
3027: d(\tau(x),\tau(y))<d(x,y)\quad\forall x,y\in\mathcal{X},\, x\neq y,\end{equation}
3028: and a \emph{strict contraction}\index{strict contraction} if and
3029: only if there exists a $k<1$ such that\begin{equation}
3030: d(\tau(x),\tau(y))\leqslant k\, d(x,y)\quad\forall x,y\in\mathcal{X}\,.\end{equation}
3031:
3032: \end{definition}
3033: \begin{remark}
3034: The notation adopted here is slightly different from the definitions
3035: used by other Authors~\cite{RAGINSKY,RUSKAI,WERNER} who use contraction
3036: to indicate our non-expansive maps. Our choice is motivated by the
3037: need to clearly distinguish between non-expansive transformation and
3038: weak contractions.
3039: \end{remark}
3040: We can generalise the above definition in the following way:
3041:
3042: \begin{definition}
3043: \label{asymptdef}A map $\tau:\mathcal{X}\rightarrow\mathcal{X}$
3044: on a metric space is called an \emph{asymptotic deformation}\index{asymptotic deformation}
3045: if and only if the sequence $d(\tau^{n}(x),\tau^{n}(y))$ converges
3046: point-wise for all $x,y\in\mathcal{X}$ and\begin{equation}
3047: \lim_{n\rightarrow\infty}d(\tau^{n}(x),\tau^{n}(y))\neq d(x,y)\quad\forall x,y\in\mathcal{X},\, x\neq y.\end{equation}
3048:
3049: \end{definition}
3050: \begin{lemma}
3051: \label{asymptlem}Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$ be
3052: a non-expansive map on a metric space $\mathcal{X},$ and let \begin{equation}
3053: d(\tau^{N}(x),\tau^{N}(y))<d(x,y)\quad\forall x,y\in\mathcal{X},\, x\neq y\end{equation}
3054: for some fixed $N\in\mathbb{N}.$ Then $\tau$ is an asymptotic deformation.
3055: Then $\tau$ is an asymptotic deformation.
3056: \end{lemma}
3057: \begin{proof}
3058: The existence of the limit $\lim_{n\rightarrow\infty}d(\tau^{n}(x),\tau^{n}(y))$
3059: follows from the monotonicity and the fact the any metric is lower
3060: bounded.
3061: \end{proof}
3062: \begin{remarknn}
3063: Any weak contraction is an asymptotic deformation (with $N=1$).
3064: \end{remarknn}
3065: \begin{framedtheorem}
3066: [Asymptotic deformations]\label{thm:weakBanach} Let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
3067: be a continuous map on a compact metric space $\mathcal{X}$ with
3068: at least one fixed point. Then $\tau$ is mixing if and only if $\tau$
3069: is an asymptotic deformation.
3070: \end{framedtheorem}
3071: \begin{proof}
3072: Firstly assume that $\tau$ is an asymptotic deformation. Let $x_{*}$
3073: be a fixed point and define $S(x)=d(x_{*},x).$\begin{eqnarray}
3074: \lim_{n\rightarrow\infty}S(\tau^{n}(x)) & = & \lim_{n\rightarrow\infty}d(x_{*},\tau^{n}(x))\nonumber \\
3075: & = & \lim_{n\rightarrow\infty}d(\tau^{n}(x_{*}),\tau^{n}(x))\neq d(x_{*},x)=S(x)\quad\forall x\neq x_{*},\end{eqnarray}
3076: hence $S(x)$ is a generalised Lyapunov function. By Theorem \ref{thm:weaklyapov}
3077: it follows that $\tau$ is mixing. Secondly, if $\tau$ is mixing,
3078: then \begin{equation}
3079: \lim_{n\rightarrow\infty}d(\tau^{n}(x),\tau^{n}(y))=d(x_{*},x_{*})=0\neq d(x,y)\quad\forall x,y\in\mathcal{X},\, x\neq y,\end{equation}
3080: so $\tau$ is an asymptotic deformation.
3081: \end{proof}
3082: \begin{remarknn}
3083: Note that the existence of a fixed point is assured if $\tau$ is
3084: a weak contraction on a compact space~\cite{STAKGOLD}, or if the
3085: metric space is convex compact \cite{DUGUNDJI}.
3086: \end{remarknn}
3087: As a special case, we get the following result:
3088:
3089: \begin{corollary}
3090: \label{cor:contraction1} Any weak contraction $\tau$ on a compact
3091: metric space is mixing.
3092: \end{corollary}
3093: \begin{proof}
3094: Since the space is compact $\tau$ has at least one fixed point. Moreover
3095: from Lemma~\ref{asymptlem} we know that $\tau$ is an asymptotic
3096: deformation. Then Theorem~\ref{thm:weakBanach} applies.
3097: \end{proof}
3098: \begin{remarknn}
3099: This result can be seen as an instance of Banach contraction principle
3100: on compact spaces. In the second part of the chapter we will present
3101: a counterexample which shows that weak contractivity is only a sufficient
3102: criterion for mixing (see Example~\ref{exm:mixing}). In the context
3103: of quantum channels an analogous criterion was suggested in~\cite{STRICTCONTRATIONS,RAGINSKY}
3104: which applied to strict contractions. We also note that for weak and
3105: strict contractions, the trivial generalised Lyapunov function (Theorem
3106: \ref{thm:glyap}) is a strict Lyapunov function.
3107: \end{remarknn}
3108: Lemma~\ref{thm:ergodic} states the ergodic theorem by Birkhoff~\cite{BIRKHOFF}
3109: which, in the context of normed vector spaces, shows the equivalence
3110: between the definition of ergodicity of Eq.~(\ref{defergo10}) and
3111: the standard time average definition.
3112:
3113: \begin{lemma}
3114: \label{thm:ergodic}Let $\mathcal{X}$ be a convex and compact subset
3115: of a normed vector space, and let $\tau:\mathcal{X}\rightarrow\mathcal{X}$
3116: be a continuous map. If $\tau$ is ergodic with fixed point $x_{*},$
3117: then \begin{eqnarray}
3118: \lim_{n\rightarrow\infty}\frac{1}{n+1}\sum_{\ell=0}^{n}\tau^{\ell}(x)=x_{*}\;.\label{average}\end{eqnarray}
3119:
3120: \end{lemma}
3121: \begin{proof}
3122: Define the sequence $A_{n}\equiv\frac{1}{n+1}\sum_{\ell=0}^{n}\tau^{\ell}(x)$.
3123: Let then $M$ be the upper bound for the norm of vectors in $\mathcal{X}$,
3124: i.e. $M\equiv\sup_{x\in\mathcal{X}}\| x\|<\infty$. which exists because
3125: $\mathcal{X}$ is compact. The sequence $A_{n}$ has a convergent
3126: subsequence $A_{n_{k}}$ with limit $\tilde{A}.$ Since $\tau$ is
3127: continuous one has $\lim_{k\rightarrow\infty}\tau(A_{n_{k}})=\tau(\tilde{A})$.
3128: On the other hand, we have\begin{equation}
3129: \|\tau(A_{n_{k}})-A_{n_{k}}\|=\frac{1}{n_{k}+1}\|\tau^{n_{k}+1}(x)-x\|\leqslant\frac{\|\tau^{n_{k}+1}(x)\|+\| x\|}{n_{k}+1}\leqslant\frac{2M}{n_{k}+1},\end{equation}
3130: so the two sequences must have the same limit, i.e. $\tau(\tilde{A})=\tilde{A}$.
3131: Since $\tau$ is ergodic, we have $\tilde{A}=x_{*}$ and $\lim_{n\rightarrow\infty}A_{n}=x_{*}$
3132: by Lemma \ref{lem:subsequences}.
3133: \end{proof}
3134: \begin{remarknn}
3135: Note that if $\tau$ has a second fixed point $y_{*}\neq x_{*}$,
3136: then for all $n$ one has $\frac{1}{n+1}\sum_{\ell=0}^{n}\tau^{\ell}(y_{*})=y_{*}$,
3137: so Eq.~(\ref{average}) would not apply.
3138: \end{remarknn}
3139:
3140: \section{Quantum Channels}
3141:
3142: \label{QC}
3143:
3144: In this Section we discuss the mixing properties of quantum channels\index{quantum channel}~\cite{NIELSEN}
3145: which account for the most general evolution a quantum system can
3146: undergo including measurements and coupling with external environments.
3147: In this context solving the mixing problem~(\ref{mixing0}) is equivalent
3148: to determine if repetitive application of a certain physical transformation
3149: will drive any input state of the system (i.e. its density matrices)
3150: into a unique output configuration. The relationship between the different
3151: mixing criteria one can obtain in this case is summarised in Fig.~\ref{fig:relations}.
3152:
3153: At a mathematical level quantum channels correspond to linear maps
3154: acting on the density operators $\rho$ of the system and satisfying
3155: the requirement of being completely positive and trace preserving
3156: (\index{CPT}CPT). For a formal definition of these properties we
3157: refer the reader to~\cite{KRAUS,WERNER,KEYL}: here we note only
3158: that a necessary and sufficient condition to being CPT is to allow
3159: Kraus decomposition~\cite{KRAUS} or, equivalently, Stinespring dilation~\cite{STINE}.
3160: Our results are applicable if the underlying Hilbert space is finite-dimensional.
3161: In such regime there is no ambiguity in defining the convergence of
3162: a sequence since all operator norms are equivalent (i.e. given two
3163: norms one can construct an upper and a lower bound for the first one
3164: by properly scaling the second one). Also the set of bounded operators
3165: and the set of operators of Hilbert-Schmidt class coincide. For the
3166: sake of definiteness, however, we will adopt the trace-norm which,
3167: given the linear operator $\Theta:\mathcal{H}\rightarrow\mathcal{H}$,
3168: is defined as $\|\Theta\|_{1}=\mbox{Tr}[\sqrt{\Theta^{\dag}\Theta}]$
3169: with $\mbox{Tr}[\cdots]$ being the trace over $\mathcal{H}$ and
3170: $\Theta^{\dag}$ being the adjoint of $\Theta$. This choice is in
3171: part motivated by the fact~\cite{RUSKAI} that any quantum channel
3172: is non-expansive with respect to the metric induced%
3173: \footnote{This is just the trace distance $d(\rho,\sigma)=\|\rho-\sigma\|_{1}$.%
3174: } by $\|\cdot\|_{1}$ (the same property does not necessarily apply
3175: to other operator norms, e.g. the Hilbert-Schmidt norm, also when
3176: these are equivalent to $\|\cdot\|_{1}$).
3177:
3178: We start by showing that the mixing criteria discussed in the first
3179: half of the chapter do apply to the case of quantum channel. Then
3180: we will analyse these maps by studying their linear extensions in
3181: the whole vector space formed by the linear operators of $\mathcal{H}$.%
3182: \begin{figure}[t]
3183: \begin{centering}\includegraphics[width=1\textwidth]{relations5}\par\end{centering}
3184:
3185:
3186: \caption{\label{fig:relations}Relations between the different properties
3187: of a quantum channel.}
3188: \end{figure}
3189:
3190:
3191:
3192: \subsection{Mixing criteria for Quantum Channels}
3193:
3194: \label{sec:mixing}
3195:
3196: Let $\mathcal{H}$ be a finite dimensional Hilbert space and let $\mathcal{S}(\mathcal{H})$
3197: be the set of its density matrices $\rho$. The latter is a convex
3198: and compact subset of the larger normed vector space $\mathcal{L}(\mathcal{H})$
3199: composed by the linear operators $\Theta:\mathcal{H}\rightarrow\mathcal{H}$
3200: of $\mathcal{H}$. From this and from the fact that CPT maps are continuous
3201: (indeed they are linear) it follows that for a quantum channel there
3202: always exists at least one density operator which is a fixed point~\cite{TERHAL}.
3203: It also follows that all the results of the previous section apply
3204: to quantum channels. In particular Lemma~\ref{lem:hausdorff} holds,
3205: implying that any mixing quantum channel must be ergodic. The following
3206: example shows, however, that it is possible to have ergodic quantum
3207: channels which are not mixing.
3208:
3209: \begin{example}
3210: \label{exm:ergodic} Consider the qubit quantum channel $\tau$ obtained
3211: by cascading a completely decoherent channel with a NOT gate. Explicitly
3212: $\tau$ is defined by the transformations $\tau(|0\rangle\langle0|)=|1\rangle\langle1|$,
3213: $\tau(|1\rangle\langle1|)=|0\rangle\langle0|$, and $\tau(|0\rangle\langle1|)=\tau(|1\rangle\langle0|)=0$
3214: with $|0\rangle,|1\rangle$ being the computational basis of the qubit.
3215: This map is ergodic with fixed point given by the completely mixed
3216: state $(|0\rangle\langle0|+|1\rangle\langle1|)/2$. However it is
3217: trivially not mixing since, for instance, repetitive application of
3218: $\tau$ on $|0\rangle\langle0|$ will oscillate between $|0\rangle\langle0|$
3219: and $|1\rangle\langle1|$.
3220: \end{example}
3221: Theorems~\ref{thm:weakBanach} implies that a quantum channel $\tau:\mathcal{S}(\mathcal{H})\rightarrow\mathcal{S}(\mathcal{H})$
3222: is mixing if and only if it is an asymptotic deformation. As already
3223: pointed out in the introduction, this property is \emph{metric independent}
3224: (as opposed to contractivity). Alternatively, if the fixed point of
3225: a quantum channel is known, then one may use the trivial generalised
3226: Lyapunov function (Theorem~\ref{thm:glyap}) to check if it is mixing.
3227: However both criteria depend on the metric distance, which usually
3228: has no easy physical interpretation. A more useful choice is the quantum
3229: relative entropy\index{quantum relative entropy}, which is defined
3230: as \begin{eqnarray}
3231: H(\rho,\sigma)\equiv\textrm{Tr}\rho(\log\rho-\log\sigma).\end{eqnarray}
3232: The quantum relative entropy is continuous in finite dimension \cite{JENS}
3233: and can be used as a measure of \emph{distance} (though it is not
3234: a metric). It is finite if the support of $\rho$ is contained in
3235: the support of $\sigma.$ To ensure that it is a continuous function
3236: on a compact space, we choose $\sigma$ to be faithful:
3237:
3238: \begin{theorem}
3239: [Relative entropy criterion]A quantum channel with faithful fixed
3240: point $\rho_{*}$ is mixing if and only if the quantum relative entropy
3241: with respect to $\rho_{*}$ is a generalised Lyapunov function.
3242: \end{theorem}
3243: \begin{proof}
3244: Because of Theorem~\ref{thm:weaklyapov} we only need to prove the
3245: second part of the thesis, i.e. that mixing channels admit the quantum
3246: relative entropy with respect to the fixed point, $S(\rho)\equiv H(\rho,\rho_{*})$,
3247: as a generalised Lyapunov function. Firstly notice that the quantum
3248: relative entropy is monotonic under quantum channels~\cite{RUSKAI2,RUSKAI2b}.
3249: Therefore the limit $S_{*}(\rho)\equiv\lim_{n\rightarrow\infty}S\left(\tau^{n}(\rho)\right)$
3250: does exist and satisfies the condition $S_{*}(\rho)\geqslant S(\rho)$.
3251: Suppose now there exists a $\rho$ such that $S_{*}(\rho)=S(\rho)$.
3252: Because $\tau$ is mixing and $S$ is continuous we have \begin{equation}
3253: S(\rho)=S_{*}(\rho)=\lim_{n\rightarrow\infty}S\left(\tau^{n}(\rho)\right)=S(\rho_{*})=0,\end{equation}
3254: and hence $H(\rho,\rho_{*})=0$. Since $H(\rho,\sigma)=0$ if and
3255: only if $\rho=\sigma$ it follows that $S$ is a Lyapunov function
3256: around $\rho_{*}$.
3257: \end{proof}
3258: Another important investigation tool is Corollary~\ref{cor:contraction1}:
3259: weak contractivity of a quantum channel is a sufficient condition
3260: for mixing. As already mentioned in the previous section, unfortunately
3261: this not a necessary condition. Here we present an explicit counterexample
3262: based on a quantum channel introduced in Ref.~\cite{TERHAL}.
3263:
3264: \begin{example}
3265: \label{exm:mixing} Consider a three-level quantum system characterised
3266: by the orthogonal vectors $|0\rangle,|1\rangle,|2\rangle$ and the
3267: quantum channel $\tau$ defined by the transformations $\tau(|2\rangle\langle2|)=|1\rangle\langle1|$,
3268: $\tau(|1\rangle\langle1|)=\tau(|0\rangle\langle0|)=|0\rangle\langle0|$,
3269: and $\tau(|i\rangle\langle j|)=0$ for all $i\neq j$. Its easy to
3270: verify that after just two iterations any input state $\rho$ will
3271: be transformed into the vector $|0\rangle\langle0|$. Therefore the
3272: map is mixing. On the other hand it is explicitly not a weak contraction
3273: with respect to the trace norm since, for instance, one has \begin{equation}
3274: \|\;\tau(|2\rangle\langle2|)-\tau(|0\rangle\langle0|)\;\|_{1}=\|\;|1\rangle\langle1|-|0\rangle\langle0|\;\|_{1}=\|\;|2\rangle\langle2|-|0\rangle\langle0|\;\|_{1}\;,\end{equation}
3275: where in the last identity we used the invariance of $\|\cdot\|_{1}$
3276: with respect to unitary transformations.
3277: \end{example}
3278:
3279: \subsection{Beyond the density matrix operator space: spectral properties}
3280:
3281: \label{sec:newmixing}
3282:
3283: Exploiting linearity quantum channels can be extended beyond the space
3284: $\mathcal{S}(\mathcal{H})$ of density operators to become maps defined
3285: on the full vector space $\mathcal{L}(\mathcal{H})$ of the linear
3286: operators of the system, in which basic linear algebra results hold.
3287: This allows one to simplify the analysis even though the mixing property~(\ref{mixing0})
3288: is still defined with respect to the density operators of the system.
3289:
3290: Mixing conditions for quantum channels can be obtained by considering
3291: the structure of their eigenvectors in the extended space $\mathcal{L}(\mathcal{H})$.
3292: For example, it is easily shown that the spectral radius~\cite{HORNJOHNSON}
3293: of any quantum channel is equal to unity~\cite{TERHAL}, so its eigenvalues
3294: are contained in the unit circle. The eigenvalues $\lambda$ on the
3295: unit circle (i.e. $|\lambda|=1$) are referred to as \emph{peripheral
3296: eigenvalues\index{peripheral eigenvalues}.} Also, as already mentioned,
3297: since $\mathcal{S}(\mathcal{H})$ is compact and convex, CPT maps
3298: have always at least one fixed point which is a density matrix~\cite{TERHAL}.
3299:
3300: \begin{theorem}
3301: [Spectral gap criterion]\label{thm:peri}\label{thm:necessary} Let
3302: $\tau$ be a quantum channel. $\tau$ is mixing if and only if its
3303: only peripheral eigenvalue is $1$ and this eigenvalue is simple.
3304: \end{theorem}
3305: \begin{proof}
3306: The ''if'' direction of the proof is a well known result from linear
3307: algebra (see for example~\cite[Lemma 8.2.7]{HORNJOHNSON}). Now let
3308: us assume $\tau$ is mixing towards $\rho_{*}.$ Let $\Theta$ be
3309: a generic operator in $\mathcal{L}(\mathcal{H})$. Then $\Theta$
3310: can be decomposed in a finite set of non-orthogonal density operators%
3311: \footnote{To show that this is possible, consider an arbitrary operator basis
3312: of $\mathcal{L}(\mathcal{H})$. If $N$ is the finite dimension of
3313: $\mathcal{H}$ the basis will contain $N^{2}$ elements. Each element
3314: of the basis can then be decomposed into two Hermitian operators,
3315: which themselves can be written as linear combinations of at most
3316: $N$ projectors. Therefore there exists a generating set of at most
3317: $2N^{3}$ positive operators, which can be normalised such that they
3318: are quantum states. There even exists a basis (i.e. a minimal generating
3319: set) consisting of density operators, but in general it cannot be
3320: orthogonalised.%
3321: }, i.e. $\Theta=\sum_{\ell}c_{\ell}\rho_{\ell}$, with $\rho_{\ell}\in\mathcal{S}(\mathcal{H})$
3322: and $c_{\ell}$ complex. Since $\textrm{Tr}\left[\rho_{\ell}\right]=1$,
3323: we have have $\textrm{Tr}\left[\Theta\right]=\sum_{\ell}c_{\ell}$.
3324: Moreover since $\tau$ is mixing we have $\lim_{n\rightarrow\infty}\tau^{n}\left(\rho_{\ell}\right)=\rho_{*}$
3325: for all $\ell$, with convergence with respect to the trace-norm.
3326: Because of linearity this implies \begin{eqnarray}
3327: \lim_{n\rightarrow\infty}\tau^{n}\left(\Theta\right)=\sum_{\ell}c_{\ell}\;\rho_{*}=\textrm{Tr}\left[\Theta\right]\;\rho_{*}\;.\label{limit}\end{eqnarray}
3328: If there existed any other eigenvector $\Theta_{*}$ of $\tau$ with
3329: eigenvalue on the unit circle, then $\lim_{n\rightarrow\infty}\tau^{n}(\Theta_{*})$
3330: would not satisfy Eq.~(\ref{limit}).
3331: \end{proof}
3332: The speed of convergence can also be estimated by~\cite{TERHAL}
3333: \begin{eqnarray}
3334: \|\tau^{n}\left(\rho\right)-\rho_{*}\|_{1}\;\leqslant C_{N}\; n^{N}\;\kappa^{n}\;,\label{speed}\end{eqnarray}
3335: where $N$ is the dimensionality of the underlying Hilbert space,
3336: $\kappa$ is the second largest eigenvalue of $\tau$, and $C_{N}$
3337: is some constant depending only on $N$ and on the chosen norm. Hence,
3338: for $n\gg N$ the convergence becomes exponentially fast. As mentioned
3339: in~\cite{RAGINSKY}, the criterion of Theorem~\ref{thm:peri} is
3340: in general difficult to check. This is because one has to find all
3341: eigenvalues of the quantum channel, which is hard especially in the
3342: high dimensional case. Also, if one only wants to check if a particular
3343: channel is mixing or not, then the amount of information obtained
3344: is much higher than the required amount.
3345:
3346: \begin{example}
3347: As an application consider the non mixing CPT map of Example~\ref{exm:ergodic}.
3348: One can verify that apart from the eigenvalue $1$ associated with
3349: its fixed point (i.e. the completely mixed state), it possess another
3350: peripheral eigenvalue. This is $\lambda=-1$ which is associated with
3351: the Pauli operator $|0\rangle\langle0|-|1\rangle\langle1|$.
3352: \end{example}
3353: \begin{corollary}
3354: \label{cor:speed} The convergence speed of any mixing quantum channel
3355: is exponentially fast for sufficiently high values of $n$.
3356: \end{corollary}
3357: \begin{proof}
3358: From Theorem~\ref{thm:necessary} mixing channels have exactly one
3359: peripheral eigenvalue, which is also simple. Therefore the derivation
3360: of Ref.~\cite{TERHAL} applies and Eq.~(\ref{speed}) holds.
3361: \end{proof}
3362: This result should be compared with the case of strictly contractive
3363: quantum channels whose convergence was shown to be exponentially fast
3364: along to whole trajectory~\cite{RAGINSKY,STRICTCONTRATIONS}.
3365:
3366:
3367: \subsection{Ergodic channels with pure fixed points\index{pure fix-points}\label{sub:Ergodic-channels-with}}
3368:
3369: \label{sec:pure}
3370:
3371: An interesting class of ergodic quantum channel is formed by those
3372: CPT maps whose fixed point is a \emph{pure} density matrix. Among
3373: them we find for instance the maps employed in the communication protocols
3374: discussed in this thesis or those of the purification schemes of Refs.~\cite{YUASA2,YUASA1}.
3375: We will now show that within this particular class, ergodicity and
3376: mixing are indeed equivalent properties.
3377:
3378: We first need the following Lemma, which discusses a useful property
3379: of quantum channels (see also~\cite{SCHRADER}).
3380:
3381: \begin{lemma}
3382: \label{lem:OBS5} Let $\tau$ be a quantum channel and $\Theta$ be
3383: an eigenvector of $\tau$ with peripheral eigenvalue $\lambda=e^{i\varphi}$.
3384: Then, given $g=\mbox{\emph{Tr}}\left[\sqrt{\Theta^{\dag}\Theta}\right]>0$,
3385: the density matrices $\rho=\sqrt{\Theta\Theta^{\dag}}/g$ and $\sigma=\sqrt{\Theta^{\dag}\Theta}/g$
3386: are fixed points of $\tau$.
3387: \end{lemma}
3388: \begin{proof}
3389: Use the left polar decomposition to write $\Theta=g\;\rho U$ where
3390: $U$ is a unitary operator. The operator $\rho U$ is clearly an eigenvector
3391: of $\tau$ with eigenvalue $e^{i\varphi}$, i.e. \begin{eqnarray}
3392: \tau(\rho U)=\lambda\;\rho U\;.\label{uuu}\end{eqnarray}
3393: Hence introducing a Kraus set $\{ K_{n}\}_{n}$ of $\tau$~\cite{KRAUS}
3394: and the spectral decomposition of the density matrix $\rho=\sum_{j}p_{j}|\psi_{j}\rangle\langle\psi_{j}|$
3395: with $p_{j}>0$ being its positive eigenvalues, one gets \begin{eqnarray}
3396: \lambda=\mbox{Tr}[\tau(\rho U)U^{\dag}]=\sum_{j,\ell,n}p_{j}\langle\phi_{\ell}|K_{n}|\psi_{j}\rangle\langle\psi_{j}|UK_{n}^{\dag}U^{\dag}|\phi_{\ell}\rangle\;,\end{eqnarray}
3397: where the trace has been performed with respect to an orthonormal
3398: basis $\{|\phi_{\ell}\rangle\}_{\ell}$ of $\mathcal{H}$. Taking
3399: the absolute values of both terms gives \begin{eqnarray}
3400: |\lambda| & = & |\sum_{j,\ell,n}p_{j}\langle\phi_{\ell}|K_{n}|\psi_{j}\rangle\langle\psi_{j}|UK_{n}^{\dag}U^{\dag}|\phi_{\ell}\rangle|\nonumber \\
3401: & \leqslant & \sqrt{\sum_{j,\ell,n}p_{j}\langle\phi_{\ell}|K_{n}|\psi_{j}\rangle\langle\psi_{j}|K_{n}^{\dag}|\phi_{\ell}\rangle}\sqrt{\sum_{j,\ell,n}p_{j}\langle\phi_{\ell}|UK_{n}U^{\dag}|\psi_{j}\rangle\langle\psi_{j}|UK_{n}^{\dag}U^{\dag}|\phi_{\ell}\rangle}\nonumber \\
3402: & = & \sqrt{\mbox{Tr}[\tau(\rho)}]\sqrt{\mbox{Tr}[\tilde{\tau}(\rho)]}=1,\end{eqnarray}
3403: where the inequality follows from the Cauchy-Schwartz inequality.
3404: The last identity instead is a consequence of the fact that the transformation
3405: $\tilde{\tau}(\rho)=U\tau(U^{\dag}\rho U)U^{\dag}$ is CPT and thus
3406: trace preserving. Since $|\lambda|=1$ it follows that the inequality
3407: must be replaced by an identity. This happens if and only if there
3408: exist $e^{i\vartheta}$ such that \begin{eqnarray}
3409: \sqrt{p_{j}}\{\langle\phi_{\ell}|K_{n}|\psi_{j}\rangle\}^{*}=\sqrt{p_{j}}\langle\psi_{j}|K_{n}^{\dag}|\phi_{\ell}\rangle=e^{i\vartheta}\sqrt{p_{j}}\langle\psi_{j}|UK_{n}^{\dag}U^{\dag}|\phi_{\ell}\rangle\;,\end{eqnarray}
3410: for all $j,\ell$ and $n$. Since the $|\phi_{\ell}\rangle$ form
3411: a basis of $\mathcal{H}$, and $p_{j}>0$ this implies \begin{eqnarray}
3412: \langle\psi_{j}|K_{n}^{\dag}=e^{i\vartheta}\;\langle\psi_{j}|UK_{n}^{\dag}U^{\dag}\quad\Rightarrow\quad\langle\psi_{j}|UK_{n}^{\dag}=e^{-i\vartheta}\;\langle\psi_{j}|K_{n}^{\dag}U\;,\end{eqnarray}
3413: for all $n$ and for all the not null eigenvectors $|\psi_{j}\rangle$
3414: of $\rho$. This yields \begin{eqnarray}
3415: \tau(\rho U) & = & \sum_{j}p_{j}\sum_{n}K_{n}|\psi_{j}\rangle\langle\psi_{j}|UK_{n}^{\dag}=e^{-i\vartheta}\;\sum_{j}p_{j}\sum_{n}K_{n}|\psi_{j}\rangle\langle\psi_{j}|K_{n}^{\dag}U\nonumber \\
3416: & = & e^{-i\vartheta}\;\tau(\rho)U\end{eqnarray}
3417: which, replaced in (\ref{uuu}) gives $e^{-i\vartheta}\;\tau(\rho)=e^{i\varphi}\;\rho$,
3418: whose only solution is $e^{-i\vartheta}=e^{i\varphi}$. Therefore
3419: $\tau(\rho)=\rho$ and $\rho$ is a fixed point of $\tau$. The proof
3420: for $\sigma$ goes along similar lines: simply consider the right
3421: polar decomposition of $\Theta$ instead of the left polar decomposition.
3422: \end{proof}
3423: \begin{corollary}
3424: Let $\tau$ be an ergodic quantum channel. It follows that its eigenvectors
3425: associated with peripheral eigenvalues are normal operators.
3426: \end{corollary}
3427: \begin{proof}
3428: Let $\Theta$ be an eigenoperator with peripheral eigenvalue $e^{i\varphi}$
3429: such that $\tau\left(\Theta\right)=e^{i\varphi}\;\Theta$. By Lemma
3430: \ref{lem:OBS5} we know that, given $g=\mbox{Tr}\left[\sqrt{\Theta^{\dag}\Theta}\right]$
3431: the density matrices $\rho=\sqrt{\Theta\Theta^{\dag}}/g$ and $\sigma=\sqrt{\Theta^{\dag}\Theta}/g$
3432: must be fixed points of $\tau$. Since the map is ergodic we must
3433: have $\rho=\sigma$, i.e. $\Theta\Theta^{\dag}=\Theta^{\dag}\Theta$.
3434: \end{proof}
3435: \begin{framedtheorem}
3436: [Purely ergodic maps]\label{thm:ergodicmixing}Let $|\psi_{1}\rangle\langle\psi_{1}|$
3437: be the pure fixed point of an ergodic quantum channel $\tau$. It
3438: follows that $\tau$ is mixing.
3439: \end{framedtheorem}
3440: \begin{proof}
3441: We will use the spectral gap criterion showing that $|\psi_{1}\rangle\langle\psi_{1}|$
3442: is the only peripheral eigenvector of $\tau$. Assume in fact that
3443: $\Theta\in L(\mathcal{H})$ is a eigenvector of $\tau$ with peripheral
3444: eigenvalue, i.e. \begin{eqnarray}
3445: \tau\left(\Theta\right)=e^{i\varphi}\Theta\;.\label{identity1}\end{eqnarray}
3446: From Lemma~\ref{lem:OBS5} we know that the density matrix \begin{equation}
3447: \rho=\sqrt{\Theta\Theta^{\dag}}/g,\end{equation}
3448: with $g=\mbox{Tr}\left[\sqrt{\Theta^{\dag}\Theta}\right]>0$, must
3449: be a fixed point of $\tau$. Since this is an ergodic map we must
3450: have $\rho=|\psi_{1}\rangle\langle\psi_{1}|$. This implies $\Theta=g|\psi_{1}\rangle\langle\psi_{2}|$,
3451: with $|\psi_{2}\rangle$ some normalised vector of $\mathcal{H}$.
3452: Replacing it into Eq.~(\ref{identity1}) and dividing both terms
3453: by $g$ yields $\tau\left(|\psi_{1}\rangle\langle\psi_{2}|\right)=e^{i\varphi}|\psi_{1}\rangle\langle\psi_{2}|$
3454: and \begin{eqnarray}
3455: |\langle\psi_{1}|\tau(|\psi_{1}\rangle\langle\psi_{2}|)|\psi_{2}\rangle|=1\;.\end{eqnarray}
3456: Introducing a Kraus set $\{ K_{n}\}_{n}$ of $\tau$ and employing
3457: Cauchy-Schwartz inequality one can then write \begin{eqnarray}
3458: 1 & = & |\langle\psi_{1}|\tau(|\psi_{1}\rangle\langle\psi_{2}|)|\psi_{2}\rangle|=|\sum_{n}\langle\psi_{1}|K_{n}|\psi_{1}\rangle\langle\psi_{2}|K_{n}^{\dag}|\psi_{2}\rangle|\\
3459: & \leqslant & \sqrt{\sum_{n}\langle\psi_{1}|K_{n}|\psi_{1}\rangle\langle\psi_{1}|K_{n}^{\dag}|\psi_{1}\rangle}\sqrt{\sum_{n}\langle\psi_{2}|K_{n}|\psi_{2}\rangle\langle\psi_{2}|K_{n}^{\dag}|\psi_{2}\rangle}\nonumber \\
3460: & = & \sqrt{\langle\psi_{1}|\tau(|\psi_{1}\rangle\langle\psi_{1}|)|\psi_{1}\rangle}\sqrt{\langle\psi_{2}|\tau(|\psi_{2}\rangle\langle\psi_{2}|)|\psi_{2}\rangle}=\sqrt{\langle\psi_{2}|\tau(|\psi_{2}\rangle\langle\psi_{2}|)|\psi_{2}\rangle}\;,\nonumber \end{eqnarray}
3461: where we used the fact that $|\psi_{1}\rangle$ is the fixed point
3462: of $\tau$. Since $\tau$ is CPT the quantity $\langle\psi_{2}|\tau(|\psi_{2}\rangle\langle\psi_{2}|)|\psi_{2}\rangle$
3463: is upper bounded by $1$. Therefore in the above expression the inequality
3464: must be replaced by an identity, i.e. \begin{eqnarray}
3465: \langle\psi_{2}|\tau(|\psi_{2}\rangle\langle\psi_{2}|)|\psi_{2}\rangle=1\qquad\Longleftrightarrow\qquad\tau(|\psi_{2}\rangle\langle\psi_{2}|)=|\psi_{2}\rangle\langle\psi_{2}|\;.\end{eqnarray}
3466: Since $\tau$ is ergodic, we must have $|\psi_{2}\rangle\langle\psi_{2}|=|\psi_{1}\rangle\langle\psi_{1}|$.
3467: Therefore $\Theta\propto|\psi_{1}\rangle\langle\psi_{1}|$ which shows
3468: that $|\psi_{1}\rangle\langle\psi_{1}|$ is the only eigenvector of
3469: $\tau$ with peripheral eigenvalue of.
3470: \end{proof}
3471: An application of the previous Theorem is obtained as follows.
3472:
3473: \begin{lemma}
3474: \label{lem:add} Let $M_{AB}=M_{A}\otimes1_{B}+1_{A}\otimes M_{B}$
3475: be an observable of the composite system $\mathcal{H}_{A}\otimes\mathcal{H}_{B}$
3476: and $\tau$ the CPT linear map on $\mathcal{H}_{A}$ of Stinespring
3477: form~\emph{\cite{STINE}} \begin{equation}
3478: \tau(\rho)=\emph{\mbox{Tr}}_{B}\left[U\left(\rho\otimes|\phi\rangle_{B}\langle\phi|\right)U^{\dag}\right]\;,\label{eq:rep}\end{equation}
3479: (here $\emph{\mbox{Tr}}_{X}\left[\cdots\right]$ is the partial trace
3480: over the system $X$, and $U$ is a unitary operator of $\mathcal{H}_{A}\otimes\mathcal{H}_{B}$).
3481: Assume that $\left[M_{AB},U\right]=0$ and that $|\phi\rangle_{B}$
3482: is the eigenvector corresponding to a non-degenerate maximal or minimal
3483: eigenvalue of $M_{B}.$ Then $\tau$ is mixing if and only if $U$
3484: has one and only one eigenstate that factorises as $|\nu\rangle_{A}\otimes|\phi\rangle_{B}.$
3485: \end{lemma}
3486: \begin{proof}
3487: Let $\rho$ be an arbitrary fixed point of $\tau$ (since $\tau$
3488: is CPT it has always at least one), i.e. $\textrm{Tr}_{B}\left[U\left(\rho\otimes|\phi\rangle_{B}\langle\phi|\right)U^{\dag}\right]=\rho$.
3489: Since $M_{AB}$ is conserved and $\textrm{Tr}_{A}\left[M_{A}\rho\right]=\textrm{Tr}_{A}\left[M_{A}\tau(\rho)\right]$,
3490: the system $B$ must remain in the maximal state, which we have assumed
3491: to be unique and pure, i.e. \begin{equation}
3492: U\left(\rho\otimes|\phi\rangle_{B}\langle\phi|\right)U^{\dag}=\rho\otimes|\phi\rangle_{B}\langle\phi|\qquad\Longrightarrow\qquad\left[U,\rho\otimes|\phi\rangle_{B}\langle\phi|\right]=0\;.\end{equation}
3493: Thus there exists a orthonormal basis $\left\{ |u_{k}\rangle\right\} _{k}$
3494: of $\mathcal{H}_{A}\otimes\mathcal{H}_{B}$ diagonalising simultaneously
3495: both $U$ and $\rho\otimes|\phi\rangle_{B}\langle\phi|$. We express
3496: the latter in this basis, i.e. $\rho\otimes|\phi\rangle_{B}\langle\phi|=\sum_{k}p_{k}|u_{k}\rangle\langle u_{k}|$
3497: with $p_{k}>0$, and compute the von Neumann entropy of subsystem
3498: $B$. This yields \begin{eqnarray}
3499: 0 & = & H(|\phi\rangle_{B}\langle\phi|)=H\left(\textrm{Tr}_{A}\left[\sum_{k}p_{k}|u_{k}\rangle\langle u_{k}|\right]\right)\geqslant\sum_{k}p_{k}\; H\left(\textrm{Tr}_{A}\left[|u_{k}\rangle\langle u_{k}|\right]\right)\;.\end{eqnarray}
3500: From the convexity of the von Neumann entropy the above inequality
3501: leads to a contradiction unless $\textrm{Tr}_{A}\left[|u_{k}\rangle\langle u_{k}|\right]=|\phi\rangle_{B}\langle\phi|$
3502: for all $k$. The $|u_{k}\rangle$ must therefore be factorising,
3503: \begin{equation}
3504: |u_{k}\rangle=|\nu_{k}\rangle_{A}\otimes|\phi\rangle_{B}.\label{eq:fact}\end{equation}
3505: If the factorising eigenstate of $U$ is unique, it must follow that
3506: $\rho=|\nu\rangle\langle\nu|$ for some $|\nu\rangle$ and that $\tau$
3507: is ergodic. By Theorem \ref{thm:ergodicmixing} it then follows that
3508: $\tau$ is also mixing. If on the other hand there exists more than
3509: one factorising eigenstate, than all states of the form of Eq. (\ref{eq:fact})
3510: correspond to a fixed point $\rho_{k}=|\nu_{k}\rangle\langle\nu_{k}|$
3511: and $\tau$ is neither ergodic nor mixing.
3512: \end{proof}
3513: \begin{remark}
3514: An application of this Lemma is the protocol for read and write access
3515: by local control discussed in the next chapter.
3516: \end{remark}
3517:
3518: \section{Conclusion}
3519:
3520: \label{CONCLUSION} In reviewing some known results on the mixing
3521: property of continuous maps, we obtained a stronger version of the
3522: direct Lyapunov method. For compact metric spaces (including quantum
3523: channels operating over density matrices) it provides a necessary
3524: and sufficient condition for mixing. Moreover it allows us to prove
3525: that asymptotic deformations with at least one fixed point must be
3526: mixing.
3527:
3528: In the specific context of quantum channels we employed the generalised
3529: Lyapunov method to analyse the mixing properties. Here we also analysed
3530: different mixing criteria. In particular we have shown that an ergodic
3531: quantum channel with a pure fixed point is also mixing.
3532:
3533:
3534: \chapter{Read and write access by local control\label{cha:Full-read-and}}
3535:
3536:
3537: \section{Introduction}
3538:
3539: The unitarity of Quantum Mechanics implies that information is conserved.
3540: Whatever happens to a quantum system - as long as it is unitary, the
3541: original state can in principle be recovered by applying the inverse
3542: unitary transformation. However it is well known that in open quantum
3543: systems~\cite{OPENQUANTUM} the reduced dynamics is no longer unitary.
3544: The reduced dynamics is described by a completely positive, trace
3545: preserving maps, and we have seen in the last chapter that there are
3546: extreme examples, namely \emph{mixing} maps, where all information
3547: about the initial state is eventually lost. Where has it gone? If
3548: the whole system evolves unitary, then this information must have
3549: been transferred in the \emph{correlations} between reduced system
3550: and environment~\cite{Hayden2004}, and/or in the environment. We
3551: can see that this may be useful for quantum state transfer, in particular
3552: the case where all information is transferred into the ''environment'',
3553: which could be another quantum system (the receiver). A particularly
3554: useful case is given by mixing maps with pure convergence points,
3555: because a pure state cannot be correlated, and because we have a simple
3556: convergence criterion in this case (Subsection~\ref{sub:Ergodic-channels-with}).
3557: This is an example of \emph{homogenisation\index{homogenisation}}~\cite{HOMOGENIZATION1,HOMOGENIZATION2}\emph{.}
3558: Furthermore, if the mixing property arises from some operations, we
3559: can expect that by applying the inverse operations, information can
3560: also be transferred back to the system. This property was used in~\cite{WELLENS,Wellens2002}
3561: to generate arbitrary states of a cavity field by sending atoms through
3562: the cavity. The crucial difference is that in our system control is
3563: only assumed to be available on a subsystem (such as, for example,
3564: the ends of a quantum chain). Hence we will show in this chapter how
3565: arbitrary quantum states can be written to (i.e. prepared on) a large
3566: system, and read from it, by \emph{local} control only. This is similar
3567: in spirit to universal quantum interfaces~\cite{Lloyd2004}, but
3568: our different approach allows us to specify explicit protocols and
3569: to give lower bounds for fidelities. We also demonstrate how this
3570: can be used to significantly improve the quantum communication between
3571: two parties if the receiver is allowed to store the received signals
3572: in a quantum memory before decoding them. In the limit of an infinite
3573: memory, the transfer is perfect. We prove that this scheme allows
3574: the transfer of arbitrary multi-partite states along Heisenberg chains
3575: of spin-$1/2$ particles with random coupling strengths.
3576:
3577: Even though the convergence of a mixing map is essentially exponentially
3578: fast (Corollary~\ref{cor:speed}), we still have to deal with infinite
3579: limits. Looking at the environment this in turn would require to study
3580: states on an infinite dimensional Hilbert space, and unfortunately
3581: this can introduce many mathematical difficulties. We are mainly interested
3582: in bounds for the finite case: if the protocol stops after finitely
3583: many steps, what is the fidelity of the reading/writing? Which encoding
3584: and decoding operations must be applied? By stressing on these questions,
3585: we can actually avoid the infinite dimensional case, but the price
3586: we have to pay is that our considerations become a bit technically
3587: involved.
3588:
3589:
3590: \section{Protocol\label{sec:Protocol}}
3591:
3592: We consider a tripartite finite dimensional Hilbert space given by
3593: $\mathcal{H}=\mathcal{H}_{C}\otimes\mathcal{H}_{\bar{C}}\otimes\mathcal{H}_{M}.$
3594: We assume that full control (the ability to prepare states and apply
3595: unitary transformations) is possible on system $C$ and $M,$ but
3596: no control is available on system $\bar{C}.$ However, we assume that
3597: $C$ and $\bar{C}$ are coupled by some time-independent Hamiltonian
3598: $H.$ We show here that under certain assumptions, if the system $C\bar{C}$
3599: is initialised in some arbitrary state we can transfer (''read'')
3600: this state into the system $M$ by applying some operations between
3601: $M$ and $C.$ Likewise, by initialising the system $M$ in the correct
3602: state, we can prepare (''write'') arbitrary states on the system
3603: $C\bar{C.}$ The system $M$ functions as a \emph{quantum memory}\index{quantum memory}
3604: and must be at least as large as the system $C\bar{C}.$ As sketched
3605: in Fig.~\ref{fig:memory} we can imagine it to be split into sectors
3606: $M_{\ell},$ I.e.. \begin{equation}
3607: \mathcal{H}_{M}=\bigotimes_{\ell=1}^{L}\mathcal{H}_{M_{\ell}}\end{equation}
3608: with\begin{equation}
3609: \textrm{dim}\mathcal{H}_{M_{\ell}}=\textrm{dim}\mathcal{H}_{C}.\end{equation}
3610: For the reading case, we assume that the memory is initialised in
3611: the state \begin{equation}
3612: |0\rangle_{M}\equiv\bigotimes_{\ell}|0\rangle_{M_{\ell}}\end{equation}
3613: where $|0\rangle$ can stand for some generic state%
3614: \footnote{Later on we will give an example where $|0\rangle$ represents a multi-qubit
3615: state with all qubits aligned, but here we don't need to assume this.%
3616: }. Like in the multi rail protocols considered in Chapter~\ref{cha:Multi-rail-and-Capacity},
3617: we let the system evolve for a while, perform an operation, let it
3618: evolve again and so forth, only that now the operation is not a measurement,
3619: but a \emph{unitary gate}. More specifically, at step $\ell$ of the
3620: protocol we perform a unitary swap $S_{\ell}$ between system $C$
3621: and systems $M_{\ell}.$ After the $L$th swap operation the protocol
3622: stops. The protocol for reading is thus represented by the unitary
3623: operator \begin{equation}
3624: W\equiv S_{L}US_{L-1}U\cdots S_{\ell}U\cdots S_{1}U,\end{equation}
3625: where $U\in\mathcal{L}(\mathcal{H}_{C\bar{C}})$ is the time-evolution
3626: operator $U=\exp\left\{ -iHt\right\} $ for some fixed time interval
3627: $t.$ As we will see in the next section, the reduced evolution of
3628: the system $\bar{C}$ under the protocol can be expressed in terms
3629: of the CPT map \begin{equation}
3630: \tau(\rho_{\bar{C}})\equiv\textrm{tr}_{C}\left[U\left(\rho_{\bar{C}}\otimes|0\rangle_{C}\langle0|\right)U^{\dag}\right],\end{equation}
3631: where $|0\rangle_{C}$ is the state that is swapped in from the memory.
3632: Our main assumption now is that $\tau$ is ergodic with a pure fixed
3633: point (which we denote as $|0\rangle_{\bar{C}}$). By Theorem~\ref{thm:ergodicmixing}
3634: this implies that $\tau$ is mixing, and therefore asymptotically
3635: all information is transferred into the memory. %
3636: \begin{figure}[t]
3637: \begin{centering}\includegraphics[width=0.8\columnwidth]{memory}\par\end{centering}
3638:
3639:
3640: \caption{\label{fig:memory}The system $C\bar{C}$ can only be controlled
3641: by acting on a (small) subsystem $C.$ However system $C$ is coupled
3642: to system $\bar{C}$ by a unitary operator $U=\exp\left\{ -iHt\right\} .$
3643: This coupling can - in some cases - \emph{mediate} the local control
3644: on $C$ to the full system $C\bar{C}.$ In our case, system $C$ is
3645: controlled by performing regular swap operations $S_{\ell}$ between
3646: it and a quantum memory $M_{\ell}.$ }
3647: \end{figure}
3648:
3649:
3650: For writing states on the system, we just make use of the unitarity
3651: of $W.$ Roughly speaking, we initialise the memory in the state that
3652: it \emph{would have ended up in} after applying $W$ if system $C\bar{C}$
3653: had started in the state we want to initialise. Then we apply the
3654: \emph{inverse} of $W$ given by\begin{equation}
3655: W^{\dag}=U^{\dag}S_{1}\cdots U^{\dag}S_{\ell}\cdots U^{\dag}S_{L-1}U^{\dag}S_{L}.\end{equation}
3656: We will see in Section~\ref{sec:Coding-transformation} how this
3657: gives rise to a unitary coding transformation on the memory system,
3658: such that arbitrary and unknown states can be initialised on the system.
3659: The reader has probably noticed that the inverse of $W$ is generally
3660: unphysical in the sense that it requires backward time evolution,
3661: i.e. one has to wait \emph{negative} time steps between the swaps.
3662: But we will see later how this can be fixed by a simple transformation.
3663: For the moment, we just assume that $W^{\dag}$ is physical.
3664:
3665:
3666: \section{Decomposition equations}
3667:
3668: In this section we give a decomposition of the state after applying
3669: the protocol which will allow us to estimate the fidelities for state
3670: transfer in terms of the mixing properties of the map $\tau.$ Let
3671: $|\psi\rangle_{C\bar{C}}\in\mathcal{H}_{C\bar{C}}$ be an arbitrary
3672: state. We notice that the $C$ component of $W|\psi\rangle_{C\bar{C}}|0\rangle_{M}$
3673: is always $|0\rangle_{C}$. Therefore we can decompose it as follows
3674: \begin{equation}
3675: W|\psi\rangle_{C\bar{C}}|0\rangle_{M}=|0\rangle_{C}\otimes\left[\sqrt{\eta}|0\rangle_{\bar{C}}|\phi\rangle_{M}+\sqrt{1-\eta}|\Delta\rangle_{\bar{C}M}\right]\label{eq:main}\end{equation}
3676: with $|\Delta\rangle_{\bar{C}M}$ being a normalised vector of $\bar{C}$
3677: and $M$ which satisfies the identity \begin{eqnarray}
3678: _{\bar{C}}\langle0|\Delta\rangle_{\bar{C}M}=0\;.\label{impo}\end{eqnarray}
3679: It is worth stressing that in the above expression $\eta$, $|\phi\rangle_{M}$
3680: and $|\Delta\rangle_{\bar{C}M}$ are depending on $|\psi\rangle_{C\bar{C}}$.
3681: We decompose $W^{\dag}$ acting on the first term of Eq.~(\ref{eq:main})
3682: as
3683:
3684: \begin{equation}
3685: W^{\dag}|0\rangle_{C\bar{C}}|\phi\rangle_{M}=\sqrt{\tilde{\eta}}\;|\psi\rangle_{C\bar{C}}|0\rangle_{M}+\sqrt{1-\tilde{\eta}}\;|\tilde{\Delta}\rangle_{C\bar{C}M},\label{eq:main2}\end{equation}
3686: where $|\tilde{\Delta}\rangle_{C\bar{C}M}$ is the orthogonal complement
3687: of $|\psi\rangle_{C\bar{C}}|0\rangle_{M},$ i.e. \begin{equation}
3688: _{\bar{C}C}\langle\psi|{}_{M}\langle0|\tilde{\Delta}\rangle_{C\bar{C}M}=0\;.\label{eq:o2}\end{equation}
3689: Multiplying Eq. (\ref{eq:main2}) from the left with $_{C\bar{C}}\langle\psi|_{M}\langle0|$
3690: and using the conjugate of Eq. (\ref{eq:main}) we find that $\eta=\tilde{\eta}.$
3691: An expression of $\eta$ in terms of $\tau$ can be obtained by noticing
3692: that for any vector $|\psi\rangle_{\bar{C}C}$ the following identity
3693: applies \begin{equation}
3694: \tau(\rho_{\bar{C}})=\textrm{tr}_{C}\left[U\left(\rho_{\bar{C}}\otimes|0\rangle_{C}\langle0|\right)U^{\dag}\right]=\textrm{tr}_{CM}\left[US_{\ell}\left(|\psi\rangle_{\bar{C}C}\langle\psi|\otimes|0\rangle_{M}\langle0|\right)S_{\ell}U^{\dag}\right]\;,\end{equation}
3695: with $\rho_{\bar{C}}$ being the reduced density matrix $\textrm{tr}_{C}\left[|\psi\rangle_{\bar{C}C}\langle\psi|\right]$.
3696: Reiterating this expression one gets \begin{equation}
3697: \textrm{tr}_{CM}\left[W(|\psi\rangle_{C\bar{C}}\langle\psi|\otimes|0\rangle_{M}\langle0|)W^{\dag}\right]=\tau^{L-1}\left(\rho_{\bar{C}}^{\prime}\right)\end{equation}
3698: with $\rho_{\bar{C}}^{\prime}=\textrm{tr}_{C}\left[U\left(|\psi\rangle_{\bar{C}C}\langle\psi|\right)U^{\dag}\right]$.
3699: Therefore from Eq.~(\ref{eq:main}) and the orthogonality relation
3700: (\ref{impo}) it follows that \begin{equation}
3701: \eta={}_{\bar{C}}\langle0|\tau^{L-1}\left(\rho_{\bar{C}}^{\prime}\right)|0\rangle_{\bar{C}},\label{eq:fid}\end{equation}
3702: which, since $\tau$ is mixing, shows that $\eta\rightarrow1$ for
3703: $L\rightarrow\infty$. Moreover we can use Eq.~(\ref{speed}) to
3704: claim that \begin{eqnarray}
3705: |\eta-1| & = & |{}_{\bar{C}}\langle0|\tau^{L-1}\left(\rho_{\bar{C}}^{\prime}\right)|0\rangle_{\bar{C}}-1|\nonumber \\
3706: & \leq & \|\tau^{L-1}\left(\rho_{\bar{C}}^{\prime}\right)-|0\rangle_{\bar{C}}\langle0|\|_{1}\leq R\;(L-1)^{d_{\bar{C}}}\;\kappa^{L-1},\label{eq:fid11}\end{eqnarray}
3707: where $R$ is a constant which depends upon $d_{\bar{C}}\equiv\mbox{dim}\mathcal{H}_{\bar{C}}$
3708: and where $\kappa\in]0,1[$ is the second largest eigenvalue of $\tau.$
3709:
3710:
3711: \section{Coding transformation\index{coding transformation}\label{sec:Coding-transformation}}
3712:
3713: Here we derive the decoding/encoding transformation that relates states
3714: on the memory $M$ to the states that are on the system $C\bar{C}.$
3715: We first apply the above decompositions Eqs.~(\ref{eq:main}) and
3716: (\ref{eq:main2}) to a fixed orthonormal basis $\left\{ |\psi_{k}\rangle_{C\bar{C}}\right\} $
3717: of $\mathcal{H}_{C\bar{C}},$ i.e. \begin{eqnarray}
3718: W|\psi_{k}\rangle_{C\bar{C}}|0\rangle_{M} & = & |0\rangle_{C}\otimes\left[\sqrt{\eta_{k}}|0\rangle_{\bar{C}}|\phi_{k}\rangle_{M}+\sqrt{1-\eta_{k}}|\Delta_{k}\rangle_{\bar{C}M}\right]\nonumber \\
3719: W^{\dag}|0\rangle_{C\bar{C}}|\phi_{k}\rangle_{M} & = & \sqrt{\eta_{k}}\;|\psi_{k}\rangle_{C\bar{C}}|0\rangle_{M}+\sqrt{1-\eta_{k}}\;|\tilde{\Delta_{k}}\rangle_{C\bar{C}M}.\label{eq:maink}\end{eqnarray}
3720: Define a linear operator $D$ on ${\cal H}_{M}$ which performs the
3721: following transformation \begin{eqnarray}
3722: D|\psi_{k}\rangle_{M}=|\phi_{k}\rangle_{M}.\label{DEFD}\end{eqnarray}
3723: Here $|\psi_{k}\rangle_{M}$ are orthonormal vectors of $M$ which
3724: represent the states $\left\{ |\psi_{k}\rangle_{C\bar{C}}\right\} $
3725: of $\mathcal{H}_{C\bar{C}}$ (formally they are obtained by a partial
3726: isometry from $\bar{C}C$ to $M$). The vectors $|\phi_{k}\rangle_{M}$
3727: are defined through Eq.~(\ref{eq:maink}) - typically they will not
3728: be orthogonal. We first show that for large $L$ they become approximately
3729: orthogonal.
3730:
3731: From the unitarity of $W^{\dag}$ and from Eq. (\ref{eq:maink}) we
3732: can establish the following identity \begin{eqnarray}
3733: & & _{M}\langle\phi_{k}|\phi_{k'}\rangle_{M}=\sqrt{\eta_{k}\;\eta_{k'}}\;\delta_{kk'}+\sqrt{\eta_{k}\;(1-\eta_{k'})}\;{}_{\bar{C}CM}\langle\psi_{k}0|\tilde{\Delta}_{k'}\rangle_{\bar{C}CM}\label{QQQ}\\
3734: & & +\sqrt{\eta_{k'}\;(1-\eta_{k})}\;{}_{\bar{C}CM}\langle\tilde{\Delta}_{k}|\psi_{k'}0\rangle_{\bar{C}CM}+\sqrt{(1-\tilde{\eta}_{k})(1-\tilde{\eta}_{k'})}\;{}_{C\bar{C}M}\langle\tilde{\Delta}_{k}|\tilde{\Delta}_{k'}\rangle_{C\bar{C}M}\;.\nonumber \end{eqnarray}
3735: Defining $\eta_{0}\equiv\min_{k}\eta_{k}$ it follows for $k\neq k'$
3736: that \begin{eqnarray}
3737: |_{M}\langle\phi_{k}|\phi_{k'}\rangle_{M}| & \le & \sqrt{\eta_{k}\;(1-\eta_{k'})}\;|{}_{\bar{C}CM}\langle\psi_{k}0|\tilde{\Delta}_{k'}\rangle_{\bar{C}CM}|\\
3738: & & +\sqrt{\eta_{k'}\;(1-\eta_{k})}\;|{}_{\bar{C}CM}\langle\tilde{\Delta}_{k}|\psi_{k'}0\rangle_{\bar{C}CM}|\nonumber \\
3739: & & +\sqrt{(1-\tilde{\eta}_{k})(1-\tilde{\eta}_{k'})}\;|{}_{C\bar{C}M}\langle\tilde{\Delta}_{k}|\tilde{\Delta}_{k'}\rangle_{C\bar{C}M}|\nonumber \\
3740: & \leq & 2\sqrt{1-\eta_{0}}+(1-\eta_{0})\;\leq\;3\sqrt{1-\eta_{0}}.\label{IMPRT}\end{eqnarray}
3741: Therefore for all $k,k'$ the inequality \begin{eqnarray}
3742: |{}_{M}\langle\phi_{k}|\phi_{k'}\rangle_{M}-\delta_{k,k'}|\le3\;\sqrt{1-\eta_{0}}\label{IMPO}\end{eqnarray}
3743: holds. It is worth noticing that, since Eq.~(\ref{eq:fid11}) applies
3744: for all input states $|\psi\rangle_{\bar{C}C}$, we have \begin{eqnarray}
3745: |\eta_{0}-1|\leq C\;(L-1)^{d_{\bar{C}}}\;\kappa^{L-1}\;.\label{eq:fid111}\end{eqnarray}
3746: Eq.~(\ref{IMPO}) allows us to make an estimation of the eigenvalues
3747: $\lambda_{k}$ of $D^{\dag}D$ as \begin{equation}
3748: |\lambda_{k}-1|\leq3\; d_{C\bar{C}}\;\sqrt{1-\eta_{0}},\end{equation}
3749: with $d_{C\bar{C}}\equiv\dim\mathcal{H}_{C\bar{C}}.$ We now take
3750: a polar decomposition $D=PV$ of $D.$ $V$ is the \emph{best unitary
3751: approximation} to $D$ \cite[p 432]{HORNJOHNSON} and we have \begin{eqnarray}
3752: ||D-V||_{2}^{2} & = & \sum_{k}\left[\sqrt{\lambda_{k}}-1\right]^{2}\nonumber \\
3753: & \le & \sum_{k}\left|\lambda_{k}-1\right|\nonumber \\
3754: & \le & 3\; d_{C\bar{C}}^{2}\;\sqrt{1-\eta_{0}}.\label{Q0}\end{eqnarray}
3755: Therefore\begin{equation}
3756: \boxed{||D-V||_{2}\le\sqrt{3}\: d_{C\bar{C}}\:(1-\eta_{0})^{1/4},}\label{Q}\end{equation}
3757: which, thanks to Eq.~(\ref{eq:fid111}), shows that $D$ can be
3758: approximated arbitrary well by a unitary operator $V$ for $L\rightarrow\infty$.
3759:
3760:
3761: \section{Fidelities for reading and writing}
3762:
3763: In what follows we will use $V^{\dag}$ and $V$ as our reading and
3764: writing transformation, respectively. In particular, $V^{\dag}$ will
3765: be used to recover the input state $|\psi\rangle_{C\bar{C}}$ of the
3766: chain after we have (partially) transferred it into $M$ through the
3767: unitary $W$ (i.e. we first act on $|\psi\rangle_{C\bar{C}}\otimes|0\rangle_{M}$
3768: with $W$, and then we apply $V^{\dag}$ on $M$). Vice-versa, in
3769: order to prepare a state $|\psi\rangle_{C\bar{C}}$ on $C\bar{C}$
3770: we first prepare $M$ into $|\psi\rangle_{M}$, then we apply to it
3771: the unitary transformation $V$ and finally we apply $W^{\dag}$.
3772: We now give bounds on the fidelities for both procedures.
3773:
3774: The fidelity for reading the state $|\psi\rangle_{M}$ is given by
3775: \begin{eqnarray}
3776: F_{r}(\psi)\;\equiv\;{}_{M}\langle\psi|V^{\dag}\; R_{M}\; V|\psi\rangle_{M}\end{eqnarray}
3777: where $R_{M}$ is the state of the memory after $W$, i.e. \begin{eqnarray}
3778: R_{M}\equiv\textrm{tr}_{C\bar{C}}\left[W(|\psi\rangle_{C\bar{C}}\langle\psi|\otimes|0\rangle_{M}\langle0|)W^{\dag}\right]=\eta\;|\phi\rangle_{M}\langle\phi|+(1-\eta)\;\sigma_{M}\;.\end{eqnarray}
3779: In the above expression we used Eqs.~(\ref{eq:main}) and~(\ref{impo})
3780: and defined $\sigma_{M}=\textrm{tr}_{\bar{C}}[|\Delta\rangle_{\bar{C}M}\langle\Delta|]$.
3781: Therefore by linearity we get \begin{eqnarray}
3782: F_{r}(\psi)=\eta\;|{}_{M}\langle\phi|V|\psi\rangle_{M}|^{2}+(1-\eta)\;{}_{M}\langle\psi|V^{\dag}\;\sigma_{M}\; V|\psi\rangle_{M}\geq\eta\;|{}_{M}\langle\phi|V|\psi\rangle_{M}|^{2}\;.\label{fin1}\end{eqnarray}
3783: Notice that \begin{eqnarray}
3784: |_{M}\langle\phi|V|\psi\rangle_{M}|=|_{M}\langle\phi|V-D+D|\psi\rangle_{M}|\geq|_{M}\langle\phi|D|\psi\rangle_{M}|-|_{M}\langle\phi|D-V|\psi\rangle_{M}|\;.\label{fin2}\end{eqnarray}
3785: Now we use the inequality~(\ref{Q}) to write \begin{eqnarray}
3786: |_{M}\langle\phi|D-V|\psi\rangle_{M}|\leq||D-V||_{2}\leq\sqrt{3}\; d_{C\bar{C}}\;(1-\eta_{0})^{1/4}\;.\end{eqnarray}
3787: If $|\psi\rangle_{M}$ was a basis state $|\psi_{k}\rangle_{M},$
3788: then $|_{M}\langle\phi|D|\psi\rangle_{M}|=1$ by the definition Eq.~(\ref{DEFD})
3789: of $D$. For \emph{generic} $|\psi\rangle_{M}$ we can use the linearity
3790: to find after some algebra that \begin{eqnarray}
3791: \sqrt{\eta}\;|_{M}\langle\phi|D|\psi\rangle_{M}|\;\geq\sqrt{\eta_{0}}\;-\;3\; d_{C\bar{C}}\;\sqrt{1-\eta_{0}}\;.\label{impo300}\end{eqnarray}
3792: Therefore Eq.~(\ref{fin2}) gives \begin{eqnarray}
3793: \sqrt{\eta}\;|{}_{M}\langle\phi|V|\psi\rangle_{M}| & > & \sqrt{\eta_{0}}\;-5\; d_{C\bar{C}}\;(1-\eta_{0})^{1/4}\;.\label{fin4}\end{eqnarray}
3794: By Eq.~(\ref{fin1}) it follows that \begin{eqnarray}
3795: F_{r} & \geq & \eta_{0}\;-10\; d_{C\bar{C}}\;(1-\eta_{0})^{1/4}\;.\label{fin1000}\end{eqnarray}
3796:
3797:
3798: The fidelity for writing a state $|\psi\rangle_{\bar{C}C}$ into $\bar{C}C$
3799: is given by \begin{eqnarray}
3800: F_{w}(\psi)\equiv{}_{C\bar{C}}\langle\psi|\textrm{tr}_{M}\left[W^{\dag}V\left(|\psi\rangle_{M}\langle\psi|\otimes|0\rangle_{\bar{C}C}\langle0|\right)V^{\dag}W\right]|\psi\rangle_{C\bar{C}}.\end{eqnarray}
3801: A lower bound for this quantity is obtained by replacing the trace
3802: over $M$ with the expectation value on $|0\rangle_{M}$, i.e. \begin{eqnarray}
3803: F_{w}(\psi) & \geq & _{C\bar{C}}\langle\psi|{}_{M}\langle0|W^{\dag}V\left(|\psi\rangle_{M}\langle\psi|\otimes|0\rangle_{\bar{C}C}\langle0|\right)V^{\dag}W|0\rangle_{M}|\psi\rangle_{C\bar{C}}\nonumber \\
3804: & = & \left|_{C\bar{C}}\langle0|{}_{M}\langle\psi|V^{\dag}W|0\rangle_{M}|\psi\rangle_{C\bar{C}}\right|^{2}\nonumber \\
3805: & = & \eta\;\left|_{M}\langle\psi|V^{\dag}|\phi\rangle_{M}\right|^{2}=\eta\;\left|_{M}\langle\phi|V|\psi\rangle_{M}\right|^{2}\label{vvv}\end{eqnarray}
3806: where Eqs.~(\ref{eq:main}) and the orthogonality relation~(\ref{impo})
3807: have been employed to derive the second identity. Notice that the
3808: last term of the inequality~(\ref{vvv}) coincides with the lower
3809: bound~(\ref{fin1}) of the reading fidelity. Therefore, by applying
3810: the same derivation of the previous section we can write \begin{equation}
3811: \boxed{F\ge\eta_{0}\;-10\; d_{C\bar{C}}\;(1-\eta_{0})^{1/4},}\label{fin3000}\end{equation}
3812: which shows that the \index{reading and writing fidelities}reading
3813: and writing fidelities converge to $1$ in the limit of large $L$.
3814: Note that this lower bound can probably be largely improved.
3815:
3816:
3817: \section{Application to spin chain communication}
3818:
3819: We now show how the above protocol can be used to improve quantum
3820: state transfer on a spin chain. The main advantage of using such a
3821: memory protocol is that - opposed to all other schemes - Alice can
3822: send arbitrary multi-qubit states with a single usage of the channel.
3823: She needs no encoding, all the work is done by Bob. The protocol proposed
3824: here can be used to improve the performances of any scheme mentioned
3825: in Section~\ref{sec:Advanced-transfer-protocols}, and it works for
3826: a large class of Hamiltonians, including Heisenberg and XY models
3827: with arbitrary (also randomly distributed) coupling strengths. %
3828: \begin{figure}[t]
3829: \begin{centering}\includegraphics[width=1\columnwidth]{chainw}\par\end{centering}
3830:
3831:
3832: \caption{\label{fig:swapping}Alice and Bob control the spins $N_{A}$ and
3833: $N_{B}$ interconnected by the spins $N_{R}.$ At time $jt$ Bob performs
3834: a swap $S_{j}$ between his spins and the memory $M_{j}$.}
3835: \end{figure}
3836:
3837:
3838: Consider a chain of spin-$1/2$ particles described by a Hamiltonian
3839: $H$ which conserves the number of excitations. The chain is assumed
3840: to be divided in three portions $A$ (Alice), $B$ (Bob) and $R$
3841: (the remainder of the chain, connecting Alice and Bob) containing
3842: respectively the first $N_{A}$ spins of the chain, the last $N_{B}$
3843: spins and the intermediate $N_{R}$ spins, and the total length of
3844: the chain is $N=N_{A}+N_{R}+N_{B}$ (see Fig \ref{fig:swapping}).
3845: Bob has access also to a collection of quantum memories $M_{1},\cdots,M_{j}\cdots,M_{L}$
3846: isomorphic with $B$, i.e. each having dimension equal to the dimension
3847: $2^{N_{B}}$ of $B$. We assume that Bob's memory is initialised in
3848: the zero excitation state $|0\rangle_{M}.$ Alice prepares an arbitrary
3849: and unknown state $|\psi\rangle_{A}$ on her $N_{A}$ qubits. By defining
3850: the (from Bob's perspective) controlled part of system $C=B$ and
3851: the uncontrolled part $\bar{C}=AR,$ we can apply the results of the
3852: last sections and get the following
3853:
3854: \begin{framedtheorem}
3855: [Memory swapping]\label{thm:Let-H-be}Let $H$ be the Hamiltonian
3856: of an open nearest-neighbour quantum chain that conserves the number
3857: of excitations. If there is a time $t$ such that $f_{1,N}(t)\neq0$
3858: (i.e. the Hamiltonian is capable of transport between Alice and Bob)
3859: then the state transfer can be made arbitrarily perfect by using the
3860: memory swapping protocol.
3861: \end{framedtheorem}
3862: \begin{proof}
3863: We only have to show that the reduced dynamics on the chain is mixing
3864: with a pure fixed point. Using the number of excitations as a conserved
3865: additive observable, we can use the criterion of Lemma~\ref{lem:add}:
3866: If there exists exactly one eigenstate $|E\rangle$ of factorising
3867: form with $|0\rangle_{B}$, i.e. \begin{equation}
3868: \exists_{1}\:|\lambda\rangle_{AR}:\quad H|\lambda\rangle_{AR}\otimes|0\rangle_{B}=E|\lambda\rangle_{AR}\otimes|0\rangle_{B},\label{eq:condition}\end{equation}
3869: then the reduced dynamics is mixing toward $|0\rangle_{AR}.$ Assume
3870: by contradiction that has an eigenvector $|E\rangle_{AR}\neq|0\rangle_{AR}$
3871: which falsifies Eq.~(\ref{eq:condition}). Such an eigenstate can
3872: be written as \begin{equation}
3873: |E\rangle_{AR}\otimes|0\rangle_{B}=a|\mu\rangle_{AR}\otimes|0\rangle_{B}+b|\bar{\mu}\rangle_{AR}\otimes|0\rangle_{B},\label{eq:eig}\end{equation}
3874: where $a$ and $b$ are complex coefficients and where the spin just
3875: before the section $B$ (with position $N_{A}+N_{R}$) is in the state
3876: $|0\rangle$ for $|\mu\rangle_{AR}$ and in the state $|1\rangle$
3877: for $|\bar{\mu}\rangle_{AR}.$ Since the interaction between this
3878: spin and the first spin of section $B$ includes an exchange term
3879: (otherwise $f_{1,N}(t)$=0 for all $t$), then the action of $H$
3880: on the second term of~(\ref{eq:eig}) yields exactly one state which
3881: contains an excitation in the sector $B.$ It cannot be compensated
3882: by the action of $H$ on the first term of~(\ref{eq:eig}). But by
3883: assumption $|E\rangle_{AR}\otimes|0\rangle_{B}$ is an eigenstate
3884: of $H$, so we conclude that $b=0.$ This argument can be repeated
3885: for the second last spin of section $R$, the third last spin, and
3886: so on, to finally yield $|E\rangle_{AR}=|0\rangle_{AR}$, as long
3887: as all the nearest neighbour interactions contain exchange parts.
3888: \end{proof}
3889: \begin{remark}
3890: Theorem~\ref{thm:Let-H-be} should be compared to Theorem \ref{thm:Let2}
3891: for the multi rail protocol. They are indeed very similar. However
3892: the current theorem is much stronger, since it allows to send arbitrary
3893: multi-excitation states, and also to write states back onto the chain.
3894: It is interesting to note that Lemma~\ref{lem:add} and Theorem~\ref{thm:Let-H-be}
3895: indicate a connection between the dynamical controllability of a system
3896: and its static entanglement properties. It may be interesting to obtain
3897: a \emph{quantitative} relation between the amount of entanglement
3898: and the convergence speed.
3899: \end{remark}
3900: Let us now come back to the question raised in Section~\ref{sec:Protocol}
3901: about the operation $W^{\dag}$ being unphysical. As mentioned before,
3902: this can be fixed using a simple transformation: if the Hamiltonian
3903: $H$ fulfils the requirements of Lemma~\ref{lem:add}, then also
3904: the Hamiltonian $-H$ fulfils them. Now derive the coding transformation
3905: $\tilde{V}$ as given in Section~\ref{sec:Coding-transformation}
3906: for the Hamiltonian $\tilde{H}=-H$. In this picture, the reading
3907: protocol $W$ is unphysical, whereas the writing protocol becomes
3908: physical. In the more general case where the condition of Lemma~\ref{lem:add}
3909: is not valid, but the map \begin{equation}
3910: \tau(\rho_{\bar{C}})\equiv\textrm{tr}_{C}\left[U\left(\rho_{\bar{C}}\otimes|0\rangle_{C}\langle0|\right)U^{\dag}\right]\end{equation}
3911: is still ergodic with a pure fixed point, we then require the map\begin{equation}
3912: \tilde{\tau}(\rho_{\bar{C}})\equiv\textrm{tr}_{C}\left[U^{\dag}\left(\rho_{\bar{C}}\otimes|0\rangle_{C}\langle0|\right)U\right]\end{equation}
3913: to be also ergodic with pure fixed point to be able to use this trick.
3914:
3915:
3916: \section{Conclusion}
3917:
3918: We have given an explicit protocol for controlling a large permanently
3919: coupled system by accessing a small subsystem only. In the context
3920: of quantum chain communication this allows us to make use of the quantum
3921: memory of the receiving party to improve the fidelity to a value limited
3922: only by the size of the memory. We have shown that this scheme can
3923: be applied to a Heisenberg spin chain. The main advantage of this
3924: method is that arbitrary multi-excitation states can be transferred.
3925: Also, our method can be applied to chains that do not conserve the
3926: number of excitations in the system, as long as the reduced dynamic
3927: is ergodic with a pure fixed point.
3928:
3929: It remains an open question how much of our results remain valid if
3930: the channel is mixing toward a \emph{mixed} state. In this case, a
3931: part of the quantum information will in general remain in the correlations
3932: between the system and the memory, and it cannot be expected that
3933: the fidelity converges to one. However, by concentrating only on the
3934: eigenstate of the fixed point density operator with the largest eigenvalue,
3935: it should be possible to derive some bounds of the amount of information
3936: that can be extracted.
3937:
3938:
3939: \chapter{A valve for probability amplitude\label{cha:Single-memory}}
3940:
3941:
3942: \section{Introduction}
3943:
3944: We have mainly discussed two methods for quantum state transfer so
3945: far. In the first one, multiple chains where used, and in the second
3946: one, a single chain was used in combination with a large quantum memory.
3947: Can we combine the best of the two schemes, i.e. is it possible to
3948: use only a single chain and a single memory qubit? In this chapter
3949: we will show that this is indeed the case and that the fidelity can
3950: be improved easily by applying in certain time-intervals two-qubit
3951: gates at the receiving end of the chain. These gates act as a \emph{valve\index{valve}}
3952: which takes probability amplitude out of the system without ever putting
3953: it back. The required sequence is determined \emph{a priori} by the
3954: Hamiltonian of the system. Such a protocol is \emph{optimal} in terms
3955: of resources, because two-qubit gates at the sending and receiving
3956: end are required in order to connect the chain to the blocks in \emph{all}
3957: above protocols (though often not mentioned explicitly). At the same
3958: time, the engineering demands are not higher then for the memory swapping
3959: protocol. Our scheme has some similarities with~\cite{Haselgrove2005},
3960: but the gates used here are much simpler, and arbitrarily high fidelity
3961: is guaranteed by a convergence theorem for arbitrary coupling strengths
3962: and all non-Ising coupling types that conserve the number of excitations.
3963: Furthermore, we show numerically that our protocol could also be realised
3964: by a simple switchable interaction.
3965:
3966:
3967: \section{Arbitrarily Perfect State Transfer}
3968:
3969: We now show how the receiver can improve the fidelity to an arbitrarily
3970: high value by applying two-qubit gates between the end of the chain
3971: and a {}``target qubit'' of the block. We label the qubits of the
3972: chain by $1,2,\cdots,N$ and the target qubit by $N+1$ (see Fig.~\ref{fig:setup}).
3973: The coupling of the chain is described by a Hamiltonian $H.$ We assume
3974: that the Hamiltonian $H$ conserves the number of excitations and
3975: that the target qubit $N+1$ is uncoupled,\begin{equation}
3976: H|\boldsymbol{N+1}\rangle=0\label{eq:extra}\end{equation}
3977: and set the energy of the ground state $|\boldsymbol{0}\rangle$ to
3978: zero. For what follows we restrict all operators to the $N+2$ dimensional
3979: Hilbert space\begin{equation}
3980: \mathcal{H}=\textrm{span}\left\{ |\boldsymbol{n}\rangle;\; n=0,1,2,\ldots,N+1\right\} .\end{equation}
3981: Our final assumption about the Hamiltonian of the system is that there
3982: exists a time $t$ such that\begin{equation}
3983: f_{N,t}(t)\equiv\langle\boldsymbol{N}|\exp\left\{ -itH\right\} |\boldsymbol{1}\rangle\neq0.\end{equation}
3984: Physically this means that the Hamiltonian has the capability of
3985: transporting from the first to the last qubit of the chain. As mentioned
3986: in the introduction, the fidelity of this transport may be very bad
3987: in practice. %
3988: \begin{figure}[htbp]
3989: \begin{centering}\includegraphics[width=1\columnwidth]{setup2}\par\end{centering}
3990:
3991:
3992: \caption{\label{fig:setup}A quantum chain (qubits $1,2,\cdots,N$) and a
3993: target qubit ($N+1$). By applying a sequence of two-qubit unitary
3994: gates $V_{k}$ on the last qubit of the chain and the target qubit,
3995: arbitrarily high fidelity can be achieved.}
3996: \end{figure}
3997:
3998:
3999: We denote the unitary evolution operator for a given time $t_{k}$
4000: as $U_{k}\equiv\exp\left\{ -it_{k}H\right\} $ and introduce the projector\begin{equation}
4001: P=1-|\boldsymbol{0}\rangle\langle\boldsymbol{0}|-|\boldsymbol{N}\rangle\langle\boldsymbol{N}|-|\boldsymbol{N+1}\rangle\langle\boldsymbol{N+1}|.\end{equation}
4002: A crucial ingredient to our protocol is the operator\begin{eqnarray}
4003: V(c,d) & \equiv & P+|\boldsymbol{0}\rangle\langle\boldsymbol{0}|+d|\boldsymbol{N}\rangle\langle\boldsymbol{N}|+d^{*}|\boldsymbol{N+1}\rangle\langle\boldsymbol{N+1}|\nonumber \\
4004: & & +c^{*}|\boldsymbol{N+1}\rangle\langle\boldsymbol{N}|-c|\boldsymbol{N}\rangle\langle\boldsymbol{N+1}|,\end{eqnarray}
4005: where $c$ and $d$ are complex normalised amplitudes. It is easy
4006: to check that\begin{equation}
4007: VV^{\dagger}=V^{\dagger}V=1,\end{equation}
4008: so $V$ is a unitary operator on $\mathcal{H}.$ $V$ acts as the
4009: identity on all but the last two qubits, and can hence be realised
4010: by \emph{a local two-qubit gate on the qubits $N$ and $N+1$.} Furthermore
4011: we have $VP=P$ and\begin{equation}
4012: V(c,d)\left[\left\{ c|\boldsymbol{N}\rangle+d|\boldsymbol{N+1}\rangle\right\} \right]=|\boldsymbol{N+1}\rangle.\label{eq:wdesign}\end{equation}
4013: The operator $V(c,d)$ has the role of moving probability amplitude
4014: $c$ from the $N$th qubit to target qubit, without moving amplitude
4015: back into the system, and can be thought of as a \emph{valve.} Of
4016: course as $V(c,d)$ is unitary, there are also states such that $V(c,d)$
4017: acting on them would move back probability amplitude into the system,
4018: but these do not occur in the protocol discussed here.
4019:
4020: Using the time-evolution operator and two-qubit unitary gates on the
4021: qubits $N$ and $N+1$ we will now develop a protocol that transforms
4022: the state $|\boldsymbol{1}\rangle$ into |$\boldsymbol{N+1}\rangle.$
4023: Let us first look at the action of $U_{1}$ on $|\boldsymbol{1}\rangle.$
4024: Using the projector $P$ we can decompose this time-evolved state
4025: as\begin{eqnarray}
4026: U_{1}|\boldsymbol{1}\rangle & = & PU_{1}|\boldsymbol{1}\rangle+|\boldsymbol{N}\rangle\langle\boldsymbol{N}|U_{1}|\boldsymbol{1}\rangle\nonumber \\
4027: & \equiv & PU_{1}|\boldsymbol{1}\rangle+\sqrt{p_{1}}\left\{ c_{1}|\boldsymbol{N}\rangle+d_{1}|\boldsymbol{N+1}\rangle\right\} ,\end{eqnarray}
4028: where $p_{1}=\left|\langle\boldsymbol{N}|U_{1}|\boldsymbol{1}\rangle\right|^{2},$
4029: $c_{1}=\langle\boldsymbol{N}|U_{1}|\boldsymbol{1}\rangle/\sqrt{p_{1}}$
4030: and $d_{1}=0.$ Let us now consider the action of $V_{1}\equiv V(c_{1},d_{1})$
4031: on the time-evolved state. By Eq. (\ref{eq:wdesign}) it follows that\begin{eqnarray}
4032: V_{1}U_{1}|\boldsymbol{1}\rangle & = & PU_{1}|\boldsymbol{1}\rangle+\sqrt{p_{1}}|\boldsymbol{N}+1\rangle.\label{eq:firsttime}\end{eqnarray}
4033: Hence with a probability of $p_{1},$ the excitation is now in the
4034: position $N+1,$ where it is {}``frozen'' (since that qubit is not
4035: coupled to the chain. We will now show that at the next step, this
4036: probability is increased. Applying $U_{2}$ to Eq. (\ref{eq:firsttime})
4037: we get \begin{eqnarray}
4038: \lefteqn{U_{2}V_{1}U_{1}|\boldsymbol{1}\rangle}\nonumber \\
4039: & = & PU_{2}PU_{1}|\boldsymbol{1}\rangle+\langle\boldsymbol{N}|U_{2}PU_{1}|\boldsymbol{1}\rangle|\boldsymbol{N}\rangle+\sqrt{p_{1}}|\boldsymbol{N}+1\rangle\nonumber \\
4040: & = & PU_{2}PU_{1}|\boldsymbol{1}\rangle+\sqrt{p_{2}}\left\{ c_{2}|\boldsymbol{N}\rangle+d_{2}|\boldsymbol{N}+1\rangle\right\} \end{eqnarray}
4041: with $c_{2}=\langle\boldsymbol{N}|U_{2}PU_{1}|\boldsymbol{1}\rangle/\sqrt{p_{2}},$
4042: $d_{2}=\sqrt{p_{1}}/\sqrt{p_{2}}$ and\begin{eqnarray}
4043: p_{2} & = & p_{1}+\left|\langle\boldsymbol{N}|U_{2}PU_{1}|\boldsymbol{1}\rangle\right|^{2}\ge p_{1}.\end{eqnarray}
4044: Applying $V_{2}\equiv V(c_{2},d_{2})$ we get \begin{equation}
4045: V_{2}U_{2}V_{1}U_{1}|\boldsymbol{1}\rangle=PU_{2}PU_{1}|\boldsymbol{1}\rangle+\sqrt{p_{2}}|\boldsymbol{N}+1\rangle.\end{equation}
4046: Repeating this strategy $\ell$ times we get\begin{equation}
4047: \left(\prod_{k=1}^{\ell}V_{k}U_{k}\right)|\boldsymbol{1}\rangle=\left(\prod_{k=1}^{\ell}PU_{k}\right)|\boldsymbol{1}\rangle+\sqrt{p_{\ell}}|\boldsymbol{N}+1\rangle,\label{eq:general}\end{equation}
4048: where the products are arranged in the time-ordered way. Using the
4049: normalisation of the r.h.s. of Eq. (\ref{eq:general}) we get\begin{equation}
4050: p_{\ell}=1-\left\Vert \left(\prod_{k=1}^{\ell}PU_{k}\right)|\boldsymbol{1}\rangle\right\Vert ^{2}.\end{equation}
4051: From Section~\ref{s:sec4} we know that there exists a $t>0$ such
4052: that for equal time intervals $t_{1}=t_{2}=\ldots=t_{k}=t$ we have
4053: $\lim_{\ell\rightarrow\infty}p_{\ell}=1.$ Therefore the limit of
4054: infinite gate operations for Eq. (\ref{eq:general}) is given by\begin{equation}
4055: \lim_{\ell\rightarrow\infty}\left(\prod_{k=1}^{\ell}V_{k}U_{k}\right)|\boldsymbol{1}\rangle=|\boldsymbol{N+1}\rangle.\label{eq:convergence}\end{equation}
4056: It is also easy to see that $\lim_{k\rightarrow\infty}d_{\ell}=1,$
4057: $\lim_{k\rightarrow\infty}c_{\ell}=0$ and hence the gates $V_{k}$
4058: converge to the identity operator. Furthermore, since $V_{k}U_{k}|\boldsymbol{0}\rangle=|\boldsymbol{0}\rangle$
4059: it also follows that arbitrary superpositions can be transferred.
4060: As discussed in Theorem~\ref{speed}, this convergence is asymptotically
4061: exponentially fast in the number of gate applied (a detailed analysis
4062: of the relevant scaling can be found in Chapter~\ref{cha:Dual-Rail}).
4063: Equation (\ref{eq:convergence}) is a surprising result, which shows
4064: that \emph{any non-perfect transfer can be made arbitrarily perfect}
4065: by only applying two-qubit gates on one end of the quantum chain.
4066: It avoids restricting the gate times to specific times (as opposed
4067: to the dual rail scheme) while requiring no additional memory qubit
4068: (as opposed to the memory swapping scheme).
4069:
4070: The sequence $V_{k}$ that needs to be applied to the end of the chain
4071: to perform the state transfer only depends on the Hamiltonian of the
4072: quantum chain. The relevant properties can in principle be determined
4073: a priori by preceding measurements and tomography on the quantum chain
4074: (as discussed in Sect.~\ref{sec:Tomography}).
4075:
4076:
4077: \section{Practical Considerations\label{sec:Practical-Considerations}}
4078:
4079: Motivated by the above result we now investigate how the above protocol
4080: may be implemented in practice, well before the realisation of the
4081: quantum computing blocks from Fig.~\ref{fig:connect2}. The two-qubit
4082: gates $V_{k}$ are essentially rotations in the $\{|01\rangle,|10\rangle\}$
4083: space of the qubits $N$ and $N+1.$ It is therefore to be expected
4084: that they can be realised (up to a irrelevant phase) by a switchable
4085: Heisenberg or $XY$ type coupling between the $Nth$ and the target
4086: qubit. However in the above, we have assumed that the gates $V_{k}$
4087: can be applied instantaneously, i.e. in a time-scale much smaller
4088: than the time-scale of the dynamics of the chain. This corresponds
4089: to a switchable coupling that is much stronger than the coupling strength
4090: of the chain. %
4091: \begin{figure}[htbp]
4092: \begin{centering}\includegraphics[width=0.8\columnwidth]{fig2}\par\end{centering}
4093:
4094:
4095: \caption{\label{fig:numerics}Numerical example for the convergence of the
4096: success probability. Simulated is a quantum chain of length $N=20$
4097: with the Hamiltonian from Eq. (\ref{eq:heis}) (dashed line) and Eq.
4098: (\ref{eq:mag}) with $B/J=20$ (solid line). Using the original protocol~\cite{Bose2003},
4099: the same chain would only reach a success probability of $0.63$ in
4100: the above time interval.}
4101: \end{figure}
4102:
4103:
4104: Here, we numerically investigate if a convergence similar to the above
4105: results is still possible when this assumption is not valid. We \emph{do}
4106: however assume that the switching of the interaction is still describable
4107: by an instantaneous switching (i.e. the sudden approximation is valid).
4108: This assumption is mainly made to keep the numerics simple. We do
4109: not expect qualitative differences when the switching times become
4110: finite as long as the time-dependent Hamiltonian is still conserving
4111: the number of excitations in the chain. In fact it has recently been
4112: shown that the finite switching time can even \emph{improve} the fidelity~\cite{Bruder}.
4113: Intuitively, this happens because by gradually decreasing the coupling,
4114: he not only receives the probability amplitude of the last qubit of
4115: the chain, but can also ''swallow'' a bit of the dispersed wave-packed
4116: (similar to the situation discussed in \cite{Haselgrove2005}).
4117:
4118: We have investigated two types of switching. For the first type, the
4119: coupling itself is switchable, i.e.\begin{equation}
4120: H(t)=J\sum_{n=1}^{N-1}\sigma_{n}^{-}\sigma_{n+1}^{+}+\Delta(t)\sigma_{N}^{-}\sigma_{N+1}^{+}+\textrm{h.c.},\label{eq:heis}\end{equation}
4121: where $\Delta(t)$ can be $0$ or $1.$ For the second type, the
4122: target qubit is \emph{permanently} coupled to the remainder of the
4123: chain, but a strong magnetic field on the last qubit can be switched,
4124: \begin{equation}
4125: H(t)=J\sum_{n=1}^{N}\sigma_{n}^{-}\sigma_{n+1}^{+}+\textrm{h.c.}+B\Delta(t)\sigma_{N+1}^{z},\label{eq:mag}\end{equation}
4126: where again $\Delta(t)$ can be $0$ or $1$ and $B\gg1.$ This suppresses
4127: the coupling between the $N$th and $N+1$th qubit due to an energy
4128: mismatch.
4129:
4130: In both cases, we first numerically optimise the times for unitary
4131: evolution $t_{k}$ over a fixed time interval such that the probability
4132: amplitude at the $N$th qubit is maximal. The algorithm then finds
4133: the optimal time interval during which $\Delta(t)=1$ such that the
4134: probability amplitude at the target qubit is increased. In some cases
4135: the phases are not correct, and switching on the interaction would
4136: result in probability amplitude floating back into the chain. In this
4137: situation, the target qubit is left decoupled and the chain is evolved
4138: to the next amplitude maximum at the $N$th qubit. Surprisingly, even
4139: when the time-scale of the gates is comparable to the dynamics, near-perfect
4140: transfer remains possible~(Fig \ref{fig:numerics}). In the case
4141: of the switched magnetic field, the achievable fidelity depends on
4142: the strength of the applied field. This is because the magnetic field
4143: does not fully suppress the coupling between the two last qubits.
4144: A small amount of probability amplitude is lost during each time evolution
4145: $U_{k},$ and when the gain by the gate is compensated by this loss,
4146: the total success probability no longer increases.
4147:
4148:
4149: \section{Conclusion}
4150:
4151: We have seen that by having a simple switchable interaction acting
4152: as a \emph{valve} for probability amplitude, arbitrarily perfect state
4153: transfer is possible on a single spin chain. In fact, by using the
4154: inverse protocol, arbitrary%
4155: \footnote{Opposed to the method for state preparation developed in the last
4156: chapter this allows the creation of \emph{known} states only (as the
4157: valve operations $V_{k}$ depend explicitly on the state that one
4158: wants to prepare).%
4159: } states in the first excitation sector can also be prepared on the
4160: chain. Furthermore, this protocol can easily be adopted to arbitrary
4161: graphs connecting multiple senders and receivers (as discussed for
4162: weakly coupled systems in~\cite{Bednarska}).
4163:
4164:
4165: \chapter{External noise\label{cha:Problems-and-Practical}}
4166:
4167:
4168: \section{Introduction}
4169:
4170: An important question that was left open so far is what happens to
4171: quantum state transfer in the presence of external noise. It is well
4172: known from the theory of open quantum systems~\cite{OPENQUANTUM}
4173: that this can lead to dissipation and decoherence, which also means
4174: that quantum information is lost. The evolution of a closed quantum
4175: system is described by the Schrödinger equation\index{Schrödinger equation}\begin{equation}
4176: \partial_{t}|\psi\rangle=-iH|\psi\rangle.\end{equation}
4177: If a system is very strongly coupled to a environment, the dynamic
4178: is completely incoherent and described by some simple rate equations
4179: for the occupation probabilities,\begin{equation}
4180: \partial_{t}P_{n}=\sum_{n}k_{n\rightarrow m}P_{n}-\sum_{n}k_{m\rightarrow n}P_{m}.\end{equation}
4181: In the more general case where the dynamic consists of coherent and
4182: incoherent parts, the evolution can sometimes be expressed as a Lindblad
4183: equation\index{Lindblad equation}~\cite{OPENQUANTUM}\begin{equation}
4184: \partial_{t}\rho=\mathcal{L}\rho\end{equation}
4185: for the reduced density matrix. These three regimes are shown in Fig.~\ref{fig:Dominant}.
4186: %
4187: \begin{figure}[tbh]
4188: \begin{centering}\includegraphics[width=0.45\paperwidth]{regimes}\par\end{centering}
4189:
4190:
4191: \caption{\label{fig:Dominant}Dominant regimes of dynamics depending on the
4192: relative strength of the system Hamiltonian and the environmental
4193: coupling~\cite{EXCITON}.}
4194: \end{figure}
4195: For quantum information theory, coherence is essential~\cite{NIELSEN},
4196: and one has to try to isolate the quantum chain as much as possible
4197: from the environment. In the partially coherent regime, typically
4198: the quantum behaviour decays exponentially with a rate depending on
4199: the temperature of the environment. Not surprisingly, this has also
4200: been found in the context of quantum state transfer~\cite{Guo2006,Karimipour2005,Sun}.
4201: From a theoretical point of view it is perhaps more interesting to
4202: look at the low temperature and strong coupling regime, where the
4203: dynamics is often non-Markovian~\cite{OPENQUANTUM} and can no longer
4204: expressed as a simple Lindblad equation. This is also interesting
4205: from a practical perspective, corresponding to effects of the environment
4206: which cannot be avoided by cooling. Here we consider a model where
4207: the system is coupled to a spin environment through an exchange interaction.
4208: This coupling offers the unique opportunity of an analytic solution
4209: of our problem without \emph{any} approximations regarding the strength
4210: of system-environment coupling (in most treatments of the effect of
4211: an environment on the evolution of a quantum system, the system-environment
4212: coupling is assumed to be weak) and allows us to include inhomogeneous
4213: interactions of the bath spins with the system. For such coupling,
4214: decoherence is possible for mixed (thermal) initial bath states~\cite{Schmidt2005,Breuer2004}.
4215: However if the system and bath are both initially cooled to their
4216: ground states, is there still a non-trivial effect of the environment
4217: on the fidelity? In this chapter we find that there are two important
4218: effects: the spin transfer functions (Eq.~\ref{eq:spintrans}) are
4219: \emph{slowed down} by a factor of two, and \emph{destabilised} by
4220: a modulation of $\left|\cos Gt\right|,$ where $G$ is the mean square
4221: coupling to the environment. This has both positive and negative implications
4222: for the use of strongly coupled spin systems as quantum communication
4223: channels. The spin transfer functions also occur in the charge and
4224: energy transfer dynamics in molecular systems~\cite{EXCITON} and
4225: in continuous time random walks~\cite{Blumen2006} to which our results
4226: equally apply.
4227:
4228:
4229: \section{Model}
4230:
4231: We choose to start with a specific spin system, i.e. an open spin
4232: chain of arbitrary length $N,$ with a Hamiltonian given by\begin{equation}
4233: H_{S}=-\frac{1}{2}\sum_{\ell=1}^{N-1}J_{\ell}\left(X_{\ell}X_{\ell+1}+Y_{\ell}Y_{\ell+1}\right),\end{equation}
4234: where $J_{\ell}$ are some arbitrary couplings and $X_{\ell}$ and
4235: $Y_{\ell}$ are the Pauli-X and Y matrices for the $\ell$th spin.
4236: Toward the end of the section we will however show that our results
4237: hold for any system where the number of excitations is conserved during
4238: dynamical evolution. In addition to the chain Hamiltonian, each spin
4239: $\ell$ of the chain interacts with an independent bath of $M_{\ell}$
4240: environmental spins (see Fig \ref{fig:spinchain}) via an inhomogeneous
4241: Hamiltonian,\begin{equation}
4242: H_{I}^{(\ell)}=-\frac{1}{2}\sum_{k=1}^{M_{\ell}}g_{k}^{(\ell)}\left(X_{\ell}X_{k}^{(\ell)}+Y_{\ell}Y_{k}^{(\ell)}\right).\end{equation}
4243:
4244:
4245: %
4246: \begin{figure}[htbp]
4247: \begin{centering}\includegraphics[width=0.6\paperwidth]{bc_fig1}\par\end{centering}
4248:
4249:
4250: \caption{\label{fig:spinchain}A spin chain of length $N=5$ coupled to independent
4251: baths of spins. }
4252: \end{figure}
4253:
4254:
4255: In the above expression, the Pauli matrices $X_{\ell}$ and $Y_{\ell}$
4256: act on the $\ell$th spin of the chain, whereas $X_{k}^{(\ell)}$
4257: and $Y_{k}^{(\ell)}$ act on the $k$th environmental spin attached
4258: to the $\ell$th spin of the chain. We denote the total interaction
4259: Hamiltonian by\begin{equation}
4260: H_{I}\equiv\sum_{\ell=1}^{N}H_{I}^{(\ell)}.\end{equation}
4261: The total Hamiltonian is given by $H=H_{S}+H_{I},$ where it is important
4262: to note that $\left[H_{S},H_{I}\right]\neq0.$ We assume that a homogeneous
4263: magnetic field along the z-axis is applied. The ground state of the
4264: system is then given by the fully polarised state $|0,0\rangle,$
4265: with all chain and bath spins aligned along the z-axis. The above
4266: Hamiltonian describes an extremely complex and disordered system with
4267: a Hilbert space of dimension $2^{N+NM}.$ In the context of state
4268: transfer however, only the dynamics of the first excitation sector
4269: is relevant. We proceed by mapping this sector to a much simpler system~\cite{Buzek2005,Chiara2005,Hutton2004,Olaya-Castro2005,Olaya-Castro2006}.
4270: For $\ell=1,2,\ldots,N$ we define the states\begin{equation}
4271: |\ell,0\rangle\equiv X_{\ell}|0,0\rangle\end{equation}
4272: and\begin{equation}
4273: |0,\ell\rangle\equiv\frac{1}{G_{\ell}}\sum_{k=1}^{M_{\ell}}g_{k}^{(\ell)}X_{k}^{(\ell)}|0,0\rangle\end{equation}
4274: with \begin{equation}
4275: G_{\ell}=\sqrt{\sum_{k=1}^{M_{\ell}}\left(g_{k}^{(\ell)}\right)^{2}}.\label{eq:effect}\end{equation}
4276: It is easily verified that (setting $J_{0}=J_{N}=0$)\begin{eqnarray}
4277: H_{S}|\ell,0\rangle & = & -J_{\ell-1}|\ell-1,0\rangle-J_{\ell}|\ell+1,0\rangle\nonumber \\
4278: H_{S}|0,\ell\rangle & = & 0,\label{eq:hc_action}\end{eqnarray}
4279: and\begin{eqnarray}
4280: H_{I}|\ell,0\rangle & = & -G_{\ell}|0,\ell\rangle\label{eq:hb_action}\\
4281: H_{I}|0,\ell\rangle & = & -G_{\ell}|\ell,0\rangle.\label{eq:hb_action2}\end{eqnarray}
4282: Hence these states define a $2N-$dimensional subspace that is invariant
4283: under the action of $H.$ This subspace is equivalent to the first
4284: excitation sector of a system of $2N$ spin $1/2$ particles, coupled
4285: as it is shown in Fig \ref{fig:spinchain_equiv}. %
4286: \begin{figure}[htbp]
4287: \begin{centering}\includegraphics[width=0.6\paperwidth]{bc_fig2}\par\end{centering}
4288:
4289:
4290: \caption{\label{fig:spinchain_equiv}In the first excitation sector, the system
4291: can be mapped into an effective spin model where the bath spins are
4292: replaced by a single effective spin, as indicated here for $N=5.$}
4293: \end{figure}
4294:
4295:
4296: Our main assumption is that the bath couplings are \emph{in effect}
4297: the same, i.e. $G_{\ell}=G$ for all $\ell$. Note however that the
4298: individual number of bath spins $M_{\ell}$ and bath couplings $g_{k}^{(\ell)}$
4299: may still depend on $\ell$ and $k$ as long as their means square
4300: average is the same. Also, our analytic solution given in the next
4301: paragraph relies on this assumption, but numerics show that our main
4302: result {[}Equation~(\ref{eq:scalingformula})] remains a good approximation
4303: if the $G_{\ell}$ slightly vary and we take $G\equiv\left\langle G_{\ell}\right\rangle .$
4304: Disorder in the vertical couplings is treated \emph{exactly} in the
4305: sense that our results hold for any choice of couplings $J_{\ell}.$
4306:
4307:
4308: \section{\label{sec:Solving-the-Schr=3D3DF6dinger}Results}
4309:
4310: In this paragraph, we solve the Schrödinger equation for the model
4311: outlined above and discuss the spin transfer functions. Firstly, let
4312: us denote the orthonormal eigenstates of $H_{S}$ alone by \begin{equation}
4313: H_{S}|\psi_{k}\rangle=\epsilon_{k}|\psi_{k}\rangle\quad(k=1,2\ldots,N)\end{equation}
4314: with\begin{equation}
4315: |\psi_{k}\rangle=\sum_{\ell=1}^{N}a_{k\ell}|\ell,0\rangle.\label{eq:eigen_h_c}\end{equation}
4316: For what follows, it is not important whether analytic expressions
4317: for the eigensystem of $H_{S}$ can be found. Our result holds even
4318: for models that are not analytically solvable, such as the randomly
4319: coupled chains considered in Section~\ref{sec:Disordered-chains}.
4320: We now make an ansatz for the eigenstates of the full Hamiltonian,
4321: motivated by the fact that the states \begin{equation}
4322: |\phi_{\ell}^{n}\rangle\equiv\frac{1}{\sqrt{2}}\left(|\ell,0\rangle+\left(-1\right)^{n}|0,\ell\rangle\right)\quad(n=1,2)\end{equation}
4323: are eigenstates of $H_{I}^{(\ell)}$ with the corresponding eigenvalues
4324: $\pm G$ {[}this follows directly from Eqs.~(\ref{eq:hb_action})
4325: and~(\ref{eq:hb_action2})]. Define the vectors\begin{eqnarray}
4326: |\Psi_{k}^{n}\rangle & \equiv & \sum_{\ell=1}^{N}a_{k\ell}|\phi_{\ell}^{n}\rangle\label{eq:ansatz}\end{eqnarray}
4327: with $k=1,2,\ldots,N$ and $n=0,1.$ The $|\Psi_{k}^{n}\rangle$
4328: form an orthonormal basis in which we express the matrix elements
4329: of the Hamiltonian. We can easily see that\begin{equation}
4330: H_{I}|\Psi_{k}^{n}\rangle=-\left(-1\right)^{n}G|\Psi_{k}^{n}\rangle\end{equation}
4331: and \begin{equation}
4332: H_{S}|\Psi_{k}^{n}\rangle=\frac{\epsilon_{k}}{\sqrt{2}}\sum_{\ell=1}^{N}a_{k\ell}|\ell,0\rangle=\frac{\epsilon_{k}}{2}\left(|\Psi_{k}^{0}\rangle+|\Psi_{k}^{1}\rangle\right).\end{equation}
4333: Therefore the matrix elements of the full Hamiltonian $H=H_{S}+H_{I}$
4334: are given by \begin{equation}
4335: \langle\Psi_{k'}^{n'}|H|\Psi_{k}^{n}\rangle=\delta_{kk'}\left(-\left(-1\right)^{n}G\delta_{nn'}+\frac{\epsilon_{k}}{2}\right).\end{equation}
4336: The Hamiltonian is not diagonal in the states of Eq.~(\ref{eq:ansatz}).
4337: But $H$ is now block diagonal consisting of $N$ blocks of size $2$,
4338: which can be easily diagonalised analytically. The orthonormal eigenstates
4339: of the Hamiltonian are given by\begin{equation}
4340: |E_{k}^{n}\rangle=c_{kn}^{-1}\left\{ \left(\left(-1\right)^{n}\Delta_{k}-2G\right)|\Psi_{k}^{0}\rangle+\epsilon_{k}|\Psi_{k}^{1}\rangle\right\} \label{eq:eigenstates}\end{equation}
4341: with the eigenvalues\begin{equation}
4342: E_{k}^{n}=\frac{1}{2}\left(\epsilon_{k}+\left(-1\right)^{n}\Delta_{k}\right)\end{equation}
4343: and the normalisation\begin{equation}
4344: c_{kn}\equiv\sqrt{\left(\left(-1\right)^{n}\Delta_{k}-2G\right)^{2}+\epsilon_{k}^{2}},\end{equation}
4345: where\begin{equation}
4346: \Delta_{k}=\sqrt{4G^{2}+\epsilon_{k}^{2}}.\label{eq:delta}\end{equation}
4347: Note that the ansatz of Eq.~(\ref{eq:ansatz}) that put $H$ in
4348: block diagonal form did not depend on the details of $H_{S}$ and
4349: $H_{I}^{(\ell)}.$ The methods presented here can be applied to a
4350: much larger class of systems, including the generalised spin star
4351: systems (which include an interaction within the bath) discussed in~\cite{Olaya-Castro2006}.
4352:
4353: After solving the Schrödinger equation, let us now turn to quantum
4354: state transfer. The relevant quantity~\cite{Bose2003,Haselgrove2005}
4355: is given by the transfer function\begin{eqnarray*}
4356: f_{N,1}(t) & \equiv & \langle N,0|\exp\left\{ -iHt\right\} |1,0\rangle\\
4357: & = & \sum_{k,n}\exp\left\{ -iE_{k}^{n}t\right\} \langle E_{k}^{n}|1,0\rangle\langle N,0|E_{k}^{n}\rangle.\end{eqnarray*}
4358: The modulus of $f_{N,1}(t)$ is between $0$ (no transfer) and $1$
4359: (perfect transfer) and fully determines the fidelity of state transfer.
4360: Since\begin{eqnarray*}
4361: \langle\ell,0|E_{k}^{n}\rangle & = & c_{kn}^{-1}\left\{ \left(\left(-1\right)^{n}\Delta_{k}-2G\right)\langle\ell,0|\Psi_{k}^{0}\rangle+\epsilon_{k}\langle\ell,0|\Psi_{k}^{1}\rangle\right\} \\
4362: & = & \frac{c_{kn}^{-1}}{\sqrt{2}}\left(\left(-1\right)^{n}\Delta_{k}-2G+\epsilon_{k}\right)a_{k\ell}\end{eqnarray*}
4363: we get\begin{eqnarray}
4364: \lefteqn{f_{N,1}(t)=}\label{eq:transfer}\\
4365: & & \frac{1}{2}\sum_{k,n}e^{\frac{-it}{2}\left(\epsilon_{k}+\left(-1\right)^{n}\Delta_{k}\right)}\frac{\left(\left(-1\right)^{n}\Delta_{k}-2G+\epsilon_{k}\right)^{2}}{\left(\left(-1\right)^{n}\Delta_{k}-2G\right)^{2}+\epsilon_{k}^{2}}a_{k1}a_{kN}^{*}.\nonumber \end{eqnarray}
4366: Eq.~(\ref{eq:transfer}) is the main result of this section, fully
4367: determining the transfer of quantum information and entanglement in
4368: the presence of the environments. In the limit $G\rightarrow0,$ we
4369: have $\Delta_{k}\approx\epsilon_{k}$ and $f_{N,1}(t)$ approaches
4370: the usual result without an environment,\begin{equation}
4371: f_{N,1}^{0}(t)\equiv\sum_{k}\exp\left\{ -it\epsilon_{k}\right\} a_{k1}a_{kN}^{*}.\end{equation}
4372: In fact, a series expansion of Eq.~(\ref{eq:transfer}) yields that
4373: the first modification of the transfer function is of the order of
4374: $G^{2},$\begin{equation}
4375: G^{2}\sum_{k}a_{k1}a_{kN}^{*}\left[\exp\left\{ -it\epsilon_{k}\right\} \left(-\frac{1}{\epsilon_{k}^{2}}-\frac{it}{\epsilon_{k}}\right)+\frac{1}{\epsilon_{k}^{2}}\right].\end{equation}
4376: Hence the effect is small for very weakly coupled baths. However,
4377: as the chains get longer, the lowest lying energy $\epsilon_{1}$
4378: usually approaches zero, so the changes become more significant (scaling
4379: as $1/\epsilon_{k}$). For intermediate $G,$ we evaluated Eq.~(\ref{eq:transfer})
4380: numerically and found that the first peak of the transfer function
4381: generally becomes slightly lower, and gets shifted to higher times
4382: (Figures \ref{cap:Example} and \ref{cap:Example2}). A numeric search
4383: in the coupling space $\left\{ J_{\ell},\ell=1,\ldots,N-1\right\} $
4384: however also revealed some rare examples where an environment can
4385: also slightly improve the peak of the transfer function (Fig \ref{cap:Example3}).
4386: %
4387: \begin{figure}[tbh]
4388: \begin{centering}\includegraphics[width=0.8\columnwidth]{bc_fig3}\par\end{centering}
4389:
4390:
4391: \caption{\label{cap:Example}The absolute value of the transport function
4392: $f_{N,1}(t)$ of an uniform spin chain (i.e. $J_{\ell}=1$) with length
4393: $N=10$ for three different values of the bath coupling $G.$ The
4394: filled grey curve is the envelope of the limiting function for $G\gg\epsilon_{k}/2$
4395: given by $|f^{0}(\frac{t}{2})|.$ We can see that Eq.~(\ref{eq:scalingformula})
4396: becomes a good approximation already at $G=4.$}
4397: \end{figure}
4398:
4399:
4400: %
4401: \begin{figure}[tbh]
4402: \begin{centering}\includegraphics[width=0.8\columnwidth]{bc_fig4}\par\end{centering}
4403:
4404:
4405: \caption{\label{cap:Example2}The same as Fig. \ref{cap:Example}, but now
4406: for an engineered spin chain {[}i.e. $J_{\ell}=\sqrt{\ell(N-\ell)}$]
4407: as in~Subsection \ref{sub:Engineered-Hamiltonians}. For comparison,
4408: we have rescaled the couplings such that $\sum_{\ell}J_{\ell}$ is
4409: the same as in the uniform coupling case.}
4410: \end{figure}
4411:
4412:
4413: In the strong coupling regime $G\gg\epsilon_{k}/2,$ we can approximate
4414: Eq.~(\ref{eq:delta}) by $\Delta_{k}\approx2G.$ Inserting it in
4415: Eq.~(\ref{eq:transfer}) then becomes\begin{eqnarray}
4416: f_{N,1}(t) & \approx & \frac{1}{2}e^{-iGt}\sum_{k}\exp\left\{ -it\epsilon_{k}\frac{1}{2}\right\} a_{k1}a_{kN}^{*}+\nonumber \\
4417: & & +\frac{1}{2}e^{iGt}\sum_{k}\exp\left\{ -it\epsilon_{k}\frac{1}{2}\right\} a_{k1}a_{kN}^{*}\nonumber \\
4418: & = & \cos(Gt)f_{N,1}^{0}(\frac{t}{2}).\end{eqnarray}
4419: This surprisingly simple result consists of the normal transfer function,
4420: slowed down by a factor of $1/2,$ and modulated by a quickly oscillating
4421: term (Figures \ref{cap:Example} and \ref{cap:Example2}). We call
4422: this effect \emph{destabilisation\index{destabilisation}.} Our derivation
4423: actually did not depend on the indexes of $f(t)$ and we get for the
4424: transfer from the $n$th to the $m$th spin of the chain that \begin{equation}
4425: \boxed{f_{n,m}(t)\approx\cos(Gt)f_{n,m}^{0}(\frac{t}{2}).}\label{eq:scalingformula}\end{equation}
4426: It may look surprising that the matrix $f_{n,m}$ is no longer unitary.
4427: This is because we are considering the dynamics of the chain only,
4428: which is an open quantum system~\cite{OPENQUANTUM}. A heuristic
4429: interpretation of Eq.~(\ref{eq:scalingformula}) is that the excitation
4430: oscillates back and forth between the chain and the bath (hence the
4431: modulation), and spends half of the time trapped in the bath (hence
4432: the slowing). If the time of the maximum of the transfer function
4433: $|f_{n,m}^{0}(t)|$ for $G=0$ is a multiple of $\pi/2G$ then this
4434: maximum is also reached in the presence of the bath. We remark that
4435: this behaviour is strongly non-Markovian\index{non-Markovian}~\cite{OPENQUANTUM}.
4436:
4437: Finally, we want to stress that Eq.~(\ref{eq:scalingformula}) is
4438: \emph{universal} for any spin Hamiltonian that conserves the number
4439: of excitations, i.e. with $\left[H_{S},\sum_{\ell}Z_{\ell}\right]=0$.
4440: Thus our restriction to chain-like topology and exchange couplings
4441: for $H_{S}$ is not necessary. In fact the only difference in the
4442: whole derivation of Eq.~(\ref{eq:scalingformula}) for a more general
4443: Hamiltonian is that Eq.~(\ref{eq:hc_action}) is replaced by\begin{eqnarray}
4444: H_{S}|\ell,0\rangle & = & \sum_{\ell'}h_{\ell'}|\ell',0\rangle.\label{eq:hc_action2}\end{eqnarray}
4445: The Hamiltonian can still be formally diagonalised in the first excitation
4446: sector as in Eq.~(\ref{eq:eigen_h_c}), and the states of Eq.~(\ref{eq:eigenstates})
4447: will still diagonalise the total Hamiltonian $H_{S}+H_{I}.$ Also,
4448: rather than considering an exchange Hamiltonian for the interaction
4449: with the bath, we could have considered a Heisenberg interaction~\cite{Khaetskii2003},
4450: but only for the special case where all bath couplings $g_{k}^{(\ell)}$
4451: are all the same~\cite{Rao2006}. Up to some irrelevant phases, this
4452: leads to the same results as for the exchange interaction.
4453:
4454: %
4455: \begin{figure}[tbh]
4456: \begin{centering}\includegraphics[width=0.8\columnwidth]{bc_fig5}\par\end{centering}
4457:
4458:
4459: \caption{\label{cap:Example3}A weakly coupled bath may even improve the transfer
4460: function for some specific choices of the $J_{\ell}.$ This plot shows
4461: the transfer function $|f_{N,1}(t)|$ for $N=10.$ The couplings $J_{\ell}$
4462: were found numerically. }
4463: \end{figure}
4464:
4465:
4466:
4467: \section{Conclusion}
4468:
4469: We found a surprisingly simple and universal scaling law for the spin
4470: transfer functions in the presence of spin environments. In the context
4471: of quantum state transfer this result is double-edged: on one hand,
4472: it shows that even for very strongly coupled baths quantum state transfer
4473: is possible, with the same fidelity and only reasonable slowing. On
4474: the other hand, it also shows that the fidelity as a function of time
4475: becomes destabilised with a quickly oscillating modulation factor.
4476: In practice, this factor will restrict the time-scale in which one
4477: has to be able to read the state from the system. The results here
4478: are very specific to the simple bath model and do not hold in more
4479: general models (such as these discussed in~\cite{Guo2006,Sun}, where
4480: true decoherence and dissipation takes place). What we intended to
4481: demonstrate is that even though a bath coupling need not introduce
4482: decoherence or dissipation to the system, it can cause other dynamical
4483: processes that can be problematic for quantum information processing.
4484: Because the effects observed here cannot be avoided by cooling the
4485: bath, they may become relevant in some systems as a low temperature
4486: limit.
4487:
4488:
4489: \chapter{Conclusion and outlook}
4490:
4491: Our research on quantum state transfer with spin chains has taken
4492: us on a journey from a very practical motivation to quite fundamental
4493: issues and back again. On one hand, our results are quite abstract
4494: and fundamental, and have related state transfer to number theory,
4495: topology and quantum convergence. On the other hand, we have developed
4496: schemes which are simple and practical, taking into account experimental
4497: hurdles such as disorder and restricted control. While the multi rail
4498: scheme and the memory swapping scheme will probably become useful
4499: only after much further progress in experimental QIT, the dual rail
4500: scheme and in particular the valve scheme have some good chances to
4501: be realised in the near future.
4502:
4503: State transfer with quantum chains has become an area of large interest,
4504: with more than seventy articles on the subject over the last three
4505: years. The most important goal now is an experiment that demonstrates
4506: coherent transfer on a short chain (say of length $N\ge5$). Such
4507: an experiment is not only useful building a quantum computer, but
4508: also from a fundamental perspective. For instance, the violation of
4509: a Bell-inequality between distant entangled solid state qubits would
4510: be a milestone in the field. Since this requires a very high transfer
4511: fidelity, the design of such an experiment would probably require
4512: system dependent theoretical research on how to overcome specific
4513: types of noise and how to improve the fidelity for specific Hamiltonians.
4514:
4515: \listoffigures
4516:
4517:
4518: \listoftables
4519:
4520:
4521:
4522: %\chapter*{List of Theorems}
4523:
4524: %\addcontentsline{toc}{chapter}{List of Theorems}
4525: %\theoremlisttype{allname}
4526: %\listtheorems{theorem,framedtheorem}
4527:
4528: \printindex{}
4529: \begin{thebibliography}{100}
4530: \expandafter\ifx\csname urlstyle\endcsname\relax
4531: \providecommand{\doi}[1]{doi:\discretionary{}{}{}#1}\else
4532: \providecommand{\doi}{doi:\discretionary{}{}{}\begingroup
4533: \urlstyle{rm}\Url}\fi
4534:
4535: \bibitem{Bose2003}
4536: S.~Bose, \emph{Quantum Communication through an Unmodulated Spin Chain}, Phys.
4537: Rev. Lett. \textbf{91}, 207901 (2003).
4538:
4539: \bibitem{NIELSEN}
4540: M.~A. Nielsen and I.~L. Chuang, \emph{Quantum Computation and Quantum
4541: Information}, Cambridge University Press, Cambridge (2000).
4542:
4543: \bibitem{Shor1997}
4544: P.~W. Shor, \emph{Polynomial-Time Algorithms for Prime Factorization and
4545: Discrete Logarithms on a Quantum Computer}, J. Sci. Statist. Comput.
4546: \textbf{26}, 1484 (1997).
4547:
4548: \bibitem{DiVincenzo2000}
4549: D.~P. DiVincenzo, \emph{The Physical Implementation of Quantum Computation},
4550: Fortsch.Phys. \textbf{48}, 9 (2000).
4551:
4552: \bibitem{WERNER}
4553: R.~F. Werner, \emph{Quantum Information Theory - an Invitation}, Springer
4554: Tracts Mod.Phys. \textbf{173}, 14, quant-ph/0101061 (2001).
4555:
4556: \bibitem{Cleve2000}
4557: R.~Cleve and J.~Watrous, \emph{Fast parallel circuits for the quantum Fourier
4558: transform}, Annual Symposium on Foundations of Computer Science \textbf{41},
4559: 526 (2000).
4560:
4561: \bibitem{Josza2006}
4562: R.~Jozsa, \emph{On the simulation of quantum circuits}, quant-ph/0603163
4563: (2006).
4564:
4565: \bibitem{Short2006}
4566: N.~Yoranand and A.~J. Short, \emph{Classical simulation of limited-width
4567: cluster-state quantum computation}, quant-ph/0601178 (2006).
4568:
4569: \bibitem{Haffner2005}
4570: H.~H{\"a}ffner, W.~H{\"a}nsel, C.~F. Roos, J.~Benhelm, D.~Chekalkar,
4571: M.~Chwalla, T.~K{\"o}rber, U.~D. Rapol, M.~Riebe, P.~O. Schmidt, C.~Becher,
4572: O.~G{\"u}hne, W.~D{\"u}r, and R.~Blatt, \emph{Scalable multiparticle
4573: entanglement of trapped ions}, Nature \textbf{438}, 643 (2005).
4574:
4575: \bibitem{Skinner2003}
4576: A.~J. Skinner, M.~E. Davenport, and B.~E. Kane, \emph{Hydrogenic Spin Quantum
4577: Computing in Silicon: A Digital Approach}, Phys. Rev. Lett. \textbf{90},
4578: 087901 (2003).
4579:
4580: \bibitem{Kielpinski2002}
4581: D.~Kielpinski, C.~Monroe, and D.~Wineland, \emph{Architecture for a Large-Scale
4582: Ion-Trap Quantum Computer}, Nature \textbf{417}, 709 (2002).
4583:
4584: \bibitem{LLOYD}
4585: S.~Lloyd, \emph{Power of Entanglement in Quantum Communication}, Phys. Rev.
4586: Lett. \textbf{90}, 167902 (2003).
4587:
4588: \bibitem{KOPPENS}
4589: F.~H.~L. Koppens, J.~A. Folk, J.~M. Elzerman, R.~Hanson, L.~H.~W. van Beveren,
4590: I.~T. Vink, H.~P. Tranitz, W.~Wegscheider, L.~P. Kouwenhoven, and L.~M.~K.
4591: Vandersypen, \emph{Control and Detection of Singlet-Triplet Mixing in a
4592: Random Nuclear Field}, Science \textbf{309}, 1346 (2005).
4593:
4594: \bibitem{HANSON}
4595: R.~Hanson, L.~H.~W. van Beveren, I.~T. Vink, J.~M. Elzerman, W.~J.~M. Naber,
4596: F.~H.~L. Koppens, L.~P. Kouwenhoven, and L.~M.~K. Vandersypen,
4597: \emph{Single-Shot Readout of Electron Spin States in a Quantum Dot Using
4598: Spin-Dependent Tunnel Rates}, Phys. Rev. Lett. \textbf{94}, 196802 (2005).
4599:
4600: \bibitem{YAMA}
4601: T.~Yamamoto, Y.~A. Pashkin, O.~Astafiev, Y.~Nakamura, and J.~S. Tsai,
4602: \emph{Demonstration of conditional gate operation using superconducting
4603: charge qubits}, Nature \textbf{425}, 941 (2003).
4604:
4605: \bibitem{CHIO}
4606: I.~Chiorescu, P.~Bertet, K.~Semba, Y.~Nakamura, C.~J. P.~M. Harmans, and J.~E.
4607: Mooij, \emph{Coherent dynamics of a flux qubit coupled to a harmonic
4608: oscillator}, Nature \textbf{431}, 159 (2004).
4609:
4610: \bibitem{Lakshminarayan2006}
4611: V.~Subrahmanyam and A.~Lakshminarayan, \emph{Transport of entanglement through
4612: a Heisenberg XY spin chain}, Phys. Lett. A \textbf{349}, 164 (2006).
4613:
4614: \bibitem{Subrahmanyam2004}
4615: V.~Subrahmanyam, \emph{Entanglement dynamics and quantum-state transport in
4616: spin chains}, Phys. Rev. A \textbf{69}, 034304 (2004).
4617:
4618: \bibitem{Eisert2004}
4619: M.~B. Plenio, J.~Hartley, and J.~Eisert, \emph{Dynamics and manipulation of
4620: entanglement in coupled harmonic systems with many degrees of freedom}, New.
4621: J. Phys. \textbf{6}, 36 (2004).
4622:
4623: \bibitem{Palma2004}
4624: L.~Amico, A.~Osterloh, F.~Plastina, R.~Fazio, and G.~M. Palma, \emph{Dynamics
4625: of entanglement in one-dimensional spin systems}, Phys. Rev. A \textbf{69},
4626: 022304 (2004).
4627:
4628: \bibitem{Plenio}
4629: M.~J. Hartmann, M.~E. Reuter, and M.~B. Plenio, \emph{Excitation and
4630: Entanglement Transfer Near Quantum Critical Points}, quant-ph/0608051 (2006).
4631:
4632: \bibitem{Plenio2006}
4633: M.~J. Hartmann, M.~E. Reuter, and M.~B. Plenio, \emph{Excitation and
4634: entanglement transfer versus spectral gap}, New. J. Phys. \textbf{8}, 94
4635: (2006).
4636:
4637: \bibitem{Verstraete}
4638: F.~Verstraete, M.~Popp, and J.~I. Cirac, \emph{Entanglement versus Correlations
4639: in Spin Systems}, Phys. Rev. Lett. \textbf{92}, 027901 (2004).
4640:
4641: \bibitem{Verstraete2004}
4642: F.~Verstraete, M.~A. Martin-Delgado, and J.~I. Cirac, \emph{Diverging
4643: Entanglement Length in Gapped Quantum Spin Systems}, Phys. Rev. Lett.
4644: \textbf{92}, 087201 (2004).
4645:
4646: \bibitem{Monteiro2006}
4647: T.~Boness, S.~Bose, and T.~S. Monteiro, \emph{Entanglement and Dynamics of Spin
4648: Chains in Periodically Pulsed Magnetic Fields: Accelerator Modes}, Phys. Rev.
4649: Lett. \textbf{96}, 187201 (2006).
4650:
4651: \bibitem{Twamley2005}
4652: J.~Fitzsimons and J.~Twamley, \emph{Superballistic diffusion of entanglement in
4653: disordered spin chains}, Phys. Rev. A \textbf{72}, 050301 (2005).
4654:
4655: \bibitem{Santos2004}
4656: L.~F. Santos, G.~Rigolin, and C.~O. Escobar, \emph{Entanglement versus chaos in
4657: disordered spin chains}, Phys. Rev. A \textbf{69}, 042304 (2004).
4658:
4659: \bibitem{Boness2006}
4660: T.~Boness, M.~Stocklin, and T.~Monteiro, \emph{Quantum chaos with spin-chains
4661: in pulsed magnetic fields}, quant-ph/0612074 (2006).
4662:
4663: \bibitem{Winter}
4664: J.~P. Keating, N.~Linden, J.~C.~F. Matthews, and A.~Winter, \emph{Localization
4665: and its consequences for quantum walk algorithms and quantum communication},
4666: quant-ph/0606205 (2006).
4667:
4668: \bibitem{Apollaro2006}
4669: T.~J.~G. Apollaro and F.~Plastina, \emph{Entanglement localization by a single
4670: defect in a spin chain}, Phys. Rev. A \textbf{74}, 062316 (2006).
4671:
4672: \bibitem{Bruder2005a}
4673: A.~Romito, R.~Fazio, and C.~Bruder, \emph{Solid-state quantum communication
4674: with Josephson arrays}, Phys. Rev. B \textbf{71}, 100501(R) (2005).
4675:
4676: \bibitem{Tsomokos2006}
4677: D.~Tsomokos, M.~Hartmann, S.~Huelga, and M.~B. Plenio, \emph{Dynamics of
4678: entanglement in realistic chains of superconducting qubits}, quant-ph/0611077
4679: (2006).
4680:
4681: \bibitem{Bruder}
4682: A.~Lyakhov and C.~Bruder, \emph{Use of dynamical coupling for improved quantum
4683: state transfer}, Phys. Rev. B \textbf{74}, 235303 (2006).
4684:
4685: \bibitem{Bruder2005}
4686: A.~Lyakhov and C.~Bruder, \emph{Quantum state transfer in arrays of flux
4687: qubits}, New. J. Phys. \textbf{7}, 181 (2005).
4688:
4689: \bibitem{Loss1998}
4690: D.~Loss and D.~P. DiVincenzo, \emph{Quantum computation with quantum dots},
4691: Phys. Rev. A \textbf{57}, 120 (1998).
4692:
4693: \bibitem{Greentree2004}
4694: A.~D. Greentree, J.~H. Cole, A.~R. Hamilton, and L.~C.~L. Hollenberg,
4695: \emph{Coherent electronic transfer in quantum dot systems using adiabatic
4696: passage}, Phys. Rev. B \textbf{70}, 235317 (2004).
4697:
4698: \bibitem{DAmico}
4699: I.~D'Amico, \emph{Quantum buses and quantum computer architecture based on
4700: quantum dots}, cond-mat/0511470 (2005).
4701:
4702: \bibitem{Lovett}
4703: T.~P. Spiller, I.~D'Amico, and B.~W. Lovett, \emph{Entanglement distribution
4704: for a practical quantum-dot-based quantum processor architecture},
4705: quant-ph/0601124 (2006).
4706:
4707: \bibitem{Suter2006}
4708: J.~Zhang, X.~Peng, and D.~Suter, \emph{Speedup of quantum-state transfer by
4709: three-qubit interactions: Implementation by nuclear magnetic resonance},
4710: Phys. Rev. A \textbf{73}, 062325 (2006).
4711:
4712: \bibitem{Jones}
4713: J.~Fitzsimons, L.~Xiao, S.~C. Benjamin, and J.~A. Jones, \emph{Quantum
4714: Information Processing with Delocalized Qubits under Global Control},
4715: quant-ph/0606188 (2006).
4716:
4717: \bibitem{Zhang2005}
4718: J.~Zhang, G.~L. Long, W.~Zhang, Z.~Deng, W.~Liu, , and Z.~Lu, \emph{Simulation
4719: of Heisenberg XY interactions and realization of a perfect state transfer in
4720: spin chains using liquid nuclear magnetic resonance}, Phys. Rev. A
4721: \textbf{72}, 012331 (2005).
4722:
4723: \bibitem{Garcia-Ripoll2003}
4724: J.~J. Garcia-Ripoll and J.~I. Cirac, \emph{Spin dynamics for bosons in an
4725: optical lattice}, New. J. Phys. \textbf{5}, 76 (2003).
4726:
4727: \bibitem{Falci2005}
4728: M.~Paternostro, G.~M. Palma, M.~S. Kim, and G.~Falci, \emph{Quantum-state
4729: transfer in imperfect artificial spin networks}, Phys. Rev. A \textbf{71},
4730: 042311 (2005).
4731:
4732: \bibitem{Paternostro2005}
4733: M.~Paternostro, \emph{Theoretical proposals on efficient schemes for quantum
4734: information processing}, Ph.D. thesis, The Queen's University of Belfast
4735: (2005).
4736:
4737: \bibitem{Bose}
4738: D.~G. Angelakis, M.~F. Santos, and S.~Bose, \emph{Photon blockade induced Mott
4739: transitions and XY spin models in coupled cavity arrays}, quant-ph/0606159
4740: (2006).
4741:
4742: \bibitem{Plenioa}
4743: M.~J. Hartmann, F.~G. S.~L. Brandao, and M.~B. Plenio, \emph{Strongly
4744: Interacting Polaritons in Coupled Arrays of Cavities}, Nature Physics
4745: \textbf{2}, 849 (2006).
4746:
4747: \bibitem{EXCITON}
4748: V.~May and O.~K{\"u}hn, \emph{Charge and Energy Transfer Dynamics in Molecular
4749: Systems}, Wiley-Interscience, Hoboken (2004).
4750:
4751: \bibitem{Motoyama1996}
4752: N.~Motoyama, H.~Eisaki, and S.~Uchida, \emph{Magnetic Susceptibility of Ideal
4753: Spin 1 /2 Heisenberg Antiferromagnetic Chain Systems, Sr2CuO3 and SrCuO2},
4754: Phys. Rev. Lett. \textbf{76}, 3212 (1996).
4755:
4756: \bibitem{Gambardella2002}
4757: P.~Gambardella, A.~Dallmeyer, K.~Maiti, M.~C. Malagoli, W.~Eberdardt, K.~Kern,
4758: and C.~Carbone, \emph{Ferromagnetism in one-dimensional monatomic metal
4759: chains}, Nature \textbf{416}, 301 (2002).
4760:
4761: \bibitem{Uhlmann1976}
4762: A.~Uhlmann, \emph{The 'transition probability' on the state space of a
4763: *-algebra}, Rep. Math. Phys. \textbf{9}, 273 (1976).
4764:
4765: \bibitem{Jozsa1994}
4766: R.~Jozsa, \emph{Fidelity for mixed quantum states}, J. Mod. Opt. \textbf{41},
4767: 2315 (1994).
4768:
4769: \bibitem{Werner2005}
4770: D.~Kretschmann and R.~F. Werner, \emph{Quantum channels with memory}, Phys.
4771: Rev. A \textbf{72}, 062323 (2005).
4772:
4773: \bibitem{Kay}
4774: A.~Kay, \emph{Unifying Quantum State Transfer and State Amplification}, Phys.
4775: Rev. Lett. \textbf{98}, 010501 (2007).
4776:
4777: \bibitem{Giovannetti2005}
4778: V.~Giovannetti and R.~Fazio, \emph{Information-capacity description of
4779: spin-chain correlations}, Phys. Rev. A \textbf{71}, 032314 (2005).
4780:
4781: \bibitem{Eberly1977}
4782: J.~H. Eberly, B.~W. Shore, Z.~Bialynicka-Birula, and I.~Bialynicki-Birula,
4783: \emph{Coherent dynamics of N-level atoms and molecules. I. Numerical
4784: experiments*}, Phys. Rev. A \textbf{16}, 2038 (1977).
4785:
4786: \bibitem{Bialynicka-Birula1977}
4787: Z.~Bialynicka-Birula, I.~Bialynicki-Birula, J.~H. Eberly, and B.~W. Shore,
4788: \emph{Coherent dynamics of N-level atoms and molecules. II. Analytic
4789: solutions*}, Phys. Rev. A \textbf{16}, 2048 (1977).
4790:
4791: \bibitem{Cook1979}
4792: R.~Cook and B.~W. Shore, \emph{Coherent dynamics of N-level atoms and
4793: molecules. III. An analytically soluble periodic case}, Phys. Rev. A
4794: \textbf{20}, 539 (1979).
4795:
4796: \bibitem{Shore1979}
4797: B.~W. Shore and R.~J. Cook, \emph{Coherent dynamics of N-level atoms and
4798: molecules. IV. Two- and three-level behavior}, Phys. Rev. A \textbf{20}, 1958
4799: (1979).
4800:
4801: \bibitem{Ekert2004}
4802: C.~Albanese, M.~Christandl, N.~Datta, and A.~Ekert, \emph{Mirror Inversion of
4803: Quantum States in Linear Registers}, Phys. Rev. Lett. \textbf{93}, 230502
4804: (2004).
4805:
4806: \bibitem{Linden2004}
4807: T.~J. Osborne and N.~Linden, \emph{Propagation of quantum information through a
4808: spin system}, Phys. Rev. A \textbf{69}, 052315 (2004).
4809:
4810: \bibitem{Abramowitz1972}
4811: M.~Abramowitz and I.~A. Stegun, \emph{Handbook of Mathematical Functions},
4812: Dover, New York (1972).
4813:
4814: \bibitem{Lang1984}
4815: S.~Lang, \emph{Algebra}, Addison-Wesley, Boston (1984).
4816:
4817: \bibitem{Hemmer1958}
4818: P.~C. Hemmer, L.~C. Maximon, and H.~Wergeland, \emph{Recurrence Time of a
4819: Dynamical System}, Phys. Rev. \textbf{111}, 689 (1958).
4820:
4821: \bibitem{Bose2005}
4822: M.-H. Yung and S.~Bose, \emph{Perfect state transfer, effective gates, and
4823: entanglement generation in engineered bosonic and fermionic networks}, Phys.
4824: Rev. A \textbf{71}, 032310 (2005).
4825:
4826: \bibitem{Landahl2005}
4827: M.~Christandl, N.~Datta, T.~C. Dorlas, A.~Ekert, A.~Kay, and A.~J. Landahl,
4828: \emph{Perfect transfer of arbitrary states in quantum spin networks}, Phys.
4829: Rev. A \textbf{71}, 032312 (2005).
4830:
4831: \bibitem{Bravyi2006}
4832: S.~Bravyi, M.~B. Hastings, and F.~Verstraete, \emph{Lieb-Robinson bounds and
4833: the generation of correlations and topological quantum order}, Phys. Rev.
4834: Lett. \textbf{97}, 050401 (2006).
4835:
4836: \bibitem{Nachtergaele2006}
4837: B.~Nachtergaele and R.~Sims, \emph{Lieb-Robinson Bounds and the Exponential
4838: Clustering Theorem}, Commun. Math. Phys. \textbf{265}, 119 (2006).
4839:
4840: \bibitem{Horodecki1997}
4841: M.~Horodecki, P.~Horodecki, and R.~Horodecki, \emph{Inseparable Two Spin- 1/2
4842: Density Matrices Can Be Distilled to a Singlet Form}, Phys. Rev. Lett.
4843: \textbf{78}, 574 (1997).
4844:
4845: \bibitem{Horodecki1999}
4846: M.~Horodecki, P.~Horodecki, and .~. R.~Horodecki, Phys. Rev. A~60,
4847: \emph{General teleportation channel, singlet fraction, and
4848: quasidistillation}, Phys. Rev. A \textbf{60}, 1888 (1999).
4849:
4850: \bibitem{SHOR}
4851: C.~H. Bennett and P.~W. Shor, \emph{Quantum information theory}, IEEE Trans.
4852: Inf. Theory \textbf{44}, 2724 (1998).
4853:
4854: \bibitem{Landahl2004}
4855: M.~Christandl, N.~Datta, A.~Ekert, and A.~J. Landahl, \emph{Perfect State
4856: Transfer in Quantum Spin Networks}, Phys. Rev. Lett. \textbf{92}, 187902
4857: (2004).
4858:
4859: \bibitem{Lambropoulos2004a}
4860: G.~M. Nikolopoulos, D.~Petrosyan, and P.~Lambropoulos, \emph{Coherent electron
4861: wavepacket propagation and entanglement in array of coupled quantum dots},
4862: Europhys. Lett. \textbf{65}, 297 (2004).
4863:
4864: \bibitem{Peres1985}
4865: A.~Peres, \emph{Reversible logic and quantum computers}, Phys. Rev. A
4866: \textbf{32}, 3266 (1985).
4867:
4868: \bibitem{Yung2006}
4869: M.-H. Yung, \emph{Quantum speed limit for perfect state transfer in one
4870: dimension}, Phys. Rev. A \textbf{74}, 030303(R) (2006).
4871:
4872: \bibitem{Sun2006a}
4873: S.~Yang, Z.~Song, and C.~P. Sun, \emph{Quantum state swapping via a qubit
4874: network with Hubbard interactions}, Phys. Rev. B \textbf{73}, 195122 (2006).
4875:
4876: \bibitem{Sun2005c}
4877: Z.~Song and C.~P. Sun, \emph{Quantum information storage and state transfer
4878: based on spin systems}, Low Temp.Phys. \textbf{31}, 907 (2005).
4879:
4880: \bibitem{Sun2005a}
4881: T.~Shi, Y.~Li, Z.~Song, and C.~P. Sun, \emph{Quantum-state transfer via the
4882: ferromagnetic chain in a spatially modulated field}, Phys. Rev. A
4883: \textbf{71}, 032309 (2005).
4884:
4885: \bibitem{Lambropoulos2006}
4886: D.~Petrosyan and P.~Lambropoulos, \emph{Coherent population transfer in a chain
4887: of tunnel coupled quantum dots}, Opt.Commun. \textbf{264}, 419 (2006).
4888:
4889: \bibitem{Lambropoulos2004}
4890: G.~M. Nikolopoulos, D.~Petrosyan, and P.~Lambropoulos, \emph{Electron
4891: wavepacket propagation in a chain of coupled quantum dots}, J. Phys. C:
4892: Condens. Matter \textbf{16}, 4991 (2004).
4893:
4894: \bibitem{Ericsson2005}
4895: A.~Kay and M.~Ericsson, \emph{Geometric effects and computation in spin
4896: networks}, New. J. Phys. \textbf{7}, 143 (2005).
4897:
4898: \bibitem{Kay2006}
4899: A.~Kay, \emph{Perfect state transfer: Beyond nearest-neighbor couplings}, Phys.
4900: Rev. A \textbf{73}, 032306 (2006).
4901:
4902: \bibitem{Kay2006a}
4903: A.~Kay, \emph{Quantum Computation with Minimal Control}, Ph.D. thesis,
4904: University of Cambridge, http://cam.qubit.org/users/Alastair/thesis.pdf
4905: (2006).
4906:
4907: \bibitem{Stolze2005}
4908: P.~Karbach and J.~Stolze, \emph{Spin chains as perfect quantum state mirrors},
4909: Phys. Rev. A \textbf{72}, 030301(R) (2005).
4910:
4911: \bibitem{Jing-Fu2006}
4912: L.~Dan and Z.~Jing-Fu, \emph{Effect of disturbance in perfect state transfer},
4913: Chin.Phys. \textbf{15}, 272 (2006).
4914:
4915: \bibitem{Semiao2005}
4916: M.~B. Plenio and F.~L. Semi{\~a}o, \emph{High efficiency transfer of quantum
4917: information and multiparticle entanglement generation in
4918: translation-invariant quantum chains}, New. J. Phys. \textbf{7}, 73 (2005).
4919:
4920: \bibitem{Bednarska}
4921: A.~W{\'o}jcik, T.~{\L}uczak, P.~Kurzy{\'n}ski, A.~Grudka, T.~Gdala, and
4922: M.~Bednarska, \emph{Multiuser quantum communication networks},
4923: quant-ph/0608107 (2006).
4924:
4925: \bibitem{Bednarska2005}
4926: A.~W{\'o}jcik, T.~{\L}uczak, P.~Kurzy{\'n}ski, A.~Grudka, T.~Gdala, and
4927: M.~Bednarska, \emph{Unmodulated spin chains as universal quantum wires},
4928: Phys. Rev. A \textbf{72}, 034303 (2005).
4929:
4930: \bibitem{Sun2005}
4931: Y.~Li, T.~Shi, B.~Chen, Z.~Song, and C.~P. Sun, \emph{Quantum-state
4932: transmission via a spin ladder as a robust data bus}, Phys. Rev. A
4933: \textbf{71}, 022301 (2005).
4934:
4935: \bibitem{Bose2006}
4936: M.~Avellino, A.~J. Fisher, and S.~Bose, \emph{Quantum communication in spin
4937: systems with long-range interactions}, Phys. Rev. A \textbf{74}, 012321
4938: (2006).
4939:
4940: \bibitem{Bose2006a}
4941: M.~Avellino, A.~J. Fisher, and S.~Bose, \emph{Erratum: Quantum communication in
4942: spin systems with long-range interactions [Phys. Rev. A 74, 012321 (2006)]},
4943: Phys. Rev. A \textbf{74}, 039901 (2006).
4944:
4945: \bibitem{Pasquale2006}
4946: S.~Paganelli, F.~de~Pasquale, and G.~L. Giorgi, \emph{Faithful state transfer
4947: through a quantum channel}, Phys. Rev. A \textbf{74}, 012316 (2006).
4948:
4949: \bibitem{Haselgrove2005}
4950: H.~L. Haselgrove, \emph{Optimal state encoding for quantum walks and quantum
4951: communication over spin systems}, Phys. Rev. A \textbf{72}, 062326 (2005).
4952:
4953: \bibitem{Fitzsimons2006}
4954: J.~Fitzsimons and J.~Twamley, \emph{Globally Controlled Quantum Wires for
4955: Perfect Qubit Transport, Mirroring, and Computing}, Phys. Rev. Lett.
4956: \textbf{97}, 090502 (2006).
4957:
4958: \bibitem{Raussendorf2005}
4959: R.~Raussendorf, \emph{Quantum computation via translation-invariant operations
4960: on a chain of qubits}, Phys. Rev. A \textbf{72}, 052301 (2005).
4961:
4962: \bibitem{Sun2006}
4963: S.~Yang, Z.~Song, and C.~P. Sun, \emph{Wave-packet transmission of Bloch
4964: electrons manipulated by magnetic field}, Phys. Rev. A \textbf{73}, 022317
4965: (2006).
4966:
4967: \bibitem{Maruyama}
4968: K.~Maruyama, T.~Iitaka, and F.~Nori, \emph{Enhancement of entanglement transfer
4969: in a spin chain by phase shift control}, quant-ph/0610103 (2006).
4970:
4971: \bibitem{Korepin2005}
4972: S.~Bose, B.~Jin, and V.~E. Korepin, \emph{Quantum communication through a spin
4973: ring with twisted boundary conditions}, Phys. Rev. A \textbf{72}, 022345
4974: (2005).
4975:
4976: \bibitem{Yang2006}
4977: S.~Yang, Z.~Song, and C.~Sun, \emph{Quantum dynamics of magnetically controlled
4978: network for Bloch electrons}, quant-ph/0602209 (2006).
4979:
4980: \bibitem{Fazio2005}
4981: G.~D. Chiara, D.~Rossini, S.~Montangero, and R.~Fazio, \emph{From perfect to
4982: fractal transmission in spin chains}, Phys. Rev. A \textbf{72}, 012323
4983: (2005).
4984:
4985: \bibitem{Fazio}
4986: D.~Rossini, V.~Giovannetti, and R.~Fazio, \emph{Information transfer rates in
4987: spin quantum channels}, quant-ph/0609022 (2006).
4988:
4989: \bibitem{Bose2006b}
4990: D.~Burgarth and S.~Bose, \emph{Universal destabilization and slowing of
4991: spin-transfer functions by a bath of spins}, Phys. Rev. A \textbf{73}, 062321
4992: (2006).
4993:
4994: \bibitem{DUALRAIL}
4995: D.~Burgarth and S.~Bose, \emph{Conclusive and arbitrarily perfect quantum-state
4996: transfer using parallel spin-chain channels}, Phys. Rev. A \textbf{71},
4997: 052315 (2005).
4998:
4999: \bibitem{RANDOMRAIL}
5000: D.~Burgarth and S.~Bose, \emph{Perfect quantum state transfer with randomly
5001: coupled quantum chains}, New. J. Phys. \textbf{7}, 135 (2005).
5002:
5003: \bibitem{Giovannetti2006}
5004: D.~Burgarth, S.~Bose, and V.~Giovannetti, \emph{Efficient and perfect state
5005: transfer in quantum chains}, Int. J. Quant. Inf. \textbf{4}, 405 (2006).
5006:
5007: \bibitem{Giovannetti}
5008: D.~Burgarth and V.~Giovannetti, \emph{The Generalized Lyapunov Theorem and its
5009: Application to Quantum Channels}, quant-ph/0605197 (2006).
5010:
5011: \bibitem{Bose2006c}
5012: D.~Burgarth, V.~Giovannetti, and S.~Bose, \emph{State Transfer in Permanently
5013: Coupled Quantum Chains}, NATO Sci. Ser. III \textbf{199}, 218 (2006).
5014:
5015: \bibitem{Burgarth}
5016: D.~Burgarth, V.~Giovannetti, and S.~Bose, \emph{Optimal quantum chain
5017: communication by end gates}, quant-ph/0610018 (2006).
5018:
5019: \bibitem{MULTIRAIL}
5020: D.~Burgarth, V.~Giovannetti, and S.~Bose, \emph{Efficient and perfect state
5021: transfer in quantum chains}, J. Phys. A: Math. Gen. \textbf{38}, 6793 (2005).
5022:
5023: \bibitem{MEMORYSWAP}
5024: V.~Giovannetti and D.~Burgarth, \emph{Improved Transfer of Quantum Information
5025: Using a Local Memory}, Phys. Rev. Lett. \textbf{96}, 030501 (2006).
5026:
5027: \bibitem{Bose2005e}
5028: B.~Vaucher, D.~Burgarth, and S.~Bose, \emph{Arbitrarily perfect quantum
5029: communication using unmodulated spin chains, a collaborative approach}, J.
5030: Opt. B: Quantum Semiclass. Opt. \textbf{7}, S356 (2005).
5031:
5032: \bibitem{Raussendorf2001}
5033: R.~Raussendorf and H.~J. Briegel, \emph{A One-Way Quantum Computer}, Phys. Rev.
5034: Lett. \textbf{86}, 5188 (2001).
5035:
5036: \bibitem{Chuang1996}
5037: I.~L. Chuang and Y.~Yamamoto, \emph{Quantum Bit Regeneration}, Phys. Rev. Lett.
5038: \textbf{76}, 4281 (1996).
5039:
5040: \bibitem{Bennett1997}
5041: C.~H. Bennett, D.~P. DiVincenzo, and J.~A. Smolin, \emph{Capacities of Quantum
5042: Erasure Channels}, Phys. Rev. Lett. \textbf{78}, 3217 (1997).
5043:
5044: \bibitem{Wan}
5045: J.~He, Q.~Chen, L.~Ding, and S.~Wan, \emph{Conclusive quantum-state transfer
5046: with a single randomly coupled spin chain}, quant-ph/0606088 (2006).
5047:
5048: \bibitem{Vion2002}
5049: D.~Vion, A.~Aassime, A.~Cottet, P.~Joyez, H.~Pothier, C.~Urbina, D.~Esteve, and
5050: M.~H. Devoret, \emph{Manipulating the Quantum State of an Electrical
5051: Circuit}, Science \textbf{296}, 886 (2002).
5052:
5053: \bibitem{Chiorescu2003}
5054: I.~Chiorescu, Y.~Nakamura, C.~J. P.~M. Harmans, and J.~E. Mooij, \emph{Coherent
5055: Quantum Dynamics of a Superconducting Flux Qubit}, Science \textbf{299}, 1869
5056: (2003).
5057:
5058: \bibitem{Palma1996}
5059: G.~M. Palma, K.~A. Suominen, and A.~K. Ekert, \emph{Quantum Computers and
5060: Dissipation}, Proc. Roy. Soc. Lond. A \textbf{452}, 567 (1996).
5061:
5062: \bibitem{Hwang2000}
5063: W.~Y. Hwang, H.~Lee, D.~D. Ahn, and S.~W. Hwang, \emph{Efficient schemes for
5064: reducing imperfect collective decoherences}, Phys. Rev. A \textbf{62}, 062305
5065: (2000).
5066:
5067: \bibitem{Beige2000}
5068: A.~Beige, D.~Braun, and P.~Knight, \emph{Driving atoms into decoherence-free
5069: states}, New. J. Phys. \textbf{2}, 22 (2000).
5070:
5071: \bibitem{Plenio1998}
5072: M.~Plenio and P.~Knight, \emph{The quantum-jump approach to dissipative
5073: dynamics in quantum optics}, Rev. Mod. Phys. \textbf{70}, 101 (1998).
5074:
5075: \bibitem{OPENQUANTUM}
5076: H.-P. Breuer and F.~Petruccione, \emph{The Theory Of Open Quantum Systems},
5077: Oxford University Press, Oxford (2002).
5078:
5079: \bibitem{Anderson1958}
5080: P.~W. Anderson, \emph{Absence of Diffusion in Certain Random Lattices}, Phys.
5081: Rev. \textbf{109}, 1492 (1958).
5082:
5083: \bibitem{Sutherland1975}
5084: B.~Sutherland, \emph{Model for a multicomponent quantum system}, Phys. Rev. B
5085: \textbf{12}, 3795 (1975).
5086:
5087: \bibitem{Bose2005a}
5088: C.~Hadley, A.~Serafini, and S.~Bose, \emph{Entanglement creation and
5089: distribution on a graph of exchange-coupled qutrits}, Phys. Rev. A
5090: \textbf{72}, 052333 (2005).
5091:
5092: \bibitem{Bayat2006}
5093: A.~Bayat and V.~Karimipour, \emph{Transfer of d-Level quantum states through
5094: spin chains by random swapping}, quant-ph/0612144 (2006).
5095:
5096: \bibitem{Gisin2005}
5097: N.~Gisin, N.~Linden, S.~Massar, and S.~Popescu, \emph{Error filtration and
5098: entanglement purification for quantum communication}, Phys. Rev. A
5099: \textbf{72}, 012338 (2005).
5100:
5101: \bibitem{Vaucher2005}
5102: B.~Vaucher, \emph{Quantum Communication of Spin-Qubits using a Collaborative
5103: Approach}, Master's thesis, Ecole Polytechnique Federale de Lausanne (2005).
5104:
5105: \bibitem{Schott1996}
5106: J.~R. Schott, \emph{Matrix Analysis for Statistics}, Wiley-Interscience,
5107: Hoboken (1996).
5108:
5109: \bibitem{Viola1998}
5110: L.~Viola and S.~Lloyd, \emph{Dynamical suppression of decoherence in two-state
5111: quantum systems}, Phys. Rev. A \textbf{58}, 2733 (1998).
5112:
5113: \bibitem{VIOLA2}
5114: L.~Viola, E.~Knill, and S.~Lloyd, \emph{Dynamical Decoupling of Open Quantum
5115: Systems}, Phys. Rev. Lett. \textbf{82}, 2417 (1999).
5116:
5117: \bibitem{VIOLA3}
5118: L.~Viola, S.~Lloyd, and E.~Knill, \emph{Universal Control of Decoupled Quantum
5119: Systems}, Phys. Rev. Lett. \textbf{83}, 4888 (1999).
5120:
5121: \bibitem{VITALI}
5122: D.~Vitali and P.~Tombesi, \emph{Using parity kicks for decoherence control},
5123: Phys. Rev. A \textbf{59}, 4178 (1999).
5124:
5125: \bibitem{SIMON}
5126: J.~J.~L. Morton, A.~M. Tyryshkin, A.~Ardavan, S.~C. Benjamin, K.~Porfyrakis,
5127: S.~A. Lyon, and G.~A.~D. Briggs, \emph{Bang-bang control of fullerene qubits
5128: using ultra-fast phase gates}, Nature Physics \textbf{2}, 40 (2006).
5129:
5130: \bibitem{FRANSON}
5131: J.~D. Franson, B.~C. Jacobs, and T.~B. Pittman, \emph{Quantum computing using
5132: single photons and the Zeno effect}, Phys. Rev. A \textbf{70}, 062302 (2004).
5133:
5134: \bibitem{PERES}
5135: A.~Peres, \emph{Quantum Theory: Concepts and Methods}, Kluwer Academic
5136: Publishers, Dordrecht (1993).
5137:
5138: \bibitem{KUMMERER}
5139: B.~K{\"u}mmerer and H.~Maassen, \emph{A scattering theory for Markov Chains},
5140: Inf. Dim. Anal. Quant. Probab. Rel. Top. \textbf{3}, 161 (2000).
5141:
5142: \bibitem{WELLENS}
5143: T.~Wellens, A.~Buchleitner, B.~K{\"u}mmerer, and H.~Maassen, \emph{Quantum
5144: State Preparation via Asymptotic Completenes}, Phys. Rev. Lett. \textbf{85},
5145: 3361 (2000).
5146:
5147: \bibitem{HOMOGENIZATION1}
5148: H.~V. Scarani, M.~Ziman, P.~\v{S}telmachovi\v{c}, N.~Gisin, and V.~Bu\v{z}ek,
5149: \emph{Thermalizing Quantum Machines: Dissipation and Entanglement}, Phys.
5150: Rev. Lett. \textbf{88}, 097905 (2002).
5151:
5152: \bibitem{HOMOGENIZATION2}
5153: M.~Ziman, P.~\v{S}telmachovi\v{c}, V.~Bu\v{z}ek, M.~Hillery, V.~Scarani, and
5154: N.~Gisin, \emph{Diluting quantum information: An analysis of information
5155: transfer in system-reservoir interactions}, Phys. Rev. A \textbf{65}, 042105
5156: (2002).
5157:
5158: \bibitem{TERHAL}
5159: B.~M. Terhal and D.~P. DiVincenzo, \emph{Problem of equilibration and the
5160: computation of correlation functions on a quantum computer}, Phys. Rev. A
5161: \textbf{61}, 022301 (2000).
5162:
5163: \bibitem{YUASA1}
5164: H.~Nakazato, T.~Takazawa, and K.~Yuasa, \emph{Purification through Zeno-Like
5165: Measurements}, Phys. Rev. Lett. \textbf{90}, 060401 (2003).
5166:
5167: \bibitem{YUASA2}
5168: H.~Nakazato, M.~Unoki, and K.~Yuasa, \emph{Preparation and entanglement
5169: purification of qubits through Zeno-like measurements}, Phys. Rev. A
5170: \textbf{70}, 012303 (2004).
5171:
5172: \bibitem{RAGINSKY}
5173: M.~Raginsky, \emph{Dynamical Aspects of Information Storage in
5174: Quantum-Mechanical Systems}, Ph.D. thesis, Department of Electrical and
5175: Computer Engineering, Northwestern University (2002).
5176:
5177: \bibitem{RICHTER}
5178: S.~Richter and R.~F. Werner, \emph{Ergodicity of quantum cellular automata}, J.
5179: Stat. Phys. \textbf{82}, 963 (1996).
5180:
5181: \bibitem{Raginsky2002}
5182: M.~Raginsky, \emph{Entropy production rates of bistochastic strictly
5183: contractive quantum channels on a matrix algebra}, J. Phys. A: Math. Gen.
5184: \textbf{35}, L585 (2002).
5185:
5186: \bibitem{STRICTCONTRATIONS}
5187: M.~Raginsky, \emph{Strictly contractive quantum channels and physically
5188: realizable quantum computers}, Phys. Rev. A \textbf{65}, 032306 (2002).
5189:
5190: \bibitem{STREATER}
5191: R.~F. Streater, \emph{Statistical Dynamics}, Imperial College Press, London
5192: (1995).
5193:
5194: \bibitem{ALICKI}
5195: R.~Alicki, \emph{Irreversible Quantum Dynamics}, volume 622 of \emph{Lect.Notes
5196: Phys.}, Springer, New York (2003).
5197:
5198: \bibitem{TOPOLOGYBOOK}
5199: K.~J{\"a}nich, \emph{Topology}, Springer, New York (1984).
5200:
5201: \bibitem{RUSKAI}
5202: M.~B. Ruskai, \emph{Beyond Strong Subadditivity? Improved Bounds On The
5203: Contractions Of Generalized Relative Entropy}, Rev. Math. Phys. \textbf{6},
5204: 1147 (1994).
5205:
5206: \bibitem{STAKGOLD}
5207: I.~Stakgold, \emph{Greens Functions and Boundary Value Problems},
5208: Wiley-Interscience, Hoboken (1998).
5209:
5210: \bibitem{DUGUNDJI}
5211: J.~Dugundji and A.~Granas, \emph{Fixed Point Theory}, Springer, New York
5212: (2003).
5213:
5214: \bibitem{BIRKHOFF}
5215: G.~D. Birkhoff, \emph{Proof of the Ergodic Theorem}, Proc. Natl. Acad. Sci.
5216: U.S.A. \textbf{17}, 656 (1931).
5217:
5218: \bibitem{KRAUS}
5219: K.~Kraus, \emph{States, Effects, and Operations}, Springer, New York (1983).
5220:
5221: \bibitem{KEYL}
5222: M.~Keyl, \emph{Fundamentals of quantum information theory}, Phys. Rept.
5223: \textbf{369}, 431 (2002).
5224:
5225: \bibitem{STINE}
5226: W.~F. Stinespring, \emph{Positive Functions on C* -Algebras}, Proc. Am. Math.
5227: Soc. \textbf{6}, 211 (1955).
5228:
5229: \bibitem{JENS}
5230: K.~M.~R. Audenaert and J.~Eisert, \emph{Continuity bounds on the quantum
5231: relative entropy}, J. Math. Phys. \textbf{46}, 102104 (2005).
5232:
5233: \bibitem{RUSKAI2}
5234: M.~B. Ruskai, \emph{Inequalities for quantum entropy: A review with conditions
5235: for equality}, J. Math. Phys. \textbf{43}, 4358 (2002).
5236:
5237: \bibitem{RUSKAI2b}
5238: M.~B. Ruskai, \emph{Erratum: Inequalities for quantum entropy: A review with
5239: conditions for equality [J. Math. Phys. 43, 4358 (2002)]}, J. Math. Phys.
5240: \textbf{46}, 019901 (E) (2005).
5241:
5242: \bibitem{HORNJOHNSON}
5243: R.~A. Horn and C.~R. Johnson, \emph{Matrix Analysis}, Cambridge University
5244: Press, Cambridge (1990).
5245:
5246: \bibitem{SCHRADER}
5247: R.~Schrader, \emph{Perron-Frobenius Theory for Positive Maps on Trace Ideals},
5248: volume~30 of \emph{Fields Inst.Commun.}, American Mathematical Society,
5249: Providence, math-ph/0007020 (2001).
5250:
5251: \bibitem{Hayden2004}
5252: P.~Hayden, D.~Leung, P.~W. Shor, and A.~Winter, \emph{Randomizing Quantum
5253: States: Constructions and Applications}, Commun. Math. Phys. \textbf{250},
5254: 371 (2004).
5255:
5256: \bibitem{Wellens2002}
5257: T.~Wellens, \emph{Entanglement and control of quantum states}, Ph.D. thesis,
5258: Ludwigs-Maximilians-Universit{\"a}t M{\"u}nchen (2002).
5259:
5260: \bibitem{Lloyd2004}
5261: S.~Lloyd, A.~J. Landahl, and J.-J.~E. Slotine, \emph{Universal quantum
5262: interfaces}, Phys. Rev. A \textbf{69}, 012305 (2004).
5263:
5264: \bibitem{Guo2006}
5265: J.~Cai, Z.~Zhou, and G.~Guo, \emph{Decoherence effects on the quantum spin
5266: channels}, Phys. Rev. A \textbf{74}, 022328 (2006).
5267:
5268: \bibitem{Karimipour2005}
5269: A.~Bayat and V.~Karimipour, \emph{Thermal effects on quantum communication
5270: through spin chains}, Phys. Rev. A \textbf{71}, 042330 (2005).
5271:
5272: \bibitem{Sun}
5273: L.~Zhou, J.~Lu, T.~Shi, and C.~P. Sun, \emph{Decoherence problem in quantum
5274: state transfer via an engineered spin chain}, quant-ph/0608135 (2006).
5275:
5276: \bibitem{Schmidt2005}
5277: H.~Schmidt and G.~Mahler, \emph{Control of local relaxation behavior in closed
5278: bipartite quantum systems}, Phys. Rev. E \textbf{72}, 016117 (2005).
5279:
5280: \bibitem{Breuer2004}
5281: H.-P. Breuer, D.~Burgarth, and F.~Petruccione, \emph{Non-Markovian dynamics in
5282: a spin star system: Exact solution and approximation techniques}, Phys. Rev.
5283: B \textbf{70}, 045323 (2004).
5284:
5285: \bibitem{Blumen2006}
5286: O.~M{\"u}lken and A.~Blumen, \emph{Continuous-time quantum walks in phase
5287: space}, Phys. Rev. A \textbf{73}, 012105 (2006).
5288:
5289: \bibitem{Buzek2005}
5290: M.~Koniorczyk, P.~Rap{\v c}an, and V.~Bu{\v z}ek, \emph{Direct versus
5291: measurement-assisted bipartite entanglement in multiqubit systems and their
5292: dynamical generation in spin systems}, Phys. Rev. A \textbf{72}, 022321
5293: (2005).
5294:
5295: \bibitem{Chiara2005}
5296: G.~D. Chiara, R.~Fazio, C.~Macchiavello, S.~Montangero, and G.~M. Palma,
5297: \emph{Cloning transformations in spin networks without external control},
5298: Phys. Rev. A \textbf{72}, 012328 (2005).
5299:
5300: \bibitem{Hutton2004}
5301: A.~Hutton and S.~Bose, \emph{Mediated entanglement and correlations in a star
5302: network of interacting spins}, Phys. Rev. A \textbf{69}, 042312 (2004).
5303:
5304: \bibitem{Olaya-Castro2005}
5305: A.~Olaya-Castro, N.~F. Johnson, and L.~Quiroga, \emph{Robust One-Step Catalytic
5306: Machine for High Fidelity Anticloning and W-State Generation in a Multiqubit
5307: System}, Phys. Rev. Lett. \textbf{94}, 110502 (2005).
5308:
5309: \bibitem{Olaya-Castro2006}
5310: A.~Olaya-Castro, C.~F. Lee, and N.~F. Johnson, \emph{Exact simulation of
5311: multi-qubit dynamics with only three qubits}, Europhys. Lett. \textbf{74},
5312: 208 (2006).
5313:
5314: \bibitem{Khaetskii2003}
5315: A.~Khaetskii, D.~Loss, and L.~Glazman, \emph{Electron spin evolution induced by
5316: interaction with nuclei in a quantum dot}, Phys. Rev. B \textbf{67}, 195329
5317: (2003).
5318:
5319: \bibitem{Rao2006}
5320: D.~D.~B. Rao, V.~Ravishankar, and V.~Subrahmanyam, \emph{Spin decoherence from
5321: Hamiltonian dynamics in quantum dots}, Phys. Rev. A \textbf{74}, 022301
5322: (2006).
5323:
5324: \end{thebibliography}
5325: \begin{theindex}
5326:
5327: \item amplitude damping, 41
5328: \item amplitude delaying channel, 33
5329: \item Anderson localisation, 43
5330: \item arbitrary and unknown qubit, 11
5331: \item asymptotic deformation, 81
5332:
5333: \indexspace
5334:
5335: \item black box, 12, 49, 53
5336:
5337: \indexspace
5338:
5339: \item chocolate, 11
5340: \item classical averaged fidelity, 27
5341: \item coding transformation, 96
5342: \item conclusively perfect state transfer., 33
5343: \item cooling protocol, 17
5344: \item coupled chains, 53
5345: \item CPT, 83
5346: \item criteria for quantum state transfer, 31
5347:
5348: \indexspace
5349:
5350: \item decoherence-free subspace, 41
5351: \item destabilisation, 116
5352: \item dispersion, 21
5353: \item distillation, 27
5354: \item dual rail, 35
5355:
5356: \indexspace
5357:
5358: \item efficiency, 58
5359: \item engineered couplings, 29
5360: \item entanglement distillation, 37
5361: \item entanglement of distillation, 27, 63
5362: \item entanglement of formation, 27
5363: \item entanglement transfer, 27
5364: \item ergodic, 76
5365: \item experiments, 15
5366:
5367: \indexspace
5368:
5369: \item fidelity, 16
5370: \item fix-point, 76
5371: \item flux qubits, 15
5372:
5373: \indexspace
5374:
5375: \item generalised Lyapunov function, 77
5376:
5377: \indexspace
5378:
5379: \item Heisenberg Hamiltonian, 20
5380: \item homogenisation, 92
5381:
5382: \indexspace
5383:
5384: \item Lindblad equation, 109
5385:
5386: \indexspace
5387:
5388: \item maximal peak, 23
5389: \item minimal fidelity, 16, 20
5390: \item mixing, 76
5391: \item multi rail, 68
5392:
5393: \indexspace
5394:
5395: \item non-expansive map, 80
5396: \item non-Markovian, 117
5397:
5398: \indexspace
5399:
5400: \item peak width, 27
5401: \item peripheral eigenvalues, 86
5402: \item phase noise, 41
5403: \item pure fix-points, 87
5404:
5405: \indexspace
5406:
5407: \item quantum capacity, 28
5408: \item quantum chain, 12
5409: \item quantum channel, 16, 83
5410: \item quantum computer, 10
5411: \item quantum erasure channel, 37
5412: \item quantum gates, 11
5413: \item quantum memory, 93
5414: \item quantum relative entropy, 84
5415: \item quantum-jump approach, 41
5416: \item qutrits, 55
5417:
5418: \indexspace
5419:
5420: \item reading and writing fidelities, 100
5421:
5422: \indexspace
5423:
5424: \item scalability, 12
5425: \item Schr\"odinger equation, 109
5426: \item Shor's algorithm, 10
5427: \item spectral radius, 67
5428: \item spin chain, 12
5429: \item strict contraction, 80
5430: \item swap gates, 26
5431:
5432: \indexspace
5433:
5434: \item time-scale, 38
5435: \item tomography, 49
5436: \item topological space, 74
5437: \item transfer functions, 18
5438:
5439: \indexspace
5440:
5441: \item valve, 103
5442:
5443: \indexspace
5444:
5445: \item weak contraction, 80
5446:
5447: \end{theindex}
5448: \end{document}
5449: