quant-ph0507141/qcc.tex
1: \documentclass{amsproc}
2: %\documentclass{amsart}
3: \usepackage{oldgerm}
4: \usepackage{latexsym}
5: \usepackage{amsmath}
6: \usepackage{amssymb}
7: \usepackage{amsgen}
8: \usepackage{amscd} 
9: \usepackage{color}
10: \usepackage{epsfig}
11: 
12: 
13: 
14: \newtheorem{thm}{Theorem}[section]
15: \newtheorem{prop}[thm]{Proposition}
16: \newtheorem{cor}[thm]{Corollary}
17: \newtheorem{lem}[thm]{Lemma}
18: 
19: 
20: \theoremstyle{definition}
21: \newtheorem{dfn}[thm]{Definition}
22: \newtheorem{example}[thm]{Example}
23: \theoremstyle{remark}
24: \newtheorem{rem}[thm]{Remark}
25: \theoremstyle{remark}
26: \newtheorem{rmks}[thm]{Remarks}
27: 
28: \newcommand{\bit}{\{0,1\}}
29: \newcommand{\bottom}{\mathbf{undefined}}
30: 
31: \newcommand{\opr}[1]{\operatorname{#1}}
32: 
33: \title[A Theory of Physical Quantum Computation]{A Theory of Physical
34: Quantum Computation: \\
35: The Quantum Computer Condition} \author[Gerald Gilbert, Michael Hamrick and
36: F. Javier Thayer]{Gerald Gilbert, Michael Hamrick and
37: F. Javier Thayer \\
38: \small \it Quantum Information Science Group$^\ast$$^\dag$\\
39: {\sc  Mitre} \\
40: \small \it 260 Industrial Way West, Eatontown, NJ 07724 USA
41: % \address{Mitre Corporation} \email{\{ggilbert,
42: % mharmrick, jt\}@mitre.org}
43: %
44: \footnote{$^\ast$Research supported under MITRE Technology Program Grant 51MSR211.
45: }
46: \footnote{$^\dag$E-mail address: \tt{\{ggilbert, mhamrick, jt\}@mitre.org}}
47: }
48: 
49: 
50: \makeindex 
51: 
52: \begin{document} 
53: 
54: \begin{abstract}
55: %
56: In this paper we present a new
57: unified theoretical framework that describes the full dynamics
58: of quantum computation.
59: Our formulation allows any questions pertaining to the physical behavior of
60: a quantum
61: computer to be framed, and in principle, answered.
62: We refer to the central organizing principle developed in this paper,
63: on which our theoretical structure is based, as the {\em Quantum Computer Condition} (QCC),
64: a rigorous mathematical statement that connects the irreversible
65: dynamics of the quantum computing machine, with the reversible operations that comprise the
66: quantum computation intended to be carried out by the quantum computing machine.
67: Armed with the QCC, we derive a powerful result that we call
68: the {\em Encoding No-Go Theorem}. This
69: theorem gives a precise mathematical statement of the conditions under which
70: fault-tolerant quantum computation becomes impossible in the presence of dissipation and/or
71: decoherence. In connection with this theorem, we
72: explicitly calculate a {\em universal} critical damping value
73: for fault-tolerant quantum computation.
74: In addition we show that the recently-discovered approach to quantum
75: error correction known as ``operator quantum error-correction" is a special case
76: of our more general formulation. Our approach furnishes what we will refer to as
77: ``operator quantum fault-tolerance." In particular, we show how the QCC allows one to
78: derive error thresholds for fault tolerance in a completely general context.
79: We prove the existence of solutions to a class of time-dependent generalizations of
80: the Lindblad equation.
81: Using the QCC, we also show that the seemingly
82: different circuit, graph- (including cluster-) state,
83: and adiabatic paradigms for quantum computing
84: are in fact all manifestations of a single, universal
85: paradigm for all physical quantum computation.
86: % Using a model for quantum components based on time dependent Lindblad
87: % equations, we investigate how reversible computing devices (for
88: % instance those modeled by quantum circuits) can be simulated on
89: % them. 
90: \end{abstract}
91: 
92: \maketitle
93: 
94: \newpage
95: 
96: \tableofcontents
97: 
98: 
99: \section{Introduction}
100: 
101: The promise inherent in quantum computing has stimulated a tremendous explosion
102: of interest in the research community.
103: Voluminous research has been carried out directed to many different problems associated
104: with the development and properties of quantum computational algorithms.
105: Parallel to these efforts, substantial investigations have been devoted to the
106: problems associated with the development of actual quantum computing machines.
107: A rigorous and fully general theory that connects quantum computing algorithms
108: and quantum computing machines would be of considerable value.
109: 
110: In this paper we present a new
111: unified theoretical framework that describes the full dynamics
112: of quantum computation.
113: Our formulation allows any questions pertaining to the physical behavior of
114: a quantum computer to be framed, and in principle, answered.
115: We refer to the central organizing principle developed in this paper,
116: on which our theoretical structure is based, as the {\em Quantum Computer Condition} (QCC),
117: a rigorous mathematical statement that connects the irreversible
118: dynamics of the quantum computing machine, with the reversible operations that comprise the
119: quantum computation intended to be carried out by the quantum computing machine.
120: 
121: The actual dynamics of the
122: system that we intend to use as a practical quantum computing machine are those of an open
123: quantum mechanical system, burdened with various dissipative and/or decoherence effects.
124: The QCC provides a set of mathematical constraints that must be satisfied by a physical
125: system if
126: we intend to use that system as a quantum computing machine.
127: 
128: Armed with the QCC, we derive a powerful result that we call
129: the {\em Encoding No-Go Theorem}. The
130: Encoding No-Go theorem gives a precise mathematical statement of the conditions under which
131: fault-tolerant quantum computation becomes impossible in the presence of dissipation and/or
132: decoherence. We provide a rigorous definition of damping, which includes the phenomena
133: of dissipation and decoherence, and explicitly
134: calculate a {\em universal} critical damping value
135: for fault-tolerant quantum computation.
136: This fundamental theorem has deep formal significance. Moreover,
137: it also furnishes criteria for solving diverse problems associated to
138: actual physical quantum computer realizations, such as
139: determining which practical design choices for 
140: quantum computing machines are not viable.
141: 
142: In addition we show that the recently-discovered approach to quantum
143: error correction known as ``operator quantum error-correction" (OQEC) is
144: actually a special case
145: of our general formulation. Our approach furnishes what we will refer to as
146: ``operator quantum fault-tolerance" (OQFT). In
147: particular, we show how the QCC allows one to
148: derive error thresholds for fault tolerance in a completely general context.
149: 
150: In this paper we define the concept of a {\em quantum component}, which allows us
151: to study realistic implementations of quantum computers,
152: in which decoherence and/or
153: dissipative effects are present, using a dynamical equation of motion suitable for
154: describing an open quantum system.
155: By using the QCC, we are able to reconcile the apparent contradiction between: (1) the
156: fact that quantum computations are specified by unitary transformations, the
157: associated dynamics of which are intrinsically reversible, and (2) the fact
158: that quantum computers,
159: {\em qua} practical machines, are inevitably characterized by irreversible
160: dynamics. The reconciliation suggests an analogy with
161: the {\em fluctuation-dissipation theorem}, which
162: relates irreversible dynamics to equilibrium properties in a large class of physical
163: systems.
164: 
165: In this paper we present an existence proof for fundamental solutions to useful
166: classes of {\em time-dependent} generalizations of the Lindblad equation. This provides
167: a useful tool in analyzing a wide variety of open quantum mechanical systems.
168: 
169: Our framework is sufficiently general to encompass,
170: and describe in a unified manner, the currently-known ``paradigms" for quantum computation,
171: including the {\em circuit-based (``two-way computing") paradigm},
172: the {\em graph state-based (``one-way computing") paradigm} and
173: the {\em adiabatic quantum computer paradigm}.
174: Using the QCC, we show by explicit construction that these seemingly different paradigms
175: are in fact all manifestations of a single, universal
176: paradigm for all physical quantum computation.\footnote{
177: In the particular case of the graph state-based paradigm (which includes cluster
178: state-based models), we not only show that the
179: paradigm is a manifestation of the unifying picture provided by the QCC,
180: but also introduce a definition of graph state-based quantum computers that
181: generalizes the graph state models previously defined in the literature.
182: }
183: 
184: 
185: 
186: \section{The Quantum Computer Condition}\label{QCC_section}
187: 
188: \subsection{Introduction} 
189: 
190: In this section we present the {\em Quantum Computer Condition}, a rigorous mathematical
191: statement of the constraints that determine the viability of any
192: practical quantum computing machine.
193: To achieve the goal of practical quantum computation we must produce an actual
194: physical device that implements a predetermined unitary operator $U$
195: acting on some Hilbert space.
196: The Quantum Computer Condition relates the unitary operator representing a quantum
197: computation to the actual physical device intended to perform that computation.
198: 
199: The specification of $U$
200: defines ideally the quantum computation to be performed by the quantum computing
201: machine.
202: Generically, 
203: the result of a quantum computation, $U$, is then used to carry out the
204: probabilistic evaluation of some classical function.
205: The complete quantum computation comprises a number of elements, including
206: %
207: \begin{itemize}
208: %
209: \item Preparation of a quantum state for initialization.
210: %
211: \item  Measurement of a quantum state for readout.
212: %
213: \item Various tasks that can be performed by classical computers, such
214: as preprocessing of the data, or postprocessing of the output into
215: some humanly comprehensible form.
216: %
217: \end{itemize}
218: %
219: However, the above list of elements are not what is ``important" about quantum
220: computers. Rather:
221: %
222: \begin{itemize}
223: %
224: \item The distinctive element of quantum computation is the
225: ``ability to perform quantum gates''(c.f.~\cite{nielsen-chuang},
226: \S4.6). 
227: %
228: \end{itemize}
229: %
230: 
231: Mathematically, a quantum gate is a unitary (hence
232: reversible) operator $U$ acting on a Hilbert space. The formally defined ``gates,"
233: as such, are not ``devices." They are concepts: they don't implement themselves.
234: A machine is required to physically
235: implement the abstractly defined unitary transformation.
236: The actual, physical computing device intended to implement the transformation
237: is described mathematically by a completely
238: positive trace-preserving map, $P$, that transforms the input state to the output state.
239: We will refer to a physically realizable device intended to implement
240: an ideal quantum computation as a {\em quantum component}.
241: In this paper we study
242: realistic quantum components, in which decoherence and/or
243: dissipative effects are present, using a dynamical equation of motion suitable for
244: describing an open quantum system.\footnote{
245: In order to analyze the effects of dissipation and/or decoherence one must use
246: {\em some} method of approximating the dynamics of the degrees-of-freedom comprising
247: the rest of the universe ``outside of" the quantum computer.
248: This is of course because the complete, detailed,
249: exact analytical solution to the Schr\"odinger equation of the universe, for all
250: degrees-of-freedom, is not known.
251: One reasonable approach is to construct a
252: Lindblad-type equation, based on a presumption of underlying Markovian dynamics,
253: in which environment degrees-of-freedom are traced over in such
254: a way as to result in a first-order (in time) differential equation. In this paper,
255: for definiteness, we
256: utilize a generalized
257: Lindblad-type equation to describe the environment: this is used merely
258: in order to exemplify
259: how one may take into account the effects of dissipation and/or decoherence. However,
260: most of
261: the results in our paper,
262: including the crucial Encoding No-Go Theorem,
263: are independent of this choice, and in particular
264: are independent of the assumption of underlying Markovian dynamics.}
265: We must reconcile the fact, and apparent paradox, that a
266: non-reversible mapping, $P$, is used to ``implement'' a reversible one, $U$.
267: 
268: 
269: \subsection{The Motivation of the Quantum Computer Condition}
270: %
271: Mathematically, a quantum computation is a unitary operator $U$ in the
272: unitary group of a Hilbert space.
273: A quantum component is described by a
274: completely positive trace-preserving map $P$ which maps the set of trace class
275: operators on the Hilbert space to itself. The map $P$ accounts for decoherence and
276: dissipation,
277: as well as unitary evolution. We will subsequently discuss in more detail the
278: actual form for $P$. In our analysis
279: we will consider the action of $P$ on density matrices
280: $\rho\mapsto P\cdot\rho$ rather than on state vectors
281: (and correspondingly the action of $U$ on density matrices $\rho
282: \mapsto U \rho U^\dag$, rather than the action of $U$ on state
283: vectors).
284: This is because, due to the presence of decoherence and/or
285: dissipation, our system will almost always evolve into a mixed state, which can only
286: be described by a density matrix $\rho$.\footnote{
287: Other reasons for utilizing density matrices rather than state vectors include
288: the generic importance in quantum information theory of trace-preserving
289: completely positive maps (which restrict to transformations on density
290: matrices), and the useful algebraic and analytic properties of density matrices
291: (density matrices for instance form a weak-$\ast$ compact convex set).}
292: 
293: In order to motivate the Quantum Computer Condition (QCC), let us first consider the
294: abstractly-defined quantum computation itself, prescribed by the unitary operator $U$.
295: This is assumed to be given, and
296: is represented by
297: %
298: \begin{equation}
299: %
300: \mathrm{abstract ~computation\!:}\phantom{x} U\rho U^\dag ~.
301: %
302: \end{equation}
303: %
304: The action of the quantum component
305: intended to effect the computation is represented by
306: %
307: \begin{equation}
308: %
309: \mathrm{practical ~implementation\!:}\phantom{x} P\cdot\rho ~.
310: %
311: \end{equation}
312: %
313: Motivated by the  notion of {\em having the machine implement the computation},
314: if we were to require that the identity
315: %
316: \begin{equation}\label{ersatz-quantum-computer-equation}
317: %
318: P\cdot\rho = U \rho U^\dag
319: %
320: \end{equation}
321: %
322: hold for all density states $\rho$, then the action of $P$
323: would in fact be {\em identical} to the action of the unitary operator.
324: This would imply that $P$ preserves von Neumann entropy~(\cite{vonNeumann},
325: \S5.3), and hence actually models a system with neither
326: decoherence nor dissipation, which is not the case for a practical quantum
327: computing machine. Thus, equation (\ref{ersatz-quantum-computer-equation}) cannot
328: furnish the correct constraints for realistic quantum computation. We will accordingly
329: refer to
330: equation (\ref{ersatz-quantum-computer-equation}) as the {\em ersatz} quantum
331: computer condition ($\mathcal {E}$QCC).\footnote{
332: Although the $\mathcal {E}$QCC does not describe practical quantum computing machines,
333: we note that it can be shown that in the finite-dimensional case, the set of $\rho$
334: which satisfy~\eqref{ersatz-quantum-computer-equation} is an
335: algebra $\mathfrak{A}_{P,U}$ which depends on both $P$ and $U$.
336: The details of how one explicitly
337: obtains the algebra $\mathfrak{A}_{P,U}$ of solutions to
338:  ~\eqref{ersatz-quantum-computer-equation}
339: are given in Appendix~\ref{alpha_0_solution}.}
340: 
341: More realistically, taking into account the inevitable presence of decoherence,
342: we can require
343: that~\eqref{ersatz-quantum-computer-equation} hold for some
344: restricted set of density states. In this case, the solution set will 
345: correspond to decoherence-free subspaces.
346: In order to analyze this, we must carefully distinguish between the two different Hilbert
347: spaces that arise in this problem. The abstract quantum computation is defined
348: on the Hilbert space of {\em logical} quantum
349: states, $H_{\mathrm{logical}}$, so that we have 
350: %
351: \begin{equation}
352: %
353: U: H_{\mathrm{logical}}\rightarrow H_{\mathrm{logical}}.
354: %
355: \end{equation}
356: %
357: In contrast, the presence of decoherence (which affects the actual device)
358: necessitates that
359: the completely positive trace-preserving map $P$ (which represents the actual device)
360: is associated to a different Hilbert space,
361: $H_{\mathrm{comp}}$, the states of which are referred to as
362: {\em computational} quantum states. The decoherence-free subspace is contained
363: within $H_{\mathrm{comp}}$.
364: (The specific decoherence is accounted for in the explicit form of $P$). As
365: noted above, a consequence
366: of the decoherence is that $P$ operates on density matrices rather than on
367: state vectors. Letting ${\mathbf T}(H_{\mathrm{comp}})$ be the Banach space of
368: trace-class operators on $H_{\mathrm{comp}}$, we have
369: %
370: \begin{equation}
371: %
372: P: {\mathbf T}(H_{\mathrm{comp}})\rightarrow {\mathbf T}(H_{\mathrm{comp}}).
373: %
374: \end{equation}
375: %
376: In order to replace $\mathcal {E}$QCC
377: (equation (\ref{ersatz-quantum-computer-equation}))
378: with an equation that properly
379: incorporates decoherence effects so that it can be used to
380: determine decoherence-free subspaces, we must introduce suitable encoding and decoding
381: maps that connect the relevant Hilbert spaces.
382: We can hope to find an encoding operator
383: $\mathcal{M}_\mathrm{enc}$ defined on a space of logical inputs and a
384: decoding operator $\mathcal{M}_\mathrm{dec}$ with values in a space of
385: logical outputs, the actions of which are given by
386: (here ${\mathbf T}(H_{\mathrm{logical}})$ is the Banach space of trace-class operators
387: on $H_{\mathrm{logical}}$, analogous to ${\mathbf T}(H_{\mathrm{comp}})$)
388: %
389: \begin{equation}
390: %
391: \mathcal M_{\mathrm{enc}}: {\mathbf T}(H_{\mathrm{logical}}) \rightarrow
392: {\mathbf T}(H_{\mathrm{comp}})
393: %
394: \end{equation}
395: %
396: and
397: %
398: \begin{equation}
399: %
400: \mathcal M_{\mathrm{dec}}: {\mathbf T}(H_{\mathrm{comp}}) \rightarrow
401: {\mathbf T}(H_{\mathrm{logical}}),
402: %
403: \end{equation}
404: %
405: such that ({\em cf} equation (\ref{ersatz-quantum-computer-equation}))
406: %
407: \begin{equation}\label{encoded-quantum-computer-equation}
408: %
409: \mathcal{M}_\mathrm{dec}(P\cdot(\mathcal{M}_\mathrm{enc}(\rho))) =
410: U\rho U^\dag
411: %
412: \end{equation}
413: %
414: for all logical inputs $\rho$. We will refer to equation
415: (\ref{encoded-quantum-computer-equation}) as the ``encoded {\em ersatz}
416: quantum computer condition (e$\mathcal {E}$QCC).\footnote{
417: Note that no encoding map is required on the right-hand side of equation
418: (\ref{encoded-quantum-computer-equation}) since the unitary $U$ by
419: definition acts on $H_{\mathrm{logical}}$.} The existence of the encoding and decoding maps
420: $\mathcal M_{\mathrm{enc}}$ and $\mathcal M_{\mathrm{dec}}$ is a consequence of
421: the presumed existence of an associated decoherence-free subspace of $H_{\mathrm{comp}}$,
422: of dimension greater than or equal to the dimension of $H_{\mathrm{logical}}$.
423: 
424: The meaning of e$\mathcal {E}$QCC given in eq.(\ref{encoded-quantum-computer-equation})
425: is as follows. Given a chosen quantum computation, $U$, we wish to construct a
426: physical ``machine," $P$, that implements $U$. In order to do this we must
427: find 
428: encoding and decoding maps $\mathcal{M}_\mathrm{enc}$ and $\mathcal{M}_\mathrm{dec}$
429: such that the equation is satisfied for all $\rho$. This is a crucial difference
430: between eqs.(\ref{encoded-quantum-computer-equation}) and
431: (\ref{ersatz-quantum-computer-equation}):
432: requiring that eq.(\ref{ersatz-quantum-computer-equation}) holds for
433: all $\rho$ implies a machine that preserves von Neumann
434: entropy, does not dissipate heat and does not decohere,
435: and thus does not describe a practical quantum computing device. In contrast,
436: eq.(\ref{encoded-quantum-computer-equation}) holds for all states $\rho$, but does
437: so by making use of the encoding and
438: decoding operators to map to a decoherence-free subspace.
439: The e$\mathcal {E}$QCC (eq.(\ref{encoded-quantum-computer-equation}))
440: formally holds for all
441: density states $\rho$, similar to eq.(\ref{ersatz-quantum-computer-equation}),
442: but the use of the encoding and decoding maps
443: in eq.(\ref{encoded-quantum-computer-equation}) effectively confines the solution
444: to a restricted set in $H_{\mathrm{comp}}$.
445: 
446: However, eq.(\ref{encoded-quantum-computer-equation}) does not in general provide
447: an acceptable condition to connect the dynamics of a practical quantum computing
448: device to the constraints implied by the unitary operator $U$ that defines the
449: abstract quantum computation. {\em {This is because the formulation presented by
450: \eqref{encoded-quantum-computer-equation} does not address
451: situations in which residual errors cannot
452: be completely eliminated}},
453: even with the use of decoherence-free subspaces,
454: and/or other error correction methods~\cite{lbw_dfs_qecc_99}, \cite{shor_ft_96}.
455: 
456: To recapitulate, eq.(\ref{ersatz-quantum-computer-equation}) describes a quantum
457: computer that performs the required computation $U$, but only if the implementing device,
458: described by the completely positive map $P$, neither dissipates heat nor decoheres. We thus
459: reject this expression as a viable quantum computer condition
460: because it describes a system that is effectively impossible to
461: achieve. In contrast, eq.(\ref{encoded-quantum-computer-equation})
462: describes a quantum computer that performs the required computation $U$,
463: but only if the device described by the completely positive map $P$ implements
464: the required decoherence-free subspace with absolutely no residual errors.
465: This does not provide a sufficiently
466: general formulation: {\em we need to consider situations in which residual errors cannot
467: be completely eliminated}.
468: 
469: 
470: \subsection{The Presentation of the Quantum Computer Condition}
471: %
472: We wish to allow for the likelihood that, even if a decoherence-free subspace
473: can be found, and even if error correction procedures are applied, there will still
474: be residual errors characterizing the operation of the
475: quantum computer. Such a situation may arise even if
476: error correction is correctly applied: for instance, in the application of concatenated
477: error codes, one iterates the concatenation process until the error probability is
478: reduced to a value that is deemed ``acceptable"~\cite{shor_ft_96}, \cite{preskill_ft_97}.
479: This final error
480: probability, though small, is not exactly zero. The important point is that
481: it is prudent to write down our
482: quantum computer condition so as to reflect the
483: inevitable survival of some amount of residual error.
484: 
485: In order to quantify the degree to which the actual computational device,
486: represented by $P$, {\em cannot exactly} (because of residual error) implement the
487: ideal quantum computation, represented by $U$, we consider the following difference
488: ({\em cf} eq.\eqref{encoded-quantum-computer-equation})
489: %
490: \begin{equation}\label{inaccuracy_preparameter}
491: %
492: \mathcal{M}_\mathrm{dec}(P\cdot(\mathcal{M}_\mathrm{enc}(\rho))) -
493: U\rho U^\dag.
494: %
495: \end{equation}
496: %
497: We now compute for this difference a suitable norm on matrices (this norm is made more
498: precise below), as
499: %
500: \begin{equation}\label{inaccuracy_parameter}
501: %
502: \|\mathcal{M}_\mathrm{dec}(P\cdot(\mathcal{M}_\mathrm{enc}(\rho))) -
503: U\rho U^\dag\|.
504: %
505: \end{equation}
506: %
507: This quantity is of fundamental importance: {\em it
508: is a measure of the inaccuracy of the implementation of $U$
509: by} $P$. It tells us how well a practical quantum computing
510: device actually implements an ideally-defined quantum computation. We will refer
511: to the scalar quantity
512: given by (\ref{inaccuracy_parameter}) as the {\em implementation inaccuracy}
513: associated to the pair $U$ and $P$.
514: In connection with this, we introduce a parameter, $\alpha$, to specify the maximum
515: tolerable implementation inaccuracy, so that we have
516: %
517: \begin{equation}\label{pre_encoded_QCC}
518: %
519: \|\mathcal{M}_\mathrm{dec}(P\cdot(\mathcal{M}_\mathrm{enc}(\rho))) -
520: U\rho U^\dag\| \leq \alpha .
521: %
522: \end{equation}
523: 
524: \subsubsection{Formal Statement of the Quantum Computer Condition (QCC)}
525: Motivated by the above considerations,
526: we now introduce and formally define an inequality of fundamental importance in the theory of
527: quantum computation that
528: we will refer to as the {\em Quantum
529: Computer Condition}. We will formally designate this condition
530: by $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
531: \mathcal{M}_\mathrm{dec},\alpha)$.
532: In the following, we will impose no constraint on the dimensionality of the Hilbert
533: space, and in particular we allow Hilbert spaces of infinite dimensions.
534: Let $U$ be a unitary on
535: $H_{\mathrm{logical}}$ and
536: $P$ be a trace-preserving completely positive map on
537: $\mathbf{T}(H_\mathrm{comp})$. 
538: Let
539: $\mathcal M_{\mathrm{enc}}: {\mathbf T}(H_{\mathrm{logical}}) \rightarrow
540: {\mathbf T}(H_{\mathrm{comp}})$ and
541: $\mathcal M_{\mathrm{dec}}: {\mathbf T}(H_{\mathrm{comp}}) \rightarrow
542: {\mathbf T}(H_{\mathrm{logical}})$ be
543: completely-positive, trace-preserving encoding and decoding maps
544: (with no further restrictions of any kind on
545: the encoding and decoding maps).
546: The Quantum Computer Condition,
547: $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
548: \mathcal{M}_\mathrm{dec},\alpha)$, holds iff
549: for all density matrices $\rho \in \mathbf{T}(H_\mathrm{logical})$, we have
550: %
551: \begin{equation}\label{encoded_QCC}
552: %
553: \|\mathcal{M}_\mathrm{dec}(P\cdot(\mathcal{M}_\mathrm{enc}(\rho))) -
554: U\rho U^\dag\|_1 \leq \alpha ,
555: %
556: \end{equation}
557: %
558: where $\mathbf{T}(H_\mathrm{comp})$ and $\mathbf{T}(H_\mathrm{logical})$ are
559: the Banach spaces of trace-class operators on $H_\mathrm{comp}$ and
560: $H_\mathrm{logical}$, respectively, and
561: $\|\cdot\|_1$ is the Schatten $1$-norm defined in
562: Appendix~\ref{banach-spaces}.
563: 
564: 
565: 
566: It should be noted that an alternate measure of distance between density matrices
567: to that provided by the Schatten $1$-norm is given by the fidelity
568: function \cite{nielsen-chuang}.
569: One could write an alternate form of the QCC in terms of the fidelity that would be
570: essentially equivalent to the form of the QCC given in \eqref{encoded_QCC} above.
571: Using an obvious notation to denote the QCC written with each of these definitions of
572: distance, the two forms are related as follows. Given a quartet
573: $\{ P,U,\mathcal{M}_\mathrm{enc},
574: \mathcal{M}_\mathrm{dec}\}$, one can show that if
575: $\mathbf{QCC}_{{\mathrm{Schatten}}}(P,U,\mathcal{M}_\mathrm{enc},
576: \mathcal{M}_\mathrm{dec},\alpha)$ is satisfied, then the fidelity-based version
577: of QCC given by
578: $\mathbf{QCC}_{{\mathrm{fidelity}}}(P,U,\mathcal{M}_\mathrm{enc},
579: \mathcal{M}_\mathrm{dec},\alpha^\prime)$ is also satisfied, where
580: $\alpha^\prime = \alpha^\prime(\alpha)$ is a function of $\alpha$.
581: The form of QCC based on the Schatten $1$-norm given above in \eqref{encoded_QCC}
582: is more convenient for our purposes for a number of 
583: mathematical reasons, including the fact that that the fidelity, as such, is not
584: a proper norm. For instance, the
585: statement and proof of the Encoding No-Go Theorem carried out in
586: \S\ref{NGT} below are more conveniently presented making use of the form of the QCC
587: based on the Schatten norm.
588: 
589: Note that $\mathcal{M}_\mathrm{enc}$ and $\mathcal{M}_\mathrm{dec}$
590: %as purely 
591: %{\em formal} operations that connect the logical and computational state spaces. They
592: do not 
593: represent {\em physical} operations: all physical operations are
594: carried out by the quantum component $P$.
595: If the proper distinction between these maps and $P$ 
596: is not observed, one could include the {\em entire} computation in the definition of the
597: maps, with 
598: the absurd conclusion that any quantum computation could be performed in the absence of any 
599: real hardware.
600: 
601: \subsubsection{Some implications of the QCC}
602: 
603: The QCC is a remarkably powerful expression. It constitutes a kind of ``master
604: expression" for physical quantum computation. The inequality \eqref{encoded_QCC}
605: concisely incorporates a complete specification of the full dissipative, decohering
606: dynamics of the actual,
607: practical device used as the quantum computing machine, a specification of
608: the ideally-defined quantum
609: computation intended to be performed by the machine, and a quantitative criterion
610: for the accuracy with which the computation must be executed given the inevitability
611: of residual errors surviving even after error correction has been applied.
612: 
613: Making use of the QCC, one can state and
614: prove (we do this is in \S\ref{NGT}, the next section of the paper)
615: a fundamental and powerful theorem in the subject of quantum computing,
616: the {\em Encoding No-Go Theorem}. This
617: no-go theorem gives a precise mathematical statement of the conditions under which
618: fault-tolerant quantum computation becomes impossible in the presence of dissipation and/or
619: decoherence. Apart from its formal significance, the theorem can be used to compare
620: different proposed physical approaches to actually building a quantum computing
621: machine, with the no-go condition furnishing a criterion for the practical
622: engineering viability of various choices.
623: 
624: As a further indication of the power and general applicability of the QCC, we
625: show that one can apply it
626: to the known, seemingly distinct ``paradigms" for quantum computing, based on (1) the
627: use of quantum circuits built up out of quantum gates (the {\em circuit-based paradigm},
628: or ``two-way" quantum computing), (2) the use of graph states or cluster states
629: (the {\em graph state-based paradigm}, or ``one-way" quantum computing) and (3) the
630: use of specially chosen Hamiltonians describing adiabatic dynamics (the {\em adiabatic
631: quantum computer paradigm}). The QCC allows one to show that these apparently
632: different definitions of a quantum computer are in fact manifestations of the same
633: underlying formulation: {\em there is only one paradigm for quantum computers}.
634: The application of the QCC to different quantum computing paradigms
635: is discussed in \S\ref{unification}.
636: 
637: The encoding-decoding pair $\mathcal{M}_\mathrm{enc}$ and $\mathcal{M}_\mathrm{dec}$
638: that appear in the QCC are defined quite generally as completely positive
639: trace-preserving maps. This formulation is sufficiently general to encompass
640: all possible encodings associated with standard quantum error correction (QECC)
641: techniques, decoherence-free subspaces and noiseless subsystems. More generally still,
642: we show below in \S\ref{KLP_reduction} that the recently discovered approach
643: known as ``operator quantum error correction" (OQEC) is in fact a special case
644: of our more general QCC formulation.
645: In addition the QCC can be used to extend OQEC to
646: what we will refer to as
647: ``operator quantum fault-tolerance" (OQFT). In particular, we show in \S\ref{et_ft} below
648: how the QCC allows one to
649: derive error thresholds for fault tolerance in a completely general context.
650: 
651: Another significant consequence of the QCC is that it resolves the apparent
652: paradox that the
653: quantum computations we wish to perform are defined by reversible operators, but
654: the actual devices that we must use to execute the computations are
655: necessarily described by irreversible maps.
656: We note that this is reminiscent of the {\em fluctuation-dissipation theorem}, which
657: relates irreversible dynamics to equilibrium properties in a large class of physical
658: systems. Inspection of \eqref{encoded_QCC} reveals
659: that the paradox is resolved through the transformations provided by the
660: encoding and decoding maps associated to the QCC. Roughly speaking,
661: reversible behavior of the actual device is enabled only on the code subspace
662: defined by $\mathcal{M}_\mathrm{enc}$ and $\mathcal{M}_\mathrm{dec}$.
663: 
664: Note that if one describes quantum computing solely in terms of the
665: unitary transformations that define the computations,
666: it is not too surprising that the resulting computational model
667: is in some way ``powerful."
668: After all, unitary transformations on finite dimensional
669: spaces include such powerful operations as the discrete Fourier
670: transform which are known to play an important role in number
671: theoretic problems.
672: Rather than regarding the power of the ideally-defined quantum computation as the
673: remarkable thing, the {\em truly} remarkable thing would be the
674: construction of an inherently irreversible
675: device that actually implements the reversible, unitary map to a
676: specified level of accuracy. It is this possibility that our QCC expresses
677: and makes mathematically precise. The QCC can thus be used to formulate and prove
678: assertions about the physical solvability of particular computational
679: problems.
680: 
681: \subsubsection{Operator quantum error correction (OQEC) as a special case
682: of QCC}\label{KLP_reduction}
683: 
684: The recently developed theory of operator quantum error correction
685: \cite{klp_05}, \cite{klpl_05}
686: unifies many apparently disparate approaches to the practical 
687: problem of dealing with errors in quantum information.  Among these approaches are 
688: quantum error correction, decoherence 
689: free subspaces and noiseless subsystems.
690: Here we show that OQEC is in fact a special case of the general statement of the
691: QCC. In this section we demonstrate that the entire formalism of OQEC can be obtained
692: from QCC by
693: choosing $\mathcal{M}_\mathrm{enc}$ and $\mathcal{M}_\mathrm{dec}$ as described below,
694: and by setting setting $U=I$ and $\alpha=0$ in the QCC, so that the reduction
695: QCC $\rightarrow$ OQEC is given by:
696: %
697: \begin{equation}
698: %
699: \mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec},\alpha)\rightarrow
700: \mathbf{QCC}(P,I,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec},0)~.
701: %
702: \end{equation}
703: 
704: In the scheme of OQEC, the techniques
705: of (standard) quantum error correction, decoherence 
706: free subspaces and noiseless subsystems are subsumed 
707: under the 
708: unified concept of ``correctability." This is defined formally as follows.    
709: Let $H_\mathrm{comp}$ be a Hilbert space with a decomposition 
710: $H_\mathrm{comp}=\left( H^A \otimes H^B\right) 
711: \oplus K$, where $H^A$, $H^B$ and $K$ are discussed in the following paragraph.
712: We identify this as the computational space $H_\mathrm{comp}$ 
713: of this paper because, as we shall see below, it describes the space on which the
714: real physical processes of error and recovery operate.
715: Let $\mathfrak{S}$ be the semigroup given by 
716: %
717: \begin{equation}
718: %
719: \mathfrak{S} = \{ \sigma \in \mathbf{L}(H_\mathrm{comp}) : 
720:    \exists \sigma^A \in \mathbf{L}(H^A), \exists \sigma^B \in \mathbf{L}(H^B), ~
721:    \mathrm{such~that} ~ \sigma = \sigma^A \otimes \sigma^B \}~,
722: %
723: \end{equation}
724: %
725: where $\mathbf{L}(\cdot)$ is the space of bounded operators\footnote{In \S\ref
726: {KLP_reduction} of this paper we assume (following \cite{klpl_05}) that
727: all Hilbert spaces are finite-dimensional.
728: Although not discussed in \cite{klpl_05},
729: it is important to note that if one makes this assumption,
730: there is then
731: no need to distinguish between bounded operators in $L(\cdot )$
732: and trace-class operators in $T(\cdot )$.
733: This distinction is important in the other sections of our paper
734: where we allow Hilbert spaces of infinite as well as finite dimensionality.} on
735: the appropriate Hilbert
736: space, with the operator norm,
737: and let $\mathcal{E}$ and $\mathcal{R}$ be completely-positive, trace preserving 
738: maps on $\mathbf{L}(H_\mathrm{comp})$ corresponding to error processes and 
739: recovery processes, 
740: respectively.  We say that $\mathfrak{S}$ is correctable for $\mathcal{E}$ iff
741: %
742: \begin{equation}\label{klpcorrectable}
743: %
744: \forall \sigma \in \mathfrak{S},~ 
745: \left( \mathrm{Tr}_A \circ \mathcal{P}_\mathfrak{S} \circ \mathcal{R} 
746:    \circ \mathcal{E} \right) \sigma = \mathrm{Tr}_A \sigma ~,
747: %
748: \end{equation}
749: %
750: where $\mathcal{P}_\mathfrak{S}$ effectively projects density matrices onto the
751: $ H^A \otimes H^B$ subspace of $H_{\mathrm{comp}}$.
752: 
753: Physically speaking, $H^B$ is the Hilbert space which carries the information to 
754: be protected from errors, the ``noiseless subsystem," 
755: while $H^A$ is the Hilbert space on which the errors are 
756: permitted to operate freely, the ``noisy subsystem."  
757: $K$ is the orthogonal complement of $H^A \otimes H^B$ in $H_\mathrm{comp}$ 
758: and is simply projected out by $\mathcal{P}_\mathfrak{S}$.
759: Thus, in \eqref{klpcorrectable}, 
760: $\mathrm{Tr}_A \sigma$ represents the quantum information which is to be protected.  
761: The left side of \eqref{klpcorrectable} describes the effect of allowing  
762: errors to operate on the full state $\sigma$, and then applying 
763: recovery procedures.  (The projection and the 
764: trace simply extract the state of the noiseless subsystem.)  According to 
765: \eqref{klpcorrectable}, correctability thus means that 
766: the recovery procedure does in fact recover the state of the noiseless subsystem, 
767: $\mathrm{Tr}_A \sigma$, without error.  
768: 
769: As we now show, the definition of correctability is actually a special case of the QCC.  
770: Note first that correctability, as 
771: defined above, applies to a quantum channel as opposed to a quantum computer.  
772: It describes the transportation of a quantum state in a noisy environment as opposed to the 
773: ``processing" of a quantum state so as to implement a quantum computation.  
774: In order to make the connection with the QCC, we may 
775: therefore think of the quantum channel as a quantum computer that implements the 
776: identity operation:
777: %
778: \begin{equation}
779: %
780: U = I_{H_{\mathrm {logical}}}
781: %
782: \end{equation}
783: %
784: 
785: In the definition of correctability, the part of the state containing the 
786: information of interest is 
787: recovered without error.  This corresponds to taking 
788: $\alpha = 0$ in the QCC.  Note that, as discussed above, this is not practically 
789: achievable for quantum computers.  
790: This is less obviously an issue for the theory of operator error correction as currently 
791: formulated \cite{klp_05}, \cite{klpl_05}, since that theory addresses a relatively
792: circumscribed set of circumstances in which one is concerned with transporting quantum 
793: states rather than implementing a quantum computation.  In particular, the error process 
794: $\mathcal{E}$, which is specified in advance, only operates {\em once} on the quantum state 
795: being transported. Error processes associated 
796: with the constituent parts of quantum computers operate each time the constituent part 
797: operates on the 
798: the quantum state being processed. The problem of fault 
799: tolerant quantum computation is inherently more complex than the problem of 
800: error correction/protection for a quantum channel.  We will
801: discuss this more fully in \S\ref{et_ft} below.
802: 
803: Setting $U=I$ and $\alpha = 0$ in the quantum computer condition, we obtain
804: %
805: \begin{equation}
806: %
807: \|\left( \mathcal{M}_\mathrm{dec}\circ P \circ \mathcal{M}_\mathrm{enc} \right)\rho -
808: \rho\|_1 = 0 ,
809: %
810: \end{equation}
811: %
812: or
813: %
814: \begin{equation}
815: %
816: \left (\mathcal{M}_\mathrm{dec} \circ P \circ \mathcal{M}_\mathrm{enc}
817: \right) \rho = \rho ~,
818: %
819: \end{equation}
820: %
821: where $\rho\in L(H_{{\mathrm{logical}}})$.\footnote{As
822: noted in the previous footnote, we are assuming in \S\ref{KLP_reduction}
823: that $H_{{\mathrm{logical}}}$ is finite-dimensional.}
824: 
825: We now proceed to define the encoding map $\mathcal{M}_\mathrm{enc}$. For this purpose
826: we define a map 
827: $W_\mathrm{enc}:\mathbf{L}(H_\mathrm{logical}) \rightarrow \mathbf{L}(H^B)$
828: that encodes the logical quantum state $\rho$ in a state $\sigma^B$ of the noiseless subsytem.  We 
829: further define a map $W_\mathrm{adj}(\sigma^A):\mathbf{L}(H^B) 
830: \rightarrow \mathbf{L}(H^A \otimes H^B)$ 
831: that adjoins an arbitrary state $\sigma^A$ of the noisy subsystem to the state $\sigma^B$, 
832: that is 
833: %
834: \begin{equation}
835: %
836: W_\mathrm{adj}(\sigma^A) : \sigma^B \mapsto \sigma \equiv \sigma^A \otimes \sigma^B
837: %
838: \end{equation}
839: %
840: We then define the full encoding map that appears in the QCC as follows:
841: %
842: \begin{equation}
843: %
844: \mathcal{M}_\mathrm{enc} \equiv W_\mathrm{adj}(\sigma^A) \circ W_\mathrm{enc}
845: %
846: \end{equation}
847: %
848: 
849: The map $P$ characterizes the dynamics of the physical computer, which in this case is just 
850: the (noisy) channel followed by the recovery procedure:
851: %
852: \begin{equation}
853: %
854: P \equiv \mathcal{R} \circ \mathcal{E}
855: %
856: \end{equation}
857: %
858: We note that the current formulation of the theory of operator quantum error correction 
859: implicitly assumes that the recovery process $\mathcal{R}$ can be implemented without error, 
860: even though this requires, in general, that coherent operations be performed on entangled 
861: quantum states.  Once again, 
862: the emphasis on error correction alone avoids the more difficult issue of 
863: achieving true fault tolerance. 
864: 
865: 
866: Finally we define the decoding map as:
867: %
868: \begin{equation}
869: %
870: \mathcal{M}_\mathrm{dec} = W_\mathrm{enc}^{-1} \circ \mathrm{Tr}_A \circ \mathcal{P}_\mathfrak{S}~,
871: %
872: \end{equation}
873: %
874: which extracts the state of the noiseless subsystem and decodes it to obtain a state in the 
875: logical space $\mathbf{L}(H_\mathrm{logical})$.  
876: 
877: With the above definitions, the QCC becomes 
878: %
879: \begin{equation}
880: %
881: \left( W_\mathrm{enc}^{-1} \circ \mathrm{Tr}_A \circ \mathcal{P}_\mathfrak{S} \circ
882:        \mathcal{R} \circ \mathcal{E} \circ W_\mathrm{adj}(\sigma^A) \circ W_\mathrm{enc} \right)
883: \rho = \rho~.
884: %
885: \end{equation}
886: %
887: Applying $W_\mathrm{enc}$ to both sides, and recalling that $W_\mathrm{enc} \rho = 
888: \sigma^B = \mathrm{Tr}_A \sigma$, we obtain
889: %
890: \begin{equation}
891: %
892: \left( \mathrm{Tr}_A \circ \mathcal{P}_\mathfrak{S} \circ
893:        \mathcal{R} \circ \mathcal{E} \circ W_\mathrm{adj}(\sigma^A) \circ W_\mathrm{enc} \right)
894: \rho = \mathrm{Tr}_A \sigma~.
895: %
896: \end{equation}
897: %
898: Noting that 
899: $W_\mathrm{adj}(\sigma^A) \circ W_\mathrm{enc} \rho = \sigma$, we obtain the correctability 
900: condition of~\cite{klp_05}, \cite{klpl_05}:
901: %
902: \begin{equation}
903: %
904: \left( \mathrm{Tr}_A \circ \mathcal{P}_\mathfrak{S} \circ
905:        \mathcal{R} \circ \mathcal{E} \right)
906: \sigma = \mathrm{Tr}_A \sigma~.  
907: %
908: \end{equation}
909: %
910: 
911: In summary, we have shown that the formalism of operator quantum error correction
912: actually arises as a special case of an underlying definition of physical quantum
913: computation given by the QCC. In addition, examination of operator quantum error correction
914: (OQEC) from the perspective of the QCC reveals limitations and restrictions implicit to the
915: formalism of operator quantum error correction, and inherited from the previously 
916: known, standard approaches to error correction (QECC). These limitations
917: render direct application of either OQEC or QECC to questions of fault tolerance somewhat
918: problematic, whereas the QCC approach is more immediately applicable.
919: Thus, the QCC enables one to generalize OQEC to operator fault tolerance (OQFT).
920: 
921: 
922: 
923: \section{The Encoding No-Go Theorem}\label{NGT}
924: 
925: \subsection{Introduction}
926: Armed with the QCC, in this section we state and prove a theorem of crucial
927: importance in the theory of physical quantum computation. This is the {\em Encoding
928: No-Go Theorem}, which
929: gives a precise mathematical statement of the conditions under which
930: fault-tolerant quantum computation becomes impossible in the presence of damping.
931: Damping, for which we provide a mathematically rigorous definition below,
932: includes the effects of dissipation and decoherence.
933: The No-Go theorem for a completely positive trace-preserving
934: map $P$ corresponding to a putative quantum
935: computing device then asserts that, in the presence of sufficient damping,
936: the quantum computer condition
937: $\mathbf{QCC}(P, U, \mathcal{M}_\mathrm{enc},
938: \mathcal{M}_\mathrm{dec},\alpha)$ ({\em cf} \eqref{encoded_QCC})
939: cannot be satisfied for any
940: encoding-decoding pair, unless $H_\mathrm{logical}$
941: has dimension $1$: there is then effectively no quantum computer.
942: (In the case that dim $H_\mathrm{logical}=1$ we are of course unable to define
943: a meaningful quantum computation at all.)
944: As part of the No-Go Theorem we explicitly calculate a {\em universal} critical damping
945: value for fault-tolerant quantum computation.
946: We precisely state and prove this theorem in the remainder of \S\ref{NGT}.
947: 
948: \subsection{Encoding and Decoding Maps}
949: An encoding-decoding pair are completely positive, trace-preserving maps
950: %
951: \begin{equation}
952: %
953: \begin{aligned}
954: %
955: & \mathcal{M}_\mathrm{enc}:\mathbf{T}(H_\mathrm{logical}) \rightarrow \mathbf{T}(H_\mathrm{comp})\\
956: %
957: & \mathcal{M}_\mathrm{dec}:\mathbf{T}(H_\mathrm{comp}) \rightarrow \mathbf{T}(H_\mathrm{logical}).
958: %
959: \end{aligned}
960: %
961: \end{equation}
962: %
963: Encoding-decoding pairs provide the link between a completely positive
964: map $P:\mathbf{T}(H_\mathrm{comp}) \rightarrow
965: \mathbf{T}(H_\mathrm{comp})$ corresponding to a physical device and a
966: unitary operator $U:H_\mathrm{logical} \rightarrow H_\mathrm{logical}$
967: corresponding to a quantum computation. We will
968: place no further restrictions on encoding-decoding pairs.
969: Now, suppose $\mathcal{M}_\mathrm{enc}, \mathcal{M}_\mathrm{dec}$ is an
970: encoding-decoding pair.  
971: Then the adjoint maps $\mathcal{M}_\mathrm{dec}^\mathrm{t}$,
972: $\mathcal{M}_\mathrm{enc}^\mathrm{t}$ are unit preserving completely
973: positive maps. (Adjoint maps of completely positive maps are defined
974: in \eqref{definition-of-dual} in Appendix \ref{mathematical-preliminaries-section}.)
975: 
976: \subsection{Damping and $\gamma$-damping}
977: Dissipative and decohering effects are consequences of {\em damping}.
978: To begin the development of the No-Go Theorem, we introduce
979: a mathematically precise definition of the damping
980: of a quantum mechanical system, to which we refer as $\gamma$-damping.
981: 
982: \begin{dfn}\label{gamma-damping}
983: %
984: Let $H$ be a Hilbert space. A completely positive trace-preserving map
985: $P:\mathbf{T}(H) \rightarrow \mathbf{T}(H)$ is {\em $\gamma$-damped}
986: iff there is an abelian von Neumann algebra $\mathfrak{A} \subseteq
987: \mathbf{L}(H)$ such that for all $T \in \mathbf{L}(H)$, there is
988: an $S_T \in \mathfrak{A}$ such that
989: %
990: \begin{equation}\label{gamping}
991: %
992: \|P^\mathrm{t}(T)- S_T\|_\infty \leq \gamma  \|T\|_\infty,
993: %
994: \end{equation}
995: %
996: \end{dfn}
997: %
998: \noindent where the operator norm $\|\cdot\|_\infty$ is
999: defined in Appendix~\ref{banach-spaces}.
1000: Note that larger values of $\gamma$ correspond to less damping of the system.
1001: 
1002: To make contact with the physically intuitive notion of
1003: damping, we apply this definition to the example of the simple
1004: harmonic oscillator subject to phase damping.
1005: For such an harmonic oscillator,
1006: the $ij$th element of the density matrix, $\rho_{ij}=\langle i|\rho
1007: |j\rangle$, decays exponentially as $e^{-\kappa (i-j)^2}$, where we are working
1008: in the basis of energy eigenstates identified by the labels $i$ and $j$.
1009: The quantity $\kappa$ is characteristic of the specific 
1010: oscillator and its coupling to the environment.
1011: The
1012: completely positive map $P$ transforms the initial, general density
1013: matrix for the system into a density matrix with exponentially-decaying
1014: off-diagonal elements. Under the influence of damping, as the
1015: off-diagonal states of the oscillator decay and approach zero, the
1016: density matrix converges to a diagonal density matrix
1017: in the $\{ ij\}$ basis specified above.
1018: This is true for all initial configurations of the oscillator,
1019: and thus the damping process tends to an abelian set of final configurations.
1020: (The damping
1021: parameter $\kappa$ that characterizes the decay of the off-diagonal
1022: elements of the density matrix is related to the quantity $\gamma$
1023: that appears in \eqref{gamping}.
1024: As noted above, larger values of $\gamma$ correspond to less damping and
1025: hence smaller values of $\kappa$.)
1026: 
1027: \subsection{No-Go Theorem for Encodings}
1028: 
1029: \noindent{We now state the main result of \S\ref{NGT}:
1030: {\em the Encoding No-Go Theorem}.}
1031: \begin{thm}[\bf {The Encoding No-Go Theorem}]\label{ngt1}~\\
1032: %
1033: Suppose
1034: that $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
1035: \mathcal{M}_\mathrm{dec},\alpha)$ holds. If
1036: $P:\mathbf{T}(H_\mathrm{comp}) \rightarrow
1037: \mathbf{T}(H_\mathrm{comp})$ is $\gamma$-damped and $2\gamma + \alpha < \sqrt{2}/4$,
1038: then $H_\mathrm{logical}$ has dimension $1$.
1039: %
1040: \end{thm}
1041: %
1042: \noindent To prove the Encoding No-Go Theorem, we first derive
1043: in \S\ref{ab_factor} a number of general mathematical
1044: results on completely positive maps. We then apply these
1045: specifically to obtain a proof of the Encoding No-Go Theorem in \S\ref{ng_proof} below.
1046: 
1047: \subsection{Completely Positive Maps with Abelian
1048: Factorizations}\label{ab_factor} 
1049: 
1050: We begin with the following lemma, which furnishes
1051: a superoperator version of the definition of $\gamma$-damping.
1052: For any abelian von Neumann algebra $\mathfrak{A} \subseteq
1053: \mathbf{L}(H)$ there is a unit-preserving completely positive
1054: projection operator $\mathbf{E}_\mathfrak{A}:\mathbf{L}(H) \rightarrow
1055: \mathfrak{A}$.  This fact is elementary, but also follows from
1056: injectivity~\cite{connes} of such algebras, in the case $H$ is
1057: separable.
1058: \begin{lem}\label{superop_gamma}
1059: %
1060: If $P$ is $\gamma$-damped, $\mathfrak{A}$ is as given in Definition~\ref{gamma-damping}
1061: and $\mathbf{E}_\mathfrak{A}:\mathbf{L}(H)
1062: \rightarrow \mathfrak{A}$ is a unit-preserving completely positive
1063: projection mapping $\mathbf{E}_\mathfrak{A}:\mathbf{L}(H) \rightarrow
1064: \mathfrak{A}$, then
1065: %
1066: \begin{equation}
1067: %
1068: \|P^\mathrm{t}(T) - \mathbf{E}_\mathfrak{A}P^\mathrm{t}(T)\|_\infty \leq 2 \gamma \| T\|_\infty.
1069: %
1070: \end{equation}
1071: %
1072: \end{lem}
1073: %
1074: \begin{proof}
1075: %
1076: Note that $\mathbf{E}_\mathfrak{A}$ is a linear mapping of norm $\leq
1077: 1$ and thus
1078: %
1079: \begin{equation}
1080: %
1081: \begin{aligned}
1082: %
1083: \|P^\mathrm{t}(T) - \mathbf{E}_\mathfrak{A}P^\mathrm{t}(T)\|_\infty & \leq \|P^\mathrm{t}(T) -
1084: S_T\|_\infty + \|S_T - \mathbf{E}_\mathfrak{A}P^\mathrm{t}(T)\|_\infty \\
1085: %
1086: &  \leq \gamma  \| T\|_\infty + \|\mathbf{E}_\mathfrak{A} S_T - \mathbf{E}_\mathfrak{A}P^\mathrm{t}(T)\|_\infty \\
1087: %
1088: & \leq 2 \gamma \| T\|_\infty,
1089: %
1090: %
1091: %
1092: \end{aligned}
1093: %
1094: \end{equation}
1095: %
1096: as claimed.
1097: \end{proof}
1098: 
1099: In the remainder of this section we prove an important
1100: technical result used in the proof of the no-go theorem, namely that
1101: completely positive maps $F: \mathbf{L}(H) \rightarrow \mathbf{L}(H)$,
1102: which factor through completely positive maps into abelian von Neumann
1103: algebras, cannot be used to approximate unitary operators $U$ on $H$
1104: if $\opr{dim}(H) \geq 2$. More precisely, we will show that for any
1105: $\beta < \sqrt{2}/4$, there is at least one non-zero operator $T$ for which
1106: %
1107: \begin{equation}\label{non-approximation-inequality}
1108: %
1109: \|U^\dagger T U - F(T) \|_\infty \geq \beta \  \| T\|_\infty.
1110: %
1111: \end{equation}
1112: %
1113: We first define what it means for a completely positive map to
1114: factor through an abelian von Neumann algebra:
1115: %
1116: \begin{dfn}\label{defn-of-abelian-factorization}
1117: %
1118: Let $A$ be a $\mathrm{C}^\ast$-algebra with a multiplicative unit.  A
1119: unit preserving completely positive map $F:A \rightarrow A$ has an
1120: {\em abelian factorization} (or briefly is abelian factorizable) iff
1121: there is an abelian von Neumann algebra $\mathfrak{B}$ and unital
1122: completely positive maps $Q:A \rightarrow \mathfrak{B}$, $R :
1123: \mathfrak{B} \rightarrow A$ such that $F$ is the composition $R \circ
1124: Q$.
1125: %
1126: \end{dfn}
1127: If $F: \mathbf{L}(H) \rightarrow \mathbf{L}(H)$ is abelian
1128: factorizable, then it follows from the definition that for any
1129: unit-preserving completely positive maps $Q, R$, the completely
1130: positive map $Q \circ F \circ R$ is abelian factorizable.
1131: We begin by providing a characterization of abelian factorizable maps.
1132: \begin{prop}
1133: %
1134: Let $A$ be an arbitrary $\mathrm{C}^\ast$-algebra with multiplicative
1135: unit.  A unit preserving completely postive map $F: A \rightarrow A$
1136: factors through a finite-dimensional abelian von-Neumann algebra iff
1137: there are unit preserving positive linear functionals $\rho_1, \ldots,
1138: \rho_m \in A^\dagger$ and positive elements $G_1,
1139: \ldots, G_m \in A$ of norm $\leq 1$, such that $\sum_{i=1}^m G_i = I$
1140: and
1141: %
1142: \begin{equation}\label{finite-factorization-equation}
1143: %
1144: F(T) = \sum_{i=1}^m \rho_i(T) G_i \quad \forall T \in A.
1145: %
1146: \end{equation}
1147: %
1148: \end{prop}
1149: %
1150: \begin{proof}
1151: %
1152: We first note that positive maps from an abelian
1153: $\mathrm{C}^\ast$-algebra or into an abelian $\mathrm{C}^\ast$-algebra
1154: are automatically completely positive (see~\cite{stinespring1955},\cite{choi1972}).
1155: Thus the map $F$ given by
1156: Equation~\eqref{finite-factorization-equation} is completely positive.
1157: Let $\mathfrak{B}$, $Q:A \rightarrow \mathfrak{B}$, $R : \mathfrak{B}
1158: \rightarrow A$ as in Definition~\ref{defn-of-abelian-factorization},
1159: but with $\mathfrak{B}$ finite dimensional and let $E_1, \ldots, E_m$
1160: be the minimal non-zero projections of $\mathfrak{B}$.  Then
1161: %
1162: \begin{equation}
1163: %
1164: F(T) = R\biggl(\sum_{i=1}^m E_i Q(T) E_i\biggr) = R\biggl(\sum_{i=1}^m
1165: \rho_i(T) E_i\biggr) = \sum_{i=1}^m \rho_i(T) R(E_i).
1166: %
1167: \end{equation}
1168: %
1169: Letting $G_i = R(E_i)$,
1170: \begin{equation}
1171: %
1172: \sum_{i=1}^m G_i = \sum_{i=1}^m R(E_i) = R\biggl(\sum_{i=1}^m E_i\biggr) = I.
1173: %
1174: \end{equation}
1175: This completes the proof.
1176: \end{proof}
1177: In general, unit-preserving completely positive maps with arbitrary
1178: abelian factorizations can be approximated by maps of the
1179: form~\eqref{finite-factorization-equation}.
1180: \begin{prop}\label{approximation-result-for-abelain-factorizable}
1181: %
1182: If a completely positive map $F:A \rightarrow A$ has an abelian
1183: factorization, then there is a generalized sequence of maps
1184: $\{F_\kappa\}_{\kappa \in K}$ each having the form $F_\kappa(T) =
1185: \sum_{i=1}^m \rho_i(T) G_i$ which converges to $F$ in the point-norm
1186: topology, that is for each $T \in A$,
1187: %
1188: $F_\kappa(T) \rightarrow F(T)$ in the norm of $A$.
1189: %
1190: \end{prop}
1191: %
1192: \begin{proof}
1193: %
1194: Let $\mathfrak{B}$, $Q:A \rightarrow \mathfrak{B}$, $R : \mathfrak{B}
1195: \rightarrow A$ as in
1196: Definition~\ref{defn-of-abelian-factorization}. Given $T_1, \ldots,
1197: T_m \in A$, let $\mathfrak{B}_0$ be a finite dimensional abelian
1198: von-Neumann subalgebra of $\mathfrak{B}$ and $\mathbf{E}$ a linear
1199: projection operator\footnote{These operators are sometimes referred to
1200: as {\em conditional expectations}.} $\mathfrak{B} \rightarrow
1201: \mathfrak{B}_0$ such that \begin{equation}
1202: %
1203: \|Q(T_i) - \mathbf{E}(Q(T_i))\|_\infty \leq \epsilon.
1204: %
1205: \end{equation}
1206: %
1207: Since $Q$ is contractive, it follows that
1208: %
1209: \begin{equation}
1210: %
1211: \|F(T_i) - R \circ \mathbf{E} \circ Q(T_i) \|_\infty \leq \epsilon.
1212: %
1213: \end{equation}
1214: %
1215: Now apply the preceding result.
1216: \end{proof}
1217: 
1218: Assume $H$ is a finite dimensional Hilbert space. We will consider
1219: trace functionals on two distinct spaces of operators: one on the
1220: space $\mathbf{L}(H)$, which we denote by $\opr{tr}$, and the other on
1221: the space $\mathbf{L}(\mathbf{L}(H))$ of linear mappings
1222: $\mathbf{L}(H) \rightarrow \mathbf{L}(H)$, which we denote
1223: $\opr{tr}_\mathrm{oper}$.
1224: We will prove that abelian factorizable completely positive maps cannot
1225: approximate the identity map on $\mathbf{L}(H)$. To do this we will show that the trace
1226: functional $\opr{tr}_\mathrm{oper}$ separates, in a sense to be made
1227: precise in the next paragraph, the identity operator on
1228: $\mathbf{L}(H)$ from abelian factorizable completely positive $F$.
1229: Note that $\opr{tr}_\mathrm{oper}(I_{\mathbf{L}(H)}) =
1230: \opr{dim}^2(H)$.
1231: 
1232: \begin{lem}\label{trace-separation-lemma} Suppose the Hilbert space
1233: $H$ has finite dimension $n$. For any unit-preserving
1234: abelian-factorizable completely positive map $F:\mathbf{L}(H)
1235: \rightarrow \mathbf{L}(H)$,
1236: %
1237: \begin{equation}\label{separation-equation}
1238: %
1239: \opr{tr}_\mathrm{oper}(F) \leq n.
1240: %
1241: \end{equation}
1242: %
1243: %
1244: \end{lem}
1245: %
1246: \begin{proof}
1247: %
1248: It suffices to prove this for $F$ which have the
1249: form~\eqref{finite-factorization-equation}. Referring to that
1250: representation, each positive functional $\rho_i$ can be represented
1251: by a non-negative operator $S_i$ as follows:
1252: %
1253: \begin{equation}
1254: %
1255: \rho_i(T) = \opr{tr}(T S_i), 
1256: %
1257: \end{equation}
1258: %
1259: Since $\rho_i(I)=1$, $S_i$ also has unit trace and in particular,
1260: $S_i\leq I$.  Moreover
1261: %
1262: $\sum_{i=1}^m G_i = I_H$. 
1263: %
1264: Now
1265: \begin{equation}
1266: %
1267: \begin{aligned}
1268: %
1269: \opr{tr}_{\mathrm{oper}}(P)& = \sum_{i=1}^m \opr{tr}(S_i G_i) \\
1270: %
1271: & = \sum_{i=1}^m \opr{tr}(G_i^{1/2} S_i G_i^{1/2}) \\
1272: %
1273: & \leq \sum_{i=1}^m \opr{tr}(G_i^{1/2} G_i^{1/2}) \\
1274: %
1275: & = \sum_{i=1}^m \opr{tr}(G_i) = \opr{tr}(I_H) = n.
1276: %
1277: \end{aligned}
1278: %
1279: \end{equation}
1280: \end{proof}
1281: %
1282: The previous lemma gives us a lower bound on the trace of $I -F$ for $F$ abelian factorizable:
1283: %
1284: \begin{equation}
1285: %
1286: \opr{tr}_\mathrm{oper}(I - F) \geq n^2 - n.
1287: %
1288: \end{equation}
1289: %
1290: 
1291: We can use the above lower bound on the trace of $I -F$ to obtain a
1292: lower bound on the norm of the operator $I - F:\mathbf{L}(H)
1293: \rightarrow \mathbf{L}(H)$, where we consider $\mathbf{L}(H)$ with the
1294: Schatten $2$-norm $\| \cdot\|_{2}$, defined in
1295: Appendix~\ref{banach-spaces}.  The Schatten $2$-norm is the norm that
1296: arises from the trace inner product on $\mathbf{L}(H)$, also known as
1297: the Hilbert-Schmidt norm. We denote the corresponding operator norm on
1298: $\mathbf{L}(H) \rightarrow \mathbf{L}(H)$ by $\| \cdot\|_{2\rightarrow
1299: 2}$.  
1300: 
1301: If $F$ is self-adjoint as an operator on the space $\mathbf{L}(H)$ with the
1302: trace inner product,  then from Lemma~\ref{trace-separation-lemma}, we
1303: immediately obtain the bound
1304: %
1305: \begin{equation} \label{2-2-norm-bounded}
1306: %
1307: \|I - F\|_{2\rightarrow 2} \geq 1 - \frac{1}{n}.
1308: %
1309: \end{equation} 
1310: %
1311: In fact, the lower bound~\eqref{2-2-norm-bounded} is true for general
1312: unit-preserving abelian factorizable completely positive maps $F$. To
1313: see this, write $F = F_\mathfrak{Re} + i F_\mathfrak{Im}$ where both
1314: $F_\mathfrak{Re}$, $F_\mathfrak{Im}$ are self-adjoint operators (not
1315: necessarily completely positive, however).  Now
1316: \begin{equation}
1317: %
1318: \opr{tr}_\mathrm{oper}(I - F_\mathfrak{Re}) = \opr{tr}_\mathrm{oper}(I - F) \geq n^2 - n.
1319: %
1320: \end{equation}
1321: %
1322: Therefore 
1323: %
1324: \begin{equation}
1325: %
1326: \|I - F\|_{2\rightarrow 2} \geq \|I - F_\mathfrak{Re}\|_{2\rightarrow 2} \geq 1 - 1/n.
1327: %
1328: \end{equation}
1329: %
1330: 
1331: In the discussion that follows, we need to consider another norm on
1332: the space $\mathbf{L}(H) \rightarrow \mathbf{L}(H)$ in addition to the
1333: norm $\| \cdot\|_{2\rightarrow 2}$ just considered.  The new norm,
1334: which we denote $\| \cdot \|_{\infty \rightarrow \infty}$, is also an
1335: operator norm on $\mathbf{L}(H) \rightarrow \mathbf{L}(H)$, but
1336: relative to the $\| \cdot \|_\infty$ norm on $\mathbf{L}(H)$. The $\|
1337: \cdot \|_{\infty \rightarrow \infty}$ norm is different from the $\|
1338: \cdot\|_{2\rightarrow 2}$ norm, but for finite dimensional spaces $H$
1339: the two norms are equivalent, which means that each norm is
1340: bounded relative to the other.  To obtain the bounding constants, note
1341: %
1342: that if $T \in \mathbf{L}(H)$,
1343: %
1344: \begin{equation}
1345: %
1346: \|T\|_\infty \leq \|T\|_2 \leq \sqrt{n} \|T\|_\infty.
1347: %
1348: \end{equation}
1349: %
1350: From this it follows that 
1351: \begin{equation}
1352: %
1353: \frac{1}{\sqrt{n}} \, \|F\|_{2\rightarrow 2}
1354: \leq \|F\|_{\infty\rightarrow\infty} \leq \sqrt{n} \,
1355: \|F\|_{2\rightarrow 2}.
1356: %
1357: \end{equation}
1358: %
1359: which is the desired relative bound.  In~\eqref{2-2-norm-bounded}, if
1360: we substitute $F$ by $I - F$, we immediately obtain the following
1361: proposition:
1362: %
1363: \begin{prop}\label{key-estimate}
1364: %
1365: Suppose $H$ is a Hilbert space of finite dimension $n$. If a
1366: unit-preserving completely positive map $F:\mathbf{L}(H) \rightarrow
1367: \mathbf{L}(H)$ has the form~\eqref{finite-factorization-equation}, then
1368: %
1369: \begin{equation}\label{minimum-conclusion}
1370: %
1371: \|I - F\|_{\infty\rightarrow\infty} \geq n^{-1/2}(1 - 1/n).
1372: %
1373: \end{equation}
1374: %
1375: \end{prop}
1376: %
1377: To prove the crucial result for the No-Go Theorem, we only use case
1378: $n=2$ of~\eqref{minimum-conclusion}.
1379: %
1380: \begin{thm}\label{quantifiable-constraint-of-factorizable}
1381: %
1382: Suppose $H$ is of dimension $\geq 2$. If $F$ is a unit-preserving
1383: completely positive map on $\mathbf{L}(H)$ with an abelian
1384: factorization and $\beta < \sqrt{2}/4$, then
1385: %
1386: \begin{equation}\label{equation-of-approximation}
1387: %
1388: \|U^\dagger T U - F(T) \|_\infty \geq \beta \  \|T\|_\infty
1389: %
1390: \end{equation}
1391: %
1392: for at least one $T \in \mathbf{L}(H)$.  
1393: %
1394: 
1395: If $H$ is finite dimensional, we can take $\beta = \sqrt{2}/4$.
1396: %
1397: \end{thm}
1398: %
1399: \begin{proof}
1400: %
1401: Replacing $F$ by the completely positive map 
1402: %
1403: $T \mapsto U F(T) U^\dagger$, we can assume without loss of generality that $U=I$.
1404: %
1405: To prove this, we will show that the assertion that 
1406: %
1407: \begin{equation} \label{wrong-assertion-to-be-refuted}
1408: %
1409: \|T - F(T) \|_\infty < \beta \ \|T\|_\infty, \quad \forall T \in \mathbf{L}(H),
1410: %
1411: \end{equation}
1412: leads to a contradiction. However,~\eqref{wrong-assertion-to-be-refuted} implies
1413: %
1414: \begin{equation}\label{why-<-becomes-leq}
1415: %
1416: \| I - F \|_{\infty \rightarrow \infty} \leq \beta.
1417: %
1418: \end{equation}
1419: %
1420: We now reduce the proof to the case $H$ has dimension $2$, by
1421: considering a Hilbert space $K$ of dimension $2$ and completely
1422: positive unit-preserving mappings $Q:\mathbf{L}(K) \rightarrow
1423: \mathbf{L}(H)$ and $R: \mathbf{L}(H) \rightarrow \mathbf{L}(K)$ such
1424: that $R \circ Q$ is the identity map on $\mathbf{L}(K)$.
1425: %
1426: If $I_H$ can be approximated to within $\beta$ of the abelian factorizable
1427: $F$, then $R  F  Q$ is also abelian factorizable and
1428: %
1429: \begin{equation}
1430: %
1431: \|I_K - R F Q\| = \|R I_H Q - R F Q\| \leq \|I_H - F\| \leq \beta < \sqrt{2}/4.
1432: %
1433: \end{equation}
1434: %
1435: However, in this contradicts Proposition~\ref{key-estimate}.
1436: %
1437: 
1438: In the finite dimensional case, the norm is actually achieved so that
1439: in~\eqref{why-<-becomes-leq} the $\leq$ sign can be replaced by $<$
1440: and so we can take $\beta \leq \sqrt{2}/4$ as claimed.\footnote{
1441: We note that it is possible to derive explicit results as well for the
1442: case $n=3$. In this case one can show that one obtains a larger numerical bound than
1443: $\sqrt{2}/4$, which applies when dim($H_{\mathrm{logical}})\geq 3$.}
1444: \end{proof}
1445: 
1446: 
1447: \subsection{Proof of Encoding No-Go Theorem}\label{ng_proof}
1448: 
1449: \subsubsection{Two Lemmas} \begin{lem}\label{triangle_lemma} Suppose
1450: that $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
1451: \mathcal{M}_\mathrm{dec},\alpha)$ holds.
1452: If $P:\mathbf{T}(H_\mathrm{comp}) \rightarrow
1453: \mathbf{T}(H_\mathrm{comp})$ is $\gamma$-damped and
1454: $\mathfrak{A}$ and $\mathbf{E}_\mathfrak{A}$ are as identified
1455: in Lemma~\ref{superop_gamma}, then for every $T
1456: \in\mathbf{L}(H_\mathrm{logical})$,
1457: %
1458: \begin{equation} 
1459: %
1460: %
1461: \|\mathcal{M}_\mathrm{enc}^\mathrm{t}   \mathbf{E}_\mathfrak{A} 
1462: P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}(T)  - U^\dagger T  U \|_\infty
1463: \leq (2 \gamma+\alpha) \|T\|_\infty. 
1464: %
1465: \end{equation}
1466: %
1467: \end{lem}
1468: %
1469: \begin{proof} The $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
1470: \mathcal{M}_\mathrm{dec},\alpha)$ implies
1471: % 
1472: that for every $\rho \in \mathbf{T}(H_\mathrm{logical})$ with
1473: $\|\rho\|_1 \leq 1$ and self-adjoint $T \in
1474: \mathbf{L}(H_\mathrm{logical})$,
1475: %
1476: \begin{equation}
1477: %
1478: \begin{aligned}
1479: %
1480: \biggl|\opr{tr}  \biggl[\biggl\{\mathcal{M}_\mathrm{dec} (P
1481: (\mathcal{M}_\mathrm{enc}(\rho)))  - U \rho U^\dagger\biggr\} T\biggr]\biggr| 
1482: %
1483: & = \biggl|\opr{tr} \biggl[\rho \biggl\{
1484: \mathcal{M}_\mathrm{enc}^\mathrm{t} (P^\mathrm{t}
1485: (\mathcal{M}_\mathrm{dec}^\mathrm{t}(T))) - U^\dagger T U \biggr\} \biggr] \biggr| \\
1486: %
1487: & \leq \alpha \|T\|_\infty.
1488: %
1489: \end{aligned}
1490: %
1491: \end{equation}
1492: % 
1493: It follows that for every $T \in \mathbf{L}(H_\mathrm{logical})$,
1494: %
1495: \begin{equation}
1496: %
1497: \|\mathcal{M}_\mathrm{enc}^\mathrm{t} (P^\mathrm{t}
1498: (\mathcal{M}_\mathrm{dec}^\mathrm{t}(T))) - U^\dagger T U \|_\infty
1499: \leq \alpha \|T\|_\infty.
1500: %
1501: \end{equation}
1502: %
1503: $\mathfrak{A}$ is commutative and by Lemma~\ref{superop_gamma}, for each $T \in
1504: \mathbf{L}(H_\mathrm{logical})$ 
1505: %
1506: \begin{equation}
1507: %
1508: \|P^\mathrm{t} (T) - \mathbf{E}_\mathfrak{A} P^\mathrm{t}(T)\|_\infty \leq 2 \gamma
1509: \|T\|_\infty.
1510: %
1511: \end{equation}
1512: %
1513: Thus using the fact that $\mathcal{M}_\mathrm{dec}$ and
1514: $\mathcal{M}_\mathrm{enc}$ have norm $\leq 1$, for every $T \in
1515: \mathbf{L}(H_\mathrm{logical})$, 
1516: %
1517: \begin{equation} \label{violated-condition}
1518: %
1519: \begin{aligned}
1520: %
1521: \|\mathcal{M}_\mathrm{enc}^\mathrm{t}  & \mathbf{E}_\mathfrak{A} 
1522: P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}(T)  - U^\dagger T  U \|_\infty \\
1523: %
1524: & \leq \|\mathcal{M}_\mathrm{enc}^\mathrm{t} \mathbf{E}_\mathfrak{A}
1525:  P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}(T) -
1526: \mathcal{M}_\mathrm{enc}^\mathrm{t} P^\mathrm{t}\mathcal{M}_\mathrm{dec}^\mathrm{t}(T)\|_\infty \\
1527: & + \|
1528: \mathcal{M}_\mathrm{enc}^\mathrm{t}
1529: P^\mathrm{t}\mathcal{M}_\mathrm{dec}^\mathrm{t}(T) - U^\dagger T U
1530: \|_\infty \\
1531: %
1532: & \leq (2 \gamma+\alpha) \|T\|_\infty. \end{aligned}
1533: %
1534: \end{equation}
1535: %
1536: \end{proof}
1537: %
1538: \begin{lem}\label{fac_lem}
1539: If $H_{\mathrm{logical}}$ is of dimension $\geq 2$, and $P$, $U$, $\mathcal{M}_\mathrm{enc}$,
1540: $\mathcal{M}_\mathrm{dec}$ and $\mathbf{E}_\mathfrak{A}$
1541: are the same as in Lemma \ref{triangle_lemma}, and if $\beta <  \sqrt{2}/4$, 
1542: then for some non-zero $T \in \mathbf{L}(H_\mathrm{logical})$,
1543: %
1544: \begin{equation}\label{special_number}
1545: %
1546: \| \mathcal{M}_\mathrm{enc}^\mathrm{t} \mathbf{E}_\mathfrak{A} 
1547: P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}(T)  - U^\dagger T  U \|_\infty
1548: \geq \beta\| T \|_\infty
1549: %
1550: \end{equation}
1551: %
1552: \end{lem}
1553: \begin{proof}
1554: The unit-preserving completely positive map
1555: $R=\mathcal{M}_\mathrm{enc}^\mathrm{t} \mathbf{E}_\mathfrak{A}
1556: P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}$ factors through the
1557: abelian von Neumann algebra $\mathfrak{A}$; this follows from the
1558: presence of the projection operator $\mathbf{E}_\mathfrak{A}$ in the
1559: expression for $R$. Then \eqref{special_number} follows from
1560: Theorem \ref{quantifiable-constraint-of-factorizable}.
1561: \end{proof}
1562: 
1563: \subsubsection{Statement of Proof of Encoding No-Go Theorem}
1564: %
1565: \begin{proof}
1566: %
1567: By the hypotheses of the Encoding No-Go Theorem (Theorem \ref{ngt1}),
1568: the quantity $2 \gamma+\alpha < \sqrt{2}/4$. Choose $\beta$ such that
1569: $2 \gamma + \alpha < \beta < \sqrt{2}/4$.  From Lemma
1570: \ref{triangle_lemma}, it follows that for every $T \in
1571: \mathbf{L}(H_\mathrm{logical})$,
1572: %
1573: \begin{equation}\label{first-step-in-proof-by-contradiction-of-no-go}
1574: %
1575: \| \mathcal{M}_\mathrm{enc}^\mathrm{t} \mathbf{E}_\mathfrak{A}
1576: P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}(T) - U^\dagger T U
1577: \|_\infty < (2 \gamma + \alpha) \| T \|_\infty.  
1578: %
1579: \end{equation}
1580: %
1581: On the other hand by \eqref{special_number} in Lemma \ref{fac_lem},
1582: there is a non-zero $T$ such that
1583: %
1584: \begin{equation}\label{second-step-in-proof-by-contradiction-of-no-go}
1585: %
1586: \| \mathcal{M}_\mathrm{enc}^\mathrm{t} \mathbf{E}_\mathfrak{A}
1587: P^\mathrm{t} \mathcal{M}_\mathrm{dec}^\mathrm{t}(T) - U^\dagger T U
1588: \|_\infty \geq \beta \| T \|_\infty.  
1589: %
1590: \end{equation}
1591: %
1592: Since $\| T\|_\infty >  0$,
1593: equations~\eqref{first-step-in-proof-by-contradiction-of-no-go}
1594: and \eqref{second-step-in-proof-by-contradiction-of-no-go}
1595: imply $\beta \leq 2 \gamma + \alpha$, which contradicts the
1596: choice of $\beta$.
1597: %
1598: \end{proof}
1599: 
1600: 
1601: \subsection{Interpretation of Encoding No-Go Theorem}
1602: 
1603: The Encoding No-Go Theorem is an extremely powerful application of the QCC.
1604: With this theorem, one can calculate the amount of damping for which
1605: fault-tolerant quantum computation becomes impossible.
1606: When the amount of damping exceeds the critical amount, which means that the
1607: value of $2\gamma$ becomes less than the critical
1608: value $2\gamma_{\mathrm{critical}} = \sqrt{2}/4 - \alpha$, we find
1609: that the only solutions to the QCC are those for which dim $H_{\mathrm{logical}}=1$.
1610: As noted above, in this case no meaningful quantum computation is possible, since
1611: not even one quantum bit can be accomodated.
1612: 
1613: In order to analyze the constraints on solutions to
1614: $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec},\alpha)$
1615: implied by the No-Go Theorem,
1616: we regard the pair $\{ U, \alpha\}$ as given, since both the
1617: desired quantum computation $U$, and the maximum acceptable implementation
1618: inaccuracy $\alpha$, are prescribed.
1619: We assume that $U$ is defined on a Hilbert space $H_{\mathrm{logical}}$ such that
1620: ${\mathrm{dim}}(H_{\mathrm{{logical}}})\geq 2$, to allow meaningful quantum
1621: computation.
1622: For practical applications, one then seeks to determine triples
1623: $\{ P,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec}\}$
1624: that satisfy $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec},\alpha)$.
1625: The No-Go Theorem, on the other hand,
1626: identifies conditions in which
1627: $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec},\alpha)$
1628: is not satisfied.
1629: Thus the No-Go Theorem provides a useful calculational tool for eliminating prospective
1630: quantum computer implementations that are guaranteed to fail.
1631: With the criterion provided by the No-Go Theorem, we can bound the space of
1632: solutions to $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec},\alpha)$.
1633: This allows exploration of trade-offs amongst the members of the triple
1634: $\{ P,\mathcal{M}_\mathrm{enc},\mathcal{M}_\mathrm{dec}\}$, with which one
1635: may construct generalized ``phase diagrams" that indicate boundaries between
1636: {\em potentially acceptable}
1637: and {\em definitely unacceptable} values for $P$, $\mathcal{M}_\mathrm{enc}$ and
1638: $\mathcal{M}_\mathrm{dec}$.
1639: 
1640: 
1641: \section{Quantum Components}\label{qm_components}
1642: 
1643: \subsection{Introduction}
1644: 
1645: As defined in Section \ref{QCC_section}, we refer to a physically realizable device
1646: intended to implement a quantum computation as a {\em quantum component.} Mathematically,
1647: a quantum component is represented by a completely positive, trace preserving map, $P$.
1648: The detailed form of $P$, as an explicit function, is dictated by underlying
1649: equations of motion. 
1650: 
1651: For a closed physical system, the quantum mechanical dynamics of the system
1652: are given by the Schr\"odinger
1653: equation associated to a particular Hamiltonian
1654: $\mathcal H$. However, for the general problem of a {\em practical}
1655: quantum computer, we must analyze realistic quantum
1656: components that interact with their environment, dissipate heat and exhibit decoherence.
1657: We must thus utilize a formulation that yields equations of motion
1658: appropriate to open quantum mechanical systems.
1659: 
1660: The connection
1661: to the Quantum Computer Condition presented in Section \ref{QCC_section} is made by
1662: starting with the appropriate equations of motion that describe the dynamical
1663: evolution of a realistic quantum component interacting with its environment.
1664: One proceeds by solving the appropriate equations of motion. The explicit
1665: solution thus obtained
1666: provides the time-dependence of the quantum mechanical state of the open
1667: system. In principle, this
1668: allows us to deduce the explicit functional form of $P$.
1669: 
1670: Formulations of equations of motion for quantum components
1671: that interact with complex
1672: environments comprised of many degrees-of-freedom
1673: necessarily involve approximations of one sort or another, and there is not
1674: in general a unique choice. In this section we illustrate the general approach by
1675: making use of a Lindblad-type equation (made more precise below) to describe how
1676: one obtains the quantum component $P$ that appears in the Quantum Computer Condition.
1677: 
1678: 
1679: \subsection{Dynamical Equations of Motion}
1680: 
1681: \subsubsection{Time-dependent generalization of the Lindblad equation} The state of a
1682: quantum mechanical system defined on a Hilbert space $H$
1683: can be modeled by a density operator $\rho(t)$ whose time dependence obeys a
1684: linear (time dependent) equation
1685: %
1686: \begin{equation} \label{generic-equation-of-evolution}
1687: %
1688: \frac{d}{dt} \rho (t) = A(t) \rho(t),
1689: %
1690: \end{equation}
1691: %
1692: where $A(t)$ is an operator acting on $\mathbf{T}(H)$, the Banach space of trace class
1693: operators on $H$.
1694: For a closed system we have the Schr\"odinger equation, and
1695: the right-hand-side of \eqref{generic-equation-of-evolution} is given by 
1696: %
1697: \begin{equation} \label{SchrodingerOperator}
1698: %
1699: A(t) \rho = -\frac{i}{\hbar} [\mathcal{H}(t), \rho],
1700: %
1701: \end{equation}
1702: %
1703: where the Hamiltonian $\mathcal{H}(t)$ is a self-adjoint operator which may depend
1704: on the parameter $t$.
1705: 
1706: However, it would be impractical to describe realistic quantum components using
1707: \eqref{SchrodingerOperator}, since writing down the detailed Hamiltonian operator
1708: to account for all of the degrees-of-freedom comprising the quantum component
1709: and its environment would be intractable.
1710: 
1711: Motivated by the work of Lindblad~\cite{lindblad76}, we make use instead of the following
1712: expression for the action of $A(t)$ on $\rho$:
1713: %
1714: \begin{equation} \label{LindbladOperator}
1715: %
1716: A(t) \rho = -\frac{i}{\hbar} [\mathcal{H}(t), \rho] + \sum_j \bigg[ L_j(t)
1717: \rho L_j^\dag(t) - \frac{1}{2}\big\{L_j^\dag(t) L_j(t), \rho\big\}\bigg]
1718: %
1719: \end{equation}
1720: %
1721: where as above the $\mathcal{H}(t)$ is a self-adjoint operator which may depend
1722: on the parameter $t$, and the $L_j$ are operators
1723: that describe effects arising from interaction with the environment, such as dissipation
1724: and decoherence.
1725: %
1726: These operators are generalizations of the Lindblad operators
1727: of~\cite{lindblad76}; unlike the treatment given by Lindblad in~\cite{lindblad76},
1728: however,
1729: which considered only time-independent Hamiltonians $\mathcal{H}$ and
1730: time-independent dissipative perturbations $L_j$, with all operators bounded,
1731: we will allow unbounded and
1732: time-dependent $\mathcal{H}(t)$ and time-dependent (but still bounded) dissipative
1733: perturbations $L_j(t)$. Our treatment generalizes that given in~\cite{lindblad76},
1734: and we will refer to the equation of evolution
1735: %
1736: \begin{equation} \label{LindbladEquation}
1737: %
1738: \frac{d}{dt}\rho(t) =  -\frac{i}{\hbar} [\mathcal{H}(t), \rho] + \sum_j \bigg[ L_j(t)
1739: \rho L_j^\dag(t) - \frac{1}{2}\big\{L_j^\dag(t) L_j(t), \rho\big\}\bigg]
1740: %
1741: \end{equation}
1742: %
1743: as the {\em time-dependent generalization of the Lindblad equation}.
1744: As an example of how one may approximate
1745: the dynamics of a quantum component interacting with a complex environment,
1746: this equation provides a starting point to the derivation of $P$ used in the
1747: Quantum Computer Condition.
1748: For this purpose we need to consider solutions to equations of the type given in
1749: \eqref{LindbladEquation} known as ``fundamental solutions."
1750: 
1751: 
1752: \subsection{Fundamental Solutions}
1753: 
1754: The concept of a {\em fundamental solution} associated to an evolution equation
1755: (see~\cite{tanabe}, \S4.4) of the
1756: general form~\eqref{generic-equation-of-evolution}, such
1757: as \eqref{LindbladEquation} in particular,
1758: %
1759: where $\rho$ is a function with values in the Banach space
1760: $\mathbf{T}(H)$ and $A(t)$ is a one-parameter family of (possibly
1761: unbounded) linear operators on $\mathbf{T}(H)$, will play an important
1762: role in this paper.
1763: A {\em fundamental solution} associated to an equation of motion is a solution of an
1764: operator version of the original equation. We show below how, given a fundamental
1765: solution, one obtains the completely positive, trace preserving map $P$ that appears
1766: in the Quantum Computer Condition.
1767: 
1768: Fundamental solutions are given by a family
1769: $\{P_{t,s}\}_{t \geq s \geq 0}$
1770: of bounded operators on $\mathbf{T}(H)$ indexed on pairs of real
1771: numbers $t$ and $s$ satisfying equations
1772: %
1773: \begin{equation}\label{fundamental-solution-def}
1774: %
1775: \frac{d}{dt}P_{t,s} = A(t) P_{t,s} 
1776: %
1777: \end{equation}
1778: %
1779: and
1780: %
1781: \begin{equation}
1782: %
1783: P_{s,s} = I.
1784: \end{equation}
1785: %
1786: The intended
1787: interpretation of $P_{t,s}$ is that if the system is in state
1788: $\rho$ at time $s$, then the system will be in state $P_{t,s}\cdot\rho$
1789: at later time $t$, that is
1790: %
1791: \begin{equation}\label{evolve}
1792: %
1793: \rho(t) = P_{t,s}\cdot\rho(s).
1794: %
1795: \end{equation}
1796: %
1797: 
1798: 
1799: \subsubsection{Existence and Positivity Properties of Fundamental Solutions}
1800: 
1801: 
1802: The problem of the existence and uniqueness of fundamental solutions is a
1803: central one in the mathematical theory of evolution equations.
1804: In addition to addressing the question of the existence of fundamental solutions, it is
1805: important for our analysis to determine the positivity properties of
1806: fundamental solutions.
1807: This is because, as we shall see,
1808: the complete positivity of the map $P$ that explicitly
1809: appears in the Quantum Computer Condition is inherited from
1810: the complete positivity of the set of fundamental solutions $\{P_{t,s}\}$.
1811: Positivity is important in ensuring
1812: that the set $\{P_{t,s}\}$, as well as $P$, carry density matrices to density
1813: matrices.
1814: 
1815: For equations of motion associated to
1816: finite dimensional systems, existence and uniqueness of solutions
1817: follows from the Lipschitz theorem on ordinary differential
1818: equations.
1819: Lindblad's analysis extended to infinite-dimensional (but bounded) systems, so,
1820: more generally, Lindblad characterized the infinitesimal
1821: generator of a {\em norm continuous} completely positive
1822: semigroup~\cite{lindblad76},
1823: which corresponds to the case of~\eqref{generic-equation-of-evolution}
1824: in which $A(t)$ is constant and norm bounded
1825: (but possibly infinite-dimensional), {\em i.e.}, for the standard,
1826: time-independent Lindblad equation.
1827: 
1828: In order to generalize Lindblad's analysis to allow us to study
1829: infinite-dimensional, unbounded,
1830: time-dependent quantum systems, we will need to consider general evolution
1831: equations~\eqref{generic-equation-of-evolution} in which $A(t)$ may be
1832: unbounded as well as time-dependent, for the infinite-dimensional case.
1833: The general theory of such
1834: equations was developed by Kato, Yosida and others in the 1950's.
1835: We will rely on results of Kato~\cite{kato} which pertain to both existence of solutions
1836: and positivity properties, and on Theorem
1837: 4.4.1 of~\cite{tanabe} which pertains to existence of solutions;
1838: moreover, we will use a constructive form of
1839: the theorem (which follows from an examination of the proof) which
1840: expresses the fundamental solution as a limit of a product of
1841: exponentials. The basic assumption of the approach is that if the $A(t)$ are generators
1842: with sufficiently smooth variation, then fundamental solutions exist.
1843: 
1844: Our system comprised of a quantum component interacting with its environment,
1845: described by \eqref{LindbladEquation}, consists of a time-dependent
1846: Hamiltonian $\mathcal{H}(t)$ characterized by a time-dependent perturbation 
1847: $V(t)$ of a (possibly unbounded) self-adjoint operator $\mathcal{H}_0$ so that
1848: we have $\mathcal{H}(t) = \mathcal{H}_0 + V(t)$. In order to
1849: apply Kato's theory for time dependent evolution equations, we will
1850: assume among other things that the perturbation $V(t)$ does not change the
1851: domain of $\mathcal{H}_0$.  For the necessary background
1852: see~(\cite{kato},\cite{tanabe}).
1853: We now state a proposition that asserts the existence and complete positivity
1854: of fundamental solutions
1855: to operator versions of the time-dependent generalization of
1856: the Lindblad equation given in \eqref{LindbladEquation}:
1857: 
1858: \begin{prop}
1859: %
1860: For suitably regular time-varying potentials $V(t)$ and dissipation
1861: operators $L_j(t)$, there exists a strongly continuous completely positive
1862: operator $P_{t,s}$ which is a fundamental solution to the time-dependent
1863: generalization of the Lindblad equation.
1864: %
1865: \end{prop}
1866: %
1867: The exact statement of the conditions for the above in the form of
1868: a theorem, and proof, are given in Appendix \ref{mathematical-preliminaries-section}
1869: in \S\ref{solve_lindblad} and \S\ref{existence-and-uniqueness}.
1870: 
1871: \subsection{Construction of Quantum Components}
1872: 
1873: Having established the existence and complete positivity of
1874: fundamental solutions to the operator
1875: form of the underlying equation of motion for our system (comprised of
1876: the quantum component interacting with its environment),
1877: it is straightforward to
1878: obtain an expression for the completely positive, trace preserving map $P$,
1879: characterizing the quantum component, that explicitly
1880: appears in the QCC. This is simply obtained by noting ({\em cf} \eqref{evolve}) that the
1881: time-evolution of the state $\rho(t)$ between an initial fixed
1882: time $\hat s$ (corresponding to
1883: the start of the quantum computation) and a final fixed time $\hat t$ (corresponding to the
1884: end of the quantum computation) is fully specified by
1885: the fundamental solution defined with respect to those time values,
1886: $P_{\hat {t}, \hat s}$, so that we have
1887: %
1888: \begin{equation}
1889: %
1890: \rho(\hat t ) = P_{\hat {t}, \hat s}\cdot \rho(\hat s ),
1891: %
1892: \end{equation}
1893: %
1894: and hence the completely positive, trace-preserving map $P$ that appears
1895: in the QCC is given by the equivalence
1896: %
1897: \begin{equation}\label{qcomponent_def}
1898: %
1899: P\equiv P_{\hat {t}, \hat s}~.
1900: %
1901: \end{equation}
1902: %
1903: 
1904: 
1905: \section{Unified Treatment of Quantum Computing Paradigms}\label{unification}
1906: 
1907: \subsection{Introduction}
1908: In this section we show that the QCC provides a unifying framework in which to
1909: describe on the same footing the currently-known ``paradigms" for quantum computation,
1910: including the {\em circuit-based paradigm}, the {\em graph state-based paradigm},
1911: the {\em adiabatic quantum computer paradigm}.
1912: The QCC subsumes all of these into a single, unifying paradigm for quantum computing.
1913: 
1914: 
1915: \subsection{Circuit-based paradigm}
1916: 
1917: In this section we describe the specification of quantum components based on
1918: the ``circuit-based" 
1919: paradigm of quantum computation.
1920: We proceed as follows:\\
1921: {\bf{(1)}} For purposes of clarity, we
1922: begin in \ref{ideal_circuit}
1923: by working in an idealization in which there is no noise present.
1924: For this idealized case we obtain the general form
1925: of the completely positive 
1926: map $P$ characterizing the quantum component.
1927: We then apply the result (for the noiseless idealization)
1928: to several specific realizations of the
1929: circuit-based paradigm.
1930: These include
1931: qubit-based quantum computers (these utilize states, the operators for which have
1932: a discrete eigenspectrum, {\em i.e.}, qubits in the case of 2-level systems),
1933: quantum continuous variable-based quantum computers (these utilize states, the
1934: operators for which have
1935: a continuous eigenspectrum, such as coherent states),
1936: and liquid state NMR-based quantum computers.\\
1937: {\bf{(2)}} Having obtained the general form for $P$ in the noiseless case, for each of the
1938: three above-mentioned realizations of the circuit-based paradigm, we then explain
1939: in \ref{noise_circuit} how
1940: to modify the analysis, in a way appropriate to all choices of circuit
1941: realization, so as to account for the effects of noise.
1942: 
1943: 
1944: \subsubsection{Idealized Circuits in the Absence of Decoherence and Dissipation}
1945: \label{ideal_circuit}
1946: A quantum circuit is described by a set $G$ of gates operating 
1947: in some specified order on elements of 
1948: a set $R$ of objects.  The quantum states of these objects constitute the information which 
1949: is ``processed" 
1950: by the gates of the quantum circuit.  
1951: We associate 
1952: with each object $i$ of $R$ a Hilbert space $H^{(i)}$ that describes the possible states 
1953: of that particular object.  The Hilbert space for the full set of objects on which the circuit 
1954: operates is then
1955: 
1956: \begin{equation}
1957: H_{\mathrm{circuit}} = \bigotimes_{i \in R} H^{(i)}~.  
1958: \end{equation}
1959: 
1960: \noindent The gates in $G$, labeled by the index $\mu$,
1961: are described by unitary operators $\hat{V}_\mu$, so that 
1962: the unitary operator describing the (noiseless) operation of the circuit is  
1963: 
1964: \begin{equation}
1965: \label{EQ:vcircuit}
1966: {V}_{\mathrm{circuit}} = \prod_{\mu \in G} \hat{V}_{\mu}~,   
1967: \end{equation}
1968: 
1969: \noindent where the product of operations is ordered in accordance
1970: with the definition of the circuit.
1971: The factors $\hat{V}_{\mu}$ that appear in the multiplicand of \eqref{EQ:vcircuit}
1972: are in principle obtained from the fundamental solution to the
1973: appropriate underlying equation of motion, following
1974: the procedure outlined in Section \ref{qm_components} above.\footnote{It is
1975: extremely important to note that we
1976: are {\em not} discussing here the abstractly defined quantum
1977: computation itself, which is prescribed in advance, and is
1978: represented by the unitary operator $U$ that appears explicitly
1979: in the second term under the norm symbol in the QCC given in \eqref{encoded_QCC}.
1980: Rather, we are discussing quantities
1981: that represent elements of a physical device that is to be used as an actual
1982: quantum computing machine. As such, the quantities under discussion here are to be regarded
1983: as ``building blocks" for the {\em first} term under the norm symbol in
1984: \eqref{encoded_QCC}\label{fnote_P_not_U}.}
1985: 
1986: Each gate operates on a subset $\Sigma_\mu$ of the information elements, leaving 
1987: the rest unaffected:
1988: 
1989: \begin{equation}
1990: \label{EQ:vsigmamu}
1991: \hat{V}_{\mu} = V_\mu^{\Sigma_\mu} \otimes
1992: {\Big(}\bigotimes_{i \notin \Sigma_\mu} I_{H^{(i)}}{\Big)}~,
1993: \end{equation}
1994: 
1995: \noindent where $I_{H^{(i)}}$ is the identity operator on $H^{(i)}$,
1996: and $V_\mu^{\Sigma_\mu}$ is the transformation effected by the $\mu$th gate,
1997: so that we may express the unitary operator describing the idealized circuit as 
1998: 
1999: \begin{eqnarray}
2000: \label{EQ:pcircuit}
2001: P\cdot\rho &=& V_{\mathrm{circuit}} \rho V_{\mathrm{circuit}}^\dag \\
2002:       &=& \left( \prod_{\mu \in G} \hat{V}_{\mu} \right) \rho 
2003:           \left( \prod_{\mu \in G}^\prime \hat{V}_{\mu}^\dag \right)~, 
2004: \end{eqnarray}
2005: 
2006: \noindent where the prime indicates that the second product reverses the order of the 
2007: factors relative to the first product.\footnote{We
2008: strongly reiterate the message given in
2009: Footnote \ref{fnote_P_not_U}: It is only coincidentally the case that the form of
2010: \eqref{EQ:pcircuit} resembles that of \eqref{ersatz-quantum-computer-equation}.
2011: {\em Both} the right- and left-hand sides of \eqref{EQ:pcircuit} describe the
2012: physical device ({\em {i.e.}}, $P$), not the abstractly-defined quantum computation $U$,
2013: and thus both sides of eq.(\ref{EQ:pcircuit}) are to be
2014: associated with the first term under the norm symbol in the QCC given
2015: in \eqref{encoded_QCC}.\label{reminder}}
2016: 
2017: %\subsubsection{Special Cases: Qubits, QCV and NMR}
2018: 
2019: Up to this point we have made no restrictions on the Hilbert spaces or 
2020: the types of gates appearing in the specification of the circuit. We now
2021: describe three special cases of interest within the circuit-based
2022: paradigm: qubits, quantum continuous 
2023: variables, and liquid state NMR.
2024: We obtain the completely positive, trace preserving map $P$ that characterizes the 
2025: implementation of the circuit for each example, thus showing that the QCC provides
2026: the proper foundational presentation of a quantum computer for all the cases.
2027: 
2028: 
2029: {\noindent{ { {(Case 1)}} {\em {Circuit-realization with qubits}}}}
2030: 
2031: The great majority 
2032: of the research in quantum computation has focussed on circuits for which the information 
2033: elements are qubits defined on a two dimensional Hilbert space $\mathbb{C}^2$, so that the 
2034: full Hilbert 
2035: space of the circuit is 
2036: 
2037: \begin{equation}
2038: H_{\mathrm{circuit}} = \left (\mathbb{C}^2 \right)^{\otimes |R|}~.  
2039: \end{equation}
2040: %
2041: where $|R|$ is the cardinality of the set of computational qubits.
2042: 
2043: Usually the set of gates is chosen from a relatively small set of operations, each 
2044: of which acts on only 1 or 2 qubits. Specializing \eqref{EQ:vsigmamu} to
2045: the case of a circuit
2046: built out of gates acting only
2047: on 1 or 2 qubits (this would be the proper description, for
2048: instance, of a machine that uses the universal set of quantum gates), we have
2049: 
2050: \begin{equation}
2051: \hat{V}_{\mu} = V_{\mu}^{(i)}\otimes \bigotimes_{k (\neq i) {\in R}} I_k,
2052: \end{equation}
2053: %
2054: or
2055: %
2056: \begin{equation}
2057: \hat{V}_{\mu} = V_{\mu}^{(ij)}\otimes \bigotimes_{k (\neq i,j)
2058: {\in R}} I_k~.  
2059: \end{equation}
2060: %
2061: where $I_k$ is the identity operator acting on the copy of $\mathbb{C}^2$ associated
2062: to the $k$th qubit.
2063: With these definitions, the operator $P$ that characterizes the implementation of
2064: the qubit-based realization of a
2065: quantum computer designed according to the circuit-based paradigm is
2066: given by (\ref{EQ:vcircuit}) and (\ref{EQ:pcircuit}). We thus see that the
2067: circuit-based paradigm, on which a large amount of the research in the field
2068: is based, is properly described by the QCC.
2069: 
2070: \noindent{{ {(Case 2)}} {\em {Circuit-realization with quantum continuous variables}}}
2071: 
2072: Alternatively, we may select different Hilbert spaces $H^{(i)}$ appropriate for  
2073: quantum computation using quantum continuous variables (QCV).  For instance, 
2074: the $H^{(i)}$ might describe the  
2075: states of simple harmonic 
2076: oscillators.  The full Hilbert space of the circuit, and the gate operations out of which it
2077: is built, are then defined 
2078: analogous to the above prescriptions for the qubit-based quantum computer.   
2079: In this way we arrive at a specification
2080: of the quantum component $P$ appropriate to the case of computation by QCV.
2081: Thus, the QCV realization of the circuit-based paradigm is also seen to be
2082: properly described by the QCC.
2083: 
2084: 
2085: \noindent{{ {(Case 3)}} {\em {Circuit-realization with NMR states}}}
2086: 
2087: The treatment of quantum computation by nuclear magnetic resonance (NMR) in liquids 
2088: requires special consideration due to the fact that the NMR sample effectively contains 
2089: many copies of the circuit carrying out the same computation.  In this case we define the 
2090: Hilbert space of the system as
2091: 
2092: \begin{equation}
2093: H_{\mathrm{NMR}} = H_{\mathrm{circuit}}^{\otimes N_s}~,
2094: \end{equation}
2095: 
2096: \noindent where $N_s$ is the number of copies of the circuit,
2097: which requires that we extend the definition of the unitary operation for 
2098: the circuit as follows:
2099: 
2100: \begin{equation}
2101: \label{EQ:vnmr}
2102: {V}_{\mathrm{NMR}} = {V}_{\mathrm{circuit}}^{\otimes N_s} ~.
2103: \end{equation}
2104: 
2105: \noindent We then obtain 
2106: 
2107: \begin{equation}\label{eq_nmr}
2108: P\rho = {V}_{\mathrm{NMR}} \rho {V}_{\mathrm{NMR}}^\dag
2109: \end{equation}
2110: 
2111: \noindent which describes the NMR-based quantum computer in the QCC.\footnote{In
2112: connection with \eqref{eq_nmr} we
2113: repeat the admonition given in Footnote \ref{reminder} .}
2114: 
2115: 
2116: \subsubsection{Quantum Circuits in the Presence of Decoherence and Dissipation}
2117: \label{noise_circuit}
2118: 
2119: Thus far in this section we have restricted ourselves to a discussion of quantum circuits 
2120: in the absence of decoherence and dissipation.  
2121: This is reflected in our description of the 
2122: gates as implementing purely unitary operations, as in (\ref{EQ:vcircuit}).  Even at this level of 
2123: description the circuit is not necessarily error-free.  Unitary errors derive from a situation in 
2124: which the evolution of the quantum circuit is unitary, but the circuit does not implement exactly 
2125: the desired unitary computation:
2126: 
2127: \begin{equation}
2128: \label{EQ:imperfectcircuit}
2129: V_{\mathrm{circuit}} \neq U~.  
2130: \end{equation}
2131: 
2132: \noindent Unitary errors can arise from either the design or the physical 
2133: implementation of the circuit.  
2134: Design errors arise, for example, due to the the fact that a universal set of 
2135: quantum gates only allows 
2136: for the implementation of an arbitrary unitary operation U to within an arbitrarily small 
2137: tolerance~\cite{nielsen-chuang}.  
2138: In general there will then be some residual error implied by the very design of the circuit.  
2139: Implementation errors result from inaccuracies in the 
2140: physical parameters governing the unitary evolution associated with a gate as compared with 
2141: the specification of those parameters by the design.  
2142: For example, an interaction 
2143: Hamiltonian may be applied for a longer time than specified, or there may be errors in the 
2144: field strengths or couplings in the interaction Hamiltonian.  
2145: 
2146: In addition to unitary errors, we also need to deal with errors resulting from decoherence and 
2147: dissipation.   
2148: We will refer to these simply as decoherent errors.\footnote{The case of liquid state 
2149: NMR admits another type of error, which arises when 
2150: the operators $\hat{V}_{\mathrm{circuit}}^{\left( k \right)}$ appearing in (\ref{EQ:vnmr}) 
2151: effect different unitary errors 
2152: on different copies ($k$) of the circuit.  These are known as incoherent errors.}
2153: In this case the gates are described 
2154: by completely positive maps $\hat{P}_{\mu}$,
2155: where the hat on the $P$ indicates that this quantum component is a single gate,
2156: and the index $\mu$ identifying the particular gate, as above.
2157: In addition to replacing unitary operations representing the gates 
2158: by completely positive maps, it is also 
2159: important to account for errors occurring in the transmission of quantum states from 
2160: one gate to the next.    
2161: This means that the transmission channels are also represented by 
2162: completely positive maps designated by the symbol $\hat{P}_{\mu}$.  
2163: In effect, the 
2164: transmission channels (including any quantum memories used to store the states) 
2165: are regarded as gates that (ideally) implement the identity transformation:
2166: $\hat{V}_\mu = I$.  If we call the set of transmission channels $C$, the completely 
2167: positive map that describes the circuit is then
2168: 
2169: \begin{equation}
2170: \label{EQ:prodpmu}
2171: P = \prod_{\mu \in G \cup C} \hat{P}_{\mu}~,  
2172: \end{equation}
2173: 
2174: \noindent where the index $\mu$ 
2175: now identifies both the transmission channels (in $C$) and the gates (in $G$). 
2176: 
2177: The operators $\hat{P}_\mu$ are obtained by explicitly solving the 
2178: equations of motion for the physical device that implements the gate.  The result may 
2179: be written formally as the sum of two terms, one of which represents the action of the 
2180: unitary operator $\hat{V}_\mu$ describing the ideal gate as specified by the circuit design, 
2181: and the other of which represents the effects of 
2182: unitary implementation errors as well as
2183: decoherence and/or dissipation:
2184: 
2185: \begin{equation}
2186: \label{EQ:gateerrormodel}
2187: \hat{P}_\mu \rho = \left( 1 - \epsilon_\mu \right) 
2188:    \left( \hat{V}_\mu \rho \hat{V}_\mu^\dag \right) + 
2189:    \epsilon_\mu \hat{Q}_\mu \rho ~,
2190: \end{equation}
2191: 
2192: \noindent where $\epsilon_\mu$ is the probability that an error occurs during the
2193: operation of the 
2194: gate, and $\hat{Q}_\mu$ is a completely positive map that represents the effects of the 
2195: error. Although we have indicated how this follows in principle
2196: from an explicit solution of the detailed 
2197: dynamics of the gate (as described in \S\ref{qm_components}),
2198: an error model of this form is often assumed from the outset.  The 
2199: latter approach necessitates the choice of some specific 
2200: operator $\hat{Q}_\mu$ to represent the errors.  For instance,
2201: in investigating the properties of quantum error correcting codes one often invokes a
2202: ``depolarization" qubit error model in which
2203: 
2204: \begin{equation}
2205: \label{EQ:depol}
2206: \hat{Q}_\mu \rho \equiv {1\over 4}\left( \rho + \sum_{j=1}^3 \sigma_j \rho
2207: \sigma_j\right)~,
2208: \end{equation}
2209: 
2210: \noindent where the  $\sigma_j$ are the Pauli matrices acting on a single qubit.
2211: The advantage of this approach is that 
2212: it abstracts the physical implementation of the quantum computer from the design of the 
2213: circuit while retaining the main features of decoherence that must be addressed in the 
2214: development of any practical quantum computer.  The disadvantage is that the abstraction 
2215: must then be justified relative to the detailed physical implementation of 
2216: the computer in order 
2217: to apply rigrously the results of any analysis based on \eqref{EQ:depol}.  
2218: 
2219: Whether we arrive at (\ref{EQ:gateerrormodel}) by detailed calculation or by abstraction, 
2220: we may use it in conjunction with (\ref{EQ:prodpmu}) to describe the operation of the 
2221: entire circuit:
2222: 
2223: \begin{equation}
2224: P\cdot\rho = \left[ \prod_\mu \left( 1 - \epsilon_\mu \right) \right]
2225:          V_{{\mathrm{circuit}}} \rho V_{{\mathrm{circuit}}}^\dag +
2226:   \left\{ 1 - \left[ \prod_\mu \left( 1 - \epsilon_\mu \right) \right] \right\}
2227:          \epsilon_f Q_f \rho~,
2228: \end{equation}
2229: 
2230: \noindent where  
2231: $Q_f$ incorporates the effects of the individual gate errors.  We define
2232: an error probability associated with the entire circuit:
2233: 
2234: \begin{equation}\label{eps_mu_f}
2235: \epsilon_f \equiv 1 - \prod_\mu \left( 1 - \epsilon_\mu \right)~,
2236: \end{equation}
2237: 
2238: \noindent with which we have 
2239: 
2240: \begin{equation}
2241: \label{EQ:circuiterrormodel}
2242: P\cdot\rho = \left( 1 - \epsilon_f \right)
2243:          V_{{\mathrm{circuit}}} \rho V_{{\mathrm{circuit}}}^\dag +
2244:   \epsilon_f Q_f \rho ~.  
2245: \end{equation}
2246: %
2247: We will make use of this error model in applying the QCC to the problem of 
2248: finding error thresholds for fault tolerance in \S\ref{et_ft}.
2249: 
2250: We note that eqs. \eqref{EQ:gateerrormodel} and \eqref{EQ:circuiterrormodel}
2251: connect theoretical analysis with experimental observations.
2252: A general theoretical method for obtaining an explicit
2253: expression for the quantum component, $P$, in terms of the underlying equation
2254: of motion for the system, has been given above in \S\ref{qm_components}. From an
2255: {\em experimental} perspective, explicitly writing
2256: \eqref{EQ:gateerrormodel} or \eqref{EQ:circuiterrormodel}
2257: is a goal of quantum process tomography,
2258: which is by definition
2259: the experimental method of determining the evolution of open
2260: quantum systems ~\cite{YSW}.
2261: 
2262: 
2263: \subsection{Adiabatic Quantum Computing Paradigm}
2264: 
2265: We now show that the QCC also provides the proper framework in which to formulate
2266: the adiabatic quantum computing paradigm.
2267: In adiabatic quantum computing, the quantum component can be described by a Lindblad 
2268: equation incorporating a Hamiltonian of the form:
2269: 
2270: \begin{equation}
2271: \mathcal{H}_{\mathrm{adiabatic}} \left( t \right) = f \left( t\right) \mathcal{H}_0 + 
2272:       g \left( t\right) \mathcal{H}_f
2273: \end{equation}
2274: 
2275: \noindent where $f$ and $g$ are smooth functions of time with 
2276: $f(0) = 1$, $f(T) = 0$, $g(0) = 0$ and $g(T) = 1$, so that the adiabatic Hamiltonian 
2277: goes smoothly from $\mathcal{H}_0$ to $\mathcal{H}_f$ over time.  
2278: The full Lindblad equation may then be written as
2279: %
2280: \begin{equation} \label{EQ:adiabaticlindblad}
2281: %
2282: \frac{d}{dt} \rho (t) = -\frac{i}{\hbar} 
2283: [\mathcal{H}_{\mathrm{adiabatic}}(t) + \mathcal{V}(t), \rho] + 
2284: \sum_j \bigg[ L_j(t) \rho L_j^\ast(t) - 
2285: \frac{1}{2}\big\{L_j^\ast(t) L_j(t), \rho\big\}\bigg]~,
2286: %
2287: \end{equation}
2288: %
2289: where $\mathcal{V}(t)$ represents unitary errors and the terms involving $L_j(t)$ 
2290: represent interactions with the environment.  
2291: 
2292: The computation begins with an an initial preparation of the ground 
2293: state of the Hamiltonian $\mathcal{H}_0$.  
2294: Provided the conditions for adiabatic 
2295: evolution are satisfied, evolution under the exact adiabatic Hamiltonian takes the 
2296: ground state of $\mathcal{H}_0$
2297: to the ground state of $\mathcal{H}_f$ 
2298: with a high degree of accuracy.  
2299: We identify the operator
2300: $U$ that appears in the QCC as a unitary operator that describes this desired behavior:
2301: 
2302: \begin{equation}
2303: U: |\phi_0 \rangle \mapsto |\psi_0 \rangle ~.  
2304: \end{equation}
2305: 
2306: \noindent where $|\phi_0 \rangle$ is the ground state of $\mathcal{H}_0$, 
2307: $|\psi_0 \rangle$ is the ground state of $\mathcal{H}_f$, and $U$ is otherwise arbitrary.  
2308: 
2309: The physical realization 
2310: of the adiabatic quantum computer is described by 
2311: (\ref{EQ:adiabaticlindblad}), and thus implements the desired operation $U$ only approximately.  
2312: This is due to the approximation inherent in the adiabatic evolution
2313: itself, as well as any additional 
2314: unitary errors, decoherence and dissipation.  
2315: We express this 
2316: fact quantitatively by using the fundamental solution of 
2317: (\ref{EQ:adiabaticlindblad}) to derive the map $P$
2318: (as described above in \S\ref{qm_components})
2319: that describes the operation of the 
2320: adiabatic quantum component. The construction of the QCC then follows.  
2321: 
2322: Note that the special case of the adiabatic 
2323: quantum computer is characterized under $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
2324: \mathcal{M}_\mathrm{dec},\alpha)$
2325: by two unique features: 
2326: 
2327: \begin{enumerate}
2328: \item the encoding and decoding maps, $\mathcal{M}_\mathrm{enc}$ and 
2329: $\mathcal{M}_\mathrm{dec}$ are identities, and
2330: \item the QCC is required to hold only for the ground state of the initial 
2331: Hamiltonian, that is, for $\rho = |\phi_0 \rangle \langle \phi_0 |$ ~.  
2332: \end{enumerate}
2333: 
2334: \subsection{Graph state-based (including cluster state-based) paradigm}
2335: 
2336: We now consider the cluster-based approach to quantum
2337: computation~(\cite{raussendorf-briegel-2002},\cite{raussendorf-briegel-2001},
2338: \cite{raussendorf-browne-briegel-graph-states}),
2339: which is formulated in terms of an array of two level quantum systems.  The
2340: systems in the array are referred to as sites and are elements of some
2341: set $L$ which has a geometrical structure such as a 1 or 2 dimensional
2342: lattice; in addition to the geometrical structure there is also a flow
2343: structure which models the flow of information through the
2344: cluster. The two-dimensional Hilbert space corresponding to a site
2345: $\mathbf{s} \in L$ is denoted $H_\mathbf{s}$ and the Hilbert space $H$
2346: of the entire cluster system is the tensor product
2347: %
2348: \begin{equation}
2349: %
2350: H_\mathbf{cluster} = \bigotimes_{\mathbf{s} \in L} H_\mathbf{s}.
2351: %
2352: \end{equation}
2353: %
2354: A key concept in the cluster approach is {\em measurement at a site}
2355: $\mathbf{s} \in L$ considered as an operation on quantum mechanical
2356: states.  As an operation on pure states, the measurement corresponds
2357: to a pair of self-adjoint projections $E_\mathbf{s}$, $1-
2358: E_\mathbf{s}$ on $H_\mathbf{s}$.  In terms of the cluster system, we
2359: identify the projection $E_\mathbf{s}$ with a projection on
2360: $H_\mathbf{cluster}$ defined by
2361: %
2362: \begin{equation}
2363: %
2364: E = E_\mathbf{s} \otimes \bigotimes_{\mathbf{m} \neq \mathbf{s}}
2365: I_{H_\mathbf{m}}
2366: %
2367: \end{equation}
2368: %
2369: where $I_{H_\mathbf{m}}$ is the identity operator acting on $H_{\mathbf{m}}$.
2370: The associated projective measurement on the cluster Hilbert space
2371: $H_\mathbf{cluster}$ is the completely positive map on density states\footnote{We
2372: note that the message of Footnote \ref{reminder} applies to this equation.}
2373: %
2374: \begin{equation}
2375: %
2376: P_E\cdot\rho = E \rho E + (1 - E) \rho (1 - E)~.
2377: %
2378: \end{equation}
2379: %
2380: The cluster scheme is illustrated in Figure \ref{cluster-configuration-figure}.
2381: 
2382: 
2383: %\begin{example}\label{graph_example}
2384: %
2385: A general class of cluster-type configurations associated to graphs has been introduced
2386: in the literature~(\cite{raussendorf-browne-briegel-graph-states},
2387: \cite{benjamin-eisert-stace-2005}). Given a graph
2388: $\mathcal{G}=(\mathbf{nodes}, \mathbf{edges})$, the lattice sites are
2389: the elements of $\mathbf{nodes}$.  Each lattice site $a \in
2390: \mathbf{nodes}$ has associated with it a two-dimensional Hilbert space
2391: $H_a$ and an ``entanglement'' projection operator $F$ that acts on the
2392: Hilbert space
2393: %
2394: \begin{equation}
2395: %
2396: H_\mathbf{graph} = \bigotimes_{a \in  \mathbf{nodes}} H_a
2397: %
2398: \end{equation}
2399: %
2400: as follows:\footnote{Such operators are considered
2401: from the more general point of view as partial
2402: isometries below in Definition \ref{isometric_def} and in the discussion preceding
2403: it.} To each $(a,b) \in \mathbf{edges}$, define the projector
2404: of the form
2405: %
2406: \begin{equation}
2407: %
2408: \bigotimes_{n \in \mathbf{nodes} \setminus \{a,b\}}I_{H_n} \otimes
2409: F_{(a,b)}.
2410: %
2411: \end{equation}
2412: %
2413: defined in terms of the Pauli matrices by
2414: %
2415: \begin{equation}
2416: %
2417: F_{(a,b)} = \frac{1}{2} \biggl\{I + \sigma^{(a)}_z+\sigma^{(b)}_z -
2418: \sigma^{(a)}_z \otimes \sigma^{(b)}_z\biggr\}.
2419: %
2420: \end{equation}
2421: %
2422: Then define 
2423: \begin{equation}
2424: %
2425: F = \prod_{(a,b) \in \mathbf{edges}} F_{(a,b)}.
2426: %
2427: \end{equation}
2428: %
2429: $F$ is a projection, since all the projectors $F_{a,b}$ pairwise
2430: commute.  
2431: %
2432: %\end{example}
2433: 
2434: 
2435: \subsubsection{The Cluster Measurement Scheme}
2436: 
2437: In the cluster-based approach as explained
2438: in~\cite{raussendorf-briegel-2002}, a computation is realized by a
2439: sequence of projective measurements performed on different sites
2440: $\mathbf{s} \in L$.  The projective measurement that is to be
2441: performed at each step of the computation is determined by a scheme
2442: that specifies at which site to make the measurement and what
2443: observable to measure at that site. These two choices depend on the
2444: outcome of the preceding measurements. In order to specify this
2445: process, a discrete time ``flow'' between the sites is also given.
2446: This flow determines the sequencing of sites to measure. It is
2447: important to note however that the specific projective measurement
2448: taken at each site depends on the outcome of the measurement taken at
2449: the preceding site.
2450: 
2451: \begin{figure}
2452: \input{cluster.pstex_t}
2453: \caption{Two-dimensional cluster configuration (the arrow denotes time flow)}
2454: \label{cluster-configuration-figure}
2455: \end{figure}
2456: 
2457: At the level of specification, we can also consider the cluster measurement
2458: scheme as given by a multi-rooted tree $\mathbb{T}$.  In
2459: Figure~\ref{cluster-scheme-figure} we illustrate such a tree which
2460: because of spatial limitations is singly rooted. The tree consists of
2461: nodes and directed branches. Each node on the tree $\mathbb{T}$
2462: corresponds to a pair $(\mathbf{s}, A)$ where $\mathbf{s} \in L$ is a
2463: cluster site and $A$ is a two-level observable on $H_\mathbf{s}$.  We
2464: will refer to $\mathbf{s}$ as the cluster site corresponding to the
2465: tree node $(\mathbf{s}, A)$.  Each node $\ell = (\mathbf{s}, A)$ of
2466: the tree has two outgoing branches corresponding to the two possible
2467: outcomes of the measurement of $A$. These two branches correspond to
2468: the spectral projections of $A$, which we denote $\opr{E}^+(A)$,
2469: $\opr{E}^-(A)$.
2470: 
2471: It is important to note that many different nodes of $\mathbb{T}$ may
2472: correspond to the same cluster site, that is two distinct computation
2473: sequences may take us to the same node but at which different
2474: measurements are taken.  In fact, in the usual cluster approach all
2475: computation sequences traverse the exact same cluster nodes.  This
2476: means that at each horizontal level of the tree $\mathbb{T}$, the
2477: nodes all have the same cluster site.  In the general scheme outlined
2478: above, no such restriction exists.  Thus we may consider schemes in
2479: which not only the subsequent measurement depends on previous
2480: outcomes, but in which the site at which the measurement is taken also
2481: dependent on previous outcomes.
2482: %
2483: \begin{figure}
2484: \input{tree.pstex_t}
2485: \caption{Example scheme of a cluster computation}
2486: \label{cluster-scheme-figure}
2487: \end{figure}
2488: 
2489: For clusters that are not arranged in a 1 dimensional array, we can
2490: avoid the use of multiply rooted trees if we allow cluster systems
2491: that are not restricted to two-level systems.  This means that the
2492: Hilbert space $H_\mathbf{s}$ corresponding to a site $\mathbf{s} \in
2493: L$ can have arbitrary dimension.  In that case, the corresponding node
2494: observables are allowed to have more than two outcomes and in
2495: particular, the specification tree may not be binary.
2496: 
2497: A complete path (that is one which originates at the root node and
2498: terminates at a leaf node) of $\mathbb{T}$ is specified by a sequence
2499: of measurement outcomes, where the observable measured is associated
2500: to a node of the tree. Mathematically, each measurement outcome is
2501: expressed by one of the spectral projections associated to the
2502: measurement and graphically represented by an edge of the tree.
2503: %
2504: A complete measurement outcome associated to a path is a sequence of
2505: projections, where each projection corresponds to a traversal of an
2506: edge of the tree as follows:
2507: %
2508: \begin{equation}
2509: %
2510: \ell_0 \stackrel{\opr{E}^\pm_0}{\longrightarrow} \ell_1
2511: \stackrel{\opr{E}^\pm_1}{\longrightarrow} \ell_2
2512: \stackrel{\opr{E}^\pm_2}{\longrightarrow} \ell_3
2513: \stackrel{\opr{E}^\pm_3}{\longrightarrow} \cdots
2514: \stackrel{\opr{E}^\pm_k}{\longrightarrow} \ell_{k+1}.
2515: %
2516: \end{equation}
2517: %
2518: 
2519: The projective measurement associated to the cluster consists of the
2520: projective measurements on sites following the branches of the cluster
2521: scheme tree.  We can explicitly write this down:
2522: %
2523: \begin{thm}
2524: %
2525: The projective measurement on a cluster is of the form
2526: %
2527: \begin{equation}
2528: %
2529: P_{\mathbb{T}}\cdot\rho = \sum_{\mathcal{P} \in \opr{Path{\mathbb{T}}}}
2530: \biggl(\prod_{\tau \in \mathcal{P}} E_\tau \biggr) \ \rho
2531: \ \biggl(\prod_{\tau \in \mathcal{P}} E_\tau\biggr)
2532: %
2533: \end{equation}
2534: %
2535: where $E_\tau$ is a projector on the cluster site corresponding to the
2536: edge of the cluster scheme tree.  Note that for
2537: each $\mathcal{P} \in \opr{Path}(\mathbb{T})$, all the
2538: $E_\tau$ with $\tau \in \mathcal{P}$ commute.
2539: \end{thm}
2540: 
2541: 
2542: \subsubsection{Quantum Components in the graph state-based paradigm} 
2543: 
2544: We now show that the graph state-based paradigm of quantum computation is properly
2545: described by the QCC. The paradigm has been described in the
2546: literature as employing an entangled substrate upon which a series of
2547: conditional projective measurements is performed.  The projective
2548: measurements are used to carry out a computational algorithm, but can
2549: also be used to read input from a macroscopically observable input
2550: register or output to a macroscopically observable output register.
2551: %
2552: %
2553: The entangled substrate corresponds to some particular Hilbert
2554: subspace of entangled vectors which may be characterized in various
2555: ways, for instance as the range of an entangling operation or as the
2556: solutions to some eigenvalue equations.
2557: 
2558: %
2559: Correspondingly, we should expect that the mathematical formulation of
2560: the graph model of a quantum component also consist of two parts
2561: (We assume as given the Hilbert space $H_{\mathbf{graph}}$):
2562: % 
2563: \begin{enumerate}
2564: %
2565: \item An initial ``entanglement producing'' operation of some kind.
2566: %
2567: % \item[]
2568: %
2569: \item The projective measurement on $H_{\mathbf{graph}}$
2570: corresponding to the graph scheme.
2571: %
2572: \end{enumerate}
2573: %
2574: In our approach, we have already discussed how to formalize the step
2575: (2) above.  However, there are several choices for the entanglement
2576: generating step (1).  In the examples discussed in the literature,
2577: these are given by a self-adjoint projection operator, but it is
2578: natural from our viewpoint to consider more generally partial
2579: isometries $V: H \rightarrow H$.
2580: %
2581: 
2582: %
2583: \begin{dfn}\label{isometric_def}
2584: %
2585: A {\em graph state-based quantum component} $\mathcal{C}$ is a pair $(V,
2586: \mathbb{T})$ given by a cluster scheme $\mathbb{T}$ and an
2587: ``entanglement'' partial isometry $V$ on $H_\mathrm{graph}$.
2588: %
2589: % \item[]
2590: %
2591: The completely positive map associated to $\mathcal{C}$ is defined by
2592: %
2593: \begin{equation}\label{clust_comp}
2594: %
2595: Q_\mathcal{C} \cdot\rho = \sum_{\mathcal{P} \in \opr{Path}(\mathbb{T})}
2596: E_\mathcal{P} V \rho V^\dag E_\mathcal{P}
2597: %
2598: \end{equation}
2599: %
2600: \end{dfn}
2601: %
2602: Thus, we see that, just as for the circuit-based paradigm and the adiabatic quantum
2603: computing paradigm, the QCC provides an over-arching framework in which to formulate
2604: the graph-state based paradigm of quantum computing. Moreover, in this section
2605: we have generalized the definition of graph state- (and cluster state-) based
2606: quantum computers that has previously appeared in the literature. Our generalization
2607: consists of two features: (1) our definition allows for {\em arbitrarily
2608: different} measurements to
2609: be carried out at different nodes at the same level of the multi-rooted tree $\mathbb{T}$,
2610: and (2) our definition replaces the use of a self-adjoint projection
2611: operator as an entanglement generator with the more general notion of a partial isometry
2612: (which includes self-adjoint projections as a special case).
2613: 
2614: 
2615: 
2616: \section{Error Thresholds and Fault Tolerance}\label{et_ft}
2617: 
2618: \subsection{Introduction}
2619: 
2620: In \S\ref{KLP_reduction} we showed that the recently discovered approach
2621: to error correction known as ``operator quantum error correction" is in fact
2622: a special case of the general QCC formulation. 
2623: Given the general applicability of the QCC to all quantum computing paradigms
2624: ({\em cf} \S\ref{unification}), as well as to all
2625: techniques for protection against errors (including
2626: quantum error correction, decoherence-free subspaces and noiseless subsystems),
2627: the QCC thus provides a unified framework for a fully general analysis of
2628: fault tolerance in quantum computing. We
2629: refer to this as operator quantum fault tolerance (OQFT).
2630: 
2631: As an example of OQFT, in this section we describe the application
2632: of the QCC to the analysis of error thresholds for fault tolerance in the circuit
2633: paradigm.
2634: To make contact with previous research, we begin by discussing this subject from 
2635: the perspective of the well established 
2636: method based on the analysis of error
2637: probabilities~\cite{shor_ft_96}, \cite{aharonov_ben_or_96}, \cite{kitaev1997}, \cite
2638: {klz_97}, \cite{preskill_ft_97}, \cite{agp_ft_05}.
2639: Since the QCC in fact provides the underlying framework in which to study
2640: any issues associated with physical quantum computation, we then reformulate 
2641: the problem in terms of the QCC. This allows us to
2642: relate the results of the two approaches, and to determine the extent to which the
2643: previously utilized approach (based on the method of error probabilities) is in fact
2644: justified based on the insight provided with the QCC.
2645: 
2646: \subsection{The Method of Error Probabilities}  
2647: 
2648: In this section we briefly describe the method of error probabilities.
2649: We begin by identifying a quantum operation that we wish to implement and specifying a
2650: circuit 
2651: that (ideally) implements the operation.  
2652: We then identify an error model for the gates in the 
2653: circuit that accounts for the inevitable effects of dissipation and decoherence 
2654: that come into play when the gates are implemented with  
2655: real devices.  By hypothesis, the probability that an error occurs in 
2656: this ``direct" implementation of the 
2657: operation 
2658: is too high for it to be useful as a component in a quantum computer.  
2659: 
2660: In order 
2661: to render the circuit more fault tolerant, we specify a 
2662: second, more complicated, circuit using quantum error correction.  
2663: The circuit now operates on the set of {\em encoded} qubits 
2664: obtained by encoding the {\em logical} qubits 
2665: of the direct implementation using a QECC.  The gates in the original 
2666: circuit are replaced by collections of gates that operate on the encoded qubits.  
2667: Additional gates are added to carry out the QECC's recovery procedure wherever an error 
2668: is detected.  The error model is then applied to the gates in this new circuit.  
2669: The error correction clearly provides some degree of protection against errors, but
2670: it also necessitates
2671: a larger number of gates, so that there are then more 
2672: opportunities for errors to occur.  In order to determine whether this 
2673: procedure has improved the fault tolerance, we compare the probability 
2674: of an uncorrected error occurring during operation of the QECC version
2675: ($\epsilon_f^{\mathrm{QECC}}$) with the 
2676: probability of any error 
2677: occurring in the ``direct" implementation ($\epsilon_f^{\mathrm{direct}}$).
2678: If the QECC error probability is smaller, 
2679: that is, if
2680: 
2681: \begin{equation}
2682: \label{EQ:stdratio}
2683: \frac {\epsilon_f^{\mathrm{QECC}}}{\epsilon_f^{\mathrm{direct}}} <1 ~,
2684: \end{equation}
2685: 
2686: \noindent then the procedure has been at least partially successful.  If the error 
2687: probability of the new circuit is still too high, we can reduce the error probability 
2688: further by concatenating the code, that is by  
2689: encoding the qubits used in the QECC version using the same QECC and by redesigning 
2690: the circuit to handle the second level of encoded qubits.
2691: By repeating the process of concatenation we can arrange that the error
2692: probability be made arbitrarily small.
2693: 
2694: The key point is that (\ref{EQ:stdratio}) can be shown to hold 
2695: only if the failure probabilities of the 
2696: individual gates are less than some threshold value that depends on the relative complexity of 
2697: the encoded {\em vs}. un-encoded versions of the circuit. 
2698: The thresholds appearing in these conditions are known as ``error thresholds." If the 
2699: threshold conditions are satisfied, then the use of concatenated codes will
2700: provide fault tolerant operation to within some 
2701: specified tolerance. The problem of achieving fault tolerant 
2702: quantum computation is reduced to the problem of constructing implementations of the 
2703: primitive gates 
2704: that satisfy the error threshold conditions.  
2705: 
2706: \subsection{Operator Quantum Fault Tolerance and the QCC}  
2707: 
2708: We now reconsider the above problem from the perspective of OQFT by making use of
2709: the QCC.
2710: Our goal is to identify a quantum component $P$, that implements 
2711: (approximately) a quantum computation, $U$, that is, we write the quantum
2712: computer condition, $\mathbf{QCC}(P,U,\mathcal{M}_\mathrm{enc},
2713: \mathcal{M}_\mathrm{dec},\alpha)$,
2714: 
2715: \begin{equation}
2716: \|\mathcal{M}_\mathrm{dec}(P\cdot
2717: (\mathcal{M}_\mathrm{enc}(\rho))) - U \rho U^\dag\|_1 \leq \alpha   
2718: \end{equation}
2719: 
2720: \noindent for some suitable choice of encoding and decoding
2721: maps, $\mathcal{M}_\mathrm{enc}$ and 
2722: $\mathcal{M}_\mathrm{dec}$. For simplicity of notation we define an operator 
2723: 
2724: \begin{equation}
2725: \tilde{P} \equiv \mathcal{M}_\mathrm{dec} \cdot P \cdot \mathcal{M}_\mathrm{enc}~,
2726: \end{equation}
2727: 
2728: \noindent so that the QCC becomes
2729: 
2730: \begin{equation}
2731: \|\tilde{P} \rho - U \rho U^\dag\|_1 \leq \alpha~.
2732: \end{equation}
2733: 
2734: We begin by developing a ``zero-th order" implementation 
2735: that does not use a QECC.  In the absence of errors, we assume that the 
2736: implementation faithfully implements the computation $U$:
2737: 
2738: \begin{equation}\label{EQ:perfectcircuit}
2739: \tilde{P}^{\left( 0 \right)}\rho = 
2740: \mathcal{M}_\mathrm{dec} \left[ V_{\mathrm{circuit}}^{\left( 0 \right)} 
2741:               \left( \mathcal{M}_\mathrm{enc}\rho \right)
2742:      V_{\mathrm{circuit}}^{\left( 0 \right)\dag} \right] = U \rho U^\dag.  
2743: \end{equation}
2744: 
2745: \noindent With the error model (\ref{EQ:circuiterrormodel}) we have
2746: 
2747: \begin{equation}
2748: \|\tilde{P}^{\left( 0 \right)}\rho - U \rho U^\dag\| = 
2749:    {\Big \|} \left( 1 - \epsilon_f^{\left( 0 \right)} \right) 
2750:            \mathcal{M}_\mathrm{dec} \left[
2751:               V_{\mathrm{circuit}}^{\left( 0 \right)} 
2752:                    \left( \mathcal{M}_\mathrm{enc}\rho \right) 
2753:               V_{\mathrm{circuit}}^{\left( 0 \right)\dag} \right] +
2754:        \epsilon_f^{\left( 0 \right)} 
2755:               \tilde{Q}_f^{\left( 0 \right)} \rho - U \rho U^\dag {\Big \|}_1 ~,
2756: \end{equation}
2757: 
2758: \noindent where, for generality, we have subsumed the encoding and decoding maps 
2759: into the definition of $\tilde{Q}_f$.  (The maps are identities for the zero-th order circuit.)  
2760: With (\ref{EQ:perfectcircuit}) the left hand side of the QCC takes the simple form
2761: 
2762: \begin{equation}
2763: \label{EQ:lhszero}
2764: \|\tilde{P}^{\left( 0 \right)}\rho - U \rho U^\dag\| = \epsilon_f^{\left( 0 \right)} 
2765:    \|  \tilde{Q}_f^{\left( 0 \right)} \rho - U \rho U^\dag  \|_1 ~.
2766: \end{equation}
2767: 
2768: \noindent Note from \eqref{eps_mu_f} that
2769: the failure probability for this circuit is, to lowest order, 
2770: linear in the error probabilities for the gates which make up the circuit:
2771: 
2772: \begin{equation}
2773: \label{EQ:epsorderzero}
2774: \epsilon_f^{\left( 0 \right)} \sim \mathcal{O} \left(\epsilon_\mu \right)~,
2775: \end{equation}
2776: 
2777: \noindent (recall that $\epsilon_\mu$ represents gate error,
2778: {\em cf} \eqref{eps_mu_f})
2779: a fact which, as we shall see, is crucial to the derivation of an 
2780: error threshold.  
2781: 
2782: We next construct the ``first order" implementation of $U$, which operates in the codespace 
2783: of the QECC.  By the preceding arguments, this implementation will satisfy
2784: 
2785: \begin{equation}
2786: \label{EQ:lhsone}
2787: \|\tilde{P}^{\left( 1 \right)}\rho - U \rho U^\dag\| = \epsilon_f^{\left( 1 \right)} 
2788:    \|  \tilde{Q}_f^{\left( 1 \right)} \rho - U \rho U^\dag  \|_1 ~.
2789: \end{equation}
2790: 
2791: Following~\cite{preskill_ft_97}, we
2792: consider the case that errors affect the qubits in the circuit independently 
2793: and that the QECC recovery procedure is sufficient to correct a single error in any one qubit.  
2794: In that case, the 
2795: encoded circuit will exhibit an unrecoverable error only if two or more single qubit errors 
2796: occur.  In this case, the error probability will be quadratic in $\epsilon_\mu$ to leading 
2797: order:  
2798: 
2799: \begin{equation}
2800: \label{EQ:epsorderone}
2801: \epsilon_f^{\left( 1 \right)} \sim \mathcal{O} \left(\epsilon_\mu \right)^2 ~.
2802: \end{equation}
2803: 
2804: \noindent  At this point, the QECC has not eliminated all errors, but has resulted in a 
2805: circuit with error probabilities that are quadratic, rather than linear, in the error 
2806: probabilities of the primitive operations.  
2807: 
2808: We have now introduced two implementations of the computation.  If either implementation 
2809: satisfies the QCC, then 
2810: there is no need to continue.  If the implementations do not satisfy the QCC, then it is meaningful 
2811: to ask whether this can be achieved by concatenating the code.  In order to answer this question, 
2812: we begin by asking another, related question: has the QECC improved the fault tolerance of the 
2813: implementation relative to the QCC?  In other words, we wish to know 
2814: under what conditions it is true that for all $\rho$
2815: 
2816: \begin{equation}
2817: \label{EQ:levelonecond}
2818: \|\tilde{P}^{\left( 1 \right)}\rho - U \rho U^\dag\|_1 <
2819:  \|\tilde{P}^{\left( 0 \right)}\rho - U \rho U^\dag\|_1 ~,
2820: \end{equation}
2821: 
2822: \noindent or alternatively\footnote{Relation \eqref{sup_bound} is
2823: actually a stronger condition
2824: than \eqref{EQ:levelonecond} for infinite-dimensional vector spaces.}
2825: 
2826: \begin{equation}\label{sup_bound}
2827: \opr{sup}_\rho \frac {\|\tilde{P}^{\left( 1 \right)}\rho - U \rho U^\dag\|_1}
2828:  {\|\tilde{P}^{\left( 0 \right)}\rho - U \rho U^\dag\|_1} < 1~.
2829: \end{equation}
2830: 
2831: \noindent Using (\ref{EQ:lhszero}) and (\ref{EQ:lhsone}), this becomes
2832: 
2833: \begin{equation}
2834: \label{EQ:epsnormratio}
2835: \opr{sup}_\rho \frac {\epsilon_f^{\left( 1 \right)}}{\epsilon_f^{\left( 0 \right)}}
2836:    \cdot \frac {\|  Q_f^{\left( 1 \right)} \rho - U \rho U^\dag  \|_1}
2837:                {\|  Q_f^{\left( 0 \right)} \rho - U \rho U^\dag  \|_1} < 1~.
2838: \end{equation}
2839: 
2840: \noindent The above equation represents the extension of \eqref{EQ:stdratio} to OQFT
2841: obtained by using the general framework provided by the QCC.
2842: Clearly it does not reduce to a simple ratio of error probabilities, as in 
2843: (\ref{EQ:stdratio}) above.  The reason for this is that this formulation of the error 
2844: threshold problem based on the QCC takes into account not only error
2845: probabilities, $\epsilon_f$, but 
2846: also expressions involving the norms of operators that characterize the ``strength" 
2847: of the errors.  To see this, note that we began by assuming an error model represented by 
2848: the operation $\tilde{Q}_f^{\left( 0 \right)}$ 
2849: for the zero-th order implementation.  The error model in the 
2850: encoded implementation is represented, in general, by a different operation, 
2851: $\tilde{Q}_f^{\left( 1 \right)}$.  
2852: There is no reason to suppose that these error models are equally effective in perturbing the 
2853: computation.  This is reflected in the difference in the norms:
2854: 
2855: \begin{equation}
2856: \|  \tilde{Q}_f^{\left( 1 \right)} \rho - U \rho U^\dag  \|_1 \neq
2857: \|  \tilde{Q}_f^{\left( 0 \right)} \rho - U \rho U^\dag  \|_1 ~.
2858: \end{equation}
2859: 
2860: To further emphasize this point, we obtain the 
2861: above result (\ref{EQ:stdratio}) based on the assumption that the error models 
2862: are ``commensurate" in 
2863: the sense that
2864: 
2865: \begin{equation}
2866: \label{EQ:strengthapprox1}
2867: \|  \tilde{Q}_f^{\left( 1 \right)} \rho - U \rho U^\dag  \|_1 \approx
2868: \|  \tilde{Q}_f^{\left( 0 \right)} \rho - U \rho U^\dag  \|_1 ~.
2869: \end{equation}
2870: 
2871: \noindent We shall shortly return to the question of whether this is a good approximation.    
2872: With this assumption, the condition (\ref{EQ:epsnormratio}) then becomes
2873: 
2874: \begin{equation}
2875: \label{EQ:epsratio10}
2876: \frac {\epsilon_f^{\left( 1 \right)}}{\epsilon_f^{\left( 0 \right)}} \lesssim 1~,
2877: \end{equation}
2878:  
2879: \noindent which is identical to the result (\ref{EQ:stdratio}) obtained by the 
2880: method of error probabilities.
2881: We obtain the form of the error threshold by noting 
2882: from (\ref{EQ:epsorderzero}) and 
2883: (\ref{EQ:epsorderone})
2884: that the numerator and 
2885: denominator of (\ref{EQ:epsratio10}) are, to lowest order, quadratic and linear,
2886: respectively, in $\epsilon_\mu$, so that
2887: %
2888: \begin{equation}
2889: {\epsilon_f^{(1)}\over\epsilon_f^{(0)}} =
2890: {\sum_{\mu\nu}{A_{\mu\nu}\epsilon_\mu\epsilon_\nu + \cdots}\over
2891: {\sum_\mu B_\mu\epsilon_\mu +\cdots}}~.
2892: \end{equation}
2893: %
2894: At this point, it is straightforward to obtain a threshold if we set
2895: the $\epsilon_\mu$ equal to each other, and take $\epsilon\equiv\epsilon_\mu$.
2896: Then one obtains the error
2897: threshold comparable to those obtained
2898: in \cite{preskill_ft_97}, \cite{gottesman_ft_97}, \cite{zalka_ft_96}, as
2899: %
2900: \begin{equation}
2901: \epsilon\lesssim{{\sum_\mu B_\mu}\over{\sum_{\mu\nu} A_{\mu\nu}}}~.
2902: \end{equation}
2903: %
2904: The desired behavior of 
2905: the concatenated QECC described above follows if we successively apply the 
2906: approximation (\ref{EQ:strengthapprox1}) 
2907: at each level of concatenation:
2908: 
2909: \begin{equation}
2910: \label{EQ:strengthapprox}
2911: \|  \tilde{Q}_f^{\left( i+1 \right)} \rho - U \rho U^\dag  \|_1 \approx
2912: \|  \tilde{Q}_f^{\left( i \right)} \rho - U \rho U^\dag  \|_1 ~.
2913: \end{equation}
2914: 
2915: We note that Aliferis, Gottesman and Preskill 
2916: \cite{agp_ft_05} also relate the ratio of error probabilities for successive 
2917: levels of concatenation to an overall measure of the ``accuracy" of the quantum computation.  
2918: Their approach differs from the one described here in three important ways: 
2919: 
2920: \noindent (Contrast 1) 
2921: Accuracy in \cite{agp_ft_05} is defined in terms of the 
2922: probabilities $p_i$ of the outcomes $i$ of measurements 
2923: on the final output state for an ideal, as compared with a noisy, circuit: 
2924: %
2925: \begin{equation}
2926: %
2927: \sum_i |p_i^\mathrm{noisy} - p_i^\mathrm{ideal}|~.  
2928: %
2929: \end{equation}
2930: %
2931: In contrast, we define the accuracy of the implementation by the QCC.  
2932: 
2933: \noindent (Contrast 2) The threshold proofs in \cite{agp_ft_05} rely on proofs that the 
2934: implementations at each level of concatenation are conditionally correct 
2935: relative to the noise model.  
2936: In this case the ratio of error probabilities $\epsilon^{(i+1)} / \epsilon^{(i)}$ 
2937: is automatically the quantity of interest in comparing performance at each level of 
2938: concatenation.  Here, in contrast, the QCC provides the figure of merit at each level
2939: of concatenation, 
2940: and the dependence on error probabilities is inferred.  
2941: 
2942: \noindent (Contrast 3) As a consequence of the two preceding
2943: points, \cite{agp_ft_05} makes contact with the notion of accuracy only 
2944: at the highest level of concatenation, at which the entire quantum component may be 
2945: viewed as a black box. Here, the QCC is applied systematically at each level 
2946: of concatenation.
2947: 
2948: At this point we have shown that we can obtain the standard form of the
2949: error threshold result 
2950: from the QCC by introducing the assumption (\ref{EQ:strengthapprox}) on the 
2951: relative strengths of the error models appropriate to successive levels of concatenation 
2952: of the QECC.  We now investigate the validity of the approximation.  
2953: We begin by making some reasonable assumptions about the error operators and the 
2954: initial state of the computer.  We note that the operators 
2955: $\tilde{Q}_f^{\left( i \right)}$ 
2956: are trace preserving, and thus describe the evolution of the quantum component if a 
2957: failure has in fact occurred, as can be seen by
2958: setting $\epsilon_f = 1$ in \eqref{EQ:circuiterrormodel}.
2959: It is then reasonable to expect that the state resulting from its operation on $\rho$ 
2960: will be ``close" to the maximum entropy state\footnote{
2961: Note that this is not a good assumption for errors that operate locally on only one qubit 
2962: in a larger set of 
2963: qubits, leaving the others unaffected.  This important case is a topic for further study.
2964: } in the sense that, for small $\delta_p$,
2965: 
2966: \begin{equation}\label{EQ:closetomax}
2967: \| \tilde{Q}_f^{\left( i \right)}\rho - \rho_I \|_p < \delta_p~,
2968: \end{equation}
2969: 
2970: \noindent where $\rho_I$ is the maximum entropy state, and the value of $p$ identifies
2971: the Schatten $p$-norm associated to the corresponding Schatten $p$-class
2972: ({\em cf} \eqref{p-norm}).  
2973: 
2974: \noindent On the other hand, it is normally the case that the input state for the quantum 
2975: computation is a pure state, and thus so is the state $U \rho U^\dag$.  In this case, it 
2976: is straightforward to show that
2977: 
2978: \begin{equation}
2979: \| \rho_I - U \rho U^\dag \|_1 = 2 - \frac{2}{N}
2980: \end{equation}
2981: 
2982: \noindent and 
2983: 
2984: \begin{equation}
2985: \| \rho_I - U \rho U^\dag \|_\infty = 1 - \frac{1}{N}~,
2986: \end{equation}
2987: 
2988: \noindent where $N$ is the dimension of $H_{\mathrm{logical}}$.  Since 
2989: 
2990: \begin{equation}
2991: \| \tilde{Q}_f^{\left( i \right)}\rho - U \rho U^\dag \| = 
2992: \| \left[ \tilde{Q}_f^{\left( i \right)}\rho - \rho_I \right] + 
2993:    \left[ \rho_I - U \rho U^\dag \right] \|~,
2994: \end{equation}
2995: 
2996: \noindent we have, by triangle inequalities, 
2997: 
2998: \begin{equation}
2999: 2 - \frac{2}{N} - \delta_1 \leq 
3000: \| \tilde{Q}_f^{\left( i \right)}\rho - U \rho U^\dag \|_1 \leq
3001: 2 - \frac{2}{N} + \delta_1
3002: \end{equation}
3003: 
3004: \noindent or
3005: 
3006: \begin{equation}
3007: 1 - \frac{1}{N} - \delta_\infty \leq 
3008: \| \tilde{Q}_f^{\left( i \right)}\rho - U \rho U^\dag \|_\infty \leq
3009: 1 - \frac{1}{N} + \delta_\infty
3010: \end{equation}
3011: 
3012: 
3013: \noindent where we have assumed implicitly that $N \gg 1$ and $\delta_p \ll 1$.  
3014: Since these expressions hold for any value of $i$, (\ref{EQ:strengthapprox}) 
3015: is a reasonable approximation with either choice of norm under the 
3016: conditions that 
3017: 
3018: \begin{enumerate}
3019: \item (\ref{EQ:closetomax}) holds for some $\delta_p \ll 1$, and 
3020: \item the initial state of the computation is a pure state. 
3021: \end{enumerate}
3022: 
3023: 
3024: \section{Conclusions}
3025: 
3026: In this paper we have presented a fundamental, unifying framework for describing
3027: physically-realizable quantum computing machines. This is concisely stated in the form
3028: of the Quantum Computer Condition (QCC), an inequality that
3029: incorporates a complete specification of the full dissipative, decohering
3030: dynamics of the actual,
3031: practical device used as the quantum computing machine, a specification of
3032: the ideally-defined quantum
3033: computation intended to be performed by the machine, and a quantitative criterion
3034: for the accuracy with which the computation must be executed. 
3035: 
3036: With the QCC we prove the fundamental Encoding No-Go Theorem that identifies the amount of 
3037: damping (including dissipative and decohering effects) for which physically-realizable
3038: fault-tolerant quantum computing is not possible.
3039: We provide a rigorous definition of damping, and
3040: explicitly calculate a {\em universal} critical damping
3041: value for fault-tolerant quantum computation.
3042: This theorem can be used in principle
3043: to solve practical problems involving quantum computer design. 
3044: 
3045: In this paper we have also presented an existence proof for fundamental solutions to useful
3046: classes of {\em time-dependent} generalizations of the Lindblad equation. This can provide
3047: a useful tool in analyzing a wide variety of open quantum mechanical systems.
3048: 
3049: We have demonstrated that the entire formalism of operator quantum error correction
3050: (OQEC) can be obtained from the QCC as a special case.
3051: By allowing for the possibility of residual errors, the general formalism of the QCC
3052: enables us to generalize OQEC to ``operator quantum fault tolerance" (OQFT).
3053: Since we have demonstrated that
3054: OQEC is in fact a particular reduction of the QCC, and since standard
3055: quantum error correction (QECC), decoherence-free subspaces (DFS) and noiseless
3056: subsystems are all special cases of OQEC, we have discovered that QCC applies in general
3057: across {\em all} these approaches.
3058: 
3059: As an initial application of the OQFT concept, we
3060: have begun the exploration of the application of QCC to the problem of establishing
3061: thresholds for fault-tolerant quantum computation by showing that the standard
3062: approaches to this problem can be motivated within the framework of the QCC.
3063: 
3064: Research in quantum information science has resulted in the discovery of seemingly different
3065: {\em paradigms} for quantum computation, including the circuit-based paradigm,
3066: graph state-based paradigm and adiabatic quantum computing paradigm. In this
3067: paper we have explicitly demonstrated
3068: that these paradigms are not in fact distinct at a fundamental level,
3069: but are all describable within the unifying framework provided by the QCC.
3070: In the particular case of the graph state-based paradigm (which includes cluster
3071: state-based models), we not only show that the
3072: paradigm is a manifestation of the unifying picture provided by the QCC,
3073: but also introduce a definition of graph state-based quantum computers that
3074: generalizes the graph state models previously defined in the literature.
3075: 
3076: Future work motivated by our results should include applications of the Encoding
3077: No-Go Theorem to diverse problems pertaining to practical quantum computer design
3078: and implementation.
3079: It would also be of interest to further explore the physics of the operator
3080: quantum fault tolerance (OQFT) generalization of OQEC presented in this paper.
3081: Specific work along theses lines should include
3082: further application of the QCC to obtaining error thresholds for fault-tolerant
3083: implementations of quantum computers.
3084: It would also be fruitful to explore applications of the quantum computer
3085: condition
3086: to situations in which quantum process tomography techniques are used to
3087: experimentally characterize the quantum component described by the completely positive
3088: map that appears in the QCC.\\
3089: 
3090: 
3091: \noindent{\em Acknowledgements}
3092: 
3093: \noindent We wish to thank Anthony Donadio and Yaakov Weinstein for helpful comments.
3094: This work was supported by MITRE under the MITRE Technology Program.
3095: 
3096: 
3097: 
3098: \appendix
3099: 
3100: \section{Banach Spaces of Operators} \label{banach-spaces} A linear
3101: map $T$ on a Banach space is a {\em contraction} iff its norm is $\leq
3102: 1$.
3103: 
3104: Let $H$ be a separable Hilbert space.  $\mathbf{L}(H)$ denotes the
3105: space of bounded operators on $H$ with the operator norm,
3106: $\mathbf{K}(H)$ denotes the normed closed subspace of compact
3107: operators of $\mathbf{L}(H)$.  In case $H$ is finite dimensional,
3108: these spaces are identical.  Assume now $H$ is infinite dimensional;
3109: we consider other Banach spaces of compact operators defined by
3110: eigenvalue decay conditions and whose norms reflect the rate of decay
3111: of the eigenvalues.  Specifically, let $T \in \mathbf{K}(H)$, then
3112: $|T| = \sqrt{T^\dag T}$ is a non-negative compact operator and so by
3113: the spectral theorem, has a complete set of eigenvectors with
3114: eigenvalues that can be ordered in a sequence
3115: %
3116: \begin{equation}
3117: %
3118: s_0(T) \geq s_1(T) \geq \cdots \geq s_n(T) \geq 0
3119: %
3120: \end{equation}
3121: %
3122: which converges to $0$.  The following properties are well-known
3123: (see~\cite{gohberg-krein}; also~\cite{connes} on which this discussion
3124: is based). 
3125: 
3126: $\mathbf{T}(H)$ is the Banach space of trace-class
3127: operators $T$ on $H$ with the norm
3128: %
3129: \begin{equation}
3130: %
3131: \|T\|_1 = \sum_{k=0}^\infty | s_k(T) |
3132: %
3133: \end{equation}
3134: %
3135: More generally, the Schatten $p$-class $\mathbf{T}_p(H)$ is defined by
3136: the condition
3137: %
3138: \begin{equation}\label{p-norm}
3139: %
3140: \|T\|_p = \bigg\{\sum_{k=0}^\infty s_k(T)^p \biggl\}^\frac{1}{p} < \infty
3141: %
3142: \end{equation}
3143: %
3144: with the norm given by $\| \cdot\|_p$.
3145: The operator norm for any $T\in \mathbf{L}(H)$, denoted by $\|
3146: T\|_\infty$, is defined as the supremum of $\| T x \|$ for $x \in H$
3147: of norm $\leq 1$.  If $T \in \mathbf{K}(H)$, the operator norm is also
3148: the supremum of the eigenvalues of $|T|$.
3149: 
3150: 
3151: \iffalse 
3152: 	We also consider a class of spaces consisting of operators such
3153: 	that the sequence $s_k(T)$ decays to $0$ at least as fast as a power
3154: 	$k^{-\theta}$ for some $\theta>1$.  These are proper subspaces of
3155: 	$\mathbf{T}_1(H)$ (and therefore are not symmetric normed ideals)
3156: 
3157: 	\begin{equation}
3158: 	%
3159: 	\opr{G}_\theta(T) = \sup_k s_k k^{\theta}
3160: 	%	
3161: 	\end{equation}
3162: 	%
3163: 	\begin{prop}
3164: 	%
3165: 	The set of $T$ for which $\opr{G}_\theta(T) < \infty$ is a two-sided ideal in $\mathbf{L}(H)$.
3166: 	%
3167: 	\end{prop}
3168: 	%
3169: 	\begin{proof}
3170: 	%
3171: 	For all $k$,
3172: 	%
3173: 	\begin{equation}
3174: 	%
3175: 	s_k(T+S) \leq \left\{\begin{array}{ll} s_{k/2}(T) + s_{k/2}(S) \leq
3176: 	\opr{G}_\theta(T) \frac{k}{2}^{-\theta} + \opr{G}_\theta(S)
3177: 	\frac{k}{2}^{-\theta}& \mbox{if $k$ is even} \\
3178: 	%
3179: 	s_{(k-1)/2}(T) + s_{(k-1)/2}(S) & \mbox{if $k$ is odd} 
3180: 			     \end{array}\right.
3181: 	%
3182: 	\end{equation}
3183: 	%
3184: 	Now
3185: 	%
3186: 	\begin{equation}
3187: 	%
3188: 	s_k(T+S) \leq 
3189: 	%
3190: 	\end{equation}
3191: 	\end{proof}
3192: 
3193: \fi
3194: 
3195: 
3196: \section{Completely Positive Maps and Fundamental Solutions} 
3197: \label{mathematical-preliminaries-section}
3198: 
3199: In the following, $H$ denotes a complex Hilbert space. We will
3200: consider various Banach spaces of bounded operators on $H$: these are
3201: discussed in Appendix~\ref{banach-spaces} above. 
3202: 
3203: We will consider completely positive maps on both $\mathbf{L}(H)$ on
3204: the Schatten classes $\mathbf{T}_p(H)$ and especially on the
3205: trace-class operators $\mathbf{T}(H) = \mathbf{T}_1(H)$.
3206: 
3207: The basic fact about completely positive maps we use is the {\em Kraus
3208: representation}.  
3209: %
3210: \begin{prop}
3211: %
3212: A $P:\mathbf{T}(H) \rightarrow \mathbf{T}(H)$ is a completely positive contraction
3213: if and only if it is of the form
3214: %
3215: \begin{equation}\label{KrausFormCPMap}
3216: %
3217: P(\rho)=\sum_{i \in I} X_i \rho X_i^\dag
3218: %
3219: \end{equation}
3220: %
3221: where $X_i \in \mathbf{L}(H)$ with
3222: %
3223: \begin{equation}
3224: %
3225: \sum_i X_i^\dag X_i \leq 1
3226: %
3227: \end{equation}
3228: %
3229: \end{prop}
3230: %
3231: We will consider only {\em trace preserving} completely positive maps $P$,
3232: that is which satisfy
3233: \begin{equation}\label{trace-norem}
3234: %
3235: \opr{tr}(P(\rho)) = \opr{tr}(\rho)
3236: %
3237: \end{equation}
3238: 
3239: \begin{prop} Suppose $P$ is a completely positive map given by the
3240: Krauss form~\eqref{KrausFormCPMap}.
3241: %
3242: \begin{enumerate}
3243: %
3244: \item A necessary and sufficient condition $P$ be trace-preserving is that
3245: %
3246: \begin{equation}\label{tr-preserving-cond}
3247: %
3248: \sum_{i \in I} X_i^\dag  X_i =1.
3249: %
3250: \end{equation}
3251: %
3252: \item  A necessary and sufficient condition that $P$ be bounded in the operator norm is that
3253: %
3254: \begin{equation}\label{norm-preserving-cond}
3255: %
3256: \sum_{i \in I} X_i  X_i^\dag \in \mathbf{L}(H).
3257: %
3258: \end{equation}
3259: %
3260: \end{enumerate}
3261: %
3262: \end{prop}
3263: %
3264: Note that if $T$ is a completely positive trace-preserving and
3265: operator norm continuous map, then by interpolation $T$ is also norm
3266: continuous on the Schatten $p$-classes.
3267: 
3268: \begin{prop}
3269: %
3270: If $P$ is a completely positive trace-preserving map, then the adjoint of
3271: $P$ on $\mathbf{L}(H)$ defined by
3272: %
3273: \begin{equation}\label{definition-of-dual}
3274: %
3275: \opr{tr}(P^{\mathrm{t}}(T) \rho) = \opr{tr}(T P(\rho))
3276: %
3277: \end{equation}
3278: %
3279: is a unit preserving completely positive map
3280: $\mathbf{L}(H) \rightarrow \mathbf{L}(H)$.  Its Kraus representation is
3281: %
3282: \begin{equation}
3283: %
3284: P^{\mathrm{t}}(T)=\sum_{i \in I} X_i^\dag T  X_i
3285: %
3286: \end{equation}
3287: %
3288: \end{prop}
3289: Note that the completely positive unit preserving
3290: maps $\mathbf{L}(H) \rightarrow \mathbf{L}(H)$ which are adjoints of
3291: completely positive trace-preserving maps on $\mathbf{T}(H)
3292: \rightarrow \mathbf{T}(H)$ can be characterized precisely as those
3293: which are continuous mappings $\mathbf{L}(H) \rightarrow
3294: \mathbf{L}(H)$, where $\mathbf{L}(H)$ has the ultraweak topology.
3295: 
3296: \subsection{Completely Positive Semigroups on $\mathbf{T}(H)$}
3297: 
3298: We need to first establish that each generalized Lindblad-type
3299: operator $A(t)$ ({\em cf} eqs.\eqref{generic-equation-of-evolution} and \eqref
3300: {LindbladOperator}) generates
3301: a semigroup of completely positive contractions on $\mathbf{T}(H)$
3302: {\em with respect to the trace-class norm $\| \cdot \|_1$}.  We will
3303: also consider boundedness properties relative to the operator norm $\|
3304: \cdot \|_\infty$, although in general the corresponding semigroups may
3305: not be contraction semigroups in this norm.
3306: 
3307: We begin with a general result characterizing infinitesimal generators
3308: of strongly continuous positive (or completely positive) semigroups on
3309: the Banach space $\mathbf{T}(H)$.  Recall that a one-parameter
3310: semigroup $\{T_t\}_{t \geq 0}$ on a Banach space $E$ is said to be of
3311: class $C_0$ iff for every $x \in E$, $\lim_{t \rightarrow s} T_t(x) =
3312: T_s(x)$.  If $\{T_t\}_{t \geq 0}$ is a $C_0$-semigroup, then there are
3313: positive constants $M$ and $\beta$ such that
3314: %
3315: \begin{equation}\label{exponential-condition}
3316: %
3317: \|T_t\| \leq M e^{t \beta}.
3318: %
3319: \end{equation}
3320: %
3321: Moreover, 
3322: \begin{equation}
3323: %
3324: A x = \lim_{h \rightarrow 0} h^{-1} (T_h x - x)
3325: %
3326: \end{equation}
3327: %
3328: is a densely-defined operator called the {\em infinitesimal generator} of
3329: $\{T_t\}_{t \geq 0}$
3330: 
3331: \begin{thm}
3332: %
3333: Suppose $A$ is a densely-defined operator on $\mathbf{T}(H)$ which
3334: generates a contractive semigroup $\{T_t\}_{t \geq 0}$ on
3335: $\mathbf{T}(H)$.  A necessary and sufficient condition the operators
3336: $\{T_t\}_{t > 0}$ be positive (respectively completely positive) is
3337: that for each $\lambda > 0$
3338: %
3339: \begin{equation}
3340: %
3341: \opr{R}(\lambda,A)= (\lambda I - A)^{-1}
3342: \end{equation}
3343: %
3344: (which is defined by the Hille-Yosida Theorem) be positive
3345: (respectively completely positive).  The operators $T_t$ are
3346: trace-preserving iff in addition for all $\lambda > 0$,
3347: %
3348: \begin{equation} \label{trace-dilation-property}
3349: %
3350: \opr{tr}(\opr{R}(\lambda,A) \rho) = \lambda^{-1} \opr{tr}(\rho)
3351: %
3352: \end{equation}
3353: %
3354: for every $\rho \in \mathbf{T}(H)$. 
3355: %
3356: 
3357: The family of operators $\{T_t\}_{t > 0}$ extends to a $C_0$-semigroup
3358: on $\mathbf{L}(H)$
3359: %
3360: iff there are constants $M'$ and $\beta'$ such that for all $\lambda
3361: >\beta'$ and for all $\rho \in \mathbf{T}(H)$ and non-negative
3362: integers $m$:
3363: %
3364: \begin{equation}
3365: %
3366: \|\opr{R}(\lambda,A)^m \rho\|_\infty \leq M'(\lambda - \beta')^{-m}
3367: \|\rho\|_\infty.
3368: %
3369: \end{equation}
3370: %
3371: In this case, we have the explicit bound
3372: %
3373: \begin{equation}
3374: %
3375: \|T_t\|_\infty \leq M' e^{t \beta'} \quad \forall t>0.
3376: %
3377: \end{equation}
3378: %
3379: %
3380: \end{thm}
3381: %
3382: %
3383: \begin{rem}
3384: %
3385: It suffices that the property~\eqref{trace-dilation-property} hold for
3386: density states $\rho$.
3387: %
3388: \end{rem}
3389: 
3390: %
3391: \begin{proof}
3392: %
3393: To avoid duplication, we refer only to the assertions for complete
3394: positivity.  By the general Hille-Yosida theory, if $A$ is an
3395: infinitesimal generator of a contractive semigroup on $\mathbf{T}(H)$,
3396: the resolvents
3397: %
3398: \begin{equation}
3399: %
3400: \opr{R}(\lambda,A) = (\lambda - A)^{-1}
3401: %
3402: \end{equation}
3403: %
3404: are defined for all $\lambda > 0$ and by Theorem 3.1.3 of~\cite{tanabe},
3405: %
3406: \begin{equation}\label{Laplace-transform}
3407: %
3408: \opr{R}(\lambda,A) = \int_0^\infty e^{-\lambda t} T_t dt.
3409: %
3410: \end{equation}
3411: %
3412: {\em i.e}., the resolvent is the Laplace transform of $\{T_t\}_{t \geq 0}$.
3413: In particular, if $T_t$ consist of completely positive operators, then
3414: the resolvent operators are all completely positive.  
3415: 
3416: Conversely, let
3417: %
3418: \begin{equation}
3419: %
3420: A_\lambda = \lambda (\lambda  \opr{R}(\lambda,A) - I)
3421: %
3422: \end{equation}
3423: %
3424: Then for each $t \geq 0$,
3425: %
3426: \begin{equation}
3427: %
3428: \exp (t A_\lambda) = e^{-\lambda t} \exp \big(t \lambda^2 \opr{R}(\lambda,A)\big)
3429: %
3430: \end{equation}
3431: %
3432: which is clearly completely positive and it is known that for each $t
3433: \geq 0$,
3434: %
3435: \begin{equation} \label{semigroup-approximation-equation}
3436: T_t = \lim_{\lambda \rightarrow \infty} \exp (t A_\lambda)
3437: \end{equation}
3438: in the strong operator topology.  Thus $T_t$ is completely positive.
3439: %
3440: 
3441: To deal with the trace preservation properties of $T_t$, note that if
3442: $S$ is a bounded operator on $\mathbf{T}(H)$ for which
3443: \begin{equation}
3444: %
3445: \opr{tr}(S \rho) = \alpha \opr{tr}(\rho)
3446: %
3447: \end{equation}
3448: %
3449: then 
3450: \begin{equation}
3451: %
3452: \opr{tr}(e^S \rho) = \sum_{k=0}^\infty \opr{tr}\bigg(\frac{S^k}{k!} \rho\bigg)=
3453: \sum_{k=0}^\infty\frac{\alpha^k}{k!} \opr{tr}(\rho) = e^{\alpha} \opr{tr}(\rho).
3454: %
3455: \end{equation}
3456: %
3457: Thus, 
3458: \begin{equation}
3459: %
3460: \opr{tr}(\exp t A_\lambda \rho) = e^{-\lambda t} e^{t \lambda^2 \ 1
3461: /\lambda} \opr{tr}(\rho) = \opr{tr}(\rho).
3462: %
3463: \end{equation}
3464: %
3465: By~\eqref{semigroup-approximation-equation}, it follows that $T_t$ is also
3466: trace preserving.  
3467: %
3468: Conversely, if $T_t$ is trace preserving,
3469: %
3470: \begin{equation}
3471: %
3472: \opr{tr}(\opr{R}(\lambda,A) \rho) = \int_0^\infty e^{-\lambda t}
3473: \opr{tr}(T_t \rho) dt = \int_0^\infty e^{-\lambda  t}\opr{tr}(\rho) dt = \lambda^{-1}\opr{tr}(\rho).
3474: %
3475: \end{equation}
3476: %
3477: \end{proof}
3478: %
3479: \subsubsection{Examples of Completely Positive Semigroups}
3480: 
3481: Our analysis of the generalized Lindblad equation will reduce to an analysis of
3482: the solution in two important cases:
3483: 
3484: %\subsubsection{Unitary Evolution} 
3485: \paragraph{(Case 1) Unitary Evolution} 
3486: 
3487: In particular, if $\mathcal{H}$ is a self-adjoint operator on a
3488: Hilbert space $H$, then the family of completely positive mappings
3489: %
3490: \begin{equation}
3491: %
3492: P^\mathcal{H}_t(\rho) = e^{i t \mathcal{H}} \rho e^{- i t \mathcal{H}}
3493: %
3494: \end{equation}
3495: %
3496: is a one-parameter group of completely positive maps. Its generator on
3497: the trace-class operators is formally given by the operator
3498: %
3499: \begin{equation} \label{formal-infinitesimal-generator}
3500: %
3501: \rho \mapsto i [\mathcal{H},\rho].
3502: %
3503: \end{equation}
3504: %
3505: This expression is only formal, because it is not defined for all
3506: $\rho$. Nevertheless, the infinitesimal generator is densely defined
3507: on the space of trace-class operators and it is an extension
3508: of~\eqref{formal-infinitesimal-generator} for the finite rank
3509: operators on the domain of $\mathcal{H}$.\\
3510: 
3511: \paragraph{(Case 2) Dissipative Operators} 
3512: %{\it 4.2.2.2 Dissipative Operators} 
3513: 
3514: Another type of infinitesimal
3515: generator we will consider are operators of the form
3516: 
3517: \begin{equation}\label{pure-lindblad-form}
3518: %
3519: \mathcal{L} \rho = \sum_j \bigg[ L_j \rho L_j^\dag  - \frac{1}{2}\big\{L_j^\dag
3520: L_j, \rho\big\}\bigg] \end{equation}
3521: %
3522: where braces denote the anti-commutator.
3523: %
3524: \begin{lem}  Suppose
3525: %
3526: \begin{equation} \label{boundedness-cond-of-lindblad}
3527: %
3528: \sum_j L_j^\dag L_j  \in \mathbf{L}(H).
3529: %
3530: \end{equation}
3531: %
3532: Then the operator given by~\eqref{pure-lindblad-form} is bounded on
3533: $\mathbf{T}(H)$. If in addition
3534: %
3535: \begin{equation} \label{norm-boundedness-cond-of-lindblad}
3536: %
3537: \sum_j L_j L_j^\dag  \in \mathbf{L}(H).
3538: %
3539: \end{equation}
3540: %
3541: %
3542: then $\mathcal{L}$ is a bounded operator on $\mathbf{L}(H)$.
3543: %
3544: \end{lem}
3545: %
3546: \begin{proof} Let $C$ be the operator norm of $\sum_j L_j^\dag L_j$.
3547: %
3548: To show the map $\mathcal{L}$ is defined and continuous on $\mathbf{T}(H)$, it
3549: suffices to show
3550: %
3551: $
3552: %
3553: \mathcal{L}_0: \rho \mapsto  \sum_j L_j \rho L_j^\dag
3554: %
3555: $
3556: %
3557: is defined and continuous on $\mathbf{T}(H)$.  However, if $\rho \geq 0$,
3558: %
3559: \begin{align}
3560: %
3561: \opr{tr}(\mathcal{L}_0(\rho)) & = \sum_i \opr{tr}(L_j \rho L_j^\dag) \\
3562: %
3563: & = \sum_j \opr{tr}(\rho L_j^\dag L_j) \\
3564: %
3565: & = \sum_j \opr{tr}\biggl(\rho^{1/2} L_j^\dag L_j\rho^{1/2}\biggr) \\
3566: %
3567: & = \opr{tr}\biggr(\rho^{1/2}  (\sum_j L_j^\dag L_j) \rho^{1/2}\biggr) \leq C \opr{
3568: tr}(\rho) = C \|\rho\|_1
3569: \end{align}
3570: %
3571: thus for arbitrary self-adjoint $\rho$,
3572: %
3573: \begin{align}
3574: %
3575: \|\mathcal{L}_0(\rho)\|_1 & = \|\mathcal{L}_0(\rho^+ - \rho^-)\|_1 \\
3576: %
3577: & \leq  \|\mathcal{L}_0(\rho^+)\|_1 +  \|\mathcal{L}_0(\rho^-)\|_1 \\
3578: %
3579: & \leq  C\biggl(\|\rho^+\|_1 + \|\rho^-\|_1\biggr) = C \|\rho\|_1.
3580: \end{align}
3581: 
3582: If~\eqref{norm-boundedness-cond-of-lindblad}, suppose $0 \leq T \leq
3583: 1$:
3584: %
3585: \begin{equation}
3586: %
3587: 0  \leq \sum_j L_j^\dag T L_j \leq \sum_j  L_j^\dag L_j \leq C 1_H
3588: %
3589: \end{equation}
3590: %
3591: Thus, for arbitrary $T$,
3592: %
3593: \begin{equation}
3594: %
3595: \|\mathcal{L} T \|_\infty \leq  \|\sum_j L_j L_j^\dag\|_\infty  + C \|T\|_\infty
3596: %
3597: \end{equation}
3598: %
3599: \end{proof}
3600: 
3601: It was established by Lindblad~\cite{lindblad76} (and not too hard to show directly) that
3602: if~\eqref{boundedness-cond-of-lindblad} holds, the
3603: %
3604: $\mathcal{L}$ generates a {\em uniformly continuous} a completely
3605: positive semigroup (relative to the operator norm on $\mathbf{T}(H)$).
3606: %
3607: %
3608: In this case the semigroup is given by 
3609: %
3610: \begin{equation}
3611: %
3612: e^{t \mathcal{L}} = \sum_{k=0}^\infty \frac{t^k}{k!}{\mathcal{L}}^k~.
3613: %
3614: \end{equation}
3615: %
3616: \iffalse
3617: It is elemenatry that the generated semigroup is given by the norm
3618: convergent exponential:
3619: %
3620: However, we are mainly concerned with semigroups on $\mathbf{T}(H)$.
3621: Based on this however, we can state:
3622: %
3623: However, we are mainly concerned with semigroups on $\mathbf{T}(H)$.
3624: Based on this however, we can state:
3625: %
3626: \begin{prop}
3627: %
3628: Suppose the resolvent
3629: %
3630: \begin{equation}
3631: %
3632: \opr{R}(\lambda,\mathcal{L}^\ast) = (\lambda - \mathcal{L}^\ast)^{-1}
3633: %
3634: \end{equation}
3635: %
3636: which is defined for $\lambda > 0$ leaves the compact operators
3637: invariant.  Then $e^{t \mathcal{L}}$ is a contractive trace preserving
3638: semigroup on $\mathbf{T}(H)$. \end{prop}
3639: \fi
3640: 
3641: \subsection{Perturbation of Completely Positive Generators}
3642: % {\it 4.2.2.3 Perturbation of Completely Positive Generators}
3643: 
3644: In order to show that the generalized Lindblad operators given in
3645: equation~\eqref{LindbladOperator} are completely positive generators,
3646: we need to establish a perturbation result analogous to the
3647: Kato-Rellich theorem. 
3648: 
3649: A linear map $B$ on $\mathbf{T}(H)$ is {\em trace annihilating} iff
3650: %
3651: \begin{equation}
3652: %
3653: \opr{tr}(B \rho) = 0
3654: %
3655: \end{equation}
3656: %
3657: for all $\rho \in \mathbf{T}(H)$.  For example, an operator of the form~\eqref{pure-lindblad-form}
3658: %
3659: is easily seen to be trace annihilating.
3660: %
3661: 
3662: \iffalse
3663: 	%
3664: 	\begin{dfn}
3665: 	%
3666: 	An operator $B$ on $\mathbf{T}(H)$ is completely bounded iff there is a $k$
3667: 	such that the norms of the operators 
3668: 	%
3669: 	\begin{equation}
3670: 	%
3671: 	B_n = \begin{bmatrix} B & B & \cdots & B \\ B & B & \cdots & B \\
3672: 	\vdots & \vdots & \ddots & \vdots
3673: 	 \\ B & B & \cdots & B
3674: 	\end{bmatrix}
3675: 	%
3676: 	\end{equation}
3677: 	%
3678: 	acting on the trace class operators of the $n$-fold tensor product
3679: 	$H^{\otimes n}$ is bounded by $k$.  \end{dfn} 
3680: 	%
3681: \fi
3682: %
3683: Using the Trotter-Kato product formula~(\cite{trotter59}, \cite{chernoff68}), we can show that generators of
3684: completely positive contractive semigroups have a sum which is also a
3685: generator of a contractive semigroup, provided the sum generates a
3686: contractive semigroup.
3687: 
3688: \begin{prop}
3689: %
3690: Suppose $B$ is a bounded operator of the
3691: form~\eqref{pure-lindblad-form}.  If $A$ is a generator of a completely
3692: positive semigroup then so is $A+B$. .
3693: %
3694: \end{prop}
3695: %
3696: \iffalse
3697: \begin{proof}
3698: %
3699: By Theorem 3.4.1 of~\cite{tanabe}, $A+B$ is an infinitesimal generator and
3700: %
3701: \begin{equation}
3702: %
3703: \lambda - (A+B) = ( 1 - B (\lambda - A)^{-1}) (\lambda - A)
3704: %
3705: \end{equation}
3706: %
3707: and
3708: %
3709: \begin{equation}
3710: %
3711: \big(\lambda - (A+B)\big)^{-1} = (\lambda - A)^{-1}
3712: \sum_{k=0}^\infty \big(B (\lambda - A)^{-1}\big)^k.
3713: %
3714: \end{equation}
3715: %
3716: Since each summand in the above infinite series is completely
3717: positive, the result concerning generation of completely positive
3718: semigroups follows.
3719: %
3720: 
3721: To prove the final assertion concerning trace preservation, note that
3722: since $B$ is trace annihilating,
3723: %
3724: \begin{equation}
3725: %
3726: \opr{tr}(\sum_{k=0}^\infty \big(B (\lambda - A)^{-1}\big)^k \rho) = 
3727: %
3728: \opr{tr}(\rho)
3729: %
3730: \end{equation}
3731: %
3732: since all terms in the infinite sum vanish except the first.  
3733: \end{proof}
3734: \fi
3735: \begin{cor}
3736: %
3737: The generalized Lindblad operators given in equation~\eqref{LindbladOperator}
3738: generate a completelty positive semigroup of contractions on $\mathbf{T}(H)$.
3739: %
3740: \end{cor}
3741: 
3742: We now extend the results of Lindblad and Davies to allow for time
3743: varying Hamiltonians by relying on results of Kato.  The need for this
3744: arises since in some circuit-based models the various gates are
3745: implemented by varying the Hamiltonian (see for
3746: instance~\cite{nielsen-chuang} \S7.7.2). We make the assumption that
3747: the dissipative effects are bounded which simplifies the analysis
3748: considerably.
3749: 
3750: \subsection{Solving the Generalized Lindblad Equation}\label{solve_lindblad}
3751: 
3752: In some cases it is possible to solve
3753: the generalized Lindblad equation~\cite{LuYangZangChen2003}. However, by solution
3754: we mean an expression for the fundamental solution $P_{t,s}$ as a
3755: limit of product of exponentials. Though this expression will almost
3756: never provide a closed form solution, it will provide enough
3757: information to obtain an estimate of how well a unitary (or partial
3758: isometry) can be implemented by one of the operators $P_{t,s}$.  The
3759: two tools we use are the Trotter-Kato product formula and the explicit
3760: form of the solution of a time-dependent equation as a time ordered
3761: product of exponentials given in the proof of  \S4.2
3762: of~\cite{tanabe}.
3763: 
3764: A precise formulation of a set of conditions which guarantees the
3765: convergence of the products in the next two theorems is given in
3766: Theorem~\ref{precise-formulation-thm}.  These results comprised
3767: by Theorems \ref{ordered-product-exponentials-thm} and
3768: \ref{ordered-product-exponentials-thm2}
3769: are restatements
3770: of assertions contained in the proofs in \S4.2 of~\cite{tanabe}.
3771: 
3772: \begin{thm} \label{ordered-product-exponentials-thm} Under suitable
3773: conditions, the fundamental solution $P_{t,s}$
3774: for~\eqref{fundamental-solution-def} is given by
3775: %
3776: \begin{equation} \label{solution-product-formula} P_{t,s} =
3777: \opr{str-lim}_{\Delta \rightarrow 0} \prod_{k=0}^{n-1}
3778: \exp\bigl((r_{k+1} - r_k) A(r_k)\bigr),
3779: %
3780: \end{equation}\label{ordered-product-exponentials}
3781: %
3782: where $s = r_0 < r_1 < \cdots < r_{n-1} < t$ and $\max | r_{k+1} -
3783: r_k| \leq \Delta$.
3784: %
3785: Each $P_{t,s}$ is completely positive and trace preserving.
3786: %
3787: \end{thm}
3788: 
3789: \begin{thm} \label{ordered-product-exponentials-thm2} Under
3790: the same assumptions as the previous
3791: theorem~\ref{ordered-product-exponentials-thm},
3792: %
3793: \begin{equation}\label{Trotter-Lie-exponentials}
3794: %
3795: \exp s A(t) = \opr{str-lim}_{n \rightarrow \infty} \exp
3796: \biggl(\frac{s}{n} \mathcal{L}(t)\biggr) \ \exp\biggl(\frac{s}{n} \mathcal{H}(t)\biggr)
3797: %
3798: \end{equation}
3799: %
3800: \end{thm}
3801: 
3802: \iffalse 
3803: 
3804: 	As an immediate consequence of the two results above, and the
3805: 	lower-semicontinuity of the entropy, we can show that a bound on the time
3806: 	dependence of entropy can be computed by the pure Lindblad equation.
3807: 	If $P$ is an operator on density matrices, define
3808: 	%
3809: 	\begin{equation}
3810: 	%
3811: 	\delta(P) = \sup{\rho,U} \frac{\mathbf{S}\bigl(P(U \rho U^\ast )\bigr)}{\mathbf{S}(\rho)}.
3812: 	%
3813: 	\end{equation}
3814: 	%
3815: 	$\delta(P)$ crudely measures how much $P$ increases entropy over
3816: 	density states.  $P$ is a unitary invariant, that is $\delta(U P
3817: 	U^\ast) = \delta(P)$.  Clearly, \begin{equation}
3818: 	%
3819: 	\delta(P Q) \leq \delta(P) \delta(Q)
3820: 	%
3821: 	\end{equation}
3822: 	%
3823: 	$\delta$ is not continuous, but is lower semicontinuous, that is if
3824: 	$P_t \rightarrow P$ in the strong topology, and $\delta(P_t) \leq c$,
3825: 	then $\delta(P) \leq c$.  This follows from lower semi-continuity of
3826: 	the entropy.
3827: 	
3828: 	If $P_{t,s}$ is a propagator, define
3829: 	%
3830: 	\begin{equation}
3831: 	%
3832: 	\bar{\delta}(P)_{t,s} = \sup_\Delta \prod_{k=0}^{n-1} \delta(P_{r_{k+1},r_k})
3833: 	%
3834: 	\end{equation}
3835: 	%
3836: 	where $s = r_0 < r_1 < \cdots < r_{n-1} < t$ and $\max | r_{k+1} -
3837: 	r_k| \leq \Delta$.
3838: 	
3839: 
3840: 	Using the exponential representation and the Trotter-Kato product formula,
3841: 	one shows that an upper bound for $\delta P_{t,s}$ for the general Lindblad
3842: 	equation with Lindblad operator~\eqref{LindbladOperator} can be
3843: 	obtained by considering the Lindblad equation in which $\mathcal{H}$
3844: 	is identically $0$. This leads to the following result:
3845: 	%
3846: 	\begin{prop}
3847: 	%
3848: 	Suppose that $P_{t,s}$, $Q_{t,s}$ are fundamental solutions to the
3849: 	Lindblad equations with the same dissipative part.  Then
3850: 	\begin{equation}
3851: 	%
3852: 	\delta(P_{t,s}) \leq \prod \delta(Q_{t,s}) \quad \forall t \geq s \geq 0.
3853: 	%
3854: 	\end{equation}
3855: 	%
3856: 	\end{prop}
3857: %
3858: \fi
3859: 
3860: \subsection{Existence of Solutions}\label{existence-and-uniqueness}
3861: 
3862: We will restrict our attention to bounded time varying perturbations
3863: of a fixed self-adjoint operator acting on $\mathbf{T}(H)$ via a
3864: commutator as in~\eqref{basic-operator} below. The result we state is not
3865: the most general possible, and the early results of Kato~\cite{kato}
3866: suffice for its proof. We follow the treatment in Chapter XIV, \S4
3867: of~\cite{yosida} which is a more readily available reference.
3868: 
3869: \begin{thm}\label{precise-formulation-thm}
3870: %
3871: Suppose $\mathcal{H}$ is a self-adjoint operator, $\{B(t)\}_{t \in [0,
3872: \infty[}$, $\{L_j(t)\}_{t \in [0, \infty[}$, $1 \leq j \leq n$ are
3873: families of bounded operators, all of which are continuously norm differentiable as
3874: a functions of $t$, then there is a fundamental solution $P_{t,s}$
3875: for~\eqref{fundamental-solution-def} where
3876: %
3877: \begin{equation} \label{basic-operator}
3878: %
3879: A(t) \rho = -i[\mathcal{H}+B(t), \rho] + \sum_{j=1}^n \bigg( L_j(t)
3880: \rho L_j^\dag(t) - \frac{1}{2}\big\{L_j^\dag(t) L_j(t),
3881: \rho\big\}\bigg).
3882: %
3883: \end{equation}
3884: %
3885: The solution is a given by a limit of a time-ordered product of
3886: exponentials~\eqref{solution-product-formula}.
3887: \end{thm}
3888: \begin{proof}
3889: %
3890: There are various technical assumptions for a family $A(t)$ of
3891: operators that need to be checked in order to apply Kato's Theorem.
3892: The first of these is the independence of $\opr{dom} A(t)$ of the
3893: parameter $t$. Under our assumptions
3894: %
3895: \begin{equation}
3896: %
3897: A(t) = A + C(t)
3898: %
3899: \end{equation}
3900: %
3901: where $C(t): \mathbf{T}(H) \rightarrow \mathbf{T}(H)$ are bounded
3902: operators and $A$ is the infinitesinal generator of a contractive
3903: semigroup.  Indeed, 
3904: %
3905: \begin{equation}
3906: %
3907: A\rho  =  -i[\mathcal{H}, \rho] 
3908: %
3909: \end{equation}
3910: %
3911: is the infinitesimal generator of a group on $\mathbf{T}(H)$ and 
3912: %
3913: \begin{equation}
3914: %
3915: C(t) \rho = -i[B(t), \rho] + \sum_{j=1}^n \bigg( L_j(t)
3916: \rho L_j^\dag(t) - \frac{1}{2}\big\{L_j^\dag(t) L_j(t),
3917: \rho\big\}\bigg).
3918: %
3919: \end{equation}
3920: %
3921: is by assumption a bounded operator on $\mathbf{T}(H)$. In particular,
3922: all the operators $A(t)$ have the same domain $\opr{dom}(A)$.
3923: 
3924: 
3925: We now address the remaining assumptions in Kato's theorem.  For any
3926: $\lambda >0$, \begin{equation}
3927: %
3928: \lambda - A(t) = \lambda - A - C(t) = (I +  C(t) \opr{R}(\lambda, A) ) (\lambda - A) 
3929: %
3930: \end{equation}
3931: %
3932: For $\lambda$ sufficiently large $\opr{R}(\lambda, A) C(t)$ has norm
3933: $<1$, so the Neumann (geometric) series for inverses
3934: (see~\cite{Dieudonne}, Chapter VIII, \S3) $I + \opr{R}(\lambda, A)
3935: C(t)$ is invertible.  Thus we can write,
3936: %
3937: \begin{equation}
3938: %
3939: (\lambda - A(t))^{-1} = \opr{R}(\lambda, A) (I + \opr{R}(\lambda, A) C(t))^{-1} 
3940: %
3941: \end{equation}
3942: %
3943: Thus,
3944: % 
3945: \begin{equation}
3946: %
3947: B(t,s) = (\lambda - A(t))(\lambda - A(s))^{-1} = (I + C(t)
3948: \opr{R}(\lambda, A) ) (I + C(s) \opr{R}(\lambda, A) )^{-1}
3949: %
3950: \end{equation}
3951: %
3952: is well defined and by our assumptions $B(t,s)$ is a norm
3953: differentiable function jointly in the variables $t,s$. This implies
3954: the remaining conditions in the hypothesis of Kato's theorem.  It only
3955: remains to observe that presence of the parameter $\lambda$, instead
3956: of $1$ as actually stated in Kato's theorem is immaterial, since
3957: solutions of equations
3958: %
3959: \begin{equation}
3960: %
3961: \frac{d}{dt} \rho (t) = A(t) \rho(t)
3962: %
3963: \end{equation}
3964: %
3965: are trivially affected by adding a constant scalar to $A(t)$.
3966: \end{proof}
3967: 
3968: 
3969: \section{Solving the Ersatz Quantum Computer Condition}\label{alpha_0_solution}
3970: 
3971: %\begin{dfn}
3972: %
3973: 
3974: Consider the ersatz quantum computer condition (${\mathcal E}$QCC) given
3975: in \eqref{ersatz-quantum-computer-equation}:
3976: %
3977: \begin{equation} \label{eqn:defining-part}
3978: P\cdot\rho = U \rho U^\dag.
3979: %
3980: \end{equation}
3981: %
3982: %\end{dfn}
3983: Note that this can be obtained
3984: by setting the encoding and decoding maps to unity, and
3985: setting $\alpha = 0$ in the QCC given in \eqref{encoded_QCC}.
3986: A simple observation shows that
3987: ``solving''~\eqref{eqn:defining-part} is completely equivalent to
3988: obtaining the noise-free part of a communication channel.
3989: 
3990: \begin{thm}\label{krebs-reduction-trick} Suppose $H$ is
3991: finite-dimensional. If $P$ is given by the Kraus
3992: representation~\eqref{KrausFormCPMap},
3993: %
3994: %
3995: given a unitary $U$, the set of $\rho \in \mathbf{L}(H)$
3996: satisfying~\eqref{eqn:defining-part} is the $\ast$-subalgebra of
3997: $\mathbf{L}(H)$ given by
3998: %
3999: \begin{equation}\label{commutant-condition}
4000: %
4001: \mathfrak{A}_{P,U}= \{\rho \in \mathbf{L}(H): \forall i \in I, \quad
4002: [\rho, U^\dag X_i] = 0 \}
4003: %
4004: \end{equation}
4005: \end{thm}
4006: %
4007: %
4008: \begin{proof}  The solutions of~\eqref{eqn:defining-part} are the
4009: %
4010: fixed points of the completely positive map $Q$ defined by the equation
4011: %
4012: \begin{equation}
4013: %
4014: Q\cdot\rho = \sum_{i \in I} U^\dag X_i \rho X_i^\dag U.
4015: %
4016: \end{equation}
4017: %
4018: Now apply~\cite{kribs03}, Theorem 2.1.
4019: %
4020: \end{proof}
4021: 
4022: In the above theorem, the assumption $H$ is finite-dimensional is
4023: essential, see~\cite{arias-at-al-2002}. Since $H$ is finite dimensional
4024: $\mathfrak{A}$ is an algebraic direct sum of algebras isomorphic to
4025: full matrix algebras.
4026: %
4027: %
4028: \begin{prop}
4029: %
4030: Let $\{E_\kappa\}_{\kappa \in I}$ be the set of finite-dimensional
4031: minimal central projections of $\mathfrak{A}$.  Then 
4032: %
4033: \begin{equation}
4034: %
4035: \mathfrak{A}_\kappa = E_\kappa \mathfrak{A} E_\kappa
4036: %
4037: \end{equation}
4038: %
4039: is an algebra of operators on the range $H_\kappa$ of $E_\kappa$
4040: (which is a finite dimensional space). It is isomorphic to a full
4041: matrix-algebra of finite multiplicity.
4042: 
4043: \end{prop}
4044: 
4045: 
4046: \bibliography{refs}
4047: \bibliographystyle{hplain}
4048: 
4049: 
4050: 
4051: \end{document}
4052: