quant-ph0511221/sf3.tex
1: \documentclass[prl,twocolumn,showpacs,amsmath,amssymb]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{dcolumn}
4: \usepackage{bm}
5: 
6: \begin{document}
7: 
8: \title{Optimal error tracking via quantum coding and continuous syndrome measurement}
9: \author{Ramon van Handel} \email{ramon@caltech.edu}
10: \author{Hideo Mabuchi} \email{hmabuchi@caltech.edu}
11: \affiliation{Physical Measurement and Control 266-33, California
12: Institute of Technology, Pasadena, CA 91125}
13: 
14: \date{\today}
15: \pacs{03.67.Pp,03.65.Yz,42.50.Lc,02.30.Yy}
16: 
17: \begin{abstract}
18: We revisit a scenario of continuous quantum error detection proposed
19: by Ahn, Doherty and Landahl [Phys.\ Rev.\ A {\bf 65}, 042301 (2002)]
20: and construct optimal filters for tracking accumulative errors.
21: These filters turn out to be of a canonical form from hybrid control
22: theory; we numerically assess their performance for the bit-flip and
23: five-qubit codes. We show that a tight upper bound on the stochastic
24: decay of encoded fidelity can be computed from the measurement
25: records. Our results provide an informative case study in
26: decoherence suppression with finite-strength measurement.
27: \end{abstract}
28: 
29: \maketitle
30: 
31: \noindent Prospects for quantum computation have motivated the
32: development of an extensive theory of quantum error
33: prevention/correction (QEC) \cite{Ali05}. Despite experimental
34: demonstrations of some key methods \cite{Lan05,Chi04,Kni01}, there
35: remains a significant gap between the abstract algebraic theory of
36: QEC and concrete physical models of real quantum systems. In
37: particular, QEC is still mainly discussed in terms of instantaneous
38: measurement and recovery operations rather than more realistic
39: continuous-time dynamical models. Bridging this gap would enable us,
40: {\it e.g.}, to deduce limits on QEC performance from the finite
41: speed of laboratory measurement and recovery operations. Appropriate
42: continuous-time formulations of QEC would likewise facilitate deeper
43: connections with control theory, which could in turn lead to the
44: discovery of new quantum memory schemes.
45: 
46: Our aim in this article will be to pursue a rigorous approach to
47: incorporating finite-strength measurement within the familiar
48: conceptual setting of discrete quantum error correcting codes. At
49: the same time, we will arrive at familiar equations from classical
50: control theory that should be amenable to characterization via
51: established methods of stochastic analysis. While the specific
52: models we consider---continuous-time relaxations of stabilizer codes
53: as introduced by Ahn, Doherty and Landahl \cite{Ahn}---are rather
54: impractical from an experimental point of view, they do provide a
55: canonical setting in which to demonstrate how essential ideas from
56: quantum error correction can be reconciled with stochastic control
57: theory. We regard this as an important step towards developing a
58: general theory of optimal decoherence suppression with resource
59: constraints on measurement and control.
60: 
61: We begin by considering a continuous-time relaxation of the bit-flip
62: code \cite{Ahn}. A logical qubit state $|\Psi\rangle =
63: c_0|0\rangle+c_1|1\rangle$ is encoded in the joint state of three
64: physical qubits as $|\Psi\rangle\mapsto
65: c_0|000\rangle+c_1|111\rangle \equiv |\Psi_E\rangle$. The system is
66: coupled to decay channels that can cause independent Markovian bit
67: flips and also to probe channels that allow for continuous
68: (finite-strength) syndrome measurement \cite{Saro04}. Following
69: \cite{Ahn,Saro04}, we assume that the syndrome measurements are
70: constructed by coupling two probe fields to the syndrome generators
71: \cite{Gottesman} $M_1=ZZI=\sigma_z\otimes\sigma_z\otimes\openone$
72: and $M_2=ZIZ=\sigma_z\otimes\openone\otimes\sigma_z$. The total
73: system dynamics is then described by the quantum stochastic
74: differential equation (e.g.\ \cite{GaZo})
75: \begin{multline}
76: \label{eq:QSDE}
77:     dU_t=\{\sqrt{\gamma}\,(X_1\,dB^{1\dag}_t+X_2\,dB^{2\dag}_t+
78:         X_3\,dB^{3\dag}_t-\mbox{h.c.})+\\
79:     \sqrt{\kappa}\,(M_1\,dA^{1\dag}_t+M_2\,dA^{2\dag}_t-\mbox{h.c.})
80:     -(\tfrac{3}{2}\gamma+\kappa)\,dt\}\,U_t
81: \end{multline}
82: where $B^i_t$ are the bit-flip channels, $A^i_t$ are the probe
83: channels and $X_1=XII=\sigma_x\otimes\openone\otimes\openone$,
84: $X_2=IXI$ {\it et cetera}. The unitary evolution $U_t$ is a
85: Schr{\"o}dinger picture propagator for the entire system including
86: the probe fields.
87: %\footnote{
88: %   This is not the only quantum stochastic differential equation
89: %   that describes such a setup; we could add scattering interactions
90: %   with the same effect.  However, this does not make any difference
91: %   in the following analysis.
92: %}.
93: 
94: For homodyne-type detection of the two probe channels we obtain two
95: observation processes $Y^i_t=U_t^\dag(A^i_t+A^{i\dag}_t)U_t$ that
96: can be written (via quantum It\^o rules) as
97: \begin{equation}
98:     dY^i_t=2\sqrt{\kappa}\,U_t^\dag M_i U_t\,dt + dA^i_t+dA^{i\dag}_t
99: \end{equation}
100: (this is called the input-output picture in \cite{GaZo}). Similarly,
101: we could in principle attempt to observe the bit-flip events
102: directly by performing direct detection of the corresponding decay
103: channels $B_t^i$; this would give rise to additional (counting)
104: observations $Z^i_t$ that are independent Poisson processes with
105: rate $\gamma$ (see, {\it e.g.}, \cite{BvH05}). Although the $Z_t^i$
106: are generally assumed to be unobservable in QEC, we will exploit
107: them below to obtain useful information on the statistics of the
108: measurement currents $Y_t^i$.
109: 
110: Our goal is to detect and to correct bit-flip errors by making use
111: of the probe currents $Y_t^i$, $i=1,2$. Hence we must understand how
112: best to extract information about error events from the noisy
113: observed signals. We begin by calculating the least mean-square
114: estimator for our system \cite{BvH05}, in the form of a (random)
115: three-qubit density matrix $\rho_t$ (the conditional state). For
116: every system observable $A$, ${\rm Tr}[A\rho_t]$ is the function of
117: the observation history that minimizes the estimation error $\langle
118: (U_t^\dag AU_t-{\rm Tr}[A\rho_t])^2\rangle$. If we observe only
119: $Y^i_t$, then $\rho_t$ obeys the quantum filtering equation
120: \cite{BvH05}
121: \begin{multline}
122: \label{eq:filter}
123:     d\rho_t=
124:     \sum_{k=1}^3\gamma\,\mathcal{T}[X_k]\rho_t\,dt
125:     +\sum_{i=1}^2\kappa\,\mathcal{T}[M_i]\rho_t\,dt \\
126:     +\sum_{i=1}^2\sqrt{\kappa}\,
127:     \mathcal{H}[M_i]\rho_t\,(dY_t^i-2\sqrt{\kappa}\,
128:         {\rm Tr}[M_i\rho_t]\,dt)
129: \end{multline}
130: where $\mathcal{T}[X]\rho=X\rho X^\dag-\rho$ and
131: $\mathcal{H}[X]\rho= X\rho+\rho X^\dag-{\rm
132: Tr}[(X+X^\dag)\rho]\rho$. This is equivalent to the stochastic
133: master equation used in \cite{Ahn}. If we were to additionally
134: observe the bit flip channels $Z_t^i$, we could obtain an improved
135: estimator $\tilde\rho_t$ that would obey a ``jump-unraveled''
136: version of Eq.~(\ref{eq:filter}) with $\gamma\mathcal{T}[X_k]\,dt$
137: replaced by $\mathcal{T}[X_k]\,dZ_t^i$.
138: %\begin{multline}
139: %\label{eq:filterjump}
140: %    d\tilde\rho_t=
141: %    \sum_{k=1}^3\mathcal{T}[X_k]\tilde\rho_t\,dZ_t^i
142: %    +\sum_{i=1}^2\kappa\,\mathcal{T}[M_i]\tilde\rho_t\,dt \\
143: %    +\sum_{i=1}^2\sqrt{\kappa}\,
144: %    \mathcal{H}[M_i]\tilde\rho_t\,(dY_t^i-2\sqrt{\kappa}\,
145: %        {\rm Tr}[M_i\tilde\rho_t]\,dt).
146: %\end{multline}
147: An important feature of these equations \cite{BvH05} is that the
148: so-called innovations processes $dW_t^i=dY_t^i-2\sqrt{\kappa}\,{\rm
149: Tr}[M_i\rho_t]\,dt$ are independent and have the law of a Wiener
150: process.
151: 
152: We can use the fact that the innovations are Wiener processes, in
153: the absence of `real' (physically generated) measurement signals
154: $Y_t^i$, to generate the latter through Monte Carlo simulations. By
155: driving Eq.~(\ref{eq:filter}) with random sample paths of a Wiener
156: process, we can reconstruct observation processes $Y_t^i$ that
157: sample the space of measurement records with the correct probability
158: measure. The evolution of the filter variables in each simulation
159: fairly represents what would have happened if physically generated
160: measurement records $Y_t^i$ had actually been presented to the
161: filter and it derived $W_t^i$ from them. Similarly, we can
162: reconstruct $Y_t^i$ from the jump-unraveled version of
163: Eq.~(\ref{eq:filter}) by driving it with Wiener processes $\tilde
164: W_t^i$ and independent Poisson processes $Z_t^i$ with rate $\gamma$.
165: The innovations theorem guarantees that both simulations will
166: generate sample paths $Y_t^i$ with the same probability measure. We
167: will invoke this property later.
168: 
169: In the usual setting of discrete quantum codes, an instantaneous
170: measurement of $M_1$ and $M_2$ after a period of free evolution is
171: used to determine the recovery operation (if any) that should be
172: applied. If $(M_1,M_2)=(+1,+1)$ no correction is necessary; outcomes
173: $(-1,+1)$ mean that $IXI$ should be applied; $(+1,-1)\Rightarrow
174: IIX$ and $(-1,-1)\Rightarrow XII$. These outcomes are called the
175: error syndromes. In the continuous setting we introduce the
176: orthogonal projectors $\Pi_0\ldots\Pi_3$ onto the eigenspaces
177: corresponding to each syndrome. The quantity $p^m_t={\rm
178: Tr}[\Pi_m\rho_t]$ is then the conditional probability (given the
179: noisy probe observations) that, had we actually measured $M_1$ and
180: $M_2$, we would have obtained the syndrome corresponding to $\Pi_m$.
181: Plugging $p^m_t$ into Eq.~(\ref{eq:filter}) we obtain the syndrome
182: filter
183: \begin{equation}
184: \label{eq:wonham}
185:     dp_t = \Lambda^T p_t\,dt+
186:     \sum_{i=1}^2 (H_i-h_i^Tp_t\,\openone)p_t\,(dY_t^i-h_i^Tp_t\,dt)
187: \end{equation}
188: where $h_i^m/2\sqrt{\kappa}$ is the outcome of $M_i$ corresponding
189: to the syndrome $\Pi_m$, $H_i=\mbox{diag}\,h_i$, and
190: $\Lambda_{mn}=\gamma(1-4\delta_{mn})$. The $p_t^i$ form a closed set
191: of equations, as the observations are uninformative on the logical
192: state of the qubit and the coherences (if any) between the
193: syndromes.
194: 
195: \begin{figure}
196: \includegraphics[width=0.45\textwidth]{Syndromes.eps}
197: 
198: \noindent (c)
199: 
200: \includegraphics[width=0.45\textwidth]{sf3closeup.eps}
201: \caption{\label{fig:syn}
202:     Markov chains associated to the three-qubit code:
203:     (a) the syndrome chain, and (b) the extended correction chain
204:     with the corresponding syndromes labeled between brackets.
205:     All transitions are independent with rate $\gamma$.
206:     (c) A Monte Carlo trajectory of $\max_i(p_t^i)$ (green noisy
207:     curve) and $\mathfrak{J}_t$ (black step-like curve) for the
208:     five-qubit code with $\kappa/\Gamma=100$.}
209: \vspace{-0.1in}\end{figure}
210: 
211: The syndrome filter, Eq.~(\ref{eq:wonham}), is a familiar equation
212: from classical probability theory; we will gain important insight by
213: reducing our problem to a classical one. Consider a system that can
214: be in one of four states labeled $0\ldots 3$. Suppose the system is
215: in some known state at time $t$. After an infinitesimal time
216: increment $dt$ the system can switch to one of the other three
217: states, each of which occurs with probability $\gamma\,dt$. This
218: defines a Markov jump process on a graph, as depicted in
219: Fig.~\ref{fig:syn}a. Unfortunately we cannot observe the state
220: directly; instead, we are given two observation processes of the
221: form $dY_t^i=h_i^{m_t}\,dt+dV_t^i$ where $V_t^{1,2}$ are two
222: independent Wiener processes that corrupt our observations, and
223: $m_t$ is the state of the Markov jump process at time $t$. As our
224: observations are noisy we cannot know the system state with
225: certainty at any time. However, we can calculate the {\it
226: conditional} probability $p_t^m$ that it is in state $m$ at time
227: $t$. This classical estimation problem is precisely solved by
228: Eq.~(\ref{eq:wonham}), known as the Wonham filter \cite{Wonham,EAM}.
229: 
230: To help interpret this result, consider the jump-unraveled version
231: of Eq.~(\ref{eq:filter}). In the same way that we obtained
232: Eq.~(\ref{eq:wonham}), we can substitute $\tilde p_t^m={\rm
233: Tr}[\Pi_m\tilde\rho_t]$ and get a closed form expression. Assuming
234: the initial state lies inside one of the syndrome spaces \footnote{
235:     This is not an essential restriction, as the probability measure
236:     on the space of measurement records can be written for any initial
237:     state as the corresponding mixture of such measures given a fixed
238:     initial syndrome.
239: }, it is readily verified that $[M_1,\tilde\rho_t]=
240: [M_2,\tilde\rho_t]=0$ for all $t$ and we obtain
241: \begin{equation}
242:     d\tilde p_t=\sum_{k=1}^3(\tilde X_k-\openone)\tilde p_t\,dZ_t^k
243: \end{equation}
244: where $\tilde p_0$ is one of the unit vectors $e_n^m=\delta_{nm}$.
245: Here $\tilde X_1$ is a matrix such that $\tilde X_1e_0=e_1$, $\tilde
246: X_1e_1=e_0$, $\tilde X_1e_2=e_3$, and $\tilde X_1e_3=e_2$, and
247: $X_{2,3}$ are defined similarly as shown in Fig.~\ref{fig:syn}a. The
248: solution of this equation is of the form $\tilde p_t=e_{m_t}$ where
249: $m_t$ is the Markov jump process defined above. Since
250: \begin{equation}
251:     d\tilde W_t^i=dY_t^i-h_i^T\tilde p_t\,dt=dY_t^i-h_i^{m_t}\,dt
252: \end{equation}
253: must be a Wiener process that is independent from all $Z_t^i$, the
254: statistics of the probe observations obtained from the quantum
255: system are precisely described by the classical model of the
256: previous paragraph. The Markov process $m_i$ is simply the error
257: syndrome obtained by observing the bit flips directly, and the
258: Wonham filter above has a natural classical interpretation as the
259: best estimate of $m_i$ given only the noisy probe observations.
260: 
261: We now turn to the problem of error correction. Suppose that we let
262: the system evolve for some time while propagating the filter
263: Eq.~(\ref{eq:wonham}) with the observations. At some time $T$ we
264: pose the question: what operation, if any, should we perform on the
265: system to maximize the probability of restoring the initial logical
266: state $|\Psi_E\rangle$? We will assume that we can pulse the system
267: sufficiently strongly (as compared to $\kappa,\gamma$) to perform
268: essentially instantaneous bit flips on any of the physical qubits.
269: The most obvious decision strategy simply mimics discrete error
270: correction---given the most likely syndrome state
271: $m_*=\mbox{arg\,max}_m\,p_T^m$, we do nothing if $m_*=0$ and
272: otherwise we perform a bit flip on physical qubit $m_*$.
273: 
274: But it is possible to do better. Recall that the discrete error
275: correction strategy is based on an assumption that at most one bit
276: flip occurs in the interval $[0,T]$. This assumption may not hold in
277: practice. With our continuous syndrome measurement, we do actually
278: have some basis for estimating the total number (and kind) of bit
279: flips that have occurred. Unfortunately this information does not
280: reside in the statistic $p_t$, which only gives the conditional
281: probabilities of the syndromes at the current time. We are seeking a
282: non-Markovian decision policy that knows something about the history
283: of the bit flips.
284: 
285: The classical machinery introduced above allows us to solve this
286: problem {\it optimally}. To do this we simply extend the Markov jump
287: process $m_t$ as shown in Fig.~\ref{fig:syn}b. The states of the
288: extended chain $\hat m_t$ are no longer the four syndromes but the
289: eight {\em error states} that may obtain at any given time. Every
290: syndrome corresponds to two error states, as is shown in
291: Fig.~\ref{fig:syn}b. We still consider the same observation
292: processes, so error states that belong to the same syndrome give
293: rise to identical observations. Thus on the basis of the
294: observations, the two Markov chains are indistinguishable.
295: Nonetheless the extended chain gives rise to a different estimator
296: that provides precisely the information we want. As by construction
297: the Wonham filter provides the optimal estimate, we conclude that
298: the optimal solution to our problem is given by the Wonham filter
299: for the extended chain, {\it i.e.}, the eight-dimensional equation
300: \begin{equation}
301: \label{eq:wonham2}
302:     d\hat p_t = \hat\Lambda^T\hat p_t\,dt+
303:     \sum_{i=1}^2 (\hat H_i-\hat h_i^T\hat p_t\,\openone)
304:         \hat p_t\,(dY_t^i-\hat h_i^T\hat p_t\,dt)
305: \end{equation}
306: where $\hat\Lambda$, $\hat h_i$ are the intensity matrix and
307: observation vector for the chain $\hat m_t$ (see \cite{Wonham,EAM}
308: for details). The optimal correction policy is now simple: at time
309: $T$, we perform the correction that corresponds to the state $\hat
310: m_*=\mbox{arg\,max}_m\,\hat p^m_T$. Hence if $\hat m_*=III$ we do
311: nothing, if $\hat m_*=IXX$ we flip physical qubits 2 and 3, etc.
312: This maximizes the probability of restoring $|\Psi_E\rangle$.
313: 
314: From Fig.~\ref{fig:syn}b we can see how information is lost from the
315: quantum memory. At time $t=0$ we begin in the no-error state $III$.
316: A bit flip might occur which puts us, {\it e.g.}, in the state
317: $XII$, then $XIX$, {\it etc}. But as we are observing these changes
318: in white noise there is always a chance that when two bit flips
319: happen in rapid succession, we ascribe the corresponding
320: observations to a fluctuation in the white noise background rather
321: than to the occurrence of two bit flips. In essence, successive bit
322: flips are resolvable only if they are separated by a sufficiently
323: long interval that the filter can average away the white noise
324: fluctuations (the signal-to-noise $4\kappa$ determines this
325: timescale). Occasionally, multiple bit flips occur too rapidly and
326: information is lost ({\it e.g.}, $III\rightarrow XII\rightarrow XIX$
327: may be mistaken for $III\rightarrow IXI$ since the two final states
328: have identical syndromes). It is evident from Fig.~\ref{fig:syn}b
329: that this rate of information loss is independent of the error
330: state. Hence there is no point in applying corrective bit flips at
331: intermediate times $t<T$, as this cannot slow the loss of
332: information. As the Wonham filter is optimal by construction, we
333: conclude that the correction policy described above is optimal
334: \footnote{
335:     This assumes that we trust Eq.~(\ref{eq:QSDE}) completely. If
336:     there is some uncertainty in the model it is sometimes
337:     advantageous to consider different estimators \cite{James}.
338: }.
339: 
340: How can we quantify the information loss from the system? By
341: construction $\hat p_t^*=\max_m\hat p_t^m$ is the probability of
342: correct recovery at time $t$. Unfortunately, as one can see in
343: Fig.~\ref{fig:syn}c, the quantity $\hat p_t^*$ fluctuates rather
344: wildly in time. Thus it is not a good measure of the information
345: content of the system, as it is very sensitive to the whims of the
346: filter: the filter may respond to fluctuations in the measurement
347: record by adjusting the state as if a bit flip had occurred, but
348: then correct itself when it becomes evident that nothing happened.
349: We actually want to find some quantity that gives a (sharp) upper
350: bound on all {\em future} values of $\hat p_t^*$. This would truly
351: measure the information content of the system, as it bounds the
352: probability of correct recovery that can be achieved.
353: 
354: We claim that the quantity $\mathfrak{J}_t=\max_m(\hat p_t^m/(\hat
355: p_t^m+\hat p_t^{\bar m}))$, which is a function of filter variables,
356: provides a suitable measure of the information content at time $t$.
357: Here $\bar m\ne m$ is the error state that corresponds to the same
358: syndrome as $m$, so $\hat p_t^m+\hat p_t^{\bar m}=P_t^m$ is the
359: probability of the syndrome corresponding to $m$. Hence we can
360: interpret $\mathfrak{J}_t$ as the conditional probability of the
361: error state $m$, given that the system is in the corresponding
362: syndrome. Define $I_t^m=\hat p_t^m/(\hat p_t^m+\hat p_t^{\bar m})$
363: so that $\mathfrak{J}_t=\max_mI_t^m$. Direct application of the
364: It\^o rules gives
365: \begin{equation}
366:     \frac{dI_t^m}{dt}=-\sum_{n\ne m}\hat\Lambda_{nm}\,
367:     \frac{P_t^n}{P_t^m}\,(I_t^m-I_t^n).
368: \end{equation}
369: If we define $m_t^*=\mbox{arg\,max}_m\,I_t^m$, then we get
370: \footnote{
371:     To make the argument completely rigorous one must check that
372:     this equation is well defined, i.e.\ that $d\mathfrak{J}_t/dt$
373:     exists.  This can be done using the methods in \cite{Chigansky}.
374: }
375: \begin{equation}
376:     \frac{d\mathfrak{J}_t}{dt}=-\sum_{n\ne m_t^*}\hat\Lambda_{nm_t^*}\,
377:     \frac{P_t^n}{P_t^{m_t^*}}\,(\mathfrak{J}_t-I_t^n)
378:     \le 0.
379: \end{equation}
380: Hence $\mathfrak{J}_t$ decreases monotonically, and moreover by
381: construction we must have $\hat p_t^m\le I_t^m$, so $\hat p_t^*\le
382: \mathfrak{J}_t\le \mathfrak{J}_s$ for all $s<t$. Thus evidently
383: $\mathfrak{J}_t$ bounds all future values of $\hat p_t^*$. One would
384: expect the bound to be tight for sufficiently high signal-to-noise
385: (as then $P_t^{m_t^*}$ will be close to one), which is indeed the
386: case as can be seen in Fig.~\ref{fig:syn}c.
387: 
388: \begin{figure}
389: \includegraphics[width=0.47\textwidth]{theplotzoom.eps}
390: \caption{\label{fig:perf}
391:     Average decay of $\mathfrak{J}_t$ for the five-qubit code (dependence
392:     on $\kappa/\Gamma$). These curves indicate upper bounds on
393:     the performance of a quantum memory based on coding and
394:     continuous syndrome measurement.}
395: 
396: \vspace{-0.1in}
397: \end{figure}
398: 
399: The procedure we have outlined can be generalized to other
400: stabilizer codes, such as the five-qubit code \cite{Gottesman}. This
401: code protects one logical qubit against single-qubit errors by
402: encoding in five physical qubits and measuring four stabilizer
403: generators. For `Pauli channel' decoherence described by Lindblad
404: terms $\sum_{k=1}^5\gamma\,(\mathcal{T}[X_k]+\mathcal{T}[Y_k]+
405: \mathcal{T}[Z_k])$, the error state graph can be constructed by
406: considering all possible assignments of an error state
407: $\in\{I,X,Y,Z\}$ to each qubit and by connecting every pair of
408: states that are related by the action of a Pauli operator
409: $\in\{X,Y,Z\}$ on one qubit. One thus has a graph with $4^5=1024$
410: nodes, with each node connected to $3\times 5=15$ other nodes (we
411: have validated this construction by comparing simulations of the
412: corresponding Wonham filter with an appropriate stochastic master
413: equation). The total error rate is $\Gamma=15\gamma$.
414: Fig.~\ref{fig:syn}c shows a portion of a single Monte Carlo
415: simulation of the Wonham filter for the five-qubit code;
416: Fig.~\ref{fig:perf} shows averages of $\mathfrak{J}_t$ over tens of
417: trajectories each for $\kappa/\Gamma\in\{10,30,100\}$.
418: 
419: In conclusion, we have shown that both the three- and five-qubit
420: codes are amenable to a classical analysis in terms of Markov jump
421: processes, which enables an optimal solution of the error tracking
422: problem in continuous time. Though the filters that must be
423: propagated for this purpose are high-dimensional, the optimal
424: procedure gives at least an upper bound on the achievable
425: performance of quantum memories based on coding and finite-strength
426: syndrome measurement.
427: %
428: % We have seen that in the optimal correction policy there is no point in
429: % performing corrections in real time; the problem completely reduces to
430: % an estimation problem, where the correction is performed at the terminal
431: % time.
432: In practical situations one might not have sufficient resources to
433: propagate the full optimal filter, so suboptimal methods are clearly
434: of interest.
435: % For example, it is much cheaper to propagate the
436: % syndrome filter than the extended filter; one could imagine implementing
437: % a suboptimal correction policy that is similar to the discrete case, in
438: % which case it becomes advantageous to perform multiple corrections (as
439: % then we do not need to remember the history of the bit flips.)
440: 
441: One could also consider different control goals. For example,
442: suppose that rather than requiring the state to be corrected at time
443: $T$, we allow ourselves a little more time afterwards. This can be
444: helpful; if the syndrome has just jumped, but we have not had enough
445: time to observe this yet, then we can avoid an incorrect recovery by
446: waiting just long enough to observe the jump. This can backfire,
447: however, as waiting too long will just cause us to lose information.
448: The optimal solution to this problem is known as an optimal stopping
449: problem \cite{Shiryaev} and is studied extensively in the
450: mathematical finance literature \cite{OksSulem}. All of these
451: problems, and many others, are subsumed under the title of {\it
452: hybrid control theory}. It is our hope that this theory will provide
453: important tools for the analysis and design of continuous quantum
454: error correction codes and suboptimal estimators, and for the
455: solution of the associated control problems.
456: 
457: \begin{acknowledgments}
458: This work was supported by the Army Research Office under Grant
459: DAAD19-03-1-0073. HM thanks D.\ Poulin and M.\ Nielsen for
460: insightful discussions.
461: \end{acknowledgments}
462: 
463: \bibliographystyle{apsrev}
464: \bibliography{QEC}
465: 
466: \end{document}
467: